Fact-checked by Grok 2 weeks ago

Phase noise

Phase noise is the random fluctuation in the phase of a signal generated by an oscillator or source, resulting in time-dependent deviations from the ideal periodic waveform and manifesting as sidebands close to the carrier . These fluctuations are primarily due to inherent mechanisms in electronic devices, such as thermal , (1/f) , and upconversion effects from the oscillator's time-varying circuit elements. In practical terms, phase noise is quantified using the single-sideband (SSB) phase noise spectral density, expressed in units of dBc/Hz (decibels relative to the carrier per hertz), which measures the noise power in a 1 Hz bandwidth at an offset frequency from the carrier. This metric is essential because phase noise dominates over amplitude noise in most oscillators due to inherent amplitude limiting mechanisms, making it the primary indicator of signal quality. Phase noise significantly impacts high-performance systems, including wireless communications where it reduces spectral efficiency and increases bit error rates by interfering with adjacent channels, radar systems where it degrades Doppler resolution, and digital circuits where it translates to timing jitter that erodes setup and hold margins. Measurement techniques, such as the phase detector method using a double-balanced mixer or cross-correlation interferometry, enable precise characterization down to levels below the thermal noise floor, guiding oscillator design optimizations like waveform symmetry to suppress low-frequency noise upconversion.

Fundamentals

Definition and Basic Concepts

Phase noise is the frequency-domain representation of short-term random phase fluctuations in a , arising from instabilities in the of an electronic signal such as that generated by an oscillator. These fluctuations manifest as a broadening of the signal's around the frequency and are typically quantified as the single-sideband () phase noise , expressed in units of /Hz at a specific frequency from the . The phase noise, denoted as L(f), measures the in a 1 Hz bandwidth on one side of the relative to the power, assuming the noise is symmetrically distributed across both sidebands. Unlike amplitude noise, which perturbs the signal's magnitude through random variations in its envelope, phase noise specifically affects the timing and angular position of the without altering its peak strength. This distinction is crucial in applications where phase integrity is paramount, as amplitude noise can often be filtered or compensated more readily than phase perturbations. The concept of phase noise emerged in the amid early studies of oscillator stability, with David B. Leeson's seminal 1966 model providing a foundational framework for understanding noise in feedback oscillators by linking noise to spectral broadening. In communication systems, phase noise degrades the (SNR) by introducing spurious sidebands that interfere with adjacent channels or demodulated data, particularly in RF transmitters where carrier signals must maintain precise phase coherence for modulation schemes like . For instance, elevated phase noise can mask weak signals or increase bit error rates in receivers, limiting overall system performance. The of phase fluctuations, S_\phi(f), quantifies these random variations and is derived from the function of the R_\phi(\tau) via the Wiener-Khinchin theorem, which relates time-domain statistics to frequency-domain power distribution for stationary processes: S_\phi(f) = \int_{-\infty}^{\infty} R_\phi(\tau) \cos(2\pi f \tau) \, d\tau where R_\phi(\tau) = \mathbb{E}[\phi(t) \phi(t + \tau)] is the of the product at lag \tau. This formulation underscores phase noise's role as a , with S_\phi(f) typically expressed in rad²/Hz to capture the variance per unit bandwidth.

Sources and Causes

Phase noise in electronic oscillators originates from fundamental noise processes inherent to the components, particularly active devices such as transistors and varactors. The primary intrinsic sources include thermal noise, , and (1/f) noise. Thermal noise, arising from the random thermal agitation of charge carriers in resistive elements, generates a white noise spectrum that scales with and . stems from the quantized, discontinuous flow of charge carriers across potential barriers, such as in junctions, contributing a similar white noise profile. , with its characteristic 1/f , predominates at low frequencies and is linked to imperfections in device materials and surfaces, often more pronounced in active . These noises are upconverted to the radio-frequency through nonlinearities in the oscillator's active elements, where the time-varying signals mix low-frequency sidebands into proximity of the . The resonator's quality factor () significantly influences this process; a higher enhances the selectivity of the , amplifying the desired oscillation while attenuating contributions, particularly for offsets near the . In quartz crystal oscillators, specifically arises from surface defects and interactions at the electrode-crystal interface, which introduce fluctuations that degrade close-in phase performance. Conversely, in LC oscillators, losses—primarily from series resistances in inductors and capacitors—dominate phase noise generation, as these resistive elements inject thermal directly into the resonant loop. Leeson's model offers a foundational overview of these mechanisms, conceptualizing noise addition within the before filtering and subsequent stages, where the process squares the and shifts it closer to the . Beyond intrinsic device effects, extrinsic environmental factors like temperature fluctuations and mechanical vibrations exacerbate phase noise by inducing instabilities; temperature variations alter material properties and dimensions, while vibrations cause microphonic through physical deformations.

