Coherent state

Last updated

In physics, specifically in quantum mechanics, a coherent state is the specific quantum state of the quantum harmonic oscillator, often described as a state that has dynamics most closely resembling the oscillatory behavior of a classical harmonic oscillator. It was the first example of quantum dynamics when Erwin Schrödinger derived it in 1926, while searching for solutions of the Schrödinger equation that satisfy the correspondence principle. [1] The quantum harmonic oscillator (and hence the coherent states) arise in the quantum theory of a wide range of physical systems. [2] For instance, a coherent state describes the oscillating motion of a particle confined in a quadratic potential well (for an early reference, see e.g. Schiff's textbook [3] ). The coherent state describes a state in a system for which the ground-state wavepacket is displaced from the origin of the system. This state can be related to classical solutions by a particle oscillating with an amplitude equivalent to the displacement.

Contents

These states, expressed as eigenvectors of the lowering operator and forming an overcomplete family, were introduced in the early papers of John R. Klauder, e.g. [4] In the quantum theory of light (quantum electrodynamics) and other bosonic quantum field theories, coherent states were introduced by the work of Roy J. Glauber in 1963 and are also known as Glauber states.

The concept of coherent states has been considerably abstracted; it has become a major topic in mathematical physics and in applied mathematics, with applications ranging from quantization to signal processing and image processing (see Coherent states in mathematical physics). For this reason, the coherent states associated to the quantum harmonic oscillator are sometimes referred to as canonical coherent states (CCS), standard coherent states, Gaussian states, or oscillator states.

Coherent states in quantum optics

Figure 1: The electric field, measured by optical homodyne detection, as a function of phase for three coherent states emitted by a Nd:YAG laser. The amount of quantum noise in the electric field is completely independent of the phase. As the field strength, i.e. the oscillation amplitude a of the coherent state is increased, the quantum noise or uncertainty is constant at 1/2, and so becomes less and less significant. In the limit of large field the state becomes a good approximation of a noiseless stable classical wave. The average photon numbers of the three states from top to bottom are <n> =4.2, 25.2, 924.5 Coherent noise compare3.png
Figure 1: The electric field, measured by optical homodyne detection, as a function of phase for three coherent states emitted by a Nd:YAG laser. The amount of quantum noise in the electric field is completely independent of the phase. As the field strength, i.e. the oscillation amplitude α of the coherent state is increased, the quantum noise or uncertainty is constant at 1/2, and so becomes less and less significant. In the limit of large field the state becomes a good approximation of a noiseless stable classical wave. The average photon numbers of the three states from top to bottom are n=4.2, 25.2, 924.5
Figure 2: The oscillating wave packet corresponding to the second coherent state depicted in Figure 1. At each phase of the light field, the distribution is a Gaussian of constant width. Coherent state wavepacket.jpg
Figure 2: The oscillating wave packet corresponding to the second coherent state depicted in Figure 1. At each phase of the light field, the distribution is a Gaussian of constant width.
Figure 3: Wigner function of the coherent state depicted in Figure 2. The distribution is centered on state's amplitude a and is symmetric around this point. The ripples are due to experimental errors. Wigner function coherent state.png
Figure 3: Wigner function of the coherent state depicted in Figure 2. The distribution is centered on state's amplitude α and is symmetric around this point. The ripples are due to experimental errors.

In quantum optics the coherent state refers to a state of the quantized electromagnetic field, etc. [2] [6] [7] that describes a maximal kind of coherence and a classical kind of behavior. Erwin Schrödinger derived it as a "minimum uncertainty" Gaussian wavepacket in 1926, searching for solutions of the Schrödinger equation that satisfy the correspondence principle. [1] It is a minimum uncertainty state, with the single free parameter chosen to make the relative dispersion (standard deviation in natural dimensionless units) equal for position and momentum, each being equally small at high energy.

Further, in contrast to the energy eigenstates of the system, the time evolution of a coherent state is concentrated along the classical trajectories. The quantum linear harmonic oscillator, and hence coherent states, arise in the quantum theory of a wide range of physical systems. They occur in the quantum theory of light (quantum electrodynamics) and other bosonic quantum field theories.

While minimum uncertainty Gaussian wave-packets had been well-known, they did not attract full attention until Roy J. Glauber, in 1963, provided a complete quantum-theoretic description of coherence in the electromagnetic field. [8] In this respect, the concurrent contribution of E.C.G. Sudarshan should not be omitted, [9] (there is, however, a note in Glauber's paper that reads: "Uses of these states as generating functions for the -quantum states have, however, been made by J. Schwinger [10] ). Glauber was prompted to do this to provide a description of the Hanbury-Brown & Twiss experiment, which generated very wide baseline (hundreds or thousands of miles) interference patterns that could be used to determine stellar diameters. This opened the door to a much more comprehensive understanding of coherence. (For more, see Quantum mechanical description.)

