Fact-checked by Grok 2 weeks ago

Quantum limit

In physics, a quantum limit refers to a fundamental bound on the precision of measurements imposed by quantum mechanics, such as the standard quantum limit (SQL) and the Heisenberg limit. The SQL, commonly referred to as the quantum limit in many contexts, is a fundamental bound imposed by quantum mechanics on the precision of continuous measurements of physical observables such as position, displacement, or force, stemming from the Heisenberg uncertainty principle. This limit arises from the unavoidable trade-off between measurement imprecision noise—which degrades the signal—and quantum back-action noise, where the act of measurement disturbs the system's momentum, setting a minimum total uncertainty achievable with classical probe states like coherent light. In mathematical terms, for a free-mass position measurement over time \tau, the SQL is expressed as \Delta x \geq \sqrt{\frac{\hbar \tau}{2m}}, where \hbar is the reduced Planck's constant and m is the mass, representing the optimal balance of these noise sources. The SQL was originally derived in the context of quantum nondemolition measurements for detection by Vladimir B. Braginsky in 1967, who highlighted its implications for sensitive mechanical systems, and was further formalized by Carlton M. Caves in 1980 through an analysis of in linear amplifiers and optical interferometers. It applies broadly to quantum-limited technologies, including laser interferometers for detecting , optomechanical sensors, and , where dominates classical thermal noise. For instance, in 's detectors, the SQL manifests as a frequency-dependent in strain measurements, limiting sensitivity to mergers of compact objects at certain frequencies until advanced quantum techniques are employed. Although the SQL represents a practical barrier for classical measurement schemes, it is not an absolute limit and can be surpassed using non-classical resources such as , which reduce uncertainty in one at the expense of the other, or multipartite entanglement to distribute noise across multiple probes—approaching the Heisenberg limit. Experimental demonstrations include optomechanical systems achieving 1.5 dB below the SQL in continuous force sensing by exploiting quantum correlations, and 's Advanced LIGO observatories reaching sensitivities beyond the SQL in 2023 through frequency-dependent squeezing (with up to 2.8 dB improvement detailed as of 2024), enabling enhanced detection of astrophysical signals. These advancements underscore the SQL's role as a for quantum-enhanced , driving progress in fields from fundamental physics to .

Fundamentals

Definition

The quantum limit encompasses the fundamental constraints imposed by on the precision of measurements for physical quantities such as , , and phase at quantum scales. These bounds arise from the wave-particle duality of and the inherent in measurement processes, including measurement back-action, where the act of observing a system inevitably disturbs it due to the non-commutative nature of quantum observables. The concept originated with Werner Heisenberg's 1927 paper introducing the , which established the theoretical foundation for these measurement limits by demonstrating the impossibility of simultaneously knowing certain pairs of properties with arbitrary precision. The specific terminology of "quantum limit" gained prominence in quantum measurement theory during the mid-20th century, particularly through early analyses of in quantum amplifiers and detectors, such as the work by and Mullen in 1962. Mathematically, quantum limits are formalized as inequalities derived from the commutation relations of operators, with the canonical example being the position-momentum uncertainty relation: \Delta x \, \Delta p \geq \hbar/2, which quantifies the minimal product of standard deviations for conjugate variables. These bounds distinguish between absolute quantum limits, which are intrinsic and fundamental like the Heisenberg limit, and conditional ones, such as the standard quantum limit, that depend on the specifics of the measurement apparatus and can potentially be overcome with optimized quantum resources.

