Dick effect

Last updated

The Dick effect (hereinafter; "the effect") is an important limitation to frequency stability for modern atomic clocks such as atomic fountains and optical lattice clocks. It is an aliasing effect: High frequency noise in a required local oscillator (LO) is aliased (heterodyned) to near zero frequency by a periodic interrogation process that locks the frequency of the LO to that of the atoms. The noise mimics and adds to the clock's inherent statistical instability, which is determined by the number of atoms or photons available. In so doing, the effect degrades the stability of the atomic clock and places new and stringent demands on LO performance.

Contents

For any given interrogation protocol, the effect can be calculated using a quantum-mechanical sensitivity function, together with the spectral properties of the LO noise. This calculational methodology, introduced by G. John Dick, is now widely used in the design of advanced microwave and optical frequency standards, as well as in the development of methodologies for atomic-wave interferometry, frequency standard comparison, and other areas of measurement science.

Background

General

Frequency stability

The frequency stability of an atomic clock is usually characterized by the Allan deviation , [1] a measure of the expected statistical variation of fractional frequency as a function of averaging time . Generally, short-term fluctuations (frequency or phase noise) in the clock output require averaging for an extended period of time in order to achieve high performance.

Allan deviation for a commercial atomic clock AllanDeviationExample.gif
Allan deviation for a commercial atomic clock

This stability is not the same as the accuracy of the clock, which estimates the expected difference of the average frequency from some absolute standard. [2]

Excellent frequency stability is crucial to a clock's usability: Even though it might have excellent accuracy, a clock with poor frequency stability may require averaging for a week or more for a single high precision test or comparison. Such a clock would not be as useful as one with a higher stability; one that could accomplish the test in hours instead of days.

Stability and operation of atomic clocks

Instability in the output from an atomic clock due to imperfect feedback between atoms and LO was previously well understood. [3] [4] This instability is of a short-term nature and typically does not impact the utility of the clock. The effect, on the other hand gives rise to frequency noise which has the same character as (and is typically much larger than) that due to the fundamental photon– or atom–counting limitation for atomic clocks.

With the exception of hydrogen and ammonia (hydrogen maser, ammonia maser), the atoms or ions in atomic clocks do not provide a usable output signal. Instead, an electronic or optical local oscillator (LO) provides the required output. The LO typically provides excellent short-term stability; long-term stability being achieved by correcting its frequency variability by feedback from the atoms.

In advanced frequency standards the atomic interrogation process is usually sequential in nature: After state-preparation, the atoms' internal clocks are allowed to oscillate in the presence of a signal from the LO for a period of time. At the end of this period, the atoms are interrogated by an optical signal to determine whether (and how much) the state has changed. This information is used to correct the frequency of the LO. Repeated again and again, this enables continuous operation with stability much higher than that of the LO itself. In fact, such feedback was previously thought to allow the stability of the LO output to approach the statistical limit for the atoms for long measuring times.

The effect

The effect [5] [6] is an additional source of instability that disrupts this happy picture. It arises from an interaction between phase noise in the LO and periodic variations in feedback gain that result from the interrogation procedure. The temporal variations in feedback gain alias (or heterodyne) LO noise at frequencies associated with the interrogation period to near zero frequency, and this results in an instability (Allan deviation) that improves only slowly with increasing measuring time. The increased instability limits the utility of the atomic clock and results in stringent requirements on performance (and associated expense) for the required LO: Not only must it provide excellent stability (so that its output can be improved by feedback to the ultra-high stability of the atoms); it must now also have excellent (low) phase noise.

A simple, but incomplete, analysis of the effect may be found by observing that any variation in LO frequency or phase during a dead time required to prepare atoms for the next interrogation is completely undetected, and so will not be corrected. However, this approach does not take into account the quantum-mechanical response of the atoms while they are exposed to pulses of signal from the LO. This is an additional time-dependent response, calculated in analysis of the effect by means of a sensitivity function. [5] [7]

Quantitative

Impact of the Dick effect on the frequency stability of an Hg Ion Clock Plot of Allan Deviation showing the impact of the Dick Effect on the frequency stability of an Atomic Clock.jpg
Impact of the Dick effect on the frequency stability of an Hg Ion Clock

The graphs here show predictions of the effect for a trapped-ion frequency standard using a quartz LO. [5] In addition to excellent stability, quartz oscillators have very well defined noise characteristics: Their frequency fluctuations are characterized as flicker frequency over a very wide range of frequency and time. Flicker frequency noise corresponds to a constant Allan deviation as shown for the quartz LO in the graphs here.

The "expected" curve on the plot shows how stability of the LO is improved by feedback from the atoms. As measuring time is increased (for times longer than an attack time) the stability steadily increases, approaching the inherent stability of the atoms for times longer than about 10,000 seconds. The "actual" curve shows how the stability is impacted by the effect. Instead of approaching the inherent stability of the atoms, the stability of the LO output now approaches a line with a much higher value. The slope of this line is identical to that of the atomic limitation (minus one half on a log-log plot) with a value that is comparable to that of the LO, measured at the cycle time, as indicated by the small blue (downwards) arrow. The value (the length of the blue arrow) depends on the details of the atomic interrogation protocol, and can be calculated using the sensitivity function methodology.