In classical optics, light is thought of as electromagnetic waves radiating from a source. Often, coherent laser light is thought of as light that is emitted by many such sources that are in phase. Actually, the picture of one photon being in-phase with another is not valid in quantum theory. Laser radiation is produced in a resonant cavity where the resonant frequency of the cavity is the same as the frequency associated with the atomic electron transitions providing energy flow into the field. As energy in the resonant mode builds up, the probability for stimulated emission, in that mode only, increases. That is a positive feedback loop in which the amplitude in the resonant mode increases exponentially until some nonlinear effects limit it. As a counter-example, a light bulb radiates light into a continuum of modes, and there is nothing that selects any one mode over the other. The emission process is highly random in space and time (see thermal light). In a laser, however, light is emitted into a resonant mode, and that mode is highly coherent. Thus, laser light is idealized as a coherent state. (Classically we describe such a state by an electric field oscillating as a stable wave. See Fig.1)

Besides describing lasers, coherent states also behave in a convenient manner when describing the quantum action of beam splitters: two coherent-state input beams will simply convert to two coherent-state beams at the output with new amplitudes given by classical electromagnetic wave formulas; [11] such a simple behaviour does not occur for other input states, including number states. Likewise if a coherent-state light beam is partially absorbed, then the remainder is a pure coherent state with a smaller amplitude, whereas partial absorption of non-coherent-state light produces a more complicated statistical mixed state. [11] Thermal light can be described as a statistical mixture of coherent states, and the typical way of defining nonclassical light is that it cannot be described as a simple statistical mixture of coherent states. [11]

The energy eigenstates of the linear harmonic oscillator (e.g., masses on springs, lattice vibrations in a solid, vibrational motions of nuclei in molecules, or oscillations in the electromagnetic field) are fixed-number quantum states. The Fock state (e.g. a single photon) is the most particle-like state; it has a fixed number of particles, and phase is indeterminate. A coherent state distributes its quantum-mechanical uncertainty equally between the canonically conjugate coordinates, position and momentum, and the relative uncertainty in phase [defined heuristically] and amplitude are roughly equal—and small at high amplitude.

Quantum mechanical definition

Mathematically, a coherent state is defined to be the (unique) eigenstate of the annihilation operator â with corresponding eigenvalue α. Formally, this reads,

Since â is not hermitian, α is, in general, a complex number. Writing |α| and θ are called the amplitude and phase of the state .

The state is called a canonical coherent state in the literature, since there are many other types of coherent states, as can be seen in the companion article Coherent states in mathematical physics.

Physically, this formula means that a coherent state remains unchanged by the annihilation of field excitation or, say, a particle. An eigenstate of the annihilation operator has a Poissonian number distribution when expressed in a basis of energy eigenstates, as shown below. A Poisson distribution is a necessary and sufficient condition that all detections are statistically independent. Contrast this to a single-particle state ( Fock state): once one particle is detected, there is zero probability of detecting another.

The derivation of this will make use of (unconventionally normalized) dimensionless operators, X and P, normally called field quadratures in quantum optics. (See Nondimensionalization.) These operators are related to the position and momentum operators of a mass m on a spring with constant k,

Figure 4: The probability of detecting n photons, the photon number distribution, of the coherent state in Figure 3. As is necessary for a Poissonian distribution the mean photon number is equal to the variance of the photon number distribution. Bars refer to theory, dots to experimental values. Photon numbers coherent state.jpg
Figure 4: The probability of detecting n photons, the photon number distribution, of the coherent state in Figure 3. As is necessary for a Poissonian distribution the mean photon number is equal to the variance of the photon number distribution. Bars refer to theory, dots to experimental values.

For an optical field,

are the real and imaginary components of the mode of the electric field inside a cavity of volume . [12]

With these (dimensionless) operators, the Hamiltonian of either system becomes

Erwin Schrödinger was searching for the most classical-like states when he first introduced minimum uncertainty Gaussian wave-packets. The quantum state of the harmonic oscillator that minimizes the uncertainty relation with uncertainty equally distributed between X and P satisfies the equation

or, equivalently,

and hence

Thus, given (∆X−∆P)2 ≥ 0, Schrödinger found that the minimum uncertainty states for the linear harmonic oscillator are the eigenstates of(X + iP). Since â is (X + iP), this is recognizable as a coherent state in the sense of the above definition.

