Quantum noise

Last updated

Quantum noise is noise arising from the indeterminate state of matter in accordance with fundamental principles of quantum mechanics, specifically the uncertainty principle and via zero-point energy fluctuations. Quantum noise is due to the apparently discrete nature of the small quantum constituents such as electrons, as well as the discrete nature of quantum effects, such as photocurrents.

Contents

Quantified noise is similar to classical noise theory and will not always return an asymmetric spectral density. [1] [ clarification needed ]

Shot noise as coined by J. Verdeyen [2] is a form of quantum noise related to the statistics of photon counting, the discrete nature of electrons, and intrinsic noise generation in electronics. In contrast to shot noise,[ clarification needed ] the quantum mechanical uncertainty principle sets a lower limit to a measurement. The uncertainty principle requires any amplifier or detector to have noise. [1]

Macroscopic manifestations of quantum phenomena are easily disturbed, so quantum noise is mainly observed in systems where conventional sources of noise are suppressed. In general, noise is uncontrolled random variation from an expected value and is typically unwanted. General causes are thermal fluctuations, mechanical vibrations, industrial noise, fluctuations of voltage from a power supply, thermal noise due to Brownian motion, instrumentation noise, a laser's output mode deviating from the desired mode of operation, etc. If present, and unless carefully controlled, these other noise sources typically dominate and mask quantum noise.

In astronomy, a device which pushes against the limits of quantum noise is the LIGO gravitational wave observatory.

A Heisenberg microscope

Quantum noise can be illustrated by considering a Heisenberg microscope where an atom's position is measured from the scattering of photons. The uncertainty principle is given as,

Where the is the uncertainty in an atom's position, and the is the uncertainty of the momentum or sometimes called the backaction (momentum transferred to the atom) when near the quantum limit. The precision of the position measurement can be increased at the expense of knowing the atom's momentum. When the position is precisely known enough backaction begins to affect the measurement in two ways. First, it will impart momentum back onto the measuring devices in extreme cases. Secondly, we have decreasing future knowledge of the atom's future position. Precise and sensitive instrumentation will approach the uncertainty principle at sufficiently control environments.

Basics of noise theory

Noise is of practical concern for precision engineering and engineered systems approaching the standard quantum limit. Typical engineered consideration of quantum noise is for quantum nondemolition measurement and quantum point contact. So quantifying noise is useful. [2] [3] [4] A signal's noise is quantified as the Fourier transform of its autocorrelation. The autocorrelation of a signal is given as

which measures when our signal is positively, negatively or not correlated at different times and . The time average, , is zero and our is a voltage signal. Its Fourier transform is

because we measure a voltage over a finite time window. The Wiener–Khinchin theorem generally states that a noise's power spectrum is given as the autocorrelation of a signal, i.e.,

The above relation is sometimes called the power spectrum or spectral density. In the above outline, we assumed that

One can show that an ideal "top-hat" signal, which may correspond to a finite measurement of a voltage over some time, will produce noise across its entire spectrum as a sinc function. Even in the classical case, noise is produced.

Classical to quantum noise

To study quantum noise, one replaces the corresponding classical measurements with quantum operators, e.g.,

where are the quantum statistical average using the density matrix in the Heisenberg picture.

Quantum noise and the uncertainty principle

The Heisenberg uncertainty implies the existence of noise. [5] An operator with a hermitian conjugate follows the relationship, . Define as where is real. The and are the quantum operators. We can show the following,

where the are the averages over the wavefunction and other statistical properties. The left terms are the uncertainty in and , the second term on the right is to covariance or which arises from coupling to an external source or quantum effects. The first term on the right corresponds to the Commutator relation and would cancel out if the x and y commuted. That is the origin of our quantum noise.

It is demonstrative to let and correspond to position and momentum that meets the well known commutator relation, . Then our new expression is,

Where the is the correlation. If the second term on the right vanishes, then we recover the Heisenberg uncertainty principle.