Surprisingly, the long-term stability of an atomic Clock depends on the short-term stability of the LO due to the Dick effect Plot of Allan Deviation showing how a quartz LO impacts the Dick Effect for Atomic Clock.jpg
Surprisingly, the long-term stability of an atomic Clock depends on the short-term stability of the LO due to the Dick effect

The second graph here indicates how various performance aspects of the LO impact achievable stability for the atomic clock. The dependence labeled "Previously Analyzed LO Impact" shows the stability improving on that of the LO with an approximately dependence for times longer than an "attack time" for the feedback loop. For increasing values of the measuring time , the stability approaches the limiting dependence due to statistical variation in the numbers of atoms and photons available for each measurement.

The effect, on the other hand, causes the available stability of the frequency standard to show a counter-intuitive dependence on high-frequency LO phase noise. Here stability of the LO at times less than the Cycle Time is shown to influence stability of the atomic standard over its entire range of operation. Furthermore, it often prevents the clock from ever approaching the stability inherent in the atomic system.

History

Within a few years of the publication of two papers [5] [6] laying out an analysis of LO aliasing, the methodology was experimentally verified, [8] [9] generally adopted by the Time and Frequency community, and applied to the design of many advanced frequency standards. It was also clarified by Lemonde et al. (1998) [7] with a derivation of the sensitivity function that used a more conventional quantum-mechanical approach, and was generalized by Santarelli et al. (1996) [9] so as to apply to interrogation protocols without even time symmetry.

Where performance limits for atomic clocks were previously characterized by accuracy and by the photon– or atom–counting limitation to stability, the effect was now a third part of the picture. This early stage culminated in 1998 in the publication of four papers [10] [8] [11] [12] in a Special Issue on the Dick effect [13] for the journal IEEE Transactions on Ultrasonics, Ferroelectrics, and Frequency Control .

Impact

Perhaps the most significant consequence of the Dick analysis is due to its presentation of a mathematical framework that enabled researchers to accurately calculate the effect based on the methodology and technology used for many very different atomic clocks. Since the effect is generally the most significant limitation to stability for advanced frequency standards, [14] [15] a great deal of work since that time has focused on amelioration strategies. Additionally, the effect methodology and the sensitivity function have enabled significant progress in a number of technical areas.

These atomic clocks typically operate by tossing a ball of laser-cooled atoms upward through a microwave cavity that acts to start the clock in each individual atom. As the atoms return downward, they again traverse this same cavity where they receive a second microwave pulse that stops their clocks. The ball then falls through an optical interrogation apparatus below the cavity that "reads out" the phase difference between the microwaves (the LO) and the atoms that developed during their flying time. This is repeated again and again; a sequential process that gives rise to the effect. [5]
On a smaller scale, the stability of Rb vapor clocks using a pulsed optically pumped (POP) technique has improved to such an extent that the effect due to a quartz USO LO has become the limiting performance factor. Developments in a combined USO--DRO (dielectric resonator oscillator) LO technology [22] now enable improved performance.

Methodology

Introduction

Modern atomic frequency standards or clocks typically comprise a local oscillator (LO), an atomic system that is periodically interrogated by the LO, and a feedback loop to correct the frequency errors in the LO based on the results of that interrogation; thus locking the frequency of the LO to that of the atomic system. [3] [4] The effect describes a process that makes for imperfect locking, one that depends on details of the atomic interrogation protocol. [5] Two steps are required in order to calculate this newly recognized impact of LO noise on the frequency stability of the locked local oscillator (LLO) that provides useful output for the frequency standard. These are:

In contrast to other examples of photonic excitation of atoms or ions, this excitation process is a slow business, taking milliseconds or even seconds to accomplish on account of the extremely high Q factors involved. Instead of a photon striking a solid and ejecting an electron, here is a process where a coherent EM field consisting of many photons (slowly) drives each atom or ion in a cloud from their ground state into a mixed state, typically one with equal amplitudes in the ground and excited states.
Functional forms for during the time that the atoms are being exposed to interrogating microwave or optical fields can be calculated using a fictitious spin model for the quantum mechanical state-transition process [5] [6] or by using an algebraic approach. [7] These forms, in combination with constant values (typically zero or unity) during times when no signal is applied, allow the sensitivity function to be well defined over the entire interrogation cycle.
The discriminating power for both Ramsey Interrogation and Rabi Interrogation of atomic systems had been previously calculated, based on the quantum-mechanical response of an atom or ion to a slightly detuned drive signal. These previously calculated values are now seen to correspond to a time average of the sensitivity function , taken over the interrogation cycle.

Calculation of the sensitivity function

Excitation amplitude
a
(
t
)
{\displaystyle a(t)}
and sensitivity function
g
(
t
)
{\displaystyle g(t)}
sequences for Atomic Clocks with Rabi and Ramsey interrogation protocols. Rabi interrogation uses a single signal pulse, offset in frequency so that its phase varies smoothly as the pulse progresses. Ramsey interrogation uses two short pulses with a
p
i
/
2
{\displaystyle pi/2}
phase shift between them. Two interrogations are shown in each box, along with three dead times (during which readout, state preparation, and other housekeeping tasks are performed). Compare Atomic Clock Time Sequences for Rabi and Ramsey Interrogation.png
Excitation amplitude and sensitivity function sequences for Atomic Clocks with Rabi and Ramsey interrogation protocols. Rabi interrogation uses a single signal pulse, offset in frequency so that its phase varies smoothly as the pulse progresses. Ramsey interrogation uses two short pulses with a phase shift between them. Two interrogations are shown in each box, along with three dead times (during which readout, state preparation, and other housekeeping tasks are performed).