Using the notation for multi-photon states, Glauber characterized the state of complete coherence to all orders in the electromagnetic field to be the eigenstate of the annihilation operator—formally, in a mathematical sense, the same state as found by Schrödinger. The name coherent state took hold after Glauber's work.

If the uncertainty is minimized, but not necessarily equally balanced between X and P, the state is called a squeezed coherent state.

The coherent state's location in the complex plane (phase space) is centered at the position and momentum of a classical oscillator of the phase θ and amplitude |α| given by the eigenvalue α (or the same complex electric field value for an electromagnetic wave). As shown in Figure 5, the uncertainty, equally spread in all directions, is represented by a disk with diameter 12. As the phase varies, the coherent state circles around the origin and the disk neither distorts nor spreads. This is the most similar a quantum state can be to a single point in phase space.

Figure 5: Phase space plot of a coherent state. This shows that the uncertainty in a coherent state is equally distributed in all directions. The horizontal and vertical axes are the X and P quadratures of the field, respectively (see text). The red dots on the x-axis trace out the boundaries of the quantum noise in Figure 1. For more detail, see the corresponding figure of the phase space formulation. Coherent state2.png
Figure 5: Phase space plot of a coherent state. This shows that the uncertainty in a coherent state is equally distributed in all directions. The horizontal and vertical axes are the X and P quadratures of the field, respectively (see text). The red dots on the x-axis trace out the boundaries of the quantum noise in Figure 1. For more detail, see the corresponding figure of the phase space formulation.

Since the uncertainty (and hence measurement noise) stays constant at 12 as the amplitude of the oscillation increases, the state behaves increasingly like a sinusoidal wave, as shown in Figure 1. Moreover, since the vacuum state is just the coherent state with α=0, all coherent states have the same uncertainty as the vacuum. Therefore, one may interpret the quantum noise of a coherent state as being due to vacuum fluctuations.

The notation does not refer to a Fock state. For example, when α = 1, one should not mistake for the single-photon Fock state, which is also denoted in its own notation. The expression with α = 1 represents a Poisson distribution of number states with a mean photon number of unity.

The formal solution of the eigenvalue equation is the vacuum state displaced to a location α in phase space, i.e., it is obtained by letting the unitary displacement operator D(α) operate on the vacuum,

,

where â = X + iP and â = X - iP.

This can be easily seen, as can virtually all results involving coherent states, using the representation of the coherent state in the basis of Fock states,

where are energy (number) eigenvectors of the Hamiltonian

and the final equality derives from the Baker-Campbell-Hausdorff formula. For the corresponding Poissonian distribution, the probability of detecting n photons is

Similarly, the average photon number in a coherent state is

and the variance is

.

That is, the standard deviation of the number detected goes like the square root of the number detected. So in the limit of large α, these detection statistics are equivalent to that of a classical stable wave.

These results apply to detection results at a single detector and thus relate to first order coherence (see degree of coherence). However, for measurements correlating detections at multiple detectors, higher-order coherence is involved (e.g., intensity correlations, second order coherence, at two detectors). Glauber's definition of quantum coherence involves nth-order correlation functions (n-th order coherence) for all n. The perfect coherent state has all n-orders of correlation equal to 1 (coherent). It is perfectly coherent to all orders.

The second-order correlation coefficient gives a direct measure of the degree of coherence of photon states in terms of the variance of the photon statistics in the beam under study. [13]

In Glauber's development, it is seen that the coherent states are distributed according to a Poisson distribution. In the case of a Poisson distribution, the variance is equal to the mean, i.e.

.

A second-order correlation coefficient of 1 means that photons in coherent states are uncorrelated.

Hanbury Brown and Twiss studied the correlation behavior of photons emitted from a thermal, incoherent source described by Bose–Einstein statistics. The variance of the Bose–Einstein distribution is

.

This corresponds to the correlation measurements of Hanbury Brown and Twiss, and illustrates that photons in incoherent Bose–Einstein states are correlated or bunched.

Quanta that obey Fermi–Dirac statistics are anti-correlated. In this case the variance is

.

Anti-correlation is characterized by a second-order correlation coefficient =0.