Mathematical Modeling

Phase Noise Spectrum

The phase noise spectrum in the is commonly characterized by the single-sideband () phase noise, denoted as \mathcal{L}(f), which quantifies the noise power in one relative to the power per unit at an offset frequency f from the . The standard definition is \mathcal{L}(f) = 10 \log_{10} \left( \frac{S_\phi(f)}{2} \right), where S_\phi(f) is the one-sided power of the phase fluctuations in radians²/Hz, and the units are /Hz. This measure assumes small phase deviations (less than 1 ) and represents half the total phase noise power, as the sidebands are symmetric and coherent. The typical phase noise spectrum exhibits distinct regions as a function of offset frequency f. Close to the (low offsets), a $1/f^3 slope dominates due to upconverted (1/f) from active s, resulting in a 30 /decade . At intermediate offsets, a $1/f^2 region appears from white phase modulated by the resonator's , yielding a 20 /decade slope. Far from the (high offsets), the flattens to a limited by thermal , remaining constant in /Hz. These characteristics arise from the interplay of sources and the oscillator's dynamics. Leeson's equation provides a foundational empirical model for the phase noise spectrum in feedback oscillators, linking device and circuit parameters to the observed shape. The equation is given by: \mathcal{L}(f) = 10 \log_{10} \left[ \left(1 + \frac{f_c^2}{4 Q_l^2 f^2}\right) \left( \frac{F k T}{P_s} \right) \left(1 + \frac{f_c}{f}\right)^3 \right], where f_c is the carrier frequency, Q_l is the loaded quality factor of the resonator, F is the device noise figure, k is Boltzmann's constant ($1.38 \times 10^{-23} J/K), T is the absolute temperature in Kelvin, P_s is the average oscillator signal power, and the (1 + f_c / f)^3 term approximates the flicker noise contribution with f_c as the flicker corner frequency. The derivation begins by modeling the oscillator as a linear amplifier with feedback through a bandpass resonator of bandwidth B = f_c / (2 Q_l). Noise at the amplifier input, including white thermal noise ($2 F k T / P_s) and flicker noise (\propto 1/f), perturbs the phase. For offsets f \ll B, the resonator's low-pass-like response on phase fluctuations amplifies the noise by the factor (f_c / (2 Q_l f))^2, leading to the $1/f^2 term from white noise and $1/f^3 from flicker after upconversion. At f \gg B, the factor approaches 1, yielding the flat floor. The SSB form divides the total phase noise by 2, and the 10 log converts to dBc/Hz. This model highlights that phase noise improves with higher Q_l (narrower bandwidth suppresses noise) and P_s (higher signal-to-noise ratio), though F captures unmodeled imperfections empirically. In phase noise plots, the close-in region (f < f_c / (2 Q_l)) shows the $1/f^3 and $1/f^2 slopes, critical for applications like narrowband communication where spurs degrade signal quality. The far-out region (f > f_c / (2 Q_l)) is the flat floor, relevant for systems. For dielectric resonator oscillators (DROs), the high Q_l (often >10,000) shifts the $1/f^2 corner to higher offsets, enabling superior close-in performance compared to oscillators, though still limits ultimate close-in levels. The total integrated phase noise over a bandwidth from f_1 to f_2 quantifies the accumulated phase error variance as \sigma_\phi^2 = 2 \int_{f_1}^{f_2} 10^{\mathcal{L}(f)/10} \, df, where the factor of 2 accounts for both sidebands in the phase spectral density. This integral provides the root-mean-square phase deviation in radians, establishing the overall noise impact without domain conversion.

