Fact-checked by Grok 2 weeks ago

Photon statistics

Photon statistics is a fundamental subfield of that examines the statistical distributions governing the number of photons and their arrival times in light fields, providing insights into the quantum versus classical behavior of . This discipline analyzes how photons, as indistinguishable bosons, exhibit correlations that deviate from classical expectations, such as bunching in thermal light or antibunching in non-classical sources. Key to photon statistics is the classification of light based on its photon number distribution and second-order correlation function g^{(2)}(\tau), which quantifies intensity fluctuations over time delays \tau. For coherent light, like that from an ideal laser, the distribution is Poissonian, where the variance equals the mean photon number \langle n \rangle, and g^{(2)}(0) = 1, indicating no correlations. In contrast, thermal or chaotic light, such as blackbody radiation, shows super-Poissonian statistics with variance greater than \langle n \rangle and g^{(2)}(0) = 2, reflecting photon bunching due to Bose-Einstein statistics. Non-classical light sources, including squeezed states from parametric down-conversion or resonance fluorescence from driven atoms, display sub-Poissonian statistics with variance less than \langle n \rangle and g^{(2)}(0) < 1, demonstrating antibunching and quantum noise reduction. The theoretical foundation of photon statistics was established by Roy J. Glauber in 1963 through his development of coherent states, which provide a quantum description of laser light's Poissonian photon counts and optical coherence. Earlier, Max Planck's 1900 derivation of blackbody radiation introduced the thermal distribution underlying super-Poissonian behavior. Experimentally, photon statistics are measured using photon-counting detectors in setups like the Hanbury Brown and Twiss interferometer, which correlates detection events to compute g^{(2)}(\tau). These measurements are crucial for verifying quantum effects, such as at beam splitters where single-photon interference reveals superposition and indistinguishability. Beyond characterization, photon statistics underpin applications in quantum technologies, including secure quantum key distribution where sub-Poissonian sources enhance security against eavesdropping, and quantum computing protocols relying on controlled photon correlations. The Mandel's Q parameter, defined as Q = (\langle (\Delta n)^2 \rangle - \langle n \rangle)/\langle n \rangle, further quantifies deviations from Poissonian statistics, with Q = 0 for coherent light, Q > 0 for classical bunched light, and Q < 0 for non-classical states. Ongoing research extends these concepts to multiphoton correlations and hybrid light-matter systems, advancing fields like quantum sensing and metrology.

Fundamentals

Photon number distribution

The photon number distribution P(n) represents the probability of detecting n photons in a specified mode of the electromagnetic field or within a given time interval during a measurement. This distribution encapsulates the statistical properties of light at the quantum level, distinguishing it from classical wave descriptions by accounting for the discrete, probabilistic nature of photon arrivals. In quantum optics, P(n) is derived from the field's density operator \rho, which describes the quantum state of the radiation, as the diagonal matrix element in the Fock basis: P(n) = \langle n | \rho | n \rangle, where |n\rangle denotes the number state (Fock state) with exactly n photons. This expression arises from the trace over the projector |n\rangle\langle n|, yielding the probability for an ideal photon-number-resolving detector to register n photons. For idealized cases, the vacuum state |0\rangle has P(0) = 1 and P(n) = 0 for all n > 0, reflecting the absence of photons with certainty. Similarly, the single-photon state |1\rangle yields P(1) = 1 and P(n) = 0 otherwise, representing a deterministic emission of one photon. The foundational ideas underlying photon number distributions trace back to the early 20th century, with Max Planck's 1900 introduction of quantized energy for blackbody radiation and Albert Einstein's 1905 explanation of the photoelectric effect via light quanta, establishing photons as discrete entities in the quantum theory of radiation. These concepts were formalized in the 1960s by Roy J. Glauber through his development of coherent state theory, which provided a rigorous quantum framework for optical coherence and photon statistics. Graphical representations of P(n) versus n for various quantum states of light visually highlight differences in photon arrival patterns, such as sharp peaks for Fock states indicating regularity versus broader spreads suggesting bunching tendencies. For example, the vacuum state's distribution is a delta function at n=0, while the single-photon state's is at n=1, aiding intuitive understanding of quantum versus classical behaviors.