The concepts and results of calculations presented below can be found in the first papers describing the effect. [5] [6]

Each interrogation cycle in an atomic clock typically begins with preparation of the atoms or ions in their ground states. Let P be the probability that any one will be found in its excited state after an interrogation. The amplitude and time of the interrogating signal are typically adjusted so that tuning the LO exactly to the atomic frequency will give , that is, all of the atoms or ions being in their excited state. P is determined for each measurement by then exposing the system to a different signal that will generate fluorescence only for the (e.g.) excited state atoms or ions.

In order to obtain effective feedback using periodic measurements of P, the protocol must be arranged so that P has a sensitivity to frequency variations. The sensitivity to frequency variation can then be defined as

where is the interrogation time, so that the value of characterizes the sensitivity of a measurement of P to a variation in frequency of the LO. Since P is maximized (at ) when the LO is exactly tuned to the atomic transition frequency, the value of would be zero for that case. Thus, for example, in a frequency standard using Rabi Interrogation, the LO is initially detuned so that , and when instability of the LO frequency causes a subsequent measurement of P to return a value different from this, the feedback loop adjusts the LO frequency to bring it back.

Experimentalists use various protocols to mitigate temporal variations in atomic number, light intensity, etc., and so to allow P to be accurately determined, but these are not discussed further here.

The sensitivity of P to variation of the LO frequency for Rabi Interrogation has been previously calculated, [45] and found to have a value of when the LO frequency has been offset by a frequency to give . This is achieved when is detuned so that .

A time-dependent form for the sensitivity of P to frequency variation can now be introduced, defining as:

,
where is the change in the probability of excitation when a phase step is introduced into the interrogating signal at time . Integrating both sides of the equation shows that the effect on the probability P of a frequency that varies during the excitation process, can be written:

.
This shows to be a sensitivity function; representing the time dependence for the effect of frequency variations on the final excitation probability.

The sensitivity function for the case of Rabi Interrogation is shown to be given by: [5] [6]


where ,
,
,
and where is detuned to half-signal amplitude .

Taking the time average of this functional form for , gives
,
exactly as previously referenced for : This shows to be a proper generalization of the previously used sensitivity .

Forms for the sensitivity function for the case of Ramsey Interrogation with a phase step between two interrogation pulses (instead of a frequency offset) are somewhat simpler, and are given by: [5] [6]


where is the pulse time, is the interrogation time and is the cycle time.

Calculation of the limitation to frequency standard stability

Block diagram for a passive atomic frequency standard using sequential interrogation with cycle time
t
c
{\displaystyle t_{c}}
. The term
d
n
(
t
)
{\displaystyle \delta \nu (t)}
has a time dependent part due to frequency noise from the local oscillator
S
y
L
O
(
f
)
{\displaystyle S_{y}^{LO}(f)}
. Block diagram for a passive atomic frequency standard using sequential interrogation with cycle time t c.jpg
Block diagram for a passive atomic frequency standard using sequential interrogation with cycle time . The term has a time dependent part due to frequency noise from the local oscillator .

The operation of a pulse-mode atomic clock can be broken into functional elements as shown in the block diagram here (for a complete analysis see Greenhall [11] ). Here, the LO is represented by its own block and the interrogated atomic system by the other four blocks. The time dependence of the atomic interrogation process is effected here by the Modulator, in which the time-dependent frequency error is multiplied by a time-dependent gain as calculated in the previous section. The signal input to the integrator is proportional to the frequency error , and this allows it to correct slow frequency errors and drift in the local oscillator.
To understand the action in the block diagram, consider the values and to be made up of their average values plus the deviations from the average. The value of (with averages taken over one cycle, ) gives rise to proper feedback operation, locking the frequency of the local oscillator to that of the discriminator, . Additionally, high frequency components of are smoothed by integration and sampling, giving rise to the already known short-term stability limit. [4] However, the term , while generating additional high-frequency noise, also gives rise to very low frequency variations. This is the aliasing effect that causes the loop to improperly correct the local oscillator and which results in additional low frequency variation in the output of the frequency standard.
Following the methodology of (Dick, 1987) [5] and (Santarelli et al., 1996), [9] the Fourier components of the sensitivity function are:
,
,
,
and ,
where is the cycle time. The locked local oscillator provides the useful output signal from any passive (non-maser) frequency standard. A lower limit to its White frequency noise is then shown to be dependent on the frequency noise of the LO at all frequencies with a value given by
,
where is the cycle time (the time between successive measurements of the atomic system).
The Allan variance for an oscillator with White frequency noise [46] is given by , so that the stability limit due to the effect is given by
.
For Ramsey interrogation with very short interrogation pulses, this becomes
,
where is the interrogation time. For the case of an LO with Flicker frequency noise [46] where is independent of , and where the duty factor has typical values , the Allan deviation can be approximated as [8]
.