Relation to Oscillator Linewidth

Oscillator linewidth, denoted as Δν, is defined as the full-width at half-maximum (FWHM) of the power of the oscillator's output signal, providing a measure of the spectral broadening due to fluctuations. In coherent sources like lasers and oscillators, this linewidth arises fundamentally from , where random perturbations lead to a finite spectral width rather than an ideal delta . For free-running oscillators, the linewidth captures the long-term average spectral spread, contrasting with , which offers detailed insight into noise characteristics at specific offsets from the . In the context of lasers, the intrinsic linewidth is governed by the Schawlow-Townes formula, which quantifies the quantum-limited broadening due to . The standard expression for the linewidth is \Delta \nu = \frac{h \nu (\Delta \nu_c)^2}{4 \pi P} n_{sp} (1 + \alpha^2), where h is Planck's constant, \nu is the optical , \Delta \nu_c is the passive cavity linewidth (FWHM), P is the output power, n_{sp} is the factor (typically near 1 for ideal cases but higher in ), and \alpha is the linewidth enhancement factor accounting for amplitude-phase coupling. This formula links directly to phase noise, as the white frequency noise regime yields a lineshape, with the phase noise S_\phi(f) relating to the linewidth via S_\phi(f) \approx \frac{2 \Delta \nu}{f^2} at offsets f in the 1/f² region. For lasers, the \alpha term often dominates, enhancing the linewidth by factors of 10–100 compared to gas or solid-state lasers. In oscillators, the connection between phase and linewidth follows similar principles but incorporates circuit-level effects modeled by Leeson's equation, which describes the upconversion of device (e.g., and ) to phase near the . The linewidth can be approximated from the phase in the 1/f² region as \Delta \nu \approx \frac{\pi f^2 S_\phi(f)}{2}, where S_\phi(f) is the one-sided phase power in rad²/Hz, reflecting the cumulative effect of upconverted from the resonator quality factor and . In contexts, however, the effective linewidth is typically very narrow (often <<1 Hz) and less commonly used as a metric compared to phase at specific offsets, which better captures performance limitations. Stabilized lasers, such as those used in optical atomic clocks, achieve narrow linewidths below 1 Hz through active feedback to high-finesse cavities, suppressing environmental perturbations like thermal drifts and vibrations that otherwise broaden the spectrum. These perturbations introduce excess phase noise, increasing the linewidth beyond the quantum limit and degrading long-term coherence essential for precision timing. Thus, while linewidth serves as a coarse, integrated metric of phase stability, phase noise measurements provide the offset-specific resolution needed to identify and mitigate such broadening mechanisms.

Time-Domain Analysis

Phase Jitter

Phase jitter represents the time-domain manifestation of phase noise, quantifying the random deviations in the timing of signal edges from their ideal positions in a periodic waveform. It arises from the same underlying noise processes that produce phase noise in the frequency domain, such as thermal noise and flicker noise in oscillators. In practical terms, phase jitter is often expressed as the root-mean-square (RMS) value, providing a measure of the standard deviation of these timing errors. The fundamental conversion from phase deviation to time-domain jitter is given by the relation \sigma_t = \frac{\sigma_\phi}{2\pi f_c}, where \sigma_t is the RMS jitter in seconds, \sigma_\phi is the RMS phase deviation in radians, and f_c is the carrier frequency. The RMS phase deviation itself is derived from the phase noise spectrum through integration: \sigma_\phi = \sqrt{2 \int_{f_1}^{f_2} 10^{L(f)/10} \, df}, where L(f) is the single-sideband phase noise spectral density in dBc/Hz, and the integral is taken over the relevant bandwidth from offset frequency f_1 to f_2. This integration typically excludes very low offsets dominated by deterministic effects and high offsets beyond the signal's bandwidth. Phase jitter manifests in several distinct types, each relevant to different aspects of system analysis. Absolute jitter measures the total timing deviation of each edge relative to an ideal reference clock, capturing all frequency components. Cycle-to-cycle jitter, also known as period jitter, quantifies the variation in consecutive clock periods, defined as the difference between the actual period and the nominal period, and is particularly useful for assessing short-term stability. In phase-locked loops (PLLs), accumulated jitter refers to the buildup of timing errors over multiple cycles, influenced by the loop's transfer function, where low-frequency components may be tracked or filtered. In digital communication systems, phase jitter introduces timing uncertainty that can shift sampling instants away from optimal points, leading to bit errors by causing incorrect symbol decisions near transition regions. For high-speed applications, such as 100 Gbps Ethernet links, stringent jitter requirements are imposed to maintain low bit error rates, with typical RMS phase jitter values held below 1 ps (often around 0.4 ps or less) over integration bands like 10 kHz to 20 MHz to ensure reliable data recovery. Flicker (1/f) noise contributes to long-term phase jitter accumulation through a random walk process, where phase errors integrate over time, resulting in unbounded growth and increased drift in free-running oscillators or wide-loop-bandwidth PLLs.