Statistical parameters

The variance of the photon number, denoted as \Delta n^2 = \langle n^2 \rangle - \langle n \rangle^2, quantifies the spread in the photon number distribution P(n), where \langle n \rangle = \sum_n n P(n) is the mean photon number and \langle n^2 \rangle = \sum_n n^2 P(n) is the second moment. This parameter is fundamental for characterizing fluctuations relative to the mean, with \Delta n^2 = \langle n \rangle indicating Poissonian statistics typical of coherent light. The Mandel Q parameter provides a normalized measure of deviation from Poissonian statistics, defined as Q = \frac{\Delta n^2 - \langle n \rangle}{\langle n \rangle}. It is derived directly from P(n) via the moments: \langle n \rangle = \sum_{n=0}^\infty n P(n) and \Delta n^2 = \sum_{n=0}^\infty n^2 P(n) - \left( \sum_{n=0}^\infty n P(n) \right)^2, yielding Q = 0 for Poissonian distributions, Q > 0 for super-Poissonian (bunched) light with enhanced fluctuations, and Q < 0 for sub-Poissonian (antibunched) light with reduced noise below the shot-noise limit. This parameter was introduced in the context of resonance fluorescence to identify nonclassical effects. Higher-order moments, such as skewness \gamma_1 = \frac{\langle (n - \langle n \rangle)^3 \rangle}{\Delta n^3} and excess kurtosis \gamma_2 = \frac{\langle (n - \langle n \rangle)^4 \rangle}{\Delta n^4} - 3, capture asymmetries and tail behaviors in P(n) that reveal multi-photon correlations beyond variance. These moments, computed from central moments of P(n), are essential for distinguishing complex nonclassical states involving three or more photons. The Fano factor F = \frac{\Delta n^2}{\langle n \rangle} = 1 + Q normalizes variance to the mean, analogous to noise measures in particle detection from . It equals 1 for Poissonian statistics and is used to analogize photon streams to particle emissions in experiments. The zero-time second-order correlation function g^{(2)}(0) = \frac{\langle n(n-1) \rangle}{\langle n \rangle^2} probes pairwise photon correlations, where \langle n(n-1) \rangle = \sum_n n(n-1) P(n), linking to Q via g^{(2)}(0) = 1 + \frac{Q}{\langle n \rangle}. Values g^{(2)}(0) > 1 indicate bunching (super-Poissonian), while g^{(2)}(0) < 1 signals antibunching (nonclassical), as formalized in .