See also

Related Research Articles

<span class="mw-page-title-main">Gravitational redshift</span> Shift of wavelength of a photon to longer wavelength

In physics and general relativity, gravitational redshift is the phenomenon that electromagnetic waves or photons travelling out of a gravitational well lose energy. This loss of energy corresponds to a decrease in the wave frequency and increase in the wavelength, known more generally as a redshift. The opposite effect, in which photons gain energy when travelling into a gravitational well, is known as a gravitational blueshift. The effect was first described by Einstein in 1907, eight years before his publication of the full theory of relativity.

<span class="mw-page-title-main">Stimulated emission</span> Release of a photon triggered by another

Stimulated emission is the process by which an incoming photon of a specific frequency can interact with an excited atomic electron, causing it to drop to a lower energy level. The liberated energy transfers to the electromagnetic field, creating a new photon with a frequency, polarization, and direction of travel that are all identical to the photons of the incident wave. This is in contrast to spontaneous emission, which occurs at a characteristic rate for each of the atoms/oscillators in the upper energy state regardless of the external electromagnetic field.

<span class="mw-page-title-main">Phase-locked loop</span> Electronic control system

A phase-locked loop or phase lock loop (PLL) is a control system that generates an output signal whose phase is related to the phase of an input signal. There are several different types; the simplest is an electronic circuit consisting of a variable frequency oscillator and a phase detector in a feedback loop. The oscillator's frequency and phase are controlled proportionally by an applied voltage, hence the term voltage-controlled oscillator (VCO). The oscillator generates a periodic signal of a specific frequency, and the phase detector compares the phase of that signal with the phase of the input periodic signal, to adjust the oscillator to keep the phases matched.

A numerically controlled oscillator (NCO) is a digital signal generator which creates a synchronous, discrete-time, discrete-valued representation of a waveform, usually sinusoidal. NCOs are often used in conjunction with a digital-to-analog converter (DAC) at the output to create a direct digital synthesizer (DDS).

<span class="mw-page-title-main">Hyperfine structure</span> Small shifts and splittings in the energy levels of atoms, molecules and ions

In atomic physics, hyperfine structure is defined by small shifts in otherwise degenerate energy levels and the resulting splittings in those energy levels of atoms, molecules, and ions, due to electromagnetic multipole interaction between the nucleus and electron clouds.

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

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

Resolved sideband cooling is a laser cooling technique allowing cooling of tightly bound atoms and ions beyond the Doppler cooling limit, potentially to their motional ground state. Aside from the curiosity of having a particle at zero point energy, such preparation of a particle in a definite state with high probability (initialization) is an essential part of state manipulation experiments in quantum optics and quantum computing.

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

<span class="mw-page-title-main">Magneto-optical trap</span> Apparatus for trapping and cooling neutral atoms

In atomic, molecular, and optical physics, a magneto-optical trap (MOT) is an apparatus which uses laser cooling and a spatially-varying magnetic field to create a trap which can produce samples of cold, neutral atoms. Temperatures achieved in a MOT can be as low as several microkelvin, depending on the atomic species, which is two or three times below the photon recoil limit. However, for atoms with an unresolved hyperfine structure, such as 7Li, the temperature achieved in a MOT will be higher than the Doppler cooling limit.

In spectroscopy, the Autler–Townes effect, is a dynamical Stark effect corresponding to the case when an oscillating electric field is tuned in resonance to the transition frequency of a given spectral line, and resulting in a change of the shape of the absorption/emission spectra of that spectral line. The AC Stark effect was discovered in 1955 by American physicists Stanley Autler and Charles Townes.

A spin exchange relaxation-free (SERF) magnetometer is a type of magnetometer developed at Princeton University in the early 2000s. SERF magnetometers measure magnetic fields by using lasers to detect the interaction between alkali metal atoms in a vapor and the magnetic field.

<span class="mw-page-title-main">High harmonic generation</span> Laser science process

High-harmonic generation (HHG) is a non-linear process during which a target is illuminated by an intense laser pulse. Under such conditions, the sample will emit the high harmonics of the generation beam. Due to the coherent nature of the process, high-harmonics generation is a prerequisite of attosecond physics.

The Kapitza–Dirac effect is a quantum mechanical effect consisting of the diffraction of matter by a standing wave of light. The effect was first predicted as the diffraction of electrons from a standing wave of light by Paul Dirac and Pyotr Kapitsa in 1933. The effect relies on the wave–particle duality of matter as stated by the de Broglie hypothesis in 1924.

<span class="mw-page-title-main">Atomic clock</span> Extremely accurate clock

An atomic clock is a clock that measures time by monitoring the resonant frequency of atoms. It is based on atoms having different energy levels. Electron states in an atom are associated with different energy levels, and in transitions between such states they interact with a very specific frequency of electromagnetic radiation. This phenomenon serves as the basis for the International System of Units' (SI) definition of a second:

The second, symbol s, is the SI unit of time. It is defined by taking the fixed numerical value of the caesium frequency, , the unperturbed ground-state hyperfine transition frequency of the caesium-133 atom, to be 9192631770 when expressed in the unit Hz, which is equal to s−1.

Laser linewidth is the spectral linewidth of a laser beam.

Self-mixing or back-injection laser interferometry is an interferometric technique in which a part of the light reflected by a vibrating target is reflected into the laser cavity, causing a modulation both in amplitude and in frequency of the emitted optical beam. In this way, the laser becomes sensitive to the distance traveled by the reflected beam thus becoming a distance, speed or vibration sensor. The advantage compared to a traditional measurement system is a lower cost thanks to the absence of collimation optics and external photodiodes.