Conversions Between Domains

Conversions between the frequency-domain representation of phase noise and the time-domain representation of jitter are essential for analyzing oscillator performance across domains. The forward conversion relates the spectral density of phase fluctuations S_\phi(f) to the spectral density of time jitter S_t(f) through the formula S_t(f) = \frac{S_\phi(f)}{(2\pi f_c)^2}, where f_c is the carrier frequency. This relationship arises because time jitter t is the phase fluctuation \phi divided by the angular carrier frequency $2\pi f_c, preserving the power spectral density scaling under linear transformation. The total root-mean-square (RMS) jitter \sigma_t is obtained by integrating the jitter spectral density over the relevant bandwidth: \sigma_t = \sqrt{ \int_{f_L}^{f_H} S_t(f) \, df }, where f_L and f_H are the lower and upper integration limits, often chosen based on system requirements such as 10 kHz to 20 MHz for digital applications. Equivalently, in terms of single-sideband phase noise L(f), where S_\phi(f) = 2 \times 10^{L(f)/10} for f > 0, the RMS phase jitter is \sigma_\phi = \sqrt{ 2 \int_{f_L}^{f_H} 10^{L(f)/10} \, df }, and \sigma_t = \sigma_\phi / (2\pi f_c). This is typically computed numerically by segmenting the phase noise curve into regions (e.g., 1/f³, 1/f², ) and summing their contributions in linear units before converting back to . Key approximations simplify these conversions under specific noise conditions. For white phase noise, where L(f) is flat, the phase variance approximates \sigma_\phi^2 \approx 2 (f_H - f_L) \times 10^{L/10}, yielding a Gaussian distribution for jitter; this holds when the noise is bandlimited and uncorrelated. In regions dominated by 1/f noise, the integral diverges at low frequencies, requiring careful selection of f_L (e.g., 1 Hz or system-specific minimum) to avoid overestimation, as flicker noise contributes cumulatively to long-term phase drift rather than bounded jitter. These approximations enable quick estimates but assume stationary processes and neglect cyclostationary effects in digital clocks. The inverse conversion derives phase noise from time-domain jitter measurements using the Wiener-Khinchin theorem, which states that the power spectral density S_\phi(f) is the of the phase autocorrelation function R_\phi(\tau) = \mathbb{E}[\phi(t) \phi(t + \tau)]. For , R_t(\tau) = R_\phi(\tau) / (2\pi f_c)^2, so S_\phi(f) = (2\pi f_c)^2 \mathcal{F}\{ R_t(\tau) \}, where \mathcal{F} denotes the ; in practice, the autocorrelation is estimated from jitter samples before . This method is particularly useful for validating frequency-domain models against oscilloscope-captured jitter histograms. Standards such as for Ethernet applications specify jitter budgets (e.g., maximum deterministic and random jitter components) that are directly derived from allowable phase noise profiles to ensure bit error rates below 10^{-12}. For instance, transmitter clock phase noise limits are set to limit integrated jitter to fractions of the unit interval, with conversions using the above integrals over bandwidths like 1 kHz to 100 MHz. Numerical methods in (CAD) tools facilitate these conversions through (FFT) algorithms. In time-domain simulations, time series are generated (e.g., via equations modeling sources), and the FFT computes the phase noise by estimating the PSD from the zero-crossing deviations; conversely, frequency-domain phase noise can be inverse-FFT'd to simulate waveforms for system-level verification. Tools like GoldenGate employ this for PLL jitter prediction, ensuring accuracy within 1-2% for .

Measurement Methods

Direct Measurement Techniques

Direct measurement techniques for phase noise involve hardware-based approaches that directly capture and analyze the phase fluctuations of a signal using specialized test equipment. These methods typically downconvert the phase noise to a measurable signal for , providing direct insight into the phase noise as the quantity being measured. The primary utilizes a to compare the device under test (DUT) against a stable reference, enabling high-sensitivity characterization across offset frequencies from 1 Hz to tens of MHz. The phase detector method employs a double-balanced mixer (DBM) as the core component to downconvert the phase noise. In this setup, the DUT signal is applied to the RF port of the DBM, while a low-noise reference oscillator of comparable frequency and stability is connected to the LO port; the two inputs are adjusted to quadrature (90 degrees) to maximize sensitivity to phase deviations. The mixer's IF output produces a voltage proportional to the phase difference, which is then amplified with a low-noise and analyzed using a () spectrum analyzer to generate the single-sideband () phase noise spectrum in /Hz. This configuration effectively isolates close-in phase noise, with the reference oscillator providing a clean comparison benchmark to suppress common-mode noise. For wideband phase noise measurement, particularly beyond 1 MHz offsets where 1/f diminishes, the delay-line discriminator variant is employed. Here, the DUT signal is split using a power divider; one path passes through a long or delay line (typically providing 10-100 ns delay), while the other serves as , and both are mixed in a DBM acting as a frequency discriminator. The delay converts () —closely related to phase noise—into (AM) at the mixer output, which is subsequently FFT-analyzed; this method excels for capturing without requiring a high-quality oscillator. A standard instrument for these techniques is the E5052B signal source analyzer, which integrates the source, , and capabilities for enhanced sensitivity, achieving a floor limited by instrument of approximately -170 /Hz at 10 kHz offset under optimal conditions. Calibration is essential to ensure accuracy in these setups. The mixer's conversion loss, typically 6-10 , must be quantified and compensated, as it directly impacts the detected phase noise level by reducing the signal ; this is done by injecting a known and measuring the IF response. Additionally, (LO) leakage from the reference into the IF port can introduce spurious noise sidebands, requiring suppression through on the IF or careful port isolation, with involving sweeps to subtract leakage contributions. These steps maintain measurement traceability and minimize systematic errors. Recent advancements in photonic methods leverage optical delay lines for improved performance in photonics applications. These approaches use intensity-modulated optical carriers transmitted over low-loss delay lines to achieve sensitivities around -160 /Hz at 10 kHz offsets and resolutions from 1 Hz, extending the delay-line discriminator principle for signals up to 50 GHz.