Classical photon statistics

Poissonian light

Poissonian light exhibits a photon number distribution given by the Poisson probability mass function, P(n) = \frac{e^{-\langle n \rangle} \langle n \rangle^n}{n!}, where n is the number of photons and \langle n \rangle is the mean photon number. This distribution arises for fields where the variance in photon number equals the mean, \Delta n^2 = \langle n \rangle, corresponding to Mandel's Q parameter of zero, which serves as the benchmark for classical coherent light. The primary physical source of Poissonian light is the output of an ideal laser operating well above threshold, which produces light in a quantum coherent state |\alpha\rangle, where \alpha is the complex amplitude related to the mean photon number by |\alpha|^2 = \langle n \rangle. The coherent state can be expanded in the photon number basis as |\alpha\rangle = e^{-|\alpha|^2/2} \sum_{n=0}^\infty \frac{\alpha^n}{\sqrt{n!}} |n\rangle, yielding the Poissonian photon number probabilities P(n) = |\langle n | \alpha \rangle|^2. This property emerges from the laser's stimulated emission process, which amplifies a coherent superposition of photon number states while maintaining minimum fluctuations. In the quantum mechanical framework, coherent states represent minimum uncertainty states for the electromagnetic field quadratures, saturating the Heisenberg uncertainty relation \Delta X \Delta P \geq 1/2, analogous to classical waves. They are eigenstates of the annihilation operator, satisfying \hat{a} |\alpha\rangle = \alpha |\alpha\rangle, which ensures that field measurements yield classical-like outcomes with Poissonian fluctuations in photon counts. This description, introduced by Glauber, captures the quantum nature of laser light while aligning with classical intensity correlations. Experimental confirmation of Poissonian statistics in coherent light has been achieved through photon counting measurements on early gas lasers, where the observed photocount distributions matched the Poisson form over a wide range of mean photon numbers. Additionally, Hanbury Brown-Twiss interferometry on coherent beams demonstrates Poissonian noise by yielding a second-order correlation function g^{(2)}(0) = 1, indicating no excess or deficit in intensity fluctuations. Such light establishes the shot-noise limit in photodetection, where the fundamental noise floor is set by the Poisson variance, providing a reference standard for evaluating the performance of classical optical systems against quantum-enhanced alternatives.

Super-Poissonian light

Super-Poissonian light refers to optical fields where the photon number variance exceeds the mean photon number, quantified by Mandel's Q parameter defined as Q = \frac{\Delta n^2 - \langle n \rangle}{\langle n \rangle} > 0, resulting in enhanced fluctuations and a tendency for photons to arrive in bunches rather than independently. This contrasts with the baseline Poissonian statistics of coherent light, where Q = 0. Such statistics arise in classical incoherent sources due to random phase variations across the field. The primary examples of super-Poissonian light include from blackbody sources and chaotic light from multimode lasers operating above threshold, where intensity fluctuations stem from noise. For a single-mode thermal state, the photon number follows a geometric form, P(n) = (1 - a) a^n, with a = \langle n \rangle / (1 + \langle n \rangle), yielding \Delta n^2 = \langle n \rangle + \langle n \rangle^2 > \langle n \rangle. In , super-Poissonian light is described by chaotic states with a density operator \rho = \int P(\alpha) |\alpha\rangle\langle\alpha| \, d^2\alpha, where P(\alpha) is a Gaussian quasi-probability distribution centered at the origin with width proportional to the mean number, reflecting the random superposition of coherent states. The second-order at zero time delay, g^{(2)}(0) = 2, quantifies the bunching effect in thermal light, indicating that the probability of detecting two s simultaneously is twice that expected for uncorrelated arrival. This phenomenon was first experimentally observed in the 1950s through stellar experiments by Hanbury Brown and Twiss, who measured intensity correlations in from stars and a mercury , revealing excess fluctuations attributable to bunching in thermal sources.

Non-classical photon statistics

Sub-Poissonian light

Sub-Poissonian refers to quantum states of the where the variance in number, Δn², is less than the mean number, ⟨n⟩, indicating reduced fluctuations below the Poissonian limit characteristic of coherent . This non-classical behavior is quantified by Mandel's parameter, defined as Q = (Δn² / ⟨n⟩) - 1, where Q < 0 signifies sub-Poissonian statistics. Such states violate classical inequalities for photon counting noise and are a hallmark of quantum correlations in the field. Amplitude-squeezed states, generated through nonlinear optical processes like degenerate optical parametric amplification in a cavity, exhibit sub-Poissonian photon number statistics in the amplitude quadrature. In these states, the uncertainty in the field amplitude is reduced below the vacuum level, leading to Δn² < ⟨n⟩ for bright fields. Fock states |n⟩, which are eigenstates of the photon number operator with exactly n photons, represent the ultimate sub-Poissonian case for n ≥ 1, as their photon number variance is precisely zero, Δn = 0, independent of n. Common generation methods for sub-Poissonian light include resonant four-wave mixing in atomic vapors, such as sodium beams, where quantum nondemolition measurements of photon number suppress intensity noise. Another approach employs semiconductor lasers with electronic feedback to regularize the pump noise, achieving experimental Mandel's Q values as low as approximately -0.5, corresponding to noise reductions of several decibels below the . These techniques leverage intracavity correlations between photons and carriers to enforce sub-Poissonian output. A fundamental quantum inequality, rooted in the Heisenberg uncertainty principle for field quadratures, prohibits simultaneous squeezing in both amplitude and phase below the standard quantum limit. Specifically, if the amplitude quadrature variance is squeezed (ΔX < 1/√2 in normalized units), the phase quadrature must be anti-squeezed (ΔY > 1/√2) to satisfy ΔX ΔY ≥ 1/2, limiting the extent of sub-Poissonian without corresponding enhancement. The first demonstration of sub-Poissonian light occurred in , using a laser with feedback to suppress intensity noise below the shot-noise level, as reported by Machida and Yamamoto. This milestone highlighted the practical generation of non-classical light from solid-state sources, paving the way for applications in precision measurements.

