Misplaced Pages

Uncertainty principle

Article snapshot taken from Wikipedia with creative commons attribution-sharealike license. Give it a read and then ask your questions in the chat. We can research this topic together.

In the physical sciences , the wavenumber (or wave number ), also known as repetency , is the spatial frequency of a wave , measured in cycles per unit distance ( ordinary wavenumber ) or radians per unit distance ( angular wavenumber ). It is analogous to temporal frequency , which is defined as the number of wave cycles per unit time ( ordinary frequency ) or radians per unit time ( angular frequency ).

#640359

129-423: 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. More formally,

258-480: A {\displaystyle a} larger we make the spread in momentum smaller, but the spread in position gets larger. This illustrates the uncertainty principle. As we let the Gaussian wave packet evolve in time, we see that its center moves through space at a constant velocity (like a classical particle with no forces acting on it). However, the wave packet will also spread out as time progresses, which means that

387-455: A + a † ) {\displaystyle {\hat {x}}={\sqrt {\frac {\hbar }{2m\omega }}}(a+a^{\dagger })} p ^ = i m ω ℏ 2 ( a † − a ) . {\displaystyle {\hat {p}}=i{\sqrt {\frac {m\omega \hbar }{2}}}(a^{\dagger }-a).} Using the standard rules for creation and annihilation operators on

516-461: 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 ) d x , {\displaystyle \langle u\mid v\rangle =\int _{-\infty }^{\infty }u^{*}(x)\cdot v(x)\,dx,} where

645-1711: 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. Consider a particle in a one-dimensional box of length L {\displaystyle L} . The eigenfunctions in position and momentum space are ψ n ( x , t ) = { A sin ⁡ ( k n x ) e − i ω n t , 0 < x < L , 0 , otherwise, {\displaystyle \psi _{n}(x,t)={\begin{cases}A\sin(k_{n}x)\mathrm {e} ^{-\mathrm {i} \omega _{n}t},&0<x<L,\\0,&{\text{otherwise,}}\end{cases}}} and φ n ( p , t ) = π L ℏ n ( 1 − ( − 1 ) n e − i k L ) e − i ω n t π 2 n 2 − k 2 L 2 , {\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}}},} where ω n = π 2 ℏ n 2 8 L 2 m {\textstyle \omega _{n}={\frac {\pi ^{2}\hbar n^{2}}{8L^{2}m}}} and we have used

774-438: A cost, namely the momentum has become less precise, having become a mixture of waves of many different momenta. One way to quantify the precision of the position and momentum is the standard deviation   σ . Since | ψ ( x ) | 2 {\displaystyle |\psi (x)|^{2}} is a probability density function for position, we calculate its standard deviation. The precision of

903-466: A definite prediction of what the quantum state ψ ( t ) {\displaystyle \psi (t)} will be at any later time. Some wave functions produce probability distributions that are independent of time, such as eigenstates of the Hamiltonian . Many systems that are treated dynamically in classical mechanics are described by such "static" wave functions. For example,

1032-510: A family of unitary operators parameterized by a variable t {\displaystyle t} . Under the evolution generated by A {\displaystyle A} , any observable B {\displaystyle B} that commutes with A {\displaystyle A} will be conserved. Moreover, if B {\displaystyle B} is conserved by evolution under A {\displaystyle A} , then A {\displaystyle A}

1161-569: A formulation for arbitrary Hermitian operator operators O ^ {\displaystyle {\hat {\mathcal {O}}}} expressed in terms of their standard deviation σ O = ⟨ O ^ 2 ⟩ − ⟨ O ^ ⟩ 2 , {\displaystyle \sigma _{\mathcal {O}}={\sqrt {\langle {\hat {\mathcal {O}}}^{2}\rangle -\langle {\hat {\mathcal {O}}}\rangle ^{2}}},} where

1290-471: A loss of information, though: knowing the reduced density matrices of the individual systems is not enough to reconstruct the state of the composite system. Just as density matrices specify the state of a subsystem of a larger system, analogously, positive operator-valued measures (POVMs) describe the effect on a subsystem of a measurement performed on a larger system. POVMs are extensively used in quantum information theory. As described above, entanglement

1419-426: A mathematical formulation of quantum mechanics and survey its application to some useful and oft-studied examples. In the mathematically rigorous formulation of quantum mechanics, the state of a quantum mechanical system is a vector ψ {\displaystyle \psi } belonging to a ( separable ) complex Hilbert space H {\displaystyle {\mathcal {H}}} . This vector

SECTION 10

#1732790441641

1548-417: A measurement of its position and also at the same time for a measurement of its momentum . Another consequence of the mathematical rules of quantum mechanics is the phenomenon of quantum interference , which is often illustrated with the double-slit experiment . In the basic version of this experiment, a coherent light source , such as a laser beam, illuminates a plate pierced by two parallel slits, and

1677-664: A particle initially has a momentum space wave function described by a normal distribution around some constant momentum p 0 according to φ ( p ) = ( x 0 ℏ π ) 1 / 2 exp ⁡ ( − x 0 2 ( p − p 0 ) 2 2 ℏ 2 ) , {\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),} where we have introduced

1806-858: A particle moving along with constant momentum at arbitrarily high precision. On the other hand, the standard deviation of the position is σ x = x 0 2 1 + ω 0 2 t 2 {\displaystyle \sigma _{x}={\frac {x_{0}}{\sqrt {2}}}{\sqrt {1+\omega _{0}^{2}t^{2}}}} such that the uncertainty product can only increase with time as σ x ( t ) σ p ( t ) = ℏ 2 1 + ω 0 2 t 2 {\displaystyle \sigma _{x}(t)\sigma _{p}(t)={\frac {\hbar }{2}}{\sqrt {1+\omega _{0}^{2}t^{2}}}} Starting with Kennard's derivation of position-momentum uncertainty, Howard Percy Robertson developed

1935-406: 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,

2064-471: A probability amplitude. Applying the Born rule to these amplitudes gives a probability density function for the position that the electron will be found to have when an experiment is performed to measure it. This is the best the theory can do; it cannot say for certain where the electron will be found. The Schrödinger equation relates the collection of probability amplitudes that pertain to one moment of time to