Indirect and Derived Methods

Indirect methods for assessing phase noise involve inferring it from related observables, such as linewidth or time-domain , rather than directly measuring the phase noise . These approaches are particularly useful when direct analysis is challenging due to limitations or when integrating data from other techniques. Linewidth measurements, for instance, provide an integrated view of phase fluctuations that can be related back to phase noise under certain approximations. One common indirect technique for lasers is the delayed self-heterodyne , where the laser output is split, one path delayed, and the two recombined to produce a beat note whose reflects the linewidth, which is tied to phase noise. This is effective for narrow-linewidth sources, achieving resolutions down to hertz levels by using long delay fibers to decorrelate the phases. Similarly, Fabry-Pérot interferometers scan the laser to resolve the linewidth directly from interference fringes, offering a passive optical approach suitable for continuous-wave lasers in the visible to near-infrared range. These linewidth values can then be used to estimate phase noise assuming a specific lineshape model. Jitter-based inference measures timing deviations in the time domain using oscilloscopes or time-to-digital converters (TDCs), which capture edge variations with picosecond resolution, and then converts these to phase noise via integration over frequency offsets. For example, root-mean-square (RMS) phase jitter \sigma_\phi relates to phase noise through \sigma_\phi^2 = 2 \int_{f_1}^{f_2} S_\phi(f) \, df, allowing estimation of the spectrum shape from accumulated jitter data. This is valuable for high-speed electronics where time-domain tools are readily available. Jitter conversions serve as a basis for such inference by linking time and frequency domains. The connection between linewidth and phase noise is formalized in approximations for the lineshape, such as the , which combines and Gaussian contributions from phase diffusion. A key relation for the full-width at half-maximum (FWHM) linewidth \Delta \nu is given by \Delta \nu = \frac{8\pi}{P_s} \int_0^\infty f^2 S_\phi(f) \, df, where P_s is the signal power and S_\phi(f) is the phase noise ; this integral captures the weighted contribution of frequency-modulated noise to the effective broadening. This equation provides a practical way to derive linewidth from modeled or measured phase noise, especially for non-white noise spectra. In atomic clocks, Allan variance serves as an indirect metric for long-term phase stability, quantifying frequency fluctuations over averaging times \tau via \sigma_y^2(\tau) = \frac{1}{2} \langle (y_{k+1} - y_k)^2 \rangle, where y_k are fractional frequency deviations; its square root, the Allan deviation, reveals phase noise dominance at longer \tau by distinguishing flicker and random-walk processes from . This is essential for assessing clock performance beyond short-term phase noise. Computational tools enable simulation-derived phase noise estimation, such as ADIsimPLL, which models voltage-controlled oscillators (VCOs) and phase-locked loops to predict noise spectra from circuit parameters, allowing indirect validation against derived measurements. These simulations integrate device models to forecast phase noise without physical testing. Despite their utility, indirect methods exhibit reduced accuracy for close-in phase noise at offsets below 1 kHz, where environmental sensitivities and integration assumptions amplify errors, often requiring hybrid approaches with direct techniques for validation.

System-Level Effects

Performance Limitations

Phase noise significantly degrades performance in radio frequency (RF) systems by introducing random phase fluctuations that distort the signal constellation, leading to increased (EVM). EVM quantifies the deviation between the ideal reference signal and the measured signal in digital schemes, and phase noise contributes to this by causing rotational errors in the constellation points, thereby reducing the overall signal quality. In (QAM) systems, phase noise exacerbates (BER) performance, particularly for higher-order constellations like 16-QAM or 64-QAM, where small phase deviations result in symbol misclassification and higher error probabilities at a given (SNR). A key metric for this degradation is the SNR penalty induced by phase noise in phase-modulated systems, approximated as \Delta \text{SNR} = 10 \log_{10} \left(1 + \left(\frac{\sigma_\phi}{\rho}\right)^2 \right), where \sigma_\phi is the root-mean-square (rms) phase error and \rho is the modulation index. This penalty arises because phase noise adds an effective noise component that scales inversely with the modulation depth, limiting the usable SNR for reliable demodulation. In practical RF applications, such as 5G New Radio (NR) systems operating in millimeter-wave bands as of 2025, phase noise imposes strict limits on beamforming accuracy, where elevated noise levels at high carrier frequencies degrade array gain and increase the required oscillator stability to maintain link reliability. Similarly, in Global Positioning System (GPS) receivers, phase noise from local oscillators can induce timing errors, compromising synchronization and position accuracy in navigation tasks. In applications, phase noise broadens the effective transmit , reducing by smearing the return signals and limiting the ability to distinguish closely spaced targets. This broadening is particularly detrimental in frequency-modulated continuous-wave (FMCW) radars, where in-band phase noise directly impacts the beat sharpness, effectively worsening the minimum resolvable . Additionally, in frequency synthesizers used for agile RF generation, phase noise experiences cumulative folding effects, where noise from divider spurs aliases back into the , amplifying close-in noise and degrading the overall purity. This folding mechanism, prominent in fractional-N architectures, can elevate in-band noise under nonlinear conditions, constraining synthesizer performance in high-precision systems.