Antibunching and correlations

Antibunching in photon statistics refers to a non-classical effect where the probability of detecting two photons at the same time is reduced compared to a coherent source, characterized by the second-order satisfying g^{(2)}(0) < 1. This indicates that photons arrive in a more regular, spaced-out manner rather than bunching randomly as in classical or thermal light. The second-order is defined as g^{(2)}(\tau) = \frac{\langle \hat{a}^\dagger(t) \hat{a}^\dagger(t+\tau) \hat{a}(t+\tau) \hat{a}(t) \rangle}{\langle \hat{a}^\dagger \hat{a} \rangle^2}, where \hat{a}^\dagger and \hat{a} are the creation and annihilation operators for the photon field, and the angle brackets denote the expectation value in the quantum state of the field. For single-photon sources, such as a Fock state |1\rangle, the derivation shows g^{(2)}(0) = 0, since the state has exactly one photon, making simultaneous detection of two photons impossible; this follows from \langle \hat{n}(\hat{n}-1) \rangle = 0 while \langle \hat{n} \rangle = 1, yielding g^{(2)}(0) = \langle \hat{n}(\hat{n}-1) \rangle / \langle \hat{n} \rangle^2 = 0. Prominent sources exhibiting antibunching include single-photon emitters such as semiconductor quantum dots coupled to microcavities and individual atoms trapped in optical cavities. In quantum dots, strong coupling to a pillar microcavity has demonstrated clear antibunching in resonance fluorescence, with g^{(2)}(0) values approaching zero under pulsed excitation. Similarly, resonance fluorescence from excited atoms, such as sodium, displays antibunching due to the quantum nature of the atomic transitions, where the atom cannot emit a second photon until it relaxes from the excited state. Higher-order antibunching extends this effect, where the n-th order correlation function satisfies g^{(n)}(0) < 1 for n > 2, particularly in s |m\rangle with m < n. For a single-photon |1\rangle, all higher-order correlations vanish, g^{(n)}(0) = 0 for n \geq 2, confirming perfect antibunching across multiple detection orders; in general, for a |m\rangle, g^{(n)}(0) = n! / [m(m-1)\cdots(m-n+1)] if m \geq n-1, and 0 otherwise, leading to antibunching when this value is less than 1. Experimentally, antibunching manifests as dips in the correlation histogram of arrival times, showing anticorrelation at short delays \tau \approx 0. This signature was first observed in the through resonance fluorescence experiments on a beam of sodium atoms, where measurements revealed g^{(2)}(0) \approx 0.2, confirming the quantum antibunching effect beyond classical predictions.