2193-2299: A reference scale x 0 = ℏ / m ω 0 {\textstyle x_{0}={\sqrt {\hbar /m\omega _{0}}}} , with ω 0 > 0 {\displaystyle \omega _{0}>0} 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 Φ ( p , t ) = ( x 0 ℏ π ) 1 / 2 exp ⁡ ( − x 0 2 ( p − p 0 ) 2 2 ℏ 2 − i p 2 t 2 m ℏ ) , {\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),} Ψ ( x , t ) = ( 1 x 0 π ) 1 / 2 e − x 0 2 p 0 2 / 2 ℏ 2 1 + i ω 0 t exp ⁡ ( − ( x − i x 0 2 p 0 / ℏ ) 2 2 x 0 2 ( 1 + i ω 0 t ) ) . {\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).} Since ⟨ p ( t ) ⟩ = p 0 {\displaystyle \langle p(t)\rangle =p_{0}} and σ p ( t ) = ℏ / ( 2 x 0 ) {\displaystyle \sigma _{p}(t)=\hbar /({\sqrt {2}}x_{0})} , this can be interpreted as

2322-1186: A right eigenstate of position with a constant eigenvalue x 0 . By definition, this means that x ^ | ψ ⟩ = x 0 | ψ ⟩ . {\displaystyle {\hat {x}}|\psi \rangle =x_{0}|\psi \rangle .} Applying the commutator to | ψ ⟩ {\displaystyle |\psi \rangle } yields [ x ^ , p ^ ] | ψ ⟩ = ( x ^ p ^ − p ^ x ^ ) | ψ ⟩ = ( x ^ − x 0 I ^ ) p ^ | ψ ⟩ = i ℏ | ψ ⟩ , {\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 ,} where Î

2451-405: A single electron in an unexcited atom is pictured classically as a particle moving in a circular trajectory around the atomic nucleus , whereas in quantum mechanics, it is described by a static wave function surrounding the nucleus. For example, the electron wave function for an unexcited hydrogen atom is a spherically symmetric function known as an s orbital ( Fig. 1 ). Analytic solutions of

2580-551: A single spatial dimension. A free particle is one which is not subject to external influences, so that its Hamiltonian consists only of its kinetic energy: The general solution of the Schrödinger equation is given by which is a superposition of all possible plane waves e i ( k x − ℏ k 2 2 m t ) {\displaystyle e^{i(kx-{\frac {\hbar k^{2}}{2m}}t)}} , which are eigenstates of

2709-451: A temporal frequency (in hertz) which has been divided by the speed of light in vacuum (usually in centimeters per second, cm⋅s ): The historical reason for using this spectroscopic wavenumber rather than frequency is that it is a convenient unit when studying atomic spectra by counting fringes per cm with an interferometer  : the spectroscopic wavenumber is the reciprocal of the wavelength of light in vacuum: which remains essentially

SECTION 20

#1732790441641

2838-603: A wave number, defined as the number of radians per unit distance, sometimes called "angular wavenumber", is more often used: When wavenumber is represented by the symbol ν , a frequency is still being represented, albeit indirectly. As described in the spectroscopy section, this is done through the relationship ν s c = 1 λ ≡ ν ~ , {\textstyle {\frac {\nu _{\text{s}}}{c}}\;=\;{\frac {1}{\lambda }}\;\equiv \;{\tilde {\nu }},} where ν s

2967-476: 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

3096-402: Is ψ ( x ) ∝ e i k 0 x = e i p 0 x / ℏ   . {\displaystyle \psi (x)\propto e^{ik_{0}x}=e^{ip_{0}x/\hbar }~.} The Born rule states that this should be interpreted as a probability density amplitude function in the sense that the probability of finding

3225-473: Is and this provides the lower bound on the product of standard deviations: Another consequence of the canonical commutation relation is that the position and momentum operators are Fourier transforms of each other, so that a description of an object according to its momentum is the Fourier transform of its description according to its position. The fact that dependence in momentum is the Fourier transform of

3354-600: Is 1 if X = x {\displaystyle X=x} 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 ψ ( x ) ∝ ∑ n A n e i p n x / ℏ   , {\displaystyle \psi (x)\propto \sum _{n}A_{n}e^{ip_{n}x/\hbar }~,} where A n represents

3483-553: Is a frequency expressed in the unit hertz . This is done for convenience as frequencies tend to be very large. Wavenumber has dimensions of reciprocal length , so its SI unit is the reciprocal of meters (m ). In spectroscopy it is usual to give wavenumbers in cgs unit (i.e., reciprocal centimeters; cm ); in this context, the wavenumber was formerly called the kayser , after Heinrich Kayser (some older scientific papers used this unit, abbreviated as K , where 1   K = 1   cm ). The angular wavenumber may be expressed in

3612-469: Is a fundamental theory that describes the behavior of nature at and below the scale of atoms . It is the foundation of all quantum physics , which includes quantum chemistry , quantum field theory , quantum technology , and quantum information science . Quantum mechanics can describe many systems that classical physics cannot. Classical physics can describe many aspects of nature at an ordinary ( macroscopic and (optical) microscopic ) scale, but

3741-478: Is a key feature of models of measurement processes in which an apparatus becomes entangled with the system being measured. Systems interacting with the environment in which they reside generally become entangled with that environment, a phenomenon known as quantum decoherence . This can explain why, in practice, quantum effects are difficult to observe in systems larger than microscopic. There are many mathematically equivalent formulations of quantum mechanics. One of

3870-731: Is a right eigenstate of the annihilation operator , a ^ | α ⟩ = α | α ⟩ , {\displaystyle {\hat {a}}|\alpha \rangle =\alpha |\alpha \rangle ,} which may be represented in terms of Fock states as | α ⟩ = e − | α | 2 2 ∑ n = 0 ∞ α n n ! | n ⟩ {\displaystyle |\alpha \rangle =e^{-{|\alpha |^{2} \over 2}}\sum _{n=0}^{\infty }{\alpha ^{n} \over {\sqrt {n!}}}|n\rangle } In

3999-424: Is a valid joint state that is not separable. States that are not separable are called entangled . If the state for a composite system is entangled, it is impossible to describe either component system A or system B by a state vector. One can instead define reduced density matrices that describe the statistics that can be obtained by making measurements on either component system alone. This necessarily causes

Uncertainty principle - Misplaced Pages Continue

4128-405: Is conserved under the evolution generated by B {\displaystyle B} . This implies a quantum version of the result proven by Emmy Noether in classical ( Lagrangian ) mechanics: for every differentiable symmetry of a Hamiltonian, there exists a corresponding conservation law . The simplest example of a quantum system with a position degree of freedom is a free particle in

4257-1066: Is considered as a sum over all possible classical and non-classical paths between the initial and final states. This is the quantum-mechanical counterpart of the action principle in classical mechanics. The Hamiltonian H {\displaystyle H} is known as the generator of time evolution, since it defines a unitary time-evolution operator U ( t ) = e − i H t / ℏ {\displaystyle U(t)=e^{-iHt/\hbar }} for each value of t {\displaystyle t} . From this relation between U ( t ) {\displaystyle U(t)} and H {\displaystyle H} , it follows that any observable A {\displaystyle A} that commutes with H {\displaystyle H} will be conserved : its expectation value will not change over time. This statement generalizes, as mathematically, any Hermitian operator A {\displaystyle A} can generate