Harmonic motion and weakly coupled heat bath

Consider the motion of a simple harmonic oscillator with mass, , and frequency, , coupled to some heat bath which keeps the system in equilibrium. The equations of motion are given as,

The quantum autocorrelation is then,

Classically, there is no correlation between position and momentum. The uncertainty principle requires the second term to be nonzero. It goes to . We can take the equipartition theorem or the fact that in equilibrium the energy is equally shared among a molecule/atoms degrees of freedom in thermal equilibrium, i.e.,

In the classical autocorrelation, we have

while in the quantum autocorrelation we have

Where the fraction terms in parentheses is the zero-point energy uncertainty. The is the Bose-Einstein population distribution. Notice that the quantum is asymmetric in the due to the imaginary autocorrelation. As we increase to higher temperature that corresponds to taking the limit of . One can show that the quantum approaches the classical . This allows

Physical interpretation of spectral density

Typically, the positive frequency of the spectral density corresponds to the flow of energy into the oscillator (for example, the photons' quantized field), while the negative frequency corresponds to the emitted of energy from the oscillator. Physically, an asymmetric spectral density would correspond to either the net flow of energy from or to our oscillator model.

Linear gain and quantum uncertainty

Most optical communications use amplitude modulation where the quantum noise is predominantly the shot noise. A laser's quantum noise, when not considering shot noise, is the uncertainty of its electric field's amplitude and phase. That uncertainty becomes observable when a quantum amplifier preserves phase. The phase noise becomes important when the energy of the frequency modulation or phase modulation is comparable to the energy of the signal (frequency modulation is more robust than amplitude modulation due to the additive noise intrinsic to amplitude modulation).

Linear amplification

An ideal noiseless gain cannot exit. [6] Consider the amplification of stream of photons, an ideal linear noiseless gain, and the Energy-Time uncertainty relation.

The photons, ignoring the uncertainty in frequency, will have an uncertainty in its overall phase and number, and assume a known frequency, i.e., and . We can substitute these relations into our energy-time uncertainty equation to find the number-phase uncertainty relation or the uncertainty in the phase and photon numbers.

Let an ideal linear noiseless gain, , act on the photon stream. We also assume a unity quantum efficiency, or every photon is converted to a photocurrent. The output will be following with no noise added.

The phase will be modified too,

where the is the overall accumulated phase as the photons traveled through the gain medium. Substituting our output gain and phase uncertainties, gives us

Our gain is , which is a contradiction to our uncertainty principles. So a linear noiseless amplifier cannot increase its signal without noise. A deeper analysis done by H. Heffner showed the minimum noise power output required to meet the Heisenberg uncertainty principle is given as [7]

where is half of the full width at half max, the frequency of the photons, and is the Planck constant. The term with is sometimes called quantum noise [6]

Shot noise and instrumentation

In precision optics with highly stabilized lasers and efficient detectors, quantum noise refers to the fluctuations of signal.

The random error of interferometric measurements of position, due to the discrete character of photons measurement, is another quantum noise. The uncertainty of position of a probe in probe microscopy may also attributable to quantum noise; but not the dominant mechanism governing resolution.

In an electric circuit, the random fluctuations of a signal due to the discrete character of electrons can be called quantum noise. [8] An experiment by S. Saraf, et .al. [9] demonstrated shot noise limited measurements as a demonstration of quantum noise measurements. Generally speaking, they amplified a Nd:YAG free space laser with minimal noise addition as it transitioned from linear to nonlinear amplification. The experiment required Fabry-Perot for filtering laser mode noises and selecting frequencies, two separate but identical probe and saturating beams to ensure uncorrelated beams, a zigzag slab gain medium, and a balanced detector for measuring quantum noise or shot-noise limited noise.

Shot Noise Power

The theory behind noise analysis of photon statistics (sometimes called the forward Kolmogorov equation) starts from the Masters equation from Shimoda et al. [10]

where corresponds to the emission cross section and upper population number product , and the is the absorption cross section . The above relation is describing the probability of finding photons in radiation mode . The dynamic only considers neighboring modes and as the photons travel through a medium of excited and ground state atoms from position to . This gives us a total of 4 photon transitions associated to one photon energy level. Two photon number adding to the field and leaving an atom, and and two photons leaving a field to the atom and . Its noise power is given as,

Where,

Sarif, et al. demonstrated quantum noise or shot noise limited measurements over a wide range of power gain that agreed with theory.