Roy J. Glauber's work was prompted by the results of Hanbury-Brown and Twiss that produced long-range (hundreds or thousands of miles) first-order interference patterns through the use of intensity fluctuations (lack of second order coherence), with narrow band filters (partial first order coherence) at each detector. (One can imagine, over very short durations, a near-instantaneous interference pattern from the two detectors, due to the narrow band filters, that dances around randomly due to the shifting relative phase difference. With a coincidence counter, the dancing interference pattern would be stronger at times of increased intensity [common to both beams], and that pattern would be stronger than the background noise.) Almost all of optics had been concerned with first order coherence. The Hanbury-Brown and Twiss results prompted Glauber to look at higher order coherence, and he came up with a complete quantum-theoretic description of coherence to all orders in the electromagnetic field (and a quantum-theoretic description of signal-plus-noise). He coined the term coherent state and showed that they are produced when a classical electric current interacts with the electromagnetic field.

At α ≫ 1, from Figure 5, simple geometry gives Δθ |α | = 1/2. From this, it appears that there is a tradeoff between number uncertainty and phase uncertainty, ΔθΔn = 1/2, which is sometimes interpreted as a number-phase uncertainty relation; but this is not a formal strict uncertainty relation: there is no uniquely defined phase operator in quantum mechanics. [14] [15] [16] [17] [18] [19] [20] [21]

The wavefunction of a coherent state

Time evolution of the probability distribution with quantum phase (color) of a coherent state with a=3. QHO-coherentstate3-animation-color.gif
Time evolution of the probability distribution with quantum phase (color) of a coherent state with α=3.

To find the wavefunction of the coherent state, the minimal uncertainty Schrödinger wave packet, it is easiest to start with the Heisenberg picture of the quantum harmonic oscillator for the coherent state . Note that

The coherent state is an eigenstate of the annihilation operator in the Heisenberg picture.

It is easy to see that, in the Schrödinger picture, the same eigenvalue

occurs,

.

In the coordinate representations resulting from operating by , this amounts to the differential equation,

which is easily solved to yield

where θ(t) is a yet undetermined phase, to be fixed by demanding that the wavefunction satisfies the Schrödinger equation.

It follows that

so that σ is the initial phase of the eigenvalue.

The mean position and momentum of this "minimal Schrödinger wave packet" ψ(α) are thus oscillating just like a classical system,

The probability density remains a Gaussian centered on this oscillating mean,

Mathematical features of the canonical coherent states

The canonical coherent states described so far have three properties that are mutually equivalent, since each of them completely specifies the state , namely,

  1. They are eigenvectors of the annihilation operator:  .
  2. They are obtained from the vacuum by application of a unitary displacement operator:  .
  3. They are states of (balanced) minimal uncertainty:  .

Each of these properties may lead to generalizations, in general different from each other (see the article "Coherent states in mathematical physics" for some of these). We emphasize that coherent states have mathematical features that are very different from those of a Fock state; for instance, two different coherent states are not orthogonal,

(linked to the fact that they are eigenvectors of the non-self-adjoint annihilation operator â).

Thus, if the oscillator is in the quantum state it is also with nonzero probability in the other quantum state (but the farther apart the states are situated in phase space, the lower the probability is). However, since they obey a closure relation, any state can be decomposed on the set of coherent states. They hence form an overcomplete basis, in which one can diagonally decompose any state. This is the premise for the Glauber–Sudarshan P representation.

This closure relation can be expressed by the resolution of the identity operator I in the vector space of quantum states,

This resolution of the identity is intimately connected to the Segal–Bargmann transform.

Another peculiarity is that has no eigenket (while â has no eigenbra). The following equality is the closest formal substitute, and turns out to be useful for technical computations, [22]

This last state is known as an "Agarwal state" or photon-added coherent state and denoted as

Normalized Agarwal states of order n can be expressed as [23]

The above resolution of the identity may be derived (restricting to one spatial dimension for simplicity) by taking matrix elements between eigenstates of position, , on both sides of the equation. On the right-hand side, this immediately gives δ(x-y). On the left-hand side, the same is obtained by inserting

from the previous section (time is arbitrary), then integrating over using the Fourier representation of the delta function, and then performing a Gaussian integral over .

In particular, the Gaussian Schrödinger wave-packet state follows from the explicit value

The resolution of the identity may also be expressed in terms of particle position and momentum. For each coordinate dimension (using an adapted notation with new meaning for ),

the closure relation of coherent states reads

This can be inserted in any quantum-mechanical expectation value, relating it to some quasi-classical phase-space integral and explaining, in particular, the origin of normalisation factors for classical partition functions, consistent with quantum mechanics.