4386-448: Is given by The operator U ( t ) = e − i H t / ℏ {\displaystyle U(t)=e^{-iHt/\hbar }} is known as the time-evolution operator, and has the crucial property that it is unitary . This time evolution is deterministic in the sense that – given an initial quantum state ψ ( 0 ) {\displaystyle \psi (0)} – it makes

4515-406: Is its associated eigenvector. More generally, the eigenvalue is degenerate and the probability is given by ⟨ ψ , P λ ψ ⟩ {\displaystyle \langle \psi ,P_{\lambda }\psi \rangle } , where P λ {\displaystyle P_{\lambda }} is the projector onto its associated eigenspace. In

4644-687: Is just the normal distribution . 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 x 0 as ψ ( x ) = ( m Ω π ℏ ) 1 / 4 exp ⁡ ( − m Ω ( x − x 0 ) 2 2 ℏ ) , {\displaystyle \psi (x)=\left({\frac {m\Omega }{\pi \hbar }}\right)^{1/4}\exp {\left(-{\frac {m\Omega (x-x_{0})^{2}}{2\hbar }}\right)},} where Ω describes

4773-726: Is known as wave–particle duality . In addition to light, electrons , atoms , and molecules are all found to exhibit the same dual behavior when fired towards a double slit. Another non-classical phenomenon predicted by quantum mechanics is quantum tunnelling : a particle that goes up against a potential barrier can cross it, even if its kinetic energy is smaller than the maximum of the potential. In classical mechanics this particle would be trapped. Quantum tunnelling has several important consequences, enabling radioactive decay , nuclear fusion in stars, and applications such as scanning tunnelling microscopy , tunnel diode and tunnel field-effect transistor . When quantum systems interact,

4902-409: Is known as a dispersion relation . For the special case of an electromagnetic wave in a vacuum, in which the wave propagates at the speed of light, k is given by: where E is the energy of the wave, ħ is the reduced Planck constant , and c is the speed of light in a vacuum. For the special case of a matter wave , for example an electron wave, in the non-relativistic approximation (in

5031-413: Is not in an eigenstate of that observable. 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

5160-444: Is not possible for the solution to be a single momentum eigenstate, or a single position eigenstate, as these are not normalizable quantum states. Instead, we can consider a Gaussian wave packet : which has Fourier transform, and therefore momentum distribution We see that as we make a {\displaystyle a} smaller the spread in position gets smaller, but the spread in momentum gets larger. Conversely, by making

5289-628: Is not sufficient for describing them at very small submicroscopic (atomic and subatomic ) scales. Most theories in classical physics can be derived from quantum mechanics as an approximation, valid at large (macroscopic/microscopic) scale. Quantum systems have bound states that are quantized to discrete values of energy , momentum , angular momentum , and other quantities, in contrast to classical systems where these quantities can be measured continuously. Measurements of quantum systems show characteristics of both particles and waves ( wave–particle duality ), and there are limits to how accurately

Uncertainty principle - Misplaced Pages Continue

5418-815: Is part of quantum communication protocols, such as quantum key distribution and superdense coding . Contrary to popular misconception, entanglement does not allow sending signals faster than light , as demonstrated by the no-communication theorem . Another possibility opened by entanglement is testing for " hidden variables ", hypothetical properties more fundamental than the quantities addressed in quantum theory itself, knowledge of which would allow more exact predictions than quantum theory provides. A collection of results, most significantly Bell's theorem , have demonstrated that broad classes of such hidden-variable theories are in fact incompatible with quantum physics. According to Bell's theorem, if nature actually operates in accord with any theory of local hidden variables, then

5547-540: Is postulated to be normalized under the Hilbert space inner product, that is, it obeys ⟨ ψ , ψ ⟩ = 1 {\displaystyle \langle \psi ,\psi \rangle =1} , and it is well-defined up to a complex number of modulus 1 (the global phase), that is, ψ {\displaystyle \psi } and e i α ψ {\displaystyle e^{i\alpha }\psi } represent

5676-466: Is replaced by − i ℏ ∂ ∂ x {\displaystyle -i\hbar {\frac {\partial }{\partial x}}} , and in particular in the non-relativistic Schrödinger equation in position space the momentum-squared term is replaced with a Laplacian times − ℏ 2 {\displaystyle -\hbar ^{2}} . When two different quantum systems are considered together,

5805-493: Is the commutator . For a pair of operators  and B ^ {\displaystyle {\hat {B}}} , one defines their commutator as [ A ^ , B ^ ] = A ^ B ^ − B ^ A ^ . {\displaystyle [{\hat {A}},{\hat {B}}]={\hat {A}}{\hat {B}}-{\hat {B}}{\hat {A}}.} In

5934-1014: Is the identity operator . Suppose, for the sake of proof by contradiction , that | ψ ⟩ {\displaystyle |\psi \rangle } is also a right eigenstate of momentum, with constant eigenvalue p 0 . If this were true, then one could write ( x ^ − x 0 I ^ ) p ^ | ψ ⟩ = ( x ^ − x 0 I ^ ) p 0 | ψ ⟩ = ( x 0 I ^ − x 0 I ^ ) p 0 | ψ ⟩ = 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.} On

6063-476: 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. It

6192-409: Is the free-space wavenumber, as above. The imaginary part of the wavenumber expresses attenuation per unit distance and is useful in the study of exponentially decaying evanescent fields . The propagation factor of a sinusoidal plane wave propagating in the positive x direction in a linear material is given by where The sign convention is chosen for consistency with propagation in lossy media. If

6321-415: Is the reduced Planck constant . The constant i ℏ {\displaystyle i\hbar } is introduced so that the Hamiltonian is reduced to the classical Hamiltonian in cases where the quantum system can be approximated by a classical system; the ability to make such an approximation in certain limits is called the correspondence principle . The solution of this differential equation

6450-455: Is the wavelength. It is sometimes called the "spectroscopic wavenumber". It equals the spatial frequency . For example, a wavenumber in inverse centimeters can be converted to a frequency expressed in the unit gigahertz by multiplying by 29.979 2458  cm/ns (the speed of light , in centimeters per nanosecond); conversely, an electromagnetic wave at 29.9792458 GHz has a wavelength of 1 cm in free space. In theoretical physics,

6579-420: Is then σ x σ p = ℏ ( n + 1 2 ) ≥ ℏ 2 .   {\displaystyle \sigma _{x}\sigma _{p}=\hbar \left(n+{\frac {1}{2}}\right)\geq {\frac {\hbar }{2}}.~} In particular, the above Kennard bound is saturated for the ground state n =0 , for which the probability density