Measurement and applications

Detection methods

Direct photon counting techniques enable the measurement of photon number distributions P(n) by registering individual s over a defined integration time, constructing histograms that reveal statistical properties such as sub- or super-Poissonian behavior. These methods typically employ photomultiplier tubes (PMTs) for their high gain and low noise in counting individual photoelectrons, or single-photon avalanche diodes (SPADs) for their compact, solid-state operation with picosecond timing resolution. SPADs, biased above their breakdown voltage, produce a macroscopic current pulse upon single- absorption, allowing arrays like silicon photomultipliers (SiPMs) to resolve photon numbers up to tens per through summed outputs from multiple microcells. For instance, in experiments, SPAD-based setups have quantified P(n) for attenuated light, confirming Poissonian statistics with variances matching mean photon numbers. Time-resolved detection methods, particularly time-correlated single-photon counting (TCSPC), measure second-order functions g^{(2)}(\tau) to probe temporal photon bunching or antibunching. In TCSPC, a pulsed synchronizes with high-speed timing that record arrival times relative to the , building histograms of inter-photon delays after splitting the beam to two detectors via a beamsplitter. This Hanbury Brown-Twiss-inspired approach achieves sub-nanosecond resolution, enabling the identification of non-classical effects like g^{(2)}(0) < 1 in single-photon sources. Representative applications include characterizing quantum dots, where TCSPC reveals antibunching dips indicating single-photon emission. Balanced homodyne detection provides an indirect route to infer photon number statistics through quadrature measurements, reconstructing phase-space distributions like the Wigner function from which P(n) can be extracted. The setup mixes the signal field with a strong local oscillator on a 50:50 beamsplitter, followed by differential detection of the output quadratures X_\theta = (a e^{i\theta} + a^\dagger e^{-i\theta})/\sqrt{2}, where \theta is the phase. By scanning \theta and accumulating quadrature histograms, the Wigner function W(x,p) is tomographically reconstructed, allowing computation of P(n) = \int W(x,p) | \langle n | x,p \rangle |^2 dx dp. This method has demonstrated non-classical P(n) for squeezed states, with quadrature variances below the vacuum limit confirming sub-Poissonian traits. Accurate estimation of parameters like Mandel's Q requires meticulous calibration, including dead-time corrections to account for detector recovery periods that suppress counts at high fluxes and background subtraction to isolate signal from noise. Dead-time effects, modeled as non-paralyzable or paralyzable, distort variance measurements; corrections via pile-up models or reference calibrations with known Poissonian sources ensure Q = \langle (\Delta n)^2 \rangle / \langle n \rangle - 1 reflects true statistics, shifting artificially negative Q values to positive in simulations of single-photon emitters. Background subtraction involves dark-count histograms subtracted from signal data, scaled by acquisition times, to prevent overestimation of bunching in low-light regimes. These steps are critical for reliable Q in time-dependent analyses, where integration time T reveals crossover from sub- to super-Poissonian regimes. Modern advances in detection leverage superconducting single-photon detectors (SNSPDs), which have achieved near-unity detection efficiencies since the early 2000s, surpassing traditional SPADs in sensitivity and low dark counts. Introduced in 2001, SNSPDs operate by absorbing photons into a biased superconducting , creating a resistive that diverts current and generates a fast electrical with below 20 ps. Fiber-coupled efficiencies exceeded 90% by the through optimized nitride designs and cryogenic cooling, enabling precise P(n) histograms for weak fields with minimal afterpulsing. These detectors have facilitated high-fidelity photon statistics in , with system efficiencies approaching 98% at 1550 nm.