In addition to being an exact eigenstate of annihilation operators, a coherent state is an approximate common eigenstate of particle position and momentum. Restricting to one dimension again,

The error in these approximations is measured by the uncertainties of position and momentum,

Thermal coherent state

A single mode thermal coherent state [24] is produced by displacing a thermal mixed state in phase space, in direct analogy to the displacement of the vacuum state in view of generating a coherent state. The density matrix of a coherent thermal state in operator representation reads

where is the displacement operator, which generates the coherent state with complex amplitude , and . The partition function is equal to

Using the expansion of the identity operator in Fock states, , the density operator definition can be expressed in the following form

where stands for the displaced Fock state. We remark that if temperature goes to zero we have

which is the density matrix for a coherent state. The average number of photons in that state can be calculated as below

where for the last term we can write

As a result, we find

where is the average of the photon number calculated with respect to the thermal state. Here we have defined, for ease of notation,

and we write explicitly

In the limit we obtain , which is consistent with the expression for the density matrix operator at zero temperature. Likewise, the photon number variance can be evaluated as

with . We deduce that the second moment cannot be uncoupled to the thermal and the quantum distribution moments, unlike the average value (first moment). In that sense, the photon statistics of the displaced thermal state is not described by the sum of the Poisson statistics and the Boltzmann statistics. The distribution of the initial thermal state in phase space broadens as a result of the coherent displacement.

Coherent states of Bose–Einstein condensates

Coherent electron states in superconductivity

Generalizations

See also

Related Research Articles

<span class="mw-page-title-main">Uncertainty principle</span> Foundational principle in quantum physics

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.

<span class="mw-page-title-main">Quantum harmonic oscillator</span> Important, well-understood quantum mechanical model

The quantum harmonic oscillator is the quantum-mechanical analog of the classical harmonic oscillator. Because an arbitrary smooth potential can usually be approximated as a harmonic potential at the vicinity of a stable equilibrium point, it is one of the most important model systems in quantum mechanics. Furthermore, it is one of the few quantum-mechanical systems for which an exact, analytical solution is known.

The fluctuation–dissipation theorem (FDT) or fluctuation–dissipation relation (FDR) is a powerful tool in statistical physics for predicting the behavior of systems that obey detailed balance. Given that a system obeys detailed balance, the theorem is a proof that thermodynamic fluctuations in a physical variable predict the response quantified by the admittance or impedance of the same physical variable, and vice versa. The fluctuation–dissipation theorem applies both to classical and quantum mechanical systems.

In physics, the S-matrix or scattering matrix relates the initial state and the final state of a physical system undergoing a scattering process. It is used in quantum mechanics, scattering theory and quantum field theory (QFT).

Creation operators and annihilation operators are mathematical operators that have widespread applications in quantum mechanics, notably in the study of quantum harmonic oscillators and many-particle systems. An annihilation operator lowers the number of particles in a given state by one. A creation operator increases the number of particles in a given state by one, and it is the adjoint of the annihilation operator. In many subfields of physics and chemistry, the use of these operators instead of wavefunctions is known as second quantization. They were introduced by Paul Dirac.

<span class="mw-page-title-main">LSZ reduction formula</span> Connection between correlation functions and the S-matrix

In quantum field theory, the Lehmann–Symanzik–Zimmermann (LSZ) reduction formula is a method to calculate S-matrix elements from the time-ordered correlation functions of a quantum field theory. It is a step of the path that starts from the Lagrangian of some quantum field theory and leads to prediction of measurable quantities. It is named after the three German physicists Harry Lehmann, Kurt Symanzik and Wolfhart Zimmermann.

<span class="mw-page-title-main">Two-state quantum system</span> Simple quantum mechanical system

In quantum mechanics, a two-state system is a quantum system that can exist in any quantum superposition of two independent quantum states. The Hilbert space describing such a system is two-dimensional. Therefore, a complete basis spanning the space will consist of two independent states. Any two-state system can also be seen as a qubit.

A quasiprobability distribution is a mathematical object similar to a probability distribution but which relaxes some of Kolmogorov's axioms of probability theory. Quasiprobabilities share several of general features with ordinary probabilities, such as, crucially, the ability to yield expectation values with respect to the weights of the distribution. However, they can violate the σ-additivity axiom: integrating over them does not necessarily yield probabilities of mutually exclusive states. Indeed, quasiprobability distributions also have regions of negative probability density, counterintuitively, contradicting the first axiom. Quasiprobability distributions arise naturally in the study of quantum mechanics when treated in phase space formulation, commonly used in quantum optics, time-frequency analysis, and elsewhere.