SECTION 50

#1732790441641

6708-469: Is then If the state for the first system is the vector ψ A {\displaystyle \psi _{A}} and the state for the second system is ψ B {\displaystyle \psi _{B}} , then the state of the composite system is Not all states in the joint Hilbert space H A B {\displaystyle {\mathcal {H}}_{AB}} can be written in this form, however, because

6837-2742: Is to evaluate these inner products. ⟨ f ∣ g ⟩ − ⟨ g ∣ f ⟩ = ∫ − ∞ ∞ ψ ∗ ( x ) x ⋅ ( − i ℏ d d x ) ψ ( x ) d x − ∫ − ∞ ∞ ψ ∗ ( x ) ( − i ℏ d d x ) ⋅ x ψ ( x ) d x = i ℏ ⋅ ∫ − ∞ ∞ ψ ∗ ( x ) [ ( − x ⋅ d ψ ( x ) d x ) + d ( x ψ ( x ) ) d x ] d x = i ℏ ⋅ ∫ − ∞ ∞ ψ ∗ ( x ) [ ( − x ⋅ d ψ ( x ) d x ) + ψ ( x ) + ( x ⋅ d ψ ( x ) d x ) ] d x = i ℏ ⋅ ∫ − ∞ ∞ ψ ∗ ( x ) ψ ( x ) d x = i ℏ ⋅ ∫ − ∞ ∞ | ψ ( x ) | 2 d x = i ℏ {\displaystyle {\begin{aligned}\langle f\mid g\rangle -\langle g\mid f\rangle &=\int _{-\infty }^{\infty }\psi ^{*}(x)\,x\cdot \left(-i\hbar {\frac {d}{dx}}\right)\,\psi (x)\,dx-\int _{-\infty }^{\infty }\psi ^{*}(x)\,\left(-i\hbar {\frac {d}{dx}}\right)\cdot x\,\psi (x)\,dx\\&=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\\&=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\\&=i\hbar \cdot \int _{-\infty }^{\infty }\psi ^{*}(x)\psi (x)\,dx\\&=i\hbar \cdot \int _{-\infty }^{\infty }|\psi (x)|^{2}\,dx\\&=i\hbar \end{aligned}}} Plugging this into

6966-430: Is vital to illustrate how the principle applies to relatively intelligible physical situations since it is indiscernible on the macroscopic 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

7095-505: The Born rule : in the simplest case the eigenvalue λ {\displaystyle \lambda } is non-degenerate and the probability is given by | ⟨ λ → , ψ ⟩ | 2 {\displaystyle |\langle {\vec {\lambda }},\psi \rangle |^{2}} , where λ → {\displaystyle {\vec {\lambda }}}

7224-713: The canonical commutation relation : Given a quantum state, the Born rule lets us compute expectation values for both X {\displaystyle X} and P {\displaystyle P} , and moreover for powers of them. Defining the uncertainty for an observable by a standard deviation , we have and likewise for the momentum: The uncertainty principle states that Either standard deviation can in principle be made arbitrarily small, but not both simultaneously. This inequality generalizes to arbitrary pairs of self-adjoint operators A {\displaystyle A} and B {\displaystyle B} . The commutator of these two operators

7353-764: The de Broglie relation p = ℏ k {\displaystyle p=\hbar k} . The variances of x {\displaystyle x} and p {\displaystyle p} can be calculated explicitly: σ x 2 = L 2 12 ( 1 − 6 n 2 π 2 ) {\displaystyle \sigma _{x}^{2}={\frac {L^{2}}{12}}\left(1-{\frac {6}{n^{2}\pi ^{2}}}\right)} σ p 2 = ( ℏ n π L ) 2 . {\displaystyle \sigma _{p}^{2}=\left({\frac {\hbar n\pi }{L}}\right)^{2}.} The product of

7482-1008: 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 σ x 2 = ∫ − ∞ ∞ x 2 ⋅ | ψ ( x ) | 2 d x {\displaystyle \sigma _{x}^{2}=\int _{-\infty }^{\infty }x^{2}\cdot |\psi (x)|^{2}\,dx} σ p 2 = ∫ − ∞ ∞ p 2 ⋅ | φ ( p ) | 2 d p   . {\displaystyle \sigma _{p}^{2}=\int _{-\infty }^{\infty }p^{2}\cdot |\varphi (p)|^{2}\,dp~.} The function f ( x ) = x ⋅ ψ ( x ) {\displaystyle f(x)=x\cdot \psi (x)} can be interpreted as

7611-423: The photoelectric effect . These early attempts to understand microscopic phenomena, now known as the " old quantum theory ", led to the full development of quantum mechanics in the mid-1920s by Niels Bohr , Erwin Schrödinger , Werner Heisenberg , Max Born , Paul Dirac and others. The modern theory is formulated in various specially developed mathematical formalisms . In one of them, a mathematical entity called

7740-562: The wave function provides information, in the form of probability amplitudes , about what measurements of a particle's energy, momentum, and other physical properties may yield. Quantum mechanics allows the calculation of properties and behaviour of physical systems. It is typically applied to microscopic systems: molecules, atoms and sub-atomic particles. It has been demonstrated to hold for complex molecules with thousands of atoms, but its application to human beings raises philosophical problems, such as Wigner's friend , and its application to

7869-431: The Hilbert space for the spin of a single proton is simply the space of two-dimensional complex vectors C 2 {\displaystyle \mathbb {C} ^{2}} with the usual inner product. Physical quantities of interest – position, momentum, energy, spin – are represented by observables, which are Hermitian (more precisely, self-adjoint ) linear operators acting on

SECTION 60

#1732790441641

7998-411: The Hilbert space of the combined system is the tensor product of the Hilbert spaces of the two components. For example, let A and B be two quantum systems, with Hilbert spaces H A {\displaystyle {\mathcal {H}}_{A}} and H B {\displaystyle {\mathcal {H}}_{B}} , respectively. The Hilbert space of the composite system

8127-432: The Hilbert space. A quantum state can be an eigenvector of an observable, in which case it is called an eigenstate , and the associated eigenvalue corresponds to the value of the observable in that eigenstate. More generally, a quantum state will be a linear combination of the eigenstates, known as a quantum superposition . When an observable is measured, the result will be one of its eigenvalues with probability given by

8256-519: The Kennard bound σ x σ p = ℏ 2 m ω ℏ m ω 2 = ℏ 2 . {\displaystyle \sigma _{x}\sigma _{p}={\sqrt {\frac {\hbar }{2m\omega }}}\,{\sqrt {\frac {\hbar m\omega }{2}}}={\frac {\hbar }{2}}.} with position and momentum each contributing an amount ℏ / 2 {\textstyle {\sqrt {\hbar /2}}} in