Zero-point fluctuations

The existence of zero-point energy fluctuations is well-established in the theory of the quantised electromagnetic field. [12] Generally speaking, at the lowest energy excitation of a quantized field that permeates all space (i.e. the field mode being in the vacuum state), the root-mean-square fluctuation of field strength is non-zero. This accounts for vacuum fluctuations that permeate all space.

This vacuum fluctuation or quantum noise will effect classical systems. This manifest as quantum decoherence in an entangled system, normally attributed to thermal differences in the conditions surrounding each entangled particle.[ clarification needed ] Because entanglement is studied intensely in simple pairs of entangled photons, for example, decoherence observed in experiments could well be synonymous with "quantum noise" as to the source of the decoherence. Vacuum fluctuation is a possible causes for a quanta of energy to spontaneously appear in a given field or spacetime, then thermal differences must be associated with this event. Hence, it would cause decoherence in an entangled system in proximity of the event.[ dubious discuss ]

Coherent states and noise of a quantum amplifier

A laser is described by the coherent state of light, or the superposition of harmonic oscillators eigenstates. Erwin Schrödinger first derived the coherent state for the Schrödinger equation to meet the correspondence principle in 1926. [12]

The laser is a quantum mechanical phenomena (see Maxwell–Bloch equations, rotating wave approximation, and semi-classical model of a two level atom). The Einstein coefficients and the laser rate equations are adequate if one is interested in the population levels and one does not need to account for population quantum coherences (the off diagonal terms in a density matrix). Photons of the order of 108 corresponds to a moderate energy. The relative error of measurement of the intensity due to the quantum noise is on the order of 10−5. This is considered to be of good precision for most of applications.

Quantum amplifier

A quantum amplifier is an amplifier which operates close to the quantum limit. Quantum noise becomes important when a small signal is amplified. A small signal's quantum uncertainties in its quadrature are also amplified; this sets a lower limit to the amplifier. A quantum amplifier's noise is its output amplitude and phase. Generally, a laser is amplified across a spread of wavelengths around a central wavelength, some mode distribution, and polarization spread. But one can consider a single mode amplification and generalize to many different modes. A phase-invariant amplifier preserves the phase of the input gain without drastic changes to the output phase mode. [13]

Quantum amplification can be represented with a unitary operator, , as stated in D. Kouznetsov 1995 paper.

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.

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. The quantum harmonic oscillator arise in the quantum theory of a wide range of physical systems. For instance, a coherent state describes the oscillating motion of a particle confined in a quadratic potential well. 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.

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.

<span class="mw-page-title-main">Rabi cycle</span> Quantum mechanical phenomenon

In physics, the Rabi cycle is the cyclic behaviour of a two-level quantum system in the presence of an oscillatory driving field. A great variety of physical processes belonging to the areas of quantum computing, condensed matter, atomic and molecular physics, and nuclear and particle physics can be conveniently studied in terms of two-level quantum mechanical systems, and exhibit Rabi flopping when coupled to an optical driving field. The effect is important in quantum optics, magnetic resonance and quantum computing, and is named after Isidor Isaac Rabi.

<span class="mw-page-title-main">Squeezed coherent state</span> Type of quantum state

In physics, a squeezed coherent state is a quantum state that is usually described by two non-commuting observables having continuous spectra of eigenvalues. Examples are position and momentum of a particle, and the (dimension-less) electric field in the amplitude and in the mode of a light wave. The product of the standard deviations of two such operators obeys the uncertainty principle:

In quantum physics, Fermi's golden rule is a formula that describes the transition rate from one energy eigenstate of a quantum system to a group of energy eigenstates in a continuum, as a result of a weak perturbation. This transition rate is effectively independent of time and is proportional to the strength of the coupling between the initial and final states of the system as well as the density of states. It is also applicable when the final state is discrete, i.e. it is not part of a continuum, if there is some decoherence in the process, like relaxation or collision of the atoms, or like noise in the perturbation, in which case the density of states is replaced by the reciprocal of the decoherence bandwidth.