Relation to Uncertainty Principle

The Heisenberg uncertainty principle establishes a fundamental limit on the simultaneous precision of measurements for conjugate observables in quantum mechanics. For position x and momentum p, it states that the product of their standard deviations satisfies \Delta x \Delta p \geq \frac{\hbar}{2}, where \hbar = h / 2\pi and h is Planck's constant. This inequality was first rigorously derived by Earle Hesse Kennard in 1927 using the formalism of wave mechanics, quantifying the spreads in and for a quantum state. Werner introduced the principle intuitively earlier that year in his seminal paper, emphasizing the conceptual impossibility of precisely knowing both observables at once through thought experiments involving microscopic measurements. The principle generalizes to other pairs of conjugate variables. For energy E and time t, the relation \Delta E \Delta t \geq \frac{\hbar}{2} captures limits on the lifetime of quantum states or the duration of processes, though time is not a true operator in standard quantum mechanics. Similarly, for angular position \theta and angular momentum L_z, the inequality \Delta \theta \Delta L_z \geq \frac{\hbar}{2} applies, reflecting periodic boundary conditions in angular variables. A more general form, known as the Robertson uncertainty relation, was established by Howard Percy Robertson in 1929 for any pair of non-commuting Hermitian operators A and B: \Delta A \Delta B \geq \frac{1}{2} \left| \langle [A, B] \rangle \right|, where [A, B] = AB - BA is the commutator and the expectation value is taken over the quantum state. This relation derives from the non-commutativity of quantum operators, such as [x, p] = i \hbar for position and momentum, which prevents simultaneous eigenstates and introduces inherent fluctuations. To outline the derivation, consider the variance \Delta A^2 = \langle (A - \langle A \rangle)^2 \rangle. Using the Cauchy-Schwarz inequality on the state vectors and incorporating the commutator, one obtains the Robertson bound, showing that non-zero commutators enforce minimal uncertainty products. For the canonical pair, this yields the familiar \frac{\hbar}{2} limit. In quantum measurements, the uncertainty principle manifests as back-action noise: precisely measuring one observable disturbs the conjugate one due to the unavoidable coupling via the , leading to trade-offs in accuracy. For instance, a measurement imparts random kicks, broadening the distribution and setting irreducible error floors in repeated or continuous observations. These effects underpin quantum limits as the practical bounds arising from such fundamental indeterminacies.

Types

Standard Quantum Limit

The Standard Quantum Limit (SQL) is a bound on the precision of quantum measurements achieved using classical resources like coherent states, particularly in position or displacement sensing with uncorrelated probes. In such measurements, the imprecision noise scales with the square root of the number of resources, leading to a total sensitivity that improves only as 1/√N, where N is the number of photons or particles employed. This limit stems from the Heisenberg uncertainty principle, which enforces a trade-off between measurement imprecision and quantum back-action in interferometric setups. It applies to classical measurement schemes without entanglement. The SQL was originally derived by Vladimir B. Braginsky in 1967 and further formalized by Carlton M. Caves in 1980, who examined quantum-mechanical radiation-pressure fluctuations in interferometers designed for detection, highlighting the role of from and back-action from momentum transfer. For the position measurement of a , the SQL arises from the balance between these shot-noise and back-action contributions, yielding a minimum position uncertainty of Δx_SQL = √(ℏ / (2 m ω)), where ℏ is the reduced , m is the oscillator mass, and ω is its . This expression equals the zero-point fluctuation amplitude of the oscillator, with the total measurement noise at the SQL equivalent to twice the zero-point level due to equal contributions from imprecision and back-action. In practice, the SQL acts as a for measurements with uncorrelated coherent resources, rather than an absolute limit. It has been a key constraint in detectors like the initial Laser Interferometer (LIGO) configurations before the implementation of quantum squeezing techniques. Similarly, in atomic interferometers, the SQL sets the precision floor for phase or displacement estimates using independent atoms, scaling as 1/√N and guiding the design of high-sensitivity sensors.