8385-728: The Robertson uncertainty relation is given by σ A σ B ≥ | 1 2 i ⟨ [ A ^ , B ^ ] ⟩ | = 1 2 | ⟨ [ 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|.} Erwin Schrödinger showed how to allow for correlation between

8514-489: The Schrödinger equation are known for very few relatively simple model Hamiltonians including the quantum harmonic oscillator , the particle in a box , the dihydrogen cation , and the hydrogen atom . Even the helium atom – which contains just two electrons – has defied all attempts at a fully analytic treatment, admitting no solution in closed form . However, there are techniques for finding approximate solutions. One method, called perturbation theory , uses

8643-497: The Schrödinger equation for the particle in a box are or, from Euler's formula , Wavenumber In multidimensional systems , the wavenumber is the magnitude of the wave vector . The space of wave vectors is called reciprocal space . Wave numbers and wave vectors play an essential role in optics and the physics of wave scattering , such as X-ray diffraction , neutron diffraction , electron diffraction , and elementary particle physics. For quantum mechanical waves,

8772-749: The above inequalities, we get σ x 2 σ p 2 ≥ | ⟨ f ∣ g ⟩ | 2 ≥ ( ⟨ f ∣ g ⟩ − ⟨ g ∣ f ⟩ 2 i ) 2 = ( i ℏ 2 i ) 2 = ℏ 2 4 {\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}}} or taking

8901-454: The amplitude of these modes and is called the wave function in momentum space . In mathematical terms, we say that φ ( p ) {\displaystyle \varphi (p)} is the Fourier transform of ψ ( x ) {\displaystyle \psi (x)} and that x and p are conjugate variables . Adding together all of these plane waves comes at

9030-403: The analytic result for a simple quantum mechanical model to create a result for a related but more complicated model by (for example) the addition of a weak potential energy . Another approximation method applies to systems for which quantum mechanics produces only small deviations from classical behavior. These deviations can then be computed based on the classical motion. One consequence of

9159-544: The asterisk denotes the complex conjugate . With this inner product defined, we note that the variance for position can be written as σ x 2 = ∫ − ∞ ∞ | f ( x ) | 2 d x = ⟨ f ∣ f ⟩   . {\displaystyle \sigma _{x}^{2}=\int _{-\infty }^{\infty }|f(x)|^{2}\,dx=\langle f\mid f\rangle ~.} We can repeat this for momentum by interpreting

9288-441: The attenuation constant is positive, then the wave amplitude decreases as the wave propagates in the x-direction. Wavelength , phase velocity , and skin depth have simple relationships to the components of the wavenumber: Here we assume that the wave is regular in the sense that the different quantities describing the wave such as the wavelength, frequency and thus the wavenumber are constants. See wavepacket for discussion of

9417-606: The basic quantum formalism is the uncertainty principle. In its most familiar form, this states that no preparation of a quantum particle can imply simultaneously precise predictions both for a measurement of its position and for a measurement of its momentum. Both position and momentum are observables, meaning that they are represented by Hermitian operators . The position operator X ^ {\displaystyle {\hat {X}}} and momentum operator P ^ {\displaystyle {\hat {P}}} do not commute, but rather satisfy

9546-864: The brackets ⟨ O ^ ⟩ {\displaystyle \langle {\hat {\mathcal {O}}}\rangle } indicate an expectation value of the observable represented by operator O ^ {\displaystyle {\hat {\mathcal {O}}}} . For a pair of operators A ^ {\displaystyle {\hat {A}}} and B ^ {\displaystyle {\hat {B}}} , define their commutator as [ A ^ , B ^ ] = A ^ B ^ − B ^ A ^ , {\displaystyle [{\hat {A}},{\hat {B}}]={\hat {A}}{\hat {B}}-{\hat {B}}{\hat {A}},} and

9675-474: The case of a free particle, that is, the particle has no potential energy): Here p is the momentum of the particle, m is the mass of the particle, E is the kinetic energy of the particle, and ħ is the reduced Planck constant . Wavenumber is also used to define the group velocity . In spectroscopy , "wavenumber" ν ~ {\displaystyle {\tilde {\nu }}} (in reciprocal centimeters , cm ) refers to

9804-509: The case of position and momentum, the commutator is the canonical commutation relation [ x ^ , p ^ ] = i ℏ . {\displaystyle [{\hat {x}},{\hat {p}}]=i\hbar .} The physical meaning of the non-commutativity can be understood by considering the effect of the commutator on position and momentum eigenstates . Let | ψ ⟩ {\displaystyle |\psi \rangle } be

9933-413: The case when these quantities are not constant. In general, the angular wavenumber k (i.e. the magnitude of the wave vector ) is given by where ν is the frequency of the wave, λ is the wavelength, ω = 2 πν is the angular frequency of the wave, and v p is the phase velocity of the wave. The dependence of the wavenumber on the frequency (or more commonly the frequency on the wavenumber)

10062-404: The collection of probability amplitudes that pertain to another. One consequence of the mathematical rules of quantum mechanics is a tradeoff in predictability between measurable quantities. The most famous form of this uncertainty principle says that no matter how a quantum particle is prepared or how carefully experiments upon it are arranged, it is impossible to have a precise prediction for

10191-626: The continuous case, these formulas give instead the probability density . After the measurement, if result λ {\displaystyle \lambda } was obtained, the quantum state is postulated to collapse to λ → {\displaystyle {\vec {\lambda }}} , in the non-degenerate case, or to P λ ψ / ⟨ ψ , P λ ψ ⟩ {\textstyle P_{\lambda }\psi {\big /}\!{\sqrt {\langle \psi ,P_{\lambda }\psi \rangle }}} , in

10320-431: The dependence in position means that the momentum operator is equivalent (up to an i / ℏ {\displaystyle i/\hbar } factor) to taking the derivative according to the position, since in Fourier analysis differentiation corresponds to multiplication in the dual space . This is why in quantum equations in position space, the momentum p i {\displaystyle p_{i}}

10449-395: The energy eigenstates, a † | n ⟩ = n + 1 | n + 1 ⟩ {\displaystyle a^{\dagger }|n\rangle ={\sqrt {n+1}}|n+1\rangle } a | n ⟩ = n | n − 1 ⟩ , {\displaystyle a|n\rangle ={\sqrt {n}}|n-1\rangle ,}