Practical implications

Understanding photon statistics has profound implications in quantum information science, particularly through the use of sub-Poissonian light sources that exhibit reduced photon number fluctuations below the limit, enhancing security in (QKD) protocols such as BB84. These sources minimize multi-photon emission events, which can be exploited by eavesdroppers, thereby improving key generation rates and tolerance to channel losses compared to classical Poissonian sources. For instance, sub-Poissonian light enables higher secure communication rates over long distances by suppressing intensity noise that limits protocol efficiency. In and sensing technologies, the distinct photon statistics of various sources enable specialized applications. Super-Poissonian thermal , characterized by bunching and higher variance in photon counts, is essential for ghost imaging schemes, where spatial correlations between correlated pairs allow reconstruction of object images without direct illumination, improving signal-to-noise ratios in low- conditions. Conversely, sub-Poissonian or squeezed reduces , facilitating low-noise systems that achieve Heisenberg-limited precision in range and velocity measurements for . Metrology benefits significantly from non-classical photon statistics, as squeezed light—exhibiting sub-Poissonian variance in one quadrature—enhances measurement precision beyond the standard quantum limit. In gravitational wave detectors like LIGO, the injection of squeezed vacuum states has improved broadband sensitivity since 2019, reducing shot noise and enabling the detection of fainter signals from cosmic events; further advancements, such as frequency-dependent squeezing implemented in 2023, have pushed sensitivities beyond previous quantum limits. Similarly, in optical atomic clocks, spin-squeezed states generated via cavity feedback or measurement-based techniques suppress phase noise, achieving stability improvements of up to 2 dB and pushing clocks toward applications in precision navigation and fundamental physics tests. Biological applications leverage single-photon statistics in fluorescence microscopy to resolve with high fidelity. By analyzing photon arrival times and number distributions, techniques like exploit antibunching and sub-Poissonian statistics from single emitters to distinguish individual fluorophores, enabling quantitative mapping of biomolecular interactions and without extensive . However, practical deployment faces scalability challenges, including efficient on-chip generation of non-classical light states using quantum dots or parametric processes, which suffer from low , mode mismatch, and complexities in compact devices. Recent progress in verifying quantum non-Gaussian photon statistics, as demonstrated in loss-resistant protocols by 2025, promises enhanced reliability for such applications in quantum sensing.