Mitigation Strategies

Mitigating phase noise in oscillators and related systems involves a combination of passive design choices and active techniques to minimize noise contributions from thermal, flicker, and active device sources. High-Q play a central role in design approaches, as their elevated quality factors reduce the resonator's susceptibility to noise upconversion. For instance, whispering-gallery mode resonators achieve quality factors greater than $10^8 at cryogenic temperatures, enabling microwave oscillators with phase noise floors limited primarily by amplifier noise rather than the resonator itself. Complementing these, low-noise amplifiers with noise figures below 2 are integrated into sustaining circuits to limit added phase noise from active elements, ensuring the overall oscillator performance remains dominated by the resonator's intrinsic stability. Active compensation methods further enhance phase noise performance by employing feedback architectures that suppress fluctuations. Phase-locked loops (PLLs) utilizing low-noise voltage-controlled oscillators (VCOs) achieve this through closed-loop , where the VCO's inherent is filtered by the , often yielding improvements of 20 dB or more in close-in phase noise compared to free-running configurations. Optoelectronic oscillators (OEOs), which combine optical delay lines with RF electronics, provide ultra-low phase noise by exploiting the high effective of fiber loops; recent implementations demonstrate levels as low as -148 /Hz at a 10 kHz offset from a 10 GHz carrier. Cryogenic cooling offers a powerful passive strategy for noise reduction, particularly in high-Q dielectric resonators, by lowering thermal noise density. The phase noise contribution from thermal sources decreases by approximately $10 \log_{10} \left( \frac{T}{77} \right) dB when cooling to 77 K from room temperature T, enabling oscillators with phase noise below -170 dBc/Hz at 10 kHz offsets. As of 2025, this approach has been integrated into chip-scale atomic clocks, achieving fractional frequency stability around 2 \times 10^{-12} at 1-second averaging times through microfabricated vapor cells and optical interrogation. Filtering and feedback techniques address phase noise in signal processing chains, such as multipliers and synthesizers. Narrowband filters suppress far-out noise tails by attenuating broadband contributions outside the desired spectrum, effectively lowering the integrated phase noise without impacting close-in performance. In frequency multipliers, which inherently degrade phase noise by $20 \log_{10} N dB for multiplication factor N, feedback loops recirculate the output to a low-noise reference, mitigating this penalty and achieving net phase noise comparable to the input source. A fundamental trade-off in these strategies arises from considerations, as outlined in Leeson's model, where phase noise L(f) scales inversely with the carrier signal P_s; increasing P_s reduces L(f) but elevates power consumption and potential loading. This relationship underscores the need to balance drive levels with efficiency in low-noise designs.