10578-469: The equation above to get | ⟨ f ∣ g ⟩ | 2 ≥ ( ⟨ f ∣ g ⟩ − ⟨ g ∣ f ⟩ 2 i ) 2   . {\displaystyle |\langle f\mid g\rangle |^{2}\geq \left({\frac {\langle f\mid g\rangle -\langle g\mid f\rangle }{2i}}\right)^{2}~.} All that remains

10707-927: The following (the right most equality holds only when Ω = ω ): σ x σ p ≥ ℏ 4 3 + 1 2 ( Ω 2 ω 2 + ω 2 Ω 2 ) − ( 1 2 ( Ω 2 ω 2 + ω 2 Ω 2 ) − 1 ) = ℏ 2 . {\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}}.} A coherent state

10836-505: The formal inequality relating the standard deviation of position σ x and the standard deviation of momentum σ p was derived by Earle Hesse Kennard later that year and by Hermann Weyl in 1928: σ x σ p ≥ ℏ 2 {\displaystyle \sigma _{x}\sigma _{p}\geq {\frac {\hbar }{2}}} where ℏ = h 2 π {\displaystyle \hbar ={\frac {h}{2\pi }}}

10965-452: The function g ~ ( p ) = p ⋅ φ ( p ) {\displaystyle {\tilde {g}}(p)=p\cdot \varphi (p)} as a vector, but we can also take advantage of the fact that ψ ( x ) {\displaystyle \psi (x)} and φ ( p ) {\displaystyle \varphi (p)} are Fourier transforms of each other. We evaluate

11094-415: The general case. The probabilistic nature of quantum mechanics thus stems from the act of measurement. This is one of the most difficult aspects of quantum systems to understand. It was the central topic in the famous Bohr–Einstein debates , in which the two scientists attempted to clarify these fundamental principles by way of thought experiments . In the decades after the formulation of quantum mechanics,

11223-427: The integration by parts, the cancelled term vanishes because the wave function vanishes at infinity, and the final two integrations re-assert the Fourier transforms. Often the term − i ℏ d d x {\textstyle -i\hbar {\frac {d}{dx}}} is called the momentum operator in position space. Applying Plancherel's theorem and then Parseval's theorem , we see that

11352-462: The interference pattern appears via the varying density of these particle hits on the screen. Furthermore, versions of the experiment that include detectors at the slits find that each detected photon passes through one slit (as would a classical particle), and not through both slits (as would a wave). However, such experiments demonstrate that particles do not form the interference pattern if one detects which slit they pass through. This behavior

11481-4366: The inverse Fourier transform through integration by parts : g ( x ) = 1 2 π ℏ ⋅ ∫ − ∞ ∞ g ~ ( p ) ⋅ e i p x / ℏ d p = 1 2 π ℏ ∫ − ∞ ∞ p ⋅ φ ( p ) ⋅ e i p x / ℏ d p = 1 2 π ℏ ∫ − ∞ ∞ [ p ⋅ ∫ − ∞ ∞ ψ ( χ ) e − i p χ / ℏ d χ ] ⋅ e i p x / ℏ d p = i 2 π ∫ − ∞ ∞ [ ψ ( χ ) e − i p χ / ℏ | − ∞ ∞ − ∫ − ∞ ∞ d ψ ( χ ) d χ e − i p χ / ℏ d χ ] ⋅ e i p x / ℏ d p = − i ∫ − ∞ ∞ d ψ ( χ ) d χ [ 1 2 π ∫ − ∞ ∞ e i p ( x − χ ) / ℏ d p ] d χ = − i ∫ − ∞ ∞ d ψ ( χ ) d χ [ δ ( x − χ ℏ ) ] d χ = − i ℏ ∫ − ∞ ∞ d ψ ( χ ) d χ [ δ ( x − χ ) ] d χ = − i ℏ d ψ ( x ) d x = ( − i ℏ d d x ) ⋅ ψ ( x ) , {\displaystyle {\begin{aligned}g(x)&={\frac {1}{\sqrt {2\pi \hbar }}}\cdot \int _{-\infty }^{\infty }{\tilde {g}}(p)\cdot e^{ipx/\hbar }\,dp\\&={\frac {1}{\sqrt {2\pi \hbar }}}\int _{-\infty }^{\infty }p\cdot \varphi (p)\cdot e^{ipx/\hbar }\,dp\\&={\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\\&={\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\\&=-i\int _{-\infty }^{\infty }{\frac {d\psi (\chi )}{d\chi }}\left[{\frac {1}{2\pi }}\int _{-\infty }^{\infty }\,e^{ip(x-\chi )/\hbar }\,dp\right]\,d\chi \\&=-i\int _{-\infty }^{\infty }{\frac {d\psi (\chi )}{d\chi }}\left[\delta \left({\frac {x-\chi }{\hbar }}\right)\right]\,d\chi \\&=-i\hbar \int _{-\infty }^{\infty }{\frac {d\psi (\chi )}{d\chi }}\left[\delta \left(x-\chi \right)\right]\,d\chi \\&=-i\hbar {\frac {d\psi (x)}{dx}}\\&=\left(-i\hbar {\frac {d}{dx}}\right)\cdot \psi (x),\end{aligned}}} where v = ℏ − i p e − i p χ / ℏ {\displaystyle v={\frac {\hbar }{-ip}}e^{-ip\chi /\hbar }} in

11610-430: The light passing through the slits is observed on a screen behind the plate. The wave nature of light causes the light waves passing through the two slits to interfere , producing bright and dark bands on the screen – a result that would not be expected if light consisted of classical particles. However, the light is always found to be absorbed at the screen at discrete points, as individual particles rather than waves;

11739-733: The momentum must be less precise. This precision may be quantified by the standard deviations, σ x = ⟨ x ^ 2 ⟩ − ⟨ x ^ ⟩ 2 {\displaystyle \sigma _{x}={\sqrt {\langle {\hat {x}}^{2}\rangle -\langle {\hat {x}}\rangle ^{2}}}} σ p = ⟨ p ^ 2 ⟩ − ⟨ p ^ ⟩ 2 . {\displaystyle \sigma _{p}={\sqrt {\langle {\hat {p}}^{2}\rangle -\langle {\hat {p}}\rangle ^{2}}}.} As in

11868-432: The momentum operator with momentum p = ℏ k {\displaystyle p=\hbar k} . The coefficients of the superposition are ψ ^ ( k , 0 ) {\displaystyle {\hat {\psi }}(k,0)} , which is the Fourier transform of the initial quantum state ψ ( x , 0 ) {\displaystyle \psi (x,0)} . It