Heisenberg Limit

The Heisenberg limit (HL) constitutes the ultimate fundamental bound on the precision of parameter estimation in quantum metrology, achievable only through the optimal use of quantum resources such as . It emerges directly from the saturation of quantum uncertainty relations, such as the , and provides a scaling of precision inversely proportional to the total number of independent quantum probes N, i.e., \Delta \theta \propto 1/N. This contrasts with suboptimal classical strategies, which are constrained to a weaker $1/\sqrt{N} scaling, highlighting the HL as the theoretical ceiling for quantum-enhanced sensing in any parameter estimation protocol. The derivation of the relies on the quantum Cramér-Rao bound, which connects the variance of an unbiased to the \mathcal{F}_Q of the probe state via \Delta \theta \geq 1/\sqrt{\nu \mathcal{F}_Q}, where \nu is the number of repetitions. In quantum metrology, the maximum \mathcal{F}_Q scales as N^2 for entangled resources, yielding the precision \Delta \theta_{HL} \geq 1/N. For the specific case of phase estimation in , this manifests as \Delta \phi_{HL} \geq \frac{1}{N}, demonstrating the quadratic improvement over classical limits when quantum correlations are fully exploited. Attaining the HL demands non-classical quantum states, including entangled probes that distribute resources across multiple particles or specific superpositions like NOON states |\text{NOON}\rangle = \frac{1}{\sqrt{2}} (|N0\rangle + |0N\rangle), which maximize the Fisher information. These conditions ensure the bound is saturated, rendering the HL an absolute limit impervious to further enhancement without modifications to underlying quantum mechanics. The was formalized as a central in quantum metrology through seminal works by Giovannetti, , and Maccone, who in 2004 outlined strategies to surpass classical bounds using and in 2006 established a rigorous confirming its optimality across diverse protocols. These contributions positioned the HL as the aspirational target for ultimate in fields reliant on precise measurements.

Applications

Precision Measurements

In gravitational wave detection, the standard quantum limit (SQL) imposes a fundamental bound on the precision of interferometric measurements, arising from the balance between and radiation-pressure backaction. The Laser Interferometer Gravitational-Wave Observatory () encountered this limit during its initial Advanced LIGO observing run (O1) in 2015–2016, where contributions approached the SQL, particularly at frequencies around 10 Hz with a of approximately 5 × 10^{-19} m/√Hz. This constraint highlighted the SQL as a key barrier to enhancing for detecting weaker astrophysical signals, such as those from distant mergers. By 2023, had surpassed the SQL in operational , underscoring the limit's role as a benchmark for quantum-limited performance in large-scale interferometers. Atomic and optical interferometers similarly confront the SQL in measuring inertial forces like accelerations and rotations, which are essential for applications in precision , geophysical surveying, and fundamental tests of . In light-pulse interferometers, the SQL manifests as quantum , limiting the to the of the number and capping measurements at levels around 10^{-10} m/s² for typical times of seconds. For rotation sensing, cold- gyroscopes achieve angular sensitivities near the SQL, enabling drift rates below 10^{-8} rad/s, far surpassing classical devices and supporting GPS-denied in or . These systems leverage matter-wave interference to probe weak fields, but the SQL sets the ultimate precision without advanced quantum correlations. In optomechanical displacement measurements, the SQL restricts the accuracy of determining mirror or resonator positions to the scale of the zero-point fluctuation, on the order of 10^{-12} m for typical microscale oscillators, where imprecision noise equals backaction-induced fluctuations. This bound is particularly relevant in cavity-based sensors, where continuous position monitoring of a mechanical element—such as a suspended mirror—encounters the SQL after integrating measurements equivalent to one zero-point motion. Experimental milestones include early demonstrations in microwave cavities during the 1980s, where quantum backaction effects were first observed in macroscopic systems, confirming the SQL's influence on mechanical readout precision. More recent optomechanical setups have routinely reached this limit, establishing it as a critical threshold for force and position sensing at atomic scales.