Ramsey interferometry, also known as the separated oscillating fields method, is a form of particle interferometry that uses the phenomenon of magnetic resonance to measure transition frequencies of particles. It was developed in 1949 by Norman Ramsey, who built upon the ideas of his mentor, Isidor Isaac Rabi, who initially developed a technique for measuring particle transition frequencies. Ramsey's method is used today in atomic clocks and in the S.I. definition of the second. Most precision atomic measurements, such as modern atom interferometers and quantum logic gates, have a Ramsey-type configuration. A more modern method, known as Ramsey–Bordé interferometry uses a Ramsey configuration and was developed by French physicist Christian Bordé and is known as the Ramsey–Bordé interferometer. Bordé's main idea was to use atomic recoil to create a beam splitter of different geometries for an atom-wave. The Ramsey–Bordé interferometer specifically uses two pairs of counter-propagating interaction waves, and another method named the "photon-echo" uses two co-propagating pairs of interaction waves.

<span class="mw-page-title-main">Cavity optomechanics</span>

Cavity optomechanics is a branch of physics which focuses on the interaction between light and mechanical objects on low-energy scales. It is a cross field of optics, quantum optics, solid-state physics and materials science. The motivation for research on cavity optomechanics comes from fundamental effects of quantum theory and gravity, as well as technological applications.

In quantum physics, light is in a squeezed state if its electric field strength Ԑ for some phases has a quantum uncertainty smaller than that of a coherent state. The term squeezing thus refers to a reduced quantum uncertainty. To obey Heisenberg's uncertainty relation, a squeezed state must also have phases at which the electric field uncertainty is anti-squeezed, i.e. larger than that of a coherent state. Since 2019, the gravitational-wave observatories LIGO and Virgo employ squeezed laser light, which has significantly increased the rate of observed gravitational-wave events.

Polarization gradient cooling is a technique in laser cooling of atoms. It was proposed to explain the experimental observation of cooling below the doppler limit. Shortly after the theory was introduced experiments were performed that verified the theoretical predictions. While Doppler cooling allows atoms to be cooled to hundreds of microkelvin, PG cooling allows atoms to be cooled to a few microkelvin or less.