In the quantum mechanics study of optical phase space, the displacement operator for one mode is the shift operator in quantum optics,

The Glauber–Sudarshan P representation is a suggested way of writing down the phase space distribution of a quantum system in the phase space formulation of quantum mechanics. The P representation is the quasiprobability distribution in which observables are expressed in normal order. In quantum optics, this representation, formally equivalent to several other representations, is sometimes preferred over such alternative representations to describe light in optical phase space, because typical optical observables, such as the particle number operator, are naturally expressed in normal order. It is named after George Sudarshan and Roy J. Glauber, who worked on the topic in 1963. Despite many useful applications in laser theory and coherence theory, the Sudarshan–Glauber P representation has the peculiarity that it is not always positive, and is not a bona-fide probability function.

<span class="mw-page-title-main">Jaynes–Cummings model</span> Model in quantum optics

The Jaynes–Cummings model is a theoretical model in quantum optics. It describes the system of a two-level atom interacting with a quantized mode of an optical cavity, with or without the presence of light. It was originally developed to study the interaction of atoms with the quantized electromagnetic field in order to investigate the phenomena of spontaneous emission and absorption of photons in a cavity.

Photon polarization is the quantum mechanical description of the classical polarized sinusoidal plane electromagnetic wave. An individual photon can be described as having right or left circular polarization, or a superposition of the two. Equivalently, a photon can be described as having horizontal or vertical linear polarization, or a superposition of the two.

The theoretical and experimental justification for the Schrödinger equation motivates the discovery of the Schrödinger equation, the equation that describes the dynamics of nonrelativistic particles. The motivation uses photons, which are relativistic particles with dynamics described by Maxwell's equations, as an analogue for all types of particles.

In many-body theory, the term Green's function is sometimes used interchangeably with correlation function, but refers specifically to correlators of field operators or creation and annihilation operators.

Resonance fluorescence is the process in which a two-level atom system interacts with the quantum electromagnetic field if the field is driven at a frequency near to the natural frequency of the atom.

<span class="mw-page-title-main">Optical phase space</span> Phase space used in quantum optics

In quantum optics, an optical phase space is a phase space in which all quantum states of an optical system are described. Each point in the optical phase space corresponds to a unique state of an optical system. For any such system, a plot of the quadratures against each other, possibly as functions of time, is called a phase diagram. If the quadratures are functions of time then the optical phase diagram can show the evolution of a quantum optical system with time.

In linear algebra, a raising or lowering operator is an operator that increases or decreases the eigenvalue of another operator. In quantum mechanics, the raising operator is sometimes called the creation operator, and the lowering operator the annihilation operator. Well-known applications of ladder operators in quantum mechanics are in the formalisms of the quantum harmonic oscillator and angular momentum.

This is a glossary for the terminology often encountered in undergraduate quantum mechanics courses.

In pure and applied mathematics, quantum mechanics and computer graphics, a tensor operator generalizes the notion of operators which are scalars and vectors. A special class of these are spherical tensor operators which apply the notion of the spherical basis and spherical harmonics. The spherical basis closely relates to the description of angular momentum in quantum mechanics and spherical harmonic functions. The coordinate-free generalization of a tensor operator is known as a representation operator.

In quantum optics, correlation functions are used to characterize the statistical and coherence properties – the ability of waves to interfere – of electromagnetic radiation, like optical light. Higher order coherence or n-th order coherence extends the concept of coherence to quantum optics and coincidence experiments. It is used to differentiate between optics experiments that require a quantum mechanical description from those for which classical fields are sufficient.