Quantum Optics

In , the standard quantum limit (SQL) arises primarily from photon shot noise, which imposes fundamental constraints on in interferometric measurements. This originates from the discrete, probabilistic nature of arrivals, leading to a phase sensitivity scaling as the inverse of the average number of photons, \sqrt{N}, where N is the total photon count. Such limits are critical in applications like , where precision in resolving transitions is hindered by this shot-noise floor, and in optical imaging, such as in , where it restricts the ability to discern fine spatial details beyond a certain . Amplitude squeezing addresses another key quantum limit in optical systems, bounding the uncertainty in quadrature measurements of the electromagnetic field. The vacuum state sets a minimum noise level for the amplitude quadrature, analogous to the Heisenberg uncertainty principle for field quadratures, where reducing noise in one quadrature increases it in the orthogonal phase quadrature. This bound is particularly relevant for homodyne detection schemes, which compare the signal field to a local oscillator to measure amplitude fluctuations with enhanced sensitivity, enabling applications in low-light detection and noise reduction in optical communications. Squeezed states can approach but not surpass these limits without additional resources. In (QKD) protocols, such as , quantum limits manifest through vacuum fluctuations, which introduce unavoidable error rates in photon transmission over optical channels. These fluctuations contribute to the quantum bit error rate (QBER), typically on the order of a few percent due to detector dark counts and channel losses, setting a for secure ; exceeding this limit compromises detection. The SQL here dictates that error correction and privacy amplification must account for shot-noise-induced uncertainties to maintain security. Historically, early explorations of these limits in trace back to H. P. Yuen's 1976 work on squeezed states, which demonstrated how coherent states approach the SQL in phase and amplitude measurements while highlighting pathways to mitigate noise through state engineering. This foundational analysis underscored the role of quantum limits in constraining optical information processing, influencing subsequent developments in .

Advanced Techniques

Quantum Squeezing

Quantum squeezing refers to the generation of quantum states of light in which the noise or uncertainty in one quadrature component—such as the amplitude or phase quadrature—is reduced below the level of vacuum fluctuations, while the noise in the conjugate quadrature increases to maintain the Heisenberg uncertainty principle, ensuring the product of the variances satisfies ΔX ΔP ≥ ħ/2. This reduction allows for surpassing the standard quantum limit in precision measurements by tailoring the quantum noise distribution. Squeezed states are typically produced through nonlinear optical processes that introduce correlations between field modes. In nonlinear optics, spontaneous parametric down-conversion in χ^(2) nonlinear crystals pumps a high-frequency photon into lower-frequency signal and idler photons, generating squeezed vacuum states. Atomic ensembles achieve squeezing via four-wave mixing, where atomic vapors interact with laser fields to produce correlated photon pairs. Optomechanical systems generate mechanical or optical squeezing through radiation pressure coupling between light and mechanical resonators, enabling noise reduction in hybrid quantum systems. The mathematical description of squeezing employs the squeezing operator acting on the state: |\xi\rangle = S(\xi) |0\rangle = \exp\left[\frac{1}{2} \left( \xi^* a^2 - \xi (a^\dagger)^2 \right) \right] |0\rangle, where \xi = r e^{i\theta} is the complex squeezing parameter, r quantifies the degree of squeezing, \theta sets the squeezing angle, and a (a^\dagger) is the () . For the squeezed quadrature aligned with \theta, the variance becomes \Delta X_\theta^2 = \frac{1}{4} e^{-2r}, reduced below the variance of \frac{1}{4} (in with \hbar = [1](/page/1)), while the anti-squeezed quadrature variance is \Delta X_{\theta + \pi/2}^2 = \frac{1}{4} e^{2r}. The squeezing level is often expressed in decibels as $10 \log_{10} (e^{-2r}) dB, with negative values indicating noise reduction. The first experimental observation of squeezing was reported in 1985 by Slusher and Walls, who used nondegenerate in a sodium atomic vapor within an to achieve approximately 0.6 dB of squeezing. In modern applications, such as the Laser Interferometer Gravitational-Wave Observatory (), frequency-dependent squeezing has been employed to inject squeezed vacuum into the interferometer, reducing quantum noise and surpassing the standard quantum limit by up to 3 dB in the 35–75 Hz band as of 2023. This approach enhances sensitivity to by minimizing noise in the phase quadrature, thereby allowing measurements that approach the Heisenberg limit along a single dimension.