References

  1. Allan, D. Statistics of Atomic Frequency Standards, pages 221–230. Proceedings of the IEEE, Vol. 54, No 2, February 1966.
  2. 1 2 Yao, J.; Sherman, J.A.; Fortier, T.; Leopardi, H.; Parker, T.; McGrew, W.; Zhang, X.; Nicolodi, D.; Fasano, R.; Shaffer, S.; Beloy, K.; Savory, J.; Romisch, S.; Oates, C.; Diddams, S.; Ludlow, A.; Levine, J. (October 30, 2019). "Optical-Clock-Based Time Scale". Physical Review Applied . 12 (44069): 044069. arXiv: 1902.06858 . Bibcode:2019PhRvP..12d4069Y. doi:10.1103/PhysRevApplied.12.044069. PMC   7580056 . PMID   33102625. S2CID   86468489.
  3. 1 2 Cutler, L. S.; Searle, C. L. (February 1966). "Some Aspects of the Theory and Measurements of Frequency Fluctuations in Frequency Standards" (PDF). Proceedings of the IEEE . 54 (2): 136–154. doi:10.1109/PROC.1966.4627.
  4. 1 2 3 Vanier, J.; Tetu, M.; Bernier, L-G (September 1979). "Transfer of Frequency Stability from an Atomic Frequency Reference to a Quartz-Crystal Oscillator". IEEE Transactions on Instrumentation and Measurement. 28 (3): 188–193. Bibcode:1979ITIM...28..188V. doi:10.1109/TIM.1979.4314803.
  5. 1 2 3 4 5 6 7 8 9 10 11 Dick, G.J. (1987). Local oscillator induced instabilities in trapped ion frequency standards (PDF). Precise Time and Time Interval (PTTI) Conference. Redondo Beach.
  6. 1 2 3 4 5 6 7 8 9 Dick, G.J.; Prestage, J.D.; Greenhall, C.A.; Maleki, L. (1990). Local Oscillator Induced Degradation of Medium-Term Stability in Passive Atomic Frequency Standards (PDF). Precise Time and Time Interval (PTTI) Conference.
  7. 1 2 3 Lemonde, P.; Santarelli, G.; Laurent, P.; Dos Santos, F.P.; Salomon, C.; Clairon, A. (May 29, 1998). The sensitivity function: a new tool for the evaluation of frequency shifts in atomic spectroscopy. 1998 IEEE International Frequency Control Symposium. Pasadena, California. doi:10.1109/FREQ.1998.717890.
  8. 1 2 3 Santarelli, G.; Audoin, C.; Makdissi, A.; Laurent, P.; Dick, G.J.; Clairon, A. (1998). "Frequency stability degradation of an oscillator slaved to a periodically interrogated atomic resonator". IEEE Transactions on Ultrasonics, Ferroelectrics, and Frequency Control . 45 (4): 887–894. doi:10.1109/58.710548. PMID   18244242. S2CID   12303876.
  9. 1 2 3 Santarelli, G.; Laurent, P.; Clairon, A.; Dick, G.J.; Greenhall, C.A.; Audoin, C. (1996). Theoretical description and Experimental Evaluation of the Effect of Local Oscillator Frequency Noise on the Stability of a Pulsed Atomic Frequency Standard. 10th International Conference on European Frequency and Time. pp. 66–71.
  10. Audoin, C.; Santarelli, G.; Makdissi, A.; Clairon, A. (1998). "Properties of an Oscillator Slaved to a Periodically Interrogated Atomic Resonator". IEEE Transactions on Ultrasonics, Ferroelectrics, and Frequency Control . 45 (4): 877–886. doi:10.1109/58.710546. PMID   18244241. S2CID   26471543.
  11. 1 2 3 Greenhall, C. (1998). "A Derivation of the Long-Term Degradation of a Pulsed Atomic Frequency Standard from a Control-Loop Model" (PDF). IEEE Transactions on Ultrasonics, Ferroelectrics, and Frequency Control . 45 (4): 895–898. doi:10.1109/58.710550. PMID   18244243. S2CID   8199487.
  12. Lo Presti, L.; Rovera, D.; De Marchi, A. (1998). "A Simple Analysis of the Dick Effect in Terms of Phase Noise Spectral Densities". IEEE Transactions on Ultrasonics, Ferroelectrics, and Frequency Control . 45 (4): 895–905. doi:10.1109/58.710552. PMID   18244244. S2CID   8341222.
  13. Maleki, L. (1998). "Introduction to the Special Issue on the Dick Effect". IEEE Transactions on Ultrasonics, Ferroelectrics, and Frequency Control . 45 (4): 876. doi:10.1109/TUFFC.1998.710542. PMID   18244240.
  14. 1 2 Katori, H. (2011). "Optical lattice clocks and quantum metrology". Nature Photonics . 5 (4): 203–210. Bibcode:2011NaPho...5..203K. doi:10.1038/nphoton.2011.45.
  15. 1 2 Ludlow, A.D.; Boyd, M.M.; Ye, J.; Peik, E.; Schmidt, P.O. (June 26, 2015). "Optical atomic clocks". Reviews of Modern Physics . 87 (2): 637–701. arXiv: 1407.3493 . Bibcode:2015RvMP...87..637L. doi:10.1103/RevModPhys.87.637. S2CID   119116973.
  16. 1 2 Schioppo, M.; Brown, R. C.; McGrew, W. F.; Hinkley, N.; Fasano, R. J.; Beloy, K.; Yoon, T. H.; Milani, G.; Nicolodi, D.; Sherman, J. A.; Phillips, N. B.; Oates, C. W.; Ludlow, A. D. (2017). "Ultrastable optical clock with two cold-atom ensembles". Nature Photonics . 11 (1): 48–52. arXiv: 1607.06867 . Bibcode:2017NaPho..11...48S. doi:10.1038/nphoton.2016.231. S2CID   118541117.
  17. Danet, J. M.; Lours, M.; Guérandel, S.; De Clercq, E. (2014). "Dick effect in a pulsed atomic clock using coherent population trapping". IEEE Transactions on Ultrasonics, Ferroelectrics, and Frequency Control . 61 (4): 567–574. arXiv: 1406.6673 . doi:10.1109/TUFFC.2014.2945. S2CID   363993.