References

  1. 1 2 Schrödinger, E. (1926). "Der stetige Übergang von der Mikro- zur Makromechanik". Die Naturwissenschaften (in German). Springer Science and Business Media LLC. 14 (28): 664–666. Bibcode:1926NW.....14..664S. doi:10.1007/bf01507634. ISSN   0028-1042. S2CID   34680073.
  2. 1 2 J.R. Klauder and B. Skagerstam, Coherent States, World Scientific, Singapore, 1985.
  3. L.I. Schiff, Quantum Mechanics, McGraw Hill, New York, 1955.
  4. Klauder, John R (1960). "The action option and a Feynman quantization of spinor fields in terms of ordinary c-numbers". Annals of Physics. Elsevier BV. 11 (2): 123–168. Bibcode:1960AnPhy..11..123K. doi:10.1016/0003-4916(60)90131-7. ISSN   0003-4916.
  5. Breitenbach, G.; Schiller, S.; Mlynek, J. (1997). "Measurement of the quantum states of squeezed light" (PDF). Nature. Springer Nature. 387 (6632): 471–475. Bibcode:1997Natur.387..471B. doi:10.1038/387471a0. ISSN   0028-0836. S2CID   4259166.
  6. Zhang, Wei-Min; Feng, Da Hsuan; Gilmore, Robert (1990-10-01). "Coherent states: Theory and some applications". Reviews of Modern Physics. American Physical Society (APS). 62 (4): 867–927. Bibcode:1990RvMP...62..867Z. doi:10.1103/revmodphys.62.867. ISSN   0034-6861.
  7. J-P. Gazeau, Coherent States in Quantum Physics, Wiley-VCH, Berlin, 2009.
  8. Glauber, Roy J. (1963-09-15). "Coherent and Incoherent States of the Radiation Field". Physical Review. American Physical Society (APS). 131 (6): 2766–2788. Bibcode:1963PhRv..131.2766G. doi:10.1103/physrev.131.2766. ISSN   0031-899X.
  9. Sudarshan, E. C. G. (1963-04-01). "Equivalence of Semiclassical and Quantum Mechanical Descriptions of Statistical Light Beams". Physical Review Letters. American Physical Society (APS). 10 (7): 277–279. Bibcode:1963PhRvL..10..277S. doi:10.1103/physrevlett.10.277. ISSN   0031-9007.
  10. Schwinger, Julian (1953-08-01). "The Theory of Quantized Fields. III". Physical Review. American Physical Society (APS). 91 (3): 728–740. Bibcode:1953PhRv...91..728S. doi:10.1103/physrev.91.728. ISSN   0031-899X.
  11. 1 2 3 Leonhardt, Ulf (1997). Measuring the Quantum State of Light. Cambridge University Press. ISBN   9780521497305.
  12. "The Energy Density of Fields". www.sjsu.edu. Archived from the original on 2016-01-02.
  13. Pearsall, Thomas P., "Quantum Photonics, 2nd ed." Springer Nature, Cham, Switzerland, 2020, pp. 287 ff
  14. L. Susskind and J. Glogower, Quantum mechanical phase and time operator,Physics1 (1963) 49.
  15. Carruthers, P.; Nieto, Michael Martin (1968-04-01). "Phase and Angle Variables in Quantum Mechanics". Reviews of Modern Physics. American Physical Society (APS). 40 (2): 411–440. Bibcode:1968RvMP...40..411C. doi:10.1103/revmodphys.40.411. ISSN   0034-6861. S2CID   121002585.
  16. Barnett, S.M.; Pegg, D.T. (1989). "On the Hermitian Optical Phase Operator". Journal of Modern Optics. Informa UK Limited. 36 (1): 7–19. Bibcode:1989JMOp...36....7B. doi:10.1080/09500348914550021. ISSN   0950-0340.
  17. Busch, P.; Grabowski, M.; Lahti, P.J. (1995). "Who Is Afraid of POV Measures? Unified Approach to Quantum Phase Observables". Annals of Physics. Elsevier BV. 237 (1): 1–11. Bibcode:1995AnPhy.237....1B. doi:10.1006/aphy.1995.1001. ISSN   0003-4916.
  18. Dodonov, V V (2002-01-08). "'Nonclassical' states in quantum optics: a 'squeezed' review of the first 75 years". Journal of Optics B: Quantum and Semiclassical Optics. IOP Publishing. 4 (1): R1–R33. doi:10.1088/1464-4266/4/1/201. ISSN   1464-4266.
  19. V.V. Dodonov and V.I.Man'ko (eds), Theory of Nonclassical States of Light, Taylor \& Francis, London, New York, 2003.
  20. Vourdas, A (2006-02-01). "Analytic representations in quantum mechanics". Journal of Physics A: Mathematical and General. IOP Publishing. 39 (7): R65–R141. doi:10.1088/0305-4470/39/7/r01. ISSN   0305-4470.
  21. J-P. Gazeau,Coherent States in Quantum Physics, Wiley-VCH, Berlin, 2009.
  22. Scully, Marlan O.; Zubairy, M. Suhail (1997). Quantum Optics. Cambridge, UK: Cambridge University Press. p. 67. ISBN   9780521435956.