<span class="mw-page-title-main">Lamb shift</span> Difference in energy of hydrogenic atom electron states not predicted by the Dirac equation

In physics the Lamb shift, named after Willis Lamb, refers to an anomalous difference in energy between two electron orbitals in a hydrogen atom. The difference was not predicted by theory and it cannot be derived from the Dirac equation, which predicts identical energies. Hence the Lamb shift refers to a deviation from theory seen in the differing energies contained by the 2S1/2 and 2P1/2 orbitals of the hydrogen atom.

In physics, a free particle is a particle that, in some sense, is not bound by an external force, or equivalently not in a region where its potential energy varies. In classical physics, this means the particle is present in a "field-free" space. In quantum mechanics, it means the particle is in a region of uniform potential, usually set to zero in the region of interest since the potential can be arbitrarily set to zero at any point in space.

In physics, the Hanbury Brown and Twiss (HBT) effect is any of a variety of correlation and anti-correlation effects in the intensities received by two detectors from a beam of particles. HBT effects can generally be attributed to the wave–particle duality of the beam, and the results of a given experiment depend on whether the beam is composed of fermions or bosons. Devices which use the effect are commonly called intensity interferometers and were originally used in astronomy, although they are also heavily used in the field of quantum optics.

<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.

The Rabi frequency is the frequency at which the probability amplitudes of two atomic energy levels fluctuate in an oscillating electromagnetic field. It is proportional to the transition dipole moment of the two levels and to the amplitude of the electromagnetic field. Population transfer between the levels of such a 2-level system illuminated with light exactly resonant with the difference in energy between the two levels will occur at the Rabi frequency; when the incident light is detuned from this energy difference then the population transfer occurs at the generalized Rabi frequency. The Rabi frequency is a semiclassical concept since it treats the atom as an object with quantized energy levels and the electromagnetic field as a continuous wave.

<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.

A vacuum Rabi oscillation is a damped oscillation of an initially excited atom coupled to an electromagnetic resonator or cavity in which the atom alternately emits photon(s) into a single-mode electromagnetic cavity and reabsorbs them. The atom interacts with a single-mode field confined to a limited volume V in an optical cavity. Spontaneous emission is a consequence of coupling between the atom and the vacuum fluctuations of the cavity field.

A quantum limit in physics is a limit on measurement accuracy at quantum scales. Depending on the context, the limit may be absolute, or it may only apply when the experiment is conducted with naturally occurring quantum states and can be circumvented with advanced state preparation and measurement schemes.

<span class="mw-page-title-main">Stimulated Raman adiabatic passage</span>

Stimulated Raman adiabatic passage (STIRAP) is a process that permits transfer of a population between two applicable quantum states via at least two coherent electromagnetic (light) pulses. These light pulses drive the transitions of the three level Ʌ atom or multilevel system. The process is a form of state-to-state coherent control.

In quantum mechanics, magnetic resonance is a resonant effect that can appear when a magnetic dipole is exposed to a static magnetic field and perturbed with another, oscillating electromagnetic field. Due to the static field, the dipole can assume a number of discrete energy eigenstates, depending on the value of its angular momentum (azimuthal) quantum number. The oscillating field can then make the dipole transit between its energy states with a certain probability and at a certain rate. The overall transition probability will depend on the field's frequency and the rate will depend on its amplitude. When the frequency of that field leads to the maximum possible transition probability between two states, a magnetic resonance has been achieved. In that case, the energy of the photons composing the oscillating field matches the energy difference between said states. If the dipole is tickled with a field oscillating far from resonance, it is unlikely to transition. That is analogous to other resonant effects, such as with the forced harmonic oscillator. The periodic transition between the different states is called Rabi cycle and the rate at which that happens is called Rabi frequency. The Rabi frequency should not be confused with the field's own frequency. Since many atomic nuclei species can behave as a magnetic dipole, this resonance technique is the basis of nuclear magnetic resonance, including nuclear magnetic resonance imaging and nuclear magnetic resonance spectroscopy.