References

  1. [1]
    [PDF] Phase Noise and AM Noise Measurements in the Frequency Domain
    Measurements of these fluctuations characterize the frequency source in terms of amplitude modulation (AM) and phase modulation. (PM) noise (frequency stability) ...Missing: scholarly | Show results with:scholarly
  2. [2]
    [PDF] A General Theory of Phase Noise in Electrical Oscillators
    Abstract— A general model is introduced which is capable of making accurate, quantitative predictions about the phase noise of different types of electrical ...
  3. [3]
    [PDF] Phase Noise Metrology - Enrico Rubiola
    Phase noise is a random process, and consequently its power spectrum density. Sϕ(f) can only be defined as the Fourier transform of the autocorrelation. Page 3 ...
  4. [4]
    What is Phase Noise? - everything RF
    Phase noise is defined as the noise arising from the rapid, short term, random phase fluctuations that occur in a signal.<|control11|><|separator|>
  5. [5]
    [PDF] Introduction to Phase Noise in Signal Generators - Spanawave
    Phase noise is the result of small random fluctuations or uncertainty in the phase of an electronic signal. We specify and measure phase noise because it is ...
  6. [6]
    [PDF] Phase Noise - Enrico Rubiola
    Feb 7, 2025 · where V0 is the nominal amplitude, ν0 the frequency, α(t) is the random fractional amplitude, and φ(t) is the random phase.
  7. [7]
    [PDF] Sources of Phase Noise and Jitter in Oscillators - Crystek Corporation
    “Shot Noise” - caused by discontinuous current (holes and electrons) flow across pn junction potential barriers. “Flicker Noise” - noise that is spectrally.
  8. [8]
    None
    ### Summary of Leeson's Model, Historical Origin, Definition, and Basic Concepts
  9. [9]
    None
    ### Summary of Leeson's Model: Noise Before and After Frequency Multiplication
  10. [10]
    [PDF] Vibration-induced PM Noise in Oscillators and its Suppression
    1. This sensitivity to vibration originates most commonly from phase fluctuations within the oscillator's positive-feedback loop, due usually to the physical ...
  11. [11]
    [PDF] Phase Noise Tutorial
    Noise in Region B of Figure 1, called “1/F” noise, is caused by semiconductor activity. Design techniques em- ployed in Vectron low noise “L2” crystal ...
  12. [12]
    A simple model of feedback oscillator noise spectrum - IEEE Xplore
    ... >Volume: 54 Issue: 2. A simple model of feedback oscillator noise spectrum. Publisher: IEEE. Cite This. PDF. D.B. Leeson. All Authors. Sign In or Purchase.
  13. [13]
    [PDF] Review of Oscillator Phase Noise Models - IAENG
    The most well-known phase noise model is Leeson's model which was proposed by D. B. Leeson in 1966 [1]. He presented a derivation of the expected spectrum of a.
  14. [14]
    Linewidth – bandwidth, laser, spectral, line width, measurement ...
    The linewidth (or line width) of a laser, eg a single-frequency laser, is the width (typically the full width at half-maximum, FWHM) of its optical spectrum.
  15. [15]
    Schawlow–Townes Linewidth – single-frequency laser, phase noise
    Here, the Schawlow–Townes formula provides an estimate for the linewidth near the center of the spectrum, whereas the linewidth in the spectral wings is ...
  16. [16]
    Phase Noise – laser noise, linewidth - RP Photonics
    The fundamental origin of phase noise is quantum noise, in particular spontaneous emission of the gain medium into the resonator modes, but also quantum noise ...
  17. [17]
    Diode laser with 1 Hz linewidth - Optica Publishing Group
    We report an ultranarrow-linewidth laser spectrometer at 657 nm , consisting of a diode laser locked in a single stage to a stable high-finesse reference ...
  18. [18]
    Chip-based laser with 1-hertz integrated linewidth | Science Advances
    Oct 28, 2022 · In this work, we demonstrate a laser system with a 1-s linewidth of 1.1 Hz and fractional frequency instability below 10 −14 to 1 s.
  19. [19]
    [PDF] MT-008: Converting Oscillator Phase Noise to Time Jitter
    The first step in calculating the equivalent rms jitter is to obtain the integrated phase noise power over the frequency range of interest, i.e., the area of ...
  20. [20]
    Basics of Jitter in Wireline Communications
    Aug 23, 2019 · From Phase Noise to Jitter rms. Ali Sheikholeslami. Jitter in Wireline Communications. a. T. = 𝝋. 2𝜋. 𝝋 = 𝜔0. a. 𝜎a. T. = 𝜎𝜑. 2𝜋. 25 of 78 ...
  21. [21]
  22. [22]
    [PDF] Jitter Basics, Advanced, and Noise Analysis - IEEE Long Island
    Jun 17, 2016 · What is Jitter? • Definitions. ◦ “The deviation of an edge from where it should be”. ◦ ITU Definition of Jitter: “Short-term variations of the.
  23. [23]
    [PDF] Understanding Phase Error and Jitter: Definitions, Implications ...
    This paper addresses this problem by presenting a comprehensive one-stop explanation of how oscillator error is quantified, simulated, and measured in practice, ...
  24. [24]
  25. [25]
    [PDF] High-Performance Clock Integration Key to 40/100G Networks
    Given these considerations, it is recommended that a jitter attenuating clock or local oscillator with ultralow jitter be used (<0.4 ps rms or less) as the 40/ ...
  26. [26]
    [PDF] Phase Noise and Jitter in Digital Electronics - arXiv