  23. Agarwal, G. S.; Tara, K. (1991-01-01). "Nonclassical properties of states generated by the excitations on a coherent state". Physical Review A. 43 (1): 492–497. Bibcode:1991PhRvA..43..492A. doi:10.1103/PhysRevA.43.492. PMID   9904801.
  24. Oz-Vogt, J.; Mann, A.; Revzen, M. (1991). "Thermal Coherent States and Thermal Squeezed States". Journal of Modern Optics. Informa UK Limited. 38 (12): 2339–2347. Bibcode:1991JMOp...38.2339O. doi:10.1080/09500349114552501. ISSN   0950-0340.
  25. Hyland, G.J.; Rowlands, G.; Cummings, F.W. (1970). "A proposal for an experimental determination of the equilibrium condensate fraction in superfluid helium". Physics Letters A. Elsevier BV. 31 (8): 465–466. Bibcode:1970PhLA...31..465H. doi:10.1016/0375-9601(70)90401-9. ISSN   0375-9601.
  26. Mayers, J. (2004-04-01). "Bose–Einstein Condensation, Phase Coherence, and Two-Fluid Behavior in 4He". Physical Review Letters. American Physical Society (APS). 92 (13): 135302. Bibcode:2004PhRvL..92m5302M. doi:10.1103/physrevlett.92.135302. ISSN   0031-9007. PMID   15089620.
  27. Mayers, J. (2006-07-26). "Bos–nstein condensation and two fluid behavior in 4He". Physical Review B. American Physical Society (APS). 74 (1): 014516. Bibcode:2006PhRvB..74a4516M. doi:10.1103/physrevb.74.014516. ISSN   1098-0121.
  28. Olinto, A. C. (1987-04-01). "Condensate fraction in superfluidHe4". Physical Review B. American Physical Society (APS). 35 (10): 4771–4774. Bibcode:1987PhRvB..35.4771O. doi:10.1103/physrevb.35.4771. ISSN   0163-1829. PMID   9940648.
  29. Penrose, Oliver; Onsager, Lars (1956-11-01). "Bose–Einstein Condensation and Liquid Helium". Physical Review. American Physical Society (APS). 104 (3): 576–584. Bibcode:1956PhRv..104..576P. doi:10.1103/physrev.104.576. ISSN   0031-899X.
  30. 1 2 Yang, C. N. (1962-10-01). "Concept of Off-Diagonal Long-Range Order and the Quantum Phases of Liquid He and of Superconductors". Reviews of Modern Physics. American Physical Society (APS). 34 (4): 694–704. Bibcode:1962RvMP...34..694Y. doi:10.1103/revmodphys.34.694. ISSN   0034-6861.
  31. [see John Bardeen's chapter in: Cooperative Phenomena, eds. H. Haken and M. Wagner (Springer-Verlag, Berlin, Heidelberg, New York, 1973)]
  32. Bardeen, J.; Cooper, L. N.; Schrieffer, J. R. (1957-12-01). "Theory of Superconductivity". Physical Review. American Physical Society (APS). 108 (5): 1175–1204. Bibcode:1957PhRv..108.1175B. doi: 10.1103/physrev.108.1175 . ISSN   0031-899X.
  33. A. M. Perelomov, Coherent states for arbitrary Lie groups, Commun. Math. Phys.26 (1972) 222-236; arXiv: math-ph/0203002.
  34. A. Perelomov, Generalized coherent states and their applications, Springer, Berlin 1986.
  35. Gilmore, Robert (1972). "Geometry of symmetrized states". Annals of Physics. Elsevier BV. 74 (2): 391–463. Bibcode:1972AnPhy..74..391G. doi:10.1016/0003-4916(72)90147-9. ISSN   0003-4916.
  36. Gilmore, R. (1974). "On properties of coherent states" (PDF). Revista Mexicana de Física. 23 (1–2): 143–187.
  37. G. Kaiser, Quantum Physics, Relativity, and Complex Spacetime: Towards a New Synthesis, North-Holland, Amsterdam, 1990.
  38. S.T. Ali, J-P. Antoine, and J-P. Gazeau, Coherent States, Wavelets and Their Generalizations, Springer-Verlag, New York, Berlin, Heidelberg, 2000.
  39. Anastopoulos, Charis (2004-08-25). "Generalized coherent states for spinning relativistic particles". Journal of Physics A: Mathematical and General. 37 (36): 8619–8637. arXiv: quant-ph/0312025 . Bibcode:2004JPhA...37.8619A. doi:10.1088/0305-4470/37/36/004. ISSN   0305-4470. S2CID   119064935.
  40. Ashtekar, Abhay; Lewandowski, Jerzy; Marolf, Donald; Mourão, José; Thiemann, Thomas (1996). "Coherent State Transforms for Spaces of Connections". Journal of Functional Analysis. 135 (2): 519–551. arXiv: gr-qc/9412014 . doi: 10.1006/jfan.1996.0018 . ISSN   0022-1236.
  41. Sahlmann, H.; Thiemann, T.; Winkler, O. (2001). "Coherent states for canonical quantum general relativity and the infinite tensor product extension". Nuclear Physics B. Elsevier BV. 606 (1–2): 401–440. arXiv: gr-qc/0102038 . Bibcode:2001NuPhB.606..401S. doi:10.1016/s0550-3213(01)00226-7. ISSN   0550-3213. S2CID   17857852.