In quantum probability, the Belavkin equation, also known as Belavkin-Schrödinger equation, quantum filtering equation, stochastic master equation, is a quantum stochastic differential equation describing the dynamics of a quantum system undergoing observation in continuous time. It was derived and henceforth studied by Viacheslav Belavkin in 1988.

In quantum computing, Mølmer–Sørensen gate scheme refers to an implementation procedure for various multi-qubit quantum logic gates used mostly in trapped ion quantum computing. This procedure is based on the original proposition by Klaus Mølmer and Anders Sørensen in 1999-2000.

References

  1. 1 2 Clark, Aashish A. "Quantum Noise and quantum measurement" (PDF). Retrieved 13 December 2021.
  2. 1 2 Verdeyen, Joseph T. (1995). Laser Electronics (3rd ed.). Prentice-Hall. ISBN   9780137066667.
  3. Clerk, A. A.; Devoret, M. H.; Girvin, S. M.; Marquardt, Florian; Schoelkopf, R. J. (2010). "Introduction to quantum noise, measurement, and amplification". Rev. Mod. Phys. 82 (2): 1155--1208. arXiv: 0810.4729 . Bibcode:2010RvMP...82.1155C. doi:10.1103/RevModPhys.82.1155. S2CID   119200464.
  4. Henry, Charles H.; Kazarinov, Rudolf F. (1996). "Quantum noise in photonics". Rev. Mod. Phys. 68 (3): 01–853. Bibcode:1996RvMP...68..801H. doi:10.1103/RevModPhys.68.801.
  5. Crispin W. Gardiner and Paul Zoller (2004). Quantum Noise: A Handbook of Markovian and Non-Markovian Quantum Stochastic Methods with Applications to Quantum Optics (3rd ed.). Springer. ISBN   978-3540223016.
  6. 1 2 Desurvire, Emmanuel (1994). Erbium-Doped Fiber Amplifiers. Principles and Applications (1st ed.). Wiley-Interscience. ISBN   978-0471589778.
  7. Heffner, Hubert (1962). "The Fundamental Noise Limit of Linear Amplifiers". Proceedings of the IRE. 50 (7): 1604-1608. doi:10.1109/JRPROC.1962.288130. S2CID   51674821.
  8. C. W. Gardiner and Peter Zoller, Quantum Noise, Springer-Verlag (1991, 2000, 2004)
  9. Saraf, Shally and Urbanek, Karel and Byer, Robert L. and King, Peter J. (2005). "Quantum noise measurements in a continuous-wave laser-diode-pumped Nd:YAG saturated amplifier". Optics Letters. 30 (10): 1195–1197. Bibcode:2005OptL...30.1195S. doi:10.1364/ol.30.001195. PMID   15943307.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  10. Shimoda, Koichi and Takahasi, Hidetosi and H. Townes, Charles (1957). "Fluctuations in Amplification of Quanta with Application to Maser Amplifiers". Journal of the Physical Society of Japan. 12 (5): 686-700. Bibcode:1957JPSJ...12..686S. doi:10.1143/JPSJ.12.686.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  11. Bishnu P. Pal, ed. (2006). Guided Wave Optical Components and Devices: Basics, Technology, and Applications (1st ed.). Academic. ISBN   978-0-12-088481-0.
  12. 1 2 John S Townsend. (2012). A Modern Approach to Quantum Mechanics (2nd ed.). University Science Books. ISBN   978-1891389788.
  13. D. Kouznetsov; D. Rohrlich; R.Ortega (1995). "Quantum limit of noise of a phase-invariant amplifier". Physical Review A . 52 (2): 1665–1669. arXiv: cond-mat/9407011 . Bibcode:1995PhRvA..52.1665K. doi:10.1103/PhysRevA.52.1665. PMID   9912406. S2CID   19495906.

Further reading

Sources