11997-1947: The notation N ( μ , σ 2 ) {\displaystyle {\mathcal {N}}(\mu ,\sigma ^{2})} to denote a normal distribution of mean μ and variance σ . Copying the variances above and applying trigonometric identities , we can write the product of the standard deviations as σ x σ p = ℏ 2 ( cos 2 ⁡ ( ω t ) + Ω 2 ω 2 sin 2 ⁡ ( ω t ) ) ( cos 2 ⁡ ( ω t ) + ω 2 Ω 2 sin 2 ⁡ ( ω t ) ) = ℏ 4 3 + 1 2 ( Ω 2 ω 2 + ω 2 Ω 2 ) − ( 1 2 ( Ω 2 ω 2 + ω 2 Ω 2 ) − 1 ) cos ⁡ ( 4 ω t ) {\displaystyle {\begin{aligned}\sigma _{x}\sigma _{p}&={\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)}}\\&={\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}}} From

12126-413: The oldest and most common is the " transformation theory " proposed by Paul Dirac , which unifies and generalizes the two earliest formulations of quantum mechanics – matrix mechanics (invented by Werner Heisenberg ) and wave mechanics (invented by Erwin Schrödinger ). An alternative formulation of quantum mechanics is Feynman 's path integral formulation , in which a quantum-mechanical amplitude

12255-412: The one-dimensional case in the x {\displaystyle x} direction, the time-independent Schrödinger equation may be written With the differential operator defined by with state ψ {\displaystyle \psi } in this case having energy E {\displaystyle E} coincident with the kinetic energy of the particle. The general solutions of

12384-1026: The operators, giving a stronger inequality, known as the Robertson-Schrödinger uncertainty relation , σ A 2 σ B 2 ≥ | 1 2 ⟨ { A ^ , B ^ } ⟩ − ⟨ A ^ ⟩ ⟨ B ^ ⟩ | 2 + | 1 2 i ⟨ [ 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},} Quantum mechanics Quantum mechanics

12513-455: The original quantum system ceases to exist as an independent entity (see Measurement in quantum mechanics ). The time evolution of a quantum state is described by the Schrödinger equation: Here H {\displaystyle H} denotes the Hamiltonian , the observable corresponding to the total energy of the system, and ℏ {\displaystyle \hbar }

12642-424: The other hand, the above canonical commutation relation requires that [ x ^ , p ^ ] | ψ ⟩ = i ℏ | ψ ⟩ ≠ 0. {\displaystyle [{\hat {x}},{\hat {p}}]|\psi \rangle =i\hbar |\psi \rangle \neq 0.} This implies that no quantum state can simultaneously be both

12771-501: The particle between a and b is P ⁡ [ a ≤ X ≤ b ] = ∫ a b | ψ ( x ) | 2 d x   . {\displaystyle \operatorname {P} [a\leq X\leq b]=\int _{a}^{b}|\psi (x)|^{2}\,\mathrm {d} x~.} In the case of the single-mode plane wave, | ψ ( x ) | 2 {\displaystyle |\psi (x)|^{2}}

12900-409: 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

13029-642: 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, σ x 2 = ℏ 2 m ω , {\displaystyle \sigma _{x}^{2}={\frac {\hbar }{2m\omega }},} σ p 2 = ℏ m ω 2 . {\displaystyle \sigma _{p}^{2}={\frac {\hbar m\omega }{2}}.} Therefore, every coherent state saturates

13158-428: The position becomes more and more uncertain. The uncertainty in momentum, however, stays constant. The particle in a one-dimensional potential energy box is the most mathematically simple example where restraints lead to the quantization of energy levels. The box is defined as having zero potential energy everywhere inside a certain region, and therefore infinite potential energy everywhere outside that region. For

13287-1679: The position is improved, i.e. reduced σ x , by using many plane waves, thereby weakening the precision of the momentum, i.e. increased σ p . Another way of stating this is that σ x and σ p 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. We are interested in the variances of position and momentum, defined as σ x 2 = ∫ − ∞ ∞ x 2 ⋅ | ψ ( x ) | 2 d x − ( ∫ − ∞ ∞ x ⋅ | ψ ( x ) | 2 d x ) 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}} σ p 2 = ∫ − ∞ ∞ p 2 ⋅ | φ ( p ) | 2 d p − ( ∫ − ∞ ∞ p ⋅ | φ ( p ) | 2 d p ) 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}~.} Without loss of generality , we will assume that

13416-465: 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,

13545-761: The quantity n 2 π 2 3 − 2 {\textstyle {\sqrt {{\frac {n^{2}\pi ^{2}}{3}}-2}}} is greater than 1, so the uncertainty principle is never violated. For numerical concreteness, the smallest value occurs when n = 1 {\displaystyle n=1} , in which case σ x σ p = ℏ 2 π 2 3 − 2 ≈ 0.568 ℏ > ℏ 2 . {\displaystyle \sigma _{x}\sigma _{p}={\frac {\hbar }{2}}{\sqrt {{\frac {\pi ^{2}}{3}}-2}}\approx 0.568\hbar >{\frac {\hbar }{2}}.} Assume

13674-400: The question of what constitutes a "measurement" has been extensively studied. Newer interpretations of quantum mechanics have been formulated that do away with the concept of " wave function collapse " (see, for example, the many-worlds interpretation ). The basic idea is that when a quantum system interacts with a measuring apparatus, their respective wave functions become entangled so that

13803-418: The relations Ω 2 ω 2 + ω 2 Ω 2 ≥ 2 , | cos ⁡ ( 4 ω t ) | ≤ 1 , {\displaystyle {\frac {\Omega ^{2}}{\omega ^{2}}}+{\frac {\omega ^{2}}{\Omega ^{2}}}\geq 2,\quad |\cos(4\omega t)|\leq 1,} we can conclude

13932-812: The relative contribution of the mode p n 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 ψ ( x ) = 1 2 π ℏ ∫ − ∞ ∞ φ ( p ) ⋅ e i p x / ℏ d p   , {\displaystyle \psi (x)={\frac {1}{\sqrt {2\pi \hbar }}}\int _{-\infty }^{\infty }\varphi (p)\cdot e^{ipx/\hbar }\,dp~,} with φ ( p ) {\displaystyle \varphi (p)} representing

14061-413: The result can be the creation of quantum entanglement : their properties become so intertwined that a description of the whole solely in terms of the individual parts is no longer possible. Erwin Schrödinger called entanglement "... the characteristic trait of quantum mechanics, the one that enforces its entire departure from classical lines of thought". Quantum entanglement enables quantum computing and

14190-566: The results of a Bell test will be constrained in a particular, quantifiable way. Many Bell tests have been performed and they have shown results incompatible with the constraints imposed by local hidden variables. It is not possible to present these concepts in more than a superficial way without introducing the mathematics involved; understanding quantum mechanics requires not only manipulating complex numbers, but also linear algebra , differential equations , group theory , and other more advanced subjects. Accordingly, this article will present