Entanglement Methods

Entanglement serves as a key resource in quantum metrology by enabling correlated measurements across multiple probes, allowing the collective phase accumulation to enhance sensitivity beyond the standard quantum limit (SQL). In multi-probe scenarios, such as , entangled states distribute quantum correlations that amplify the signal while suppressing uncorrelated noise, achieving phase estimation precisions that scale inversely with the number of particles N, approaching the Heisenberg limit (HL). A prominent example is the , defined as \frac{1}{\sqrt{2}} (|N0\rangle + |0N\rangle), where all N photons are either in one mode or the other in superposition. This state facilitates collective interferometric measurements, yielding a phase sensitivity of \Delta \phi \sim 1/N, which surpasses the SQL's \Delta \phi \sim 1/\sqrt{N} by leveraging the path-entangled superposition. Among techniques employing entanglement, Greenberger-Horne-Zeilinger (GHZ) states are widely used in Ramsey interferometry for quantum sensing applications like atomic clocks and gravitational wave detection. GHZ states, of the form \frac{1}{\sqrt{2}} (|00\dots0\rangle + |11\dots1\rangle), enable all probes to accumulate phase coherently, providing quadratic enhancement in precision for frequency or phase estimation. Cluster states, another class of multipartite entangled resources, support distributed sensing protocols where spatially separated nodes share correlations for estimating global parameters, such as in network-based metrology. These states allow flexible measurement-based schemes, adapting to varying sensor geometries without requiring full state reconfiguration. Theoretically, entanglement enhances through the (QFI), which quantifies the maximum extractable information about a from a . For entangled resources involving probes, the QFI scales as F \sim [N](/page/N+)^2, enabling the Cramér-Rao bound to reach \Delta \theta \sim 1/[N](/page/N+), the , whereas separable states limit it to F \sim [N](/page/N+) and \Delta \theta \sim 1/\sqrt{[N](/page/N+)}. This scaling arises from the multipartite correlations that amplify sensitivity in the . Experimental achievements include entanglement-enhanced magnetometry using spin-squeezed states derived from entangled ensembles, demonstrating a 15% improvement in sensitivity over the SQL using a spin-squeezed state implying entanglement of approximately 170 atoms, as reported in a (preprint ). In optical setups during the 2010s, progress toward the featured NOON-state achieving precisions within 4% of the exact HL for N=3 photons in , using adaptive protocols. Additionally, GHZ states with up to 8 photons in 2011 optical experiments validated the N^2 QFI scaling in tasks. More recent progress includes entanglement-enhanced sensing with neutral atom arrays achieving metrological gains beyond the SQL in 2024 experiments.

Comparison to Classical Limits

Key Differences

In classical physics, measurement precision is constrained primarily by statistical fluctuations, such as shot noise arising from the Poissonian statistics of particle counts or signal averaging, which scales inversely with the square root of the number of independent trials N (i.e., \sim 1/\sqrt{N}). Unlike quantum measurements, classical processes involve no fundamental back-action, where the act of observation disturbs the system; instead, they are deterministic in principle, with noise reducible indefinitely by increasing resources without encountering irreducible barriers. Quantum mechanics introduces additional, irreducible noise sources absent in classical descriptions, including measurement back-action from wavefunction collapse upon observation and manifested as fluctuations. These effects stem from the Heisenberg uncertainty principle and the non-commutativity of quantum observables, imposing fundamental limits on simultaneous knowledge of like and . For instance, fluctuations represent quantum zero-point motion in the , producing noise in fields or oscillators that has no classical analog, as classical states lack such inherent fluctuations. A key distinction lies in : classical precision can improve as $1/\sqrt{N} without bound by optimizing strategies or resources, whereas quantum measurements face a hard cap at the Heisenberg limit, scaling as $1/N, beyond which no further enhancement is possible due to these irreducible noises. The standard quantum limit serves as a quantum-specific bound analogous to the classical shot-noise scaling for uncorrelated probes but incorporates back-action effects. Philosophically, quantum limits preclude perfect, simultaneous knowledge of a system's state, enforcing complementarity between measurement precision and disturbance, yet they enable breakthroughs like sub-standard quantum limit sensing through quantum correlations, transforming fundamental constraints into tools for enhanced metrology.