References

  1. [1]
    [PDF] Photon Counting Statistics of Classical and Quantum Light Sources
    Photon statistics is a discipline that lies at the heart of theoretical and experimental quan- tum optics. Results obtained in the general field theoretical ...
  2. [2]
    [PDF] Photon Statistics - KTH
    There would be 3 photons on average in 30 cm of a 1 nW light beam. Since photons are the smallest quantum of light and are discrete 10.33 mean photons does not ...
  3. [3]
    Coherent and Incoherent States of the Radiation Field | Phys. Rev.
    Methods are developed for discussing the photon statistics of arbitrary fields in fully quantum-mechanical terms. ... Glauber, Phys. Rev. Letters 10, 84 ...Missing: original | Show results with:original
  4. [4]
    [PDF] 1 Photon statistics at beam splitters
    The statistical behaviour of photons at beam splitters elucidates some of the most fundamental quantum phenomena, such as quantum superposition and randomness.
  5. [5]
    Quantum-optical description of photon statistics and cross ...
    Sep 7, 2021 · In quantum-optical experiments, the most relevant auto- and cross-correlation functions are between photon numbers. Usually, semiconductor ...
  6. [6]
    [PDF] Roy J. Glauber - Nobel Lecture
    The states α are fully coherent states of the field mode. The squared moduli of the coefficients of the states n in Eq. (28) tell us the probability for the ...Missing: original | Show results with:original
  7. [7]
    Coherent-State Representations for the Photon Density Operator
    The "diagonal" P representation of the photon density operator in terms of the coherent states of the radiation field is studied.
  8. [8]
    Photon statistics of light fields based on single-photon-counting ...
    Feb 11, 2005 · It shows that for some quantum states, such as the single-photon state and squeezed vacuum state, the statistical properties are strongly ...
  9. [9]
    [PDF] Photon Statistics - KTH
    We can calculate the probability distribution P for n photons in a given mode i. This is the characteristic Poisson statistics used to describe coherent light.<|control11|><|separator|>
  10. [10]
    Sub-Poissonian photon statistics in resonance fluorescence
    - **Abstract**: Expressions for the probability \( p(n) \) of \( n \) photons emitted in a steady state by a two-level atom in a resonant, coherent field are derived. The distribution is narrower than Poissonian, with the ratio \( [\langle(\Delta n)^2\rangle - \langle n\rangle]/\langle n\rangle \) being negative and peaking at an absolute maximum of 3/4. Sub-Poissonian statistics observation is briefly discussed.
  11. [11]
  12. [12]
    Hanbury Brown--Twiss Interferometry at a Free-Electron Laser
    We present measurements of second- and higher-order intensity correlation functions (so-called Hanbury Brown—Twiss experiment) performed at the free-electron ...
  13. [13]
    Photon-number-resolved detection of photon-subtracted thermal light
    We demonstrate experimentally that the photon number distribution transforms from a Bose–Einstein distribution to a Poisson distribution as the number of ...
  14. [14]
    Photon Antibunching in Resonance Fluorescence | Phys. Rev. Lett.
    Sep 12, 1977 · The phenomenon of antibunching of photoelectric counts has been observed in resonance fluorescence experiments in which sodium atoms are continuously excited ...Missing: first | Show results with:first
  15. [15]
    Nonquantum information gain from higher-order correlation functions
    May 8, 2020 · The k th -order correlation function for Fock states reads. g ( k ) ... higher-order correlation function, which is not bound by nonclassical ...
  16. [16]
    Photon Antibunching from a Single Quantum-Dot-Microcavity ...
    Mar 16, 2007 · We observe antibunching in the photons emitted from a strongly coupled single quantum dot and pillar microcavity in resonance.
  17. [17]
    [PDF] Measurement of photon correlations with multipixel photon counters
    We consider direct detection of multiphoton states by using specialized photon number resolving detectors (PNRD) – devices where the produced outcome ...
  18. [18]
    Single-photon avalanche diode imagers in biophotonics - Nature
    Sep 18, 2019 · Single-photon avalanche diode (SPAD) arrays are solid-state detectors that offer imaging capabilities at the level of individual photons.
  19. [19]
    Characterization of Silicon Photomultiplier Photon Detection ... - arXiv
    May 20, 2025 · A SiPM contains an array of Single Photon Avalanche Diodes (SPADs), each containing a quenching resistor, and is operated a few volts above the ...
  20. [20]
    Effective second-order correlation function and single-photon ... - arXiv
    Nov 16, 2017 · Any quantum state, for which the second-order correlation function falls below 1/2, has a nonzero projection on the single-photon Fock state.
  21. [21]
    readPTU - IOP Science
    Jun 27, 2019 · Time-correlated single-photon counting (TCSPC) experiments have found widespread applications in different disciplines, including the ...
  22. [22]
    [PDF] arXiv:1703.02786v1 [quant-ph] 8 Mar 2017
    Mar 8, 2017 · Balanced homodyne detection is applied to reconstruct the Wigner function, also yielding the state's photon number distribution. The heralding ...