  18. Al Masoudi, A.; Dorcher, S.; Hafner, S.; Sterr, U.; Lisdat, C. (December 2015). "Noise and instability of an optical lattice clock". Physical Review A . 92 (63814): 063814. arXiv: 1507.04949 . Bibcode:2015PhRvA..92f3814A. doi:10.1103/PhysRevA.92.063814. S2CID   119217208.
  19. Mann, A.G.; Santarelli, G.; Chang, S.; Liuten, A.N.; Laurent, P.; Salomon, C.; Blair, D.G.; Clairon, A. (May 29, 1998). A High Stability Atomic Fountain Clock using a Cryogenic Sapphire Interrogation Oscillator. 1998 IEEE International Frequency Control Symposium. Pasadena, California. doi:10.1109/FREQ.1998.717871.
    Santarelli, G.; Laurent, P.; Lemonde, P.; Clairon, A.; Mann, A.G.; Chang, S.; Liuten, A.N.; Salomon, C. (1999). "Quantum Projection Noise in an Atomic Fountain: A High Stability Cesium Frequency Standard". Physical Review Letters . 82 (23): 4619–4622. Bibcode:1999PhRvL..82.4619S. doi:10.1103/PhysRevLett.82.4619. S2CID   26451919.
  20. Takamizawa, A.; Yanagimachi, S.; Tanabe, T.; Hagimoto, K.; Hirano, I.; Watabe, K.; Ikegami, T.; Hartnett, J. (2014). "Atomic fountain clock with very high frequency stability employing a pulse-tube-cryocooled sapphire oscillator". IEEE Transactions on Ultrasonics, Ferroelectrics, and Frequency Control . 61 (9): 1463–1469. doi:10.1109/TUFFC.2014.3060. PMID   25167146. S2CID   27535524.
  21. Weyers, S.; Lipphardt, B.; Schnatz, H. (March 11, 2009). "Reaching the quantum limit in a fountain clock using a microwave oscillator phase locked to an ultrastable laser" (PDF). Physical Review A . 79 (31803): 031803. arXiv: 0901.2788 . Bibcode:2009PhRvA..79c1803W. doi:10.1103/PhysRevA.79.031803. S2CID   119267225.
  22. Li, W.B.; Hao, Q.; Du, Y.B.; Huang, S.Q.; Yun, P.; Lu, Z.H. (2019). "Demonstration of a Sub-Sampling Phase Lock Loop Based Microwave Source for Reducing Dick Effect in Atomic Clocks". Chinese Physics Letters . 36 (7): 070601. arXiv: 1810.03803 . Bibcode:2019ChPhL..36g0601L. doi:10.1088/0256-307X/36/7/070601. S2CID   250853211.
  23. Quessada, A.; Kovacich, R. P.; Courtillot, I.; Clairon, A.; Santarelli, G.; Lemonde, P. (April 2, 2003). "The Dick effect for an optical frequency standard". Journal of Optics B: Quantum and Semiclassical Optics . 5 (2): S150–S154. Bibcode:2003QuSOp...5S.150Q. doi:10.1088/1464-4266/5/2/373.
  24. Westergaard, P.G.; Lodewyck, J.; Lemonde, P. (March 2010). "Minimizing the Dick effect in an optical lattice clock". IEEE Transactions on Ultrasonics, Ferroelectrics, and Frequency Control . 57 (3): 623–628. arXiv: 0909.0909 . doi:10.1109/TUFFC.2010.1457. PMID   20211780. S2CID   10581032.
  25. Cheng, P.; Sun, X.; Zhang, J.; Wang, L. (2018). "Suppression of Dick Effect in Ramsey-CPT Atomic Clock by Interleaving Lock". IEEE Transactions on Ultrasonics, Ferroelectrics, and Frequency Control . 65 (11): 2195–2200. arXiv: 1707.07559 . doi:10.1109/TUFFC.2018.2864622. PMID   30106720. S2CID   52006126.
  26. Biedermann, Grant (2008). Gravity tests, differential accelerometry and interleaved clocks with cold atom interferometers (PDF) (PhD). Stanford University.
  27. Biedermann, G.W.; Takese, K.; Wu, X.; Deslauriers, K.; Roy, S.; Kasevich, M. A. (2013). "Zero-Dead-Time Operation of Interleaved Atomic Clocks". Physical Review Letters . 341 (17): 1215–1218. Bibcode:2013PhRvL.111q0802B. doi:10.1103/PhysRevLett.111.170802. PMID   24206471.
  28. Hinkley, N.; Sherman, J. A.; Phillips, N. B.; Schioppo, M.; Lemke, N.D.; Beloy, K.; Pizzocaro, M.; Oates, C. W.; Ludlow, A. D. (2013). "An atomic clock with 10-18 instability". Science . 341 (6151): 1215–1218. arXiv: 1305.5869 . Bibcode:2013Sci...341.1215H. doi:10.1126/science.1240420. PMID   23970562. S2CID   206549862.
  29. Lin, H.; Lin, J.; Deng, J.; Zhang, S.; Wang, W. (2017). "Pulsed optically pumped atomic clock with zero-dead-time". Review of Scientific Instruments . 88 (123103): 123103. Bibcode:2017RScI...88l3103L. doi:10.1063/1.5008627. PMID   29289225.
  30. Meunier, M.; Dutta, I.; Geiger, R.; Guerlin, C.; Garrido Alzar, C.L.; Landragin, A. (December 22, 2014). "Stability enhancement by joint phase measurements in a single cold atomic fountain". Physical Review A . 90 (63633): 063633. arXiv: 1501.01943 . Bibcode:2014PhRvA..90f3633M. doi:10.1103/PhysRevA.90.063633. S2CID   26876528.
  31. Li, W.; Wu, S.; Smerzi, A.; Pezzè, L. (2022). "Improved absolute clock stability by the joint interrogation of two atomic ensembles". Physical Review A . 105 (5): 053116. arXiv: 2104.14309 . doi:10.1103/PhysRevA.105.053116. S2CID   233444191.