Misconceptions

A common misconception portrays quantum limits as artifacts confined to the classical regime, where the reduced Planck constant ħ approaches zero, or solely to microscopic scales, thereby dismissing their relevance in large-scale systems. In reality, these limits arise from inherent quantum fluctuations and persist in macroscopic apparatuses, such as the Laser Interferometer Gravitational-Wave Observatory (LIGO), where quantum shot noise and radiation pressure noise degrade sensitivity despite the detectors' kilometer-scale arms and involvement of kilogram-mass mirrors. For instance, LIGO's measurement of spacetime distortions—on the order of 10^{-19} meters—remains constrained by quantum uncertainty in photon arrival times and mirror displacements, demonstrating that quantum effects are not negligible even in human-engineered, macroscopic environments. Another frequent misunderstanding equates the Standard Quantum Limit (SQL) with an absolute, unbreakable ceiling on measurement precision, overlooking its status as a technical rather than fundamental constraint. The SQL, which scales as the inverse of the number of probes, emerges from uncorrelated, classical-like resources and can be exceeded through quantum enhancements like entanglement or squeezing, as demonstrated in various interferometric setups. By contrast, the (HL), scaling inversely with the number of probes, embodies a truly fundamental bound dictated by the , beyond which no quantum strategy can venture without additional resources. This distinction clarifies that while the SQL represents an achievable baseline with conventional methods, it is surmountable, whereas the HL sets the ultimate theoretical frontier. Historically, prior to the , quantum in precision measurements was often conflated with classical thermal noise, leading to underappreciation of distinct quantum contributions like back-action effects in interferometers. Early analyses treated fluctuations in optical and mechanical systems as primarily thermal in origin, without fully accounting for the irreducible quantum or arising from the wave-particle duality of light. This oversight persisted until seminal work explicitly separated these noise sources, revealing quantum as a separate, non-classical phenomenon that imposes unique limits independent of temperature. In modern quantum since the early , quantum limits have been reframed not merely as prohibitive barriers but as exploitable resources for surpassing classical performance in sensing applications. Advances in entanglement-based protocols and non-classical states have shown that correlations inherent to can convert these limits into tools for quadratic precision gains, as seen in atomic ensemble experiments and optical networks. For example, as of 2024, the Livingston detector reduced below the SQL by up to 3 decibels between 35 and 75 Hz using frequency-dependent squeezing.