  23. [23]
    Tuning between photon-number and quadrature measurements with ...
    Mar 20, 2020 · Here, we experimentally implement a weak-field homodyne detector that can continuously tune between measuring photon numbers and field quadratures.
  24. [24]
    Characterizing photon number statistics using conjugate optical ...
    We study the problem of determining the photon number statistics of an unknown quantum state using conjugate optical homodyne detection.
  25. [25]
    [PDF] Time-dependent Mandel Q parameter analysis for a hexagonal ...
    Abstract: The time-dependent Mandel Q parameter, Q(T), provides a measure of photon number variance for a light source as a function of integration time.
  26. [26]
    [PDF] Calibrating photon-counting detectors to high accuracy
    May 4, 2006 · We present a method for separating the correlated signal from the background signal that appro- priately handles deadtime effects. This method ...Missing: Mandel Q parameter
  27. [27]
    [PDF] A New Method for Dead Time Calibration and a New Expression for ...
    May 17, 2023 · Observed photon count rates must be corrected for detector dead time effects for accurate quantification, especially at high count rates. We.
  28. [28]
    Superconducting nanowire single-photon detectors: A perspective ...
    May 13, 2021 · Soon after these pioneering works, fiber-coupled SNSPDs reached a detection efficiency of 24% at 1550 nm (Ref. 35) and were further improved to ...Single-photon detection and... · SNSPD detection mechanisms · Future applications
  29. [29]
    Superconducting nanowire single-photon detectors - IOP Science
    Apr 4, 2012 · We give an overview of the evolution of SNSPD device design and the improvements in performance which have been achieved. We also evaluate ...
  30. [30]
    Saturating Intrinsic Detection Efficiency of Superconducting ...
    Oct 17, 2019 · A superconducting nanowire single-photon detector (SNSPD) has played a significant role in numerous applications for visible and near ...
  31. [31]
    Security aspects of quantum key distribution with sub-Poisson light
    Jul 29, 2025 · It was shown that sub-Poisson light sources offer significant improvements in communication rate over Poisson light. The amount of channel loss ...
  32. [32]
    Superior decoy state and purification quantum key distribution ...
    Sep 12, 2024 · Here, we propose and experimentally emulate two simple-to-implement protocols that allow practical, far from ideal sub-Poissonian photon sources ...
  33. [33]
    Role of photon statistics of light source in ghost imaging
    It is proved that, generally, Poissonian light cannot produce a ghost image. ... It is demonstrated that conventional thermal ghost imaging is related to the ...
  34. [34]
    An introduction to ghost imaging: quantum and classical - Journals
    Jun 26, 2017 · 'Ghost imaging' is often understood as imaging using light that has never physically interacted with the object to be imaged.
  35. [35]
    Heisenberg-Limited Quantum Lidar for Joint Range and Velocity ...
    We propose a quantum lidar protocol to jointly estimate the range and velocity of a target by illuminating it with a single beam of pulsed displaced squeezed ...Abstract · Article Text · Coherent state lidar · Displaced squeezed state...
  36. [36]
    Quantum-enhanced Doppler lidar | npj Quantum Information - Nature
    Dec 19, 2022 · We propose a quantum-enhanced lidar system to estimate a target's radial velocity, which employs squeezed and frequency-entangled signal and idler beams.
  37. [37]
    Enhanced sensitivity of the LIGO gravitational wave detector by ...
    Jul 21, 2013 · With the injection of squeezed states, this LIGO detector demonstrated the best broadband sensitivity to gravitational waves ever achieved, with ...
  38. [38]
    Broadband Quantum Enhancement of the LIGO Detectors with ...
    In their third observing run, LIGO and other gravitational-wave detectors employed frequency-independent squeezing, which reduced the shot noise from quantum ...Abstract · Popular Summary · Article Text
  39. [39]
    Prospects and challenges for squeezing-enhanced optical atomic ...
    Nov 24, 2020 · Here, we investigate the benefits of spin squeezed states for clocks operated with typical Brownian frequency noise-limited laser sources.
  40. [40]
    Fluorescence Microscopy: a statistics-optics perspective - PMC
    Apr 4, 2023 · Here we review the optics responsible for generating fluorescent images, fluorophore properties, microscopy modalities leveraging properties of both light and ...
  41. [41]
    Using photon statistics to boost microscopy resolution - PNAS
    These approaches yield exquisite images but are still rather complex to implement and oftentimes require a very photostable sample to collect enough photons.
  42. [42]
    Generation of Non‐Classical Light Using Semiconductor Quantum ...
    Dec 26, 2019 · Recent progress is discussed in the generation of non-classical light for off-chip applications as well as implementations for scalable on-chip ...
  43. [43]
    Recent developments in the generation of non-classical and ...
    Non-classical and entangled light states are of fundamental interest in quantum mechanics and they are a powerful tool for the emergence of new quantum ...