  32. Joyet, A.; Mileti, G.; Dudle, G.; Thomann, P. (2001). "Theoretical study of the Dick effect in a continuously operated Ramsey resonator" (PDF). IEEE Transactions on Instrumentation and Measurement. 50 (1): 150–156. Bibcode:2001ITIM...50..150J. doi:10.1109/19.903893.
  33. Devenoges, L.; Stefanov, A.; Joyet, A.; Thomann, P.; Di Domenico, G. (2012). "Improvement of the Frequency Stability Below the Dick Limit With a Continuous Atomic Fountain Clock" (PDF). IEEE Transactions on Ultrasonics, Ferroelectrics, and Frequency Control . 59 (2): 211–216. doi:10.1109/TUFFC.2012.2181. PMID   24626029. S2CID   15285482.
  34. 1 2 Fang, B.; Mielec, M.; Savole, D.; Altorio, A.; Landragin, A.; Geiger, R. (February 20, 2018). "Improving the phase response of an atom interferometer by means of temporal pulse shaping". New Journal of Physics . 20 (2): 023020. arXiv: 1712.08110 . Bibcode:2018NJPh...20b3020F. doi:10.1088/1367-2630/aaa37c. S2CID   54042350.
  35. Takamoto, M.; Takano, T.; Katori, H. (2011). "Frequency comparison of optical lattice clocks beyond the Dick limit". Nature Photonics . 5 (5): 288–292. Bibcode:2011NaPho...5..288T. doi:10.1038/nphoton.2011.34.
  36. Bize, S.; Sortais, Y.; Lemonde, P.; Zhang, S.; Laurent, P.; Santarelli, G.; Salomon, C.; Clairon, A. (September 2000). "Interrogation oscillator noise rejection in the comparison of atomic fountains". IEEE Transactions on Ultrasonics, Ferroelectrics, and Frequency Control . 47 (5): 1253–1255. doi:10.1109/58.869073. PMID   18238668. S2CID   11952522.
  37. Oelker, E.; Hutson, R.B; Kennedy, C.J.; Sonderhouse, L.; Bothwell, T.; Goban, A.; Kedar, D.; Sanner, C.; Robinson, J.M.; Marti, G.E.; Matei, D.G.; Legero, T.; Giunta, M.; Holzwarth, R.; Riehle, F.; Sterr, U.; Ye, J. (2019). "Demonstration of 4.8 × 10−17 stability at 1 s for two independent optical clocks". Nature Photonics. 13 (10): 714–719. arXiv: 1902.02741 . doi:10.1038/s41566-019-0493-4. S2CID   201255893.
  38. Lu. X.; Zhou, C.; Li, T.; Wang, Y.; Chang, H. (2020). "Synchronous frequency comparison beyond the Dick limit based on dual-excitation spectrum in an optical lattice clock". Applied Physics Letters . 117 (23): 231101. doi: 10.1063/5.0025097 .
  39. Li, L.; Qui, Q.Z.; Wang, B.; Li, T.; Zhao, J.B.; Ji, J.W.; Ren, W.; Zhao, X.; Ye, M.F; Yao, Y.Y.; Lu, D.S.; Liu, L. (2016). "Initial Tests of a Rubidium Space Cold Atom Clock". Chin Phys Lett . 33 (6): 063201. Bibcode:2016ChPhL..33f3201L. doi:10.1088/0256-307X/33/6/063201. S2CID   250872294.
  40. Derevianko, A.; Gibble, K.; Hollberg, L.; Newbury, N. R.; Oates, C.; Safronova, M. S.; Sinclair, L. C.; Yu, N. (2022). "Fundamental physics with a state-of-the-art optical clock in space". IOP Publishing . 7 (4): 044002. arXiv: 2112.10817 . Bibcode:2022QS&T....7d4002D. doi:10.1088/2058-9565/ac7df9. S2CID   245353766.
  41. Laurent, P.; Esnaut, F. X.; Gibble, K.; Peterman, P.; Lévèque, T.; Delaroche, C.; Grosjean, O.; Moric, I.; Abgrall, M.; Massonnet, D.; Salomon, C. (2020). "Qualification and frequency accuracy of the space-based primary frequency standard PHARAO" (PDF). Metrologia . 57 (5): 055005. Bibcode:2020Metro..57e5005L. doi:10.1088/1681-7575/ab948b. S2CID   219450624.
  42. Chienet, P.; Canuel, B.; Dos Santos, F. P.; Gauguet, F.; Yver=Leduc, F.; Landragin, A. (May 2, 2008). "Measurement of the Sensitivity Function in a Time-Domain Atomic Interferometer" (PDF). IEEE Transactions on Instrumentation and Measurement. 57 (6): 1141–1148. arXiv: physics/0510197 . Bibcode:2008ITIM...57.1141C. doi:10.1109/TIM.2007.915148. S2CID   5651070.
  43. Le Gouet, J.; Mehlstäubler, T.E.; Kim, J.; Merlet, S.; Clairon, A.; Landragin, A.; Pereira Dos Santos, F. (August 2008). "Limits to the sensitivity of a low noise compact atomic gravimeter". Applied Physics B . 92 (2): 133–144. arXiv: 0801.1270 . Bibcode:2008ApPhB..92..133L. doi:10.1007/s00340-008-3088-1. S2CID   52357750.
  44. Tang, B.; Zhang, B.C.; Zhou, L.; Wang, J.; Zhan, M.S. (2015). "Sensitivity function analysis of gravitational wave detection with single-laser and large-momentum-transfer atomic sensors". Research in Astronomy and Astrophysics. 15 (333): 333–347. arXiv: 1312.7652 . Bibcode:2015RAA....15..333T. doi:10.1088/1674-4527/15/3/004. S2CID   118689204.
  45. Kusch, P.; Hughes, V. W. (1959). "Atomic and Molecular Beam Spectroscopy". Atoms III — Molecules I / Atome III — Moleküle I. Encyclopedia of Physics. Vol. 7/37/1. Springer Science+Business Media. pp. 1–172. doi:10.1007/978-3-642-45917-7_1. ISBN   978-3-642-45919-1.
  46. 1 2 J. A. Barnes, A. R. Chi, L. S. Cutler, D. J. Healey, D. B. Leeson, T. E. McGunigal, J. A. Mullen, W. L. Smith, R. Sydnor, R. F. C. Vessot, G. M. R. Winkler: Characterization of Frequency Stability , NBS Technical Note 394, 1970.