References

  1. [1]
    Continuous force and displacement measurement below ... - Nature
    May 27, 2019 · In the case of interferometric displacement measurements, these restrictions impose a standard quantum limit (SQL), which optimally balances the ...
  2. [2]
    Standard Quantum Limit - RP Photonics
    The standard quantum limit is a limit for noise levels set by quantum mechanics. With certain techniques, noise below the standard quantum limit can be ...Missing: physics | Show results with:physics
  3. [3]
    Squeezing the quantum noise of a gravitational-wave detector ...
    Sep 19, 2024 · The minimal repeatable uncertainty is achieved when Δ x = Δ x ′ = ℏ τ / 2 m , which is known as the standard quantum limit (SQL) (2, 3). The SQL ...
  4. [4]
    LIGO Surpasses the Quantum Limit | LIGO Lab | Caltech
    Oct 23, 2023 · "Now that we have surpassed this quantum limit, we can do a lot more astronomy," explains Lee McCuller, assistant professor of physics at ...
  5. [5]
    Standard Quantum Limit (SQL) - AEI 10 m Prototype
    This is the Standard Quantum Limit (SQL), which describes the minimum level of total quantum noise in a certain interferometer at each frequency.Missing: definition | Show results with:definition
  6. [6]
    Surpassing the Standard Quantum Limit Using an Optical Spring
    Sep 13, 2024 · The standard quantum limit (SQL) is a theoretical limit imposed on precision measurements by the Heisenberg uncertainty principle [1, 2] . While ...Abstract · Article Text · Introduction · Results and discussion
  7. [7]
    None
    Below is a merged summary of the quantum limits in measurements based on the provided segments from arXiv:0810.4729. To retain all information in a dense and organized manner, I will use a combination of narrative text and a table in CSV format for key details. The narrative provides an overview and context, while the table captures specific definitions, mathematical forms, historical references, and distinctions between absolute and technical limits across the segments.
  8. [8]
    Quantum Noise in Linear Amplifiers | Phys. Rev.
    H. A. Haus, Internal Memorandum, Massachusetts Institute of Technology Research Laboratory of Electronics, 1962 (unpublished); H. A. Haus and R. B. Adler ...
  9. [9]
    Quantum limits on noise in linear amplifiers | Phys. Rev. D
    Oct 15, 1982 · How much noise does quantum mechanics require a linear ... H. A. Haus and J. A. Mullen, Phys. Rev. 128, 2407 (1962); H. Takahasi ...
  10. [10]
    The Uncertainty Principle (Stanford Encyclopedia of Philosophy)
    Oct 8, 2001 · The uncertainty principle (for position and momentum) states that one cannot assign exact simultaneous values to the position and momentum of a physical system.
  11. [11]
    Uncertainty Principle for Non-Commuting Operators
    Let us now derive the uncertainty relation for non-commuting operators \bgroup\color{black}$A$\egroup and \bgroup\color{black}$B$\egroup.
  12. [12]
    [PDF] Quantum Noise and Quantum Measurement - Aashish Clerk
    How small can we make the added noise? Two parts to the noise: Quantum Limit. • If our detector has a “large” gain ...
  13. [13]
    Quantum-Mechanical Radiation-Pressure Fluctuations in an ...
    The interferometers now being developed to detect gravitational vaves work by measuring small changes in the positions of free masses.Missing: noise | Show results with:noise
  14. [14]
    [2404.14569] Squeezing the quantum noise of a gravitational-wave ...
    Apr 22, 2024 · The LIGO A+ upgrade reduced quantum noise below the Standard Quantum Limit (SQL) by up to 3 dB, which is a limitation from quantum mechanics.
  15. [15]
    Differential atom interferometry beyond the standard quantum limit
    Jan 18, 2006 · We analyze methods designed to go beyond the standard quantum limit for a class of atomic interferometers, where the quantity of interest is ...Ii. Coherent Input States · C. Qnd Output Measurement · V. Decoherence<|control11|><|separator|>
  16. [16]
    Quantum-Enhanced Measurements: Beating the Standard ... - Science
    Nov 19, 2004 · Conventional bounds to the precision of measurements such as the shot noise limit or the standard quantum limit are not as fundamental as the Heisenberg limits.
  17. [17]
    Quantum Metrology | Phys. Rev. Lett. - Physical Review Link Manager
    Our analysis clarifies that, indeed, the Heisenberg limit is the bound to interferometric precision. Our bound also applies to quantum phase-estimation ...
  18. [18]
    None
    Nothing is retrieved...<|control11|><|separator|>
  19. [19]
    Advances in Atom Interferometry and their Impacts on the ...
    Oct 1, 2024 · Detection noise has two components: Quantum projection noise (QPN), which is the quantum standard limit in quantum inertial sensors and clocks, ...Advances In Atom... · 2. Modeling · 3. Sensitivity Analysis
  20. [20]
    Developments for quantum inertial navigation systems employing ...