14319-725: The same in air, and so the spectroscopic wavenumber is directly related to the angles of light scattered from diffraction gratings and the distance between fringes in interferometers , when those instruments are operated in air or vacuum. Such wavenumbers were first used in the calculations of Johannes Rydberg in the 1880s. The Rydberg–Ritz combination principle of 1908 was also formulated in terms of wavenumbers. A few years later spectral lines could be understood in quantum theory as differences between energy levels, energy being proportional to wavenumber, or frequency. However, spectroscopic data kept being tabulated in terms of spectroscopic wavenumber rather than frequency or energy. For example,

14448-463: The same physical system. In other words, the possible states are points in the projective space of a Hilbert space, usually called the complex projective space . The exact nature of this Hilbert space is dependent on the system – for example, for describing position and momentum the Hilbert space is the space of complex square-integrable functions L 2 ( C ) {\displaystyle L^{2}(\mathbb {C} )} , while

14577-405: The same time. 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

14706-524: The spectroscopic wavenumbers of the emission spectrum of atomic hydrogen are given by the Rydberg formula : where R is the Rydberg constant , and n i and n f are the principal quantum numbers of the initial and final levels respectively ( n i is greater than n f for emission). A spectroscopic wavenumber can be converted into energy per photon E by Planck's relation : It can also be converted into wavelength of light: where n

14835-821: The square root σ x σ p ≥ ℏ 2   . {\displaystyle \sigma _{x}\sigma _{p}\geq {\frac {\hbar }{2}}~.} with equality if and only if p and x are linearly dependent. Note that the only physics involved in this proof was that ψ ( x ) {\displaystyle \psi (x)} and φ ( p ) {\displaystyle \varphi (p)} 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. In matrix mechanics, observables such as position and momentum are represented by self-adjoint operators. When considering pairs of observables, an important quantity

14964-460: The standard deviations is therefore σ x σ p = ℏ 2 n 2 π 2 3 − 2 . {\displaystyle \sigma _{x}\sigma _{p}={\frac {\hbar }{2}}{\sqrt {{\frac {n^{2}\pi ^{2}}{3}}-2}}.} For all n = 1 , 2 , 3 , … {\displaystyle n=1,\,2,\,3,\,\ldots } ,

15093-407: 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

15222-625: The superposition principle implies that linear combinations of these "separable" or "product states" are also valid. For example, if ψ A {\displaystyle \psi _{A}} and ϕ A {\displaystyle \phi _{A}} are both possible states for system A {\displaystyle A} , and likewise ψ B {\displaystyle \psi _{B}} and ϕ B {\displaystyle \phi _{B}} are both possible states for system B {\displaystyle B} , then

15351-441: The theory is that it usually cannot predict with certainty what will happen, but only give probabilities. Mathematically, a probability is found by taking the square of the absolute value of a complex number , known as a probability amplitude. This is known as the Born rule , named after physicist Max Born . For example, a quantum particle like an electron can be described by a wave function, which associates to each point in space

15480-450: The uncertainty principle expresses the relationship between conjugate variables in the transform. 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 k 0 or momentum p 0

15609-410: 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 . Such paired-variables are known as complementary variables or canonically conjugate variables . First introduced in 1927 by German physicist Werner Heisenberg ,

15738-632: The unit radian per meter (rad⋅m ), or as above, since the radian is dimensionless . For electromagnetic radiation in vacuum, wavenumber is directly proportional to frequency and to photon energy. Because of this, wavenumbers are used as a convenient unit of energy in spectroscopy. A complex-valued wavenumber can be defined for a medium with complex-valued relative permittivity ε r {\displaystyle \varepsilon _{r}} , relative permeability μ r {\displaystyle \mu _{r}} and refraction index n as: where k 0

15867-437: The universe as a whole remains speculative. Predictions of quantum mechanics have been verified experimentally to an extremely high degree of accuracy . For example, the refinement of quantum mechanics for the interaction of light and matter, known as quantum electrodynamics (QED), has been shown to agree with experiment to within 1 part in 10 when predicting the magnetic properties of an electron. A fundamental feature of

15996-526: The value of a physical quantity can be predicted prior to its measurement, given a complete set of initial conditions (the uncertainty principle ). Quantum mechanics arose gradually from theories to explain observations that could not be reconciled with classical physics, such as Max Planck 's solution in 1900 to the black-body radiation problem, and the correspondence between energy and frequency in Albert Einstein 's 1905 paper , which explained

16125-2026: The variance for momentum can be written as σ p 2 = ∫ − ∞ ∞ | g ~ ( p ) | 2 d p = ∫ − ∞ ∞ | g ( x ) | 2 d x = ⟨ 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 .} The Cauchy–Schwarz inequality asserts that σ x 2 σ p 2 = ⟨ 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}~.} 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 ∗ 2 i ) 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}.} we let z = ⟨ f | g ⟩ {\displaystyle z=\langle f|g\rangle } and z ∗ = ⟨ g ∣ f ⟩ {\displaystyle z^{*}=\langle g\mid f\rangle } and substitute these into

16254-544: The variances may be computed directly, σ x 2 = ℏ m ω ( n + 1 2 ) {\displaystyle \sigma _{x}^{2}={\frac {\hbar }{m\omega }}\left(n+{\frac {1}{2}}\right)} σ p 2 = ℏ m ω ( n + 1 2 ) . {\displaystyle \sigma _{p}^{2}=\hbar m\omega \left(n+{\frac {1}{2}}\right)\,.} The product of these standard deviations

16383-421: The wave mechanics interpretation above, one sees a tradeoff between the respective precisions of the two, quantified by the uncertainty principle. 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 ^ = ℏ 2 m ω (

16512-461: The wavenumber multiplied by the reduced Planck constant is the canonical momentum . Wavenumber can be used to specify quantities other than spatial frequency. For example, in optical spectroscopy , it is often used as a unit of temporal frequency assuming a certain speed of light . Wavenumber, as used in spectroscopy and most chemistry fields, is defined as the number of wavelengths per unit distance, typically centimeters (cm ): where λ

16641-1622: 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 | Ψ ( x , t ) | 2 ∼ N ( x 0 cos ⁡ ( ω t ) , ℏ 2 m Ω ( cos 2 ⁡ ( ω t ) + Ω 2 ω 2 sin 2 ⁡ ( ω t ) ) ) {\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)} | Φ ( p , t ) | 2 ∼ N ( − m x 0 ω sin ⁡ ( ω t ) , ℏ m Ω 2 ( cos 2 ⁡ ( ω t ) + ω 2 Ω 2 sin 2 ⁡ ( ω t ) ) ) , {\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),} where we have used

#640359