    Dec 31, 2016 · Abstract. This article explains phase noise, jitter, and some slower phenomena in digital integrated circuits, focusing on high-demanding, ...<|control11|><|separator|>
  27. [27]
    AN-1067: The Power Spectral Density of Phase Noise and Jitter
    The power spectral density of phase noise and jitter is developed, time domain and frequency domain measurement techniques are described, limitations of ...Missing: S_t( S_φ( f_c)^
  28. [28]
    10GEPON Jitter Budget - IEEE 802
    What is Jitter Budget? ○ Jitter Budget is the collection of the Total Jitter ... - Phase noise and jitter. - Acquisition time. - Acquisition range. ○. VCO Noise ...
  29. [29]
    [PDF] Computing Jitter From Phase Noise in the GoldenGate Simulator
    Jitter defines the uncertainties in the zero-crossing times (timing events) of a periodic signal, or, in other terms, the random variations in the signal ...<|separator|>
  30. [30]
    Self-heterodyne Linewidth Measurement - RP Photonics
    Self-heterodyne linewidth measurement is a technique for measuring laser linewidths, based on a beat note between the beam and a delayed version of itself.
  31. [31]
    Scanning Fabry-Perot Interferometers - Thorlabs.com
    Thorlabs' Scanning Fabry-Perot Interferometers are spectrum analyzers that are ideal for examining fine spectral characteristics of CW lasers.
  32. [32]
    Clock (CLK) Jitter and Phase Noise Conversion - Analog Devices
    Dec 10, 2004 · This article demonstrates the exact mathematical relationship between jitter measured in time and the phase-noise measured in frequency.
  33. [33]
    Simple approach to the relation between laser frequency noise and ...
    The relation between frequency noise spectral density and laser linewidth has already been addressed in many papers dealing with general theoretical ...
  34. [34]
    [PDF] Characterization of Frequency and Phase Noise
    This occurs typically at times ranging from hours to days, depending on the particular kind of standard. A “modified Allan variance”, MOD. U:(T), has been - ...
  35. [35]
    ADIsimPLL - Analog Devices
    All key nonlinear effects that can impact PLL performance can be simulated, including phase noise, fractional-N spurs, and anti-backlash pulse.
  36. [36]
  37. [37]
    A Study of BER and EVM Degradation in Digital Modulation ...
    May 12, 2022 · This paper presents an analytical study of EVM and BER degradation for a variety of widely used digital modulation schemes. Phase noise and ...
  38. [38]
    A General Conditional BER Expression of Rectangular QAM in the ...
    Numerical results show that the impact of phase noise on the conditional BER performance is proportional to the constellation size.
  39. [39]
    [PDF] Signal-to-Noise Ratio Penalties for Continuous-Time Phase Noise ...
    Abstract—Signal-to-noise ratio (SNR) penalties are studied for continuous-time additive white Gaussian noise channels with white and Wiener phase noise.Missing: (σ_φ / | Show results with:(σ_φ /
  40. [40]
    [PDF] Time and Frequency Measurements Using the Global Positioning ...
    When used between two stations in the continental. United States, the timing uncertainty (2σ) of the single- channel common-view technique should be < 10 ns at ...
  41. [41]
    Modeling and analysis of the effects of PLL phase noise on FMCW ...
    The radar model shows that PLL in-band phase noise affects the spatial resolution of the FMCW radar, whereas PLL out-of-band phase noise limits the maximum ...
  42. [42]
    Phase Noise Reduction in Microwave Oscillators
    Oct 1, 2009 · This article briefly summarizes the research and development effort in the area of low-noise signal generation.
  43. [43]
    [PDF] Design techniques and noise properties of ultrastable cryogenically ...
    We review the techniques used in the design and construction of cryogenic sapphire oscillators at the University of Western Australia over the 18 year ...
  44. [44]
    [PDF] Low Phase Noise Oscillators: Theory and Application
    This paper reviews a linear model and theory for low phase noise oscillators, describing how they work and important parameters. Design rules are developed for ...<|separator|>
  45. [45]
    Phase-Locked Loop (PLL) Fundamentals - Analog Devices
    The low-pass filter then needs to be carefully placed to ensure that the in-band PLL noise will intersect with the VCO noise—this ensures lowest rms jitter. A ...
  46. [46]
    Parity-time symmetric optoelectronic oscillator with high side-mode ...
    Nov 1, 2025 · ... phase noise can reach -148.88 dBc/Hz at 10-kHz offset via a 4.4-km OEO loop. Furthermore, the significant contrast between the generated 10 ...
  47. [47]
    (PDF) Next-Generation Chip Scale Atomic Clocks - ResearchGate
    The use of optical transitions in microfabricated vapor cells improves both short- and long-term frequency stability to near 10 ⁻¹³ at the cost of the added ...
  48. [48]
    [PDF] A filtering technique to lower LC oscillator phase noise
    Abstract—Based on a physical understanding of phase-noise mechanisms, a passive LC filter is found to lower the phase-noise.Missing: out | Show results with:out
  49. [49]
    [PDF] L-Band Frequency Multipliers: Phase Noise - IPN Progress Report
    From the re- search performed so far, the improvement in phase noise is the most likely due to a reduction in the overall positive feedback present in the ...<|separator|>