    Jul 1, 2025 · Atom interferometry (AI) can therefore be a versatile tool for the measurement of inertial quantities, such as accelerations or rotations, ...Iv. Hybridization With... · B. Compact Vacuum... · Data Availability
  21. [21]
    Enhancing the sensitivity of atom-interferometric inertial sensors ...
    Nov 22, 2023 · Quantum sensors based on light-pulse atom interferometry have performed inertial measurements of unparalleled sensitivity and stability in ...
  22. [22]
    Optomechanics - Physics Magazine
    May 18, 2009 · The standard quantum limit is reached when both effects are equally strong. It corresponds to resolving the mirror's position to within its ...
  23. [23]
    Quantum limits for stationary force sensing | Phys. Rev. A
    Apr 26, 2021 · In particular, the sensitivity of the best optomechanical devices has reached the standard quantum limit (SQL), which directly follows from the ...Abstract · Article Text · LINEAR MEASUREMENT... · QUANTUM SENSITIVITY...
  24. [24]
    [PDF] A short walk through quantum optomechanics - arXiv
    Oct 12, 2012 · atomic ensemble [81], the first observation of quantum back-action on a 'macroscopic' mechanical resonator at the standard quantum limit.
  25. [25]
    Quantum-Enhanced Sensing with Squeezed Light - MDPI
    A quadrature-squeezed state is defined as a state where the variance of one quadrature component is reduced below the coherent-state level (i.e., below the SQL ...
  26. [26]
    Structured light analogy of quantum squeezed states - Nature
    Oct 21, 2024 · The quantum squeezed vacuum state is routinely generated in the spontaneous parametric down-conversion (SPDC) process, taking place in a non- ...Missing: seminal | Show results with:seminal
  27. [27]
    Observation of Squeezed States Generated by Four-Wave Mixing in ...
    Nov 25, 1985 · Squeezed states of the electromagnetic field have been generated by nondegenerate four-wave mixing due to Na atoms in an optical cavity.
  28. [28]
    [PDF] Note for Quantum Optics: Squeezed states
    The degree of squeezing is determined by r = |ξ| which is called the squeezed parameter.
  29. [29]
    Quantum Optical Metrology -- The Lowdown on High-N00N States
    Apr 1, 2009 · We will review some of the recent theoretical and experimental advances in this exciting new field of quantum optical metrology.
  30. [30]
    Graph States as a Resource for Quantum Metrology | Phys. Rev. Lett.
    Mar 17, 2020 · In this Letter we investigate the practicality of graph states for quantum metrology. Graph states are a natural resource for much of quantum information.
  31. [31]
    Experimental optical phase measurement approaching the exact ...
    Nov 2, 2018 · Our results are very close to the Heisenberg limit for N = 3, giving substantial experimental justification to the theoretical prediction that ...
  32. [32]
    Entanglement-enhanced quantum metrology: From standard ...
    Jul 2, 2024 · This article aims to review and illustrate the fundamental principles and experimental progresses that demonstrate multi-particle entanglement for quantum ...<|control11|><|separator|>
  33. [33]
    beating the standard quantum limit - arXiv
    Dec 10, 2004 · Abstract: Quantum mechanics, through the Heisenberg uncertainty principle, imposes limits to the precision of measurement.Missing: origin | Show results with:origin
  34. [34]
    Quantum measurements of sums | Phys. Rev. A
    Jul 13, 2020 · While in classical systems measurements can, in principle, be performed with unlimited precision and without any influence on the measured ...Article Text · I. Introduction · Vii. Example: Qubit
  35. [35]
    Coherence and multimode correlations from vacuum fluctuations in ...
    Aug 26, 2016 · The existence of vacuum fluctuations is one of the most important predictions of modern quantum field theory. In the vacuum state, ...
  36. [36]
  37. [37]
    What limits limits? - PMC - PubMed Central - NIH
    However, the standard quantum limit is not a fundamental limit. It can be viewed as a technical limit of a system measured in the coherent state. After ...
  38. [38]
    Quantum-mechanical noise in an interferometer | Phys. Rev. D
    Apr 15, 1981 · This paper presents an analysis of the two types of quantum-mechanical noise, and it proposes a new technique—the "squeezed-state" technique ...
  39. [39]
    Introduction to quantum noise, measurement, and amplification
    Apr 15, 2010 · This review gives a pedagogical introduction to the physics of quantum noise and its connections to quantum measurement and quantum amplification.<|control11|><|separator|>
  40. [40]
    Quantum metrology with nonclassical states of atomic ensembles
    Sep 5, 2018 · The goal of this article is to review and illustrate the theory and the experiments with atomic ensembles that have demonstrated many-particle entanglement.Missing: post- | Show results with:post-
  41. [41]
    Flexible resources for quantum metrology - IOPscience
    We show that for a wide range of tasks in metrology, 2D cluster states (a particular family of states useful for measurement-based quantum computation) can ...