Next Article in Journal
Series Arc Fault Detection Based on Multimodal Feature Fusion
Next Article in Special Issue
Compact and Fully Integrated LED Quantum Sensor Based on NV Centers in Diamond
Previous Article in Journal
Drone Detection and Tracking Using RF Identification Signals
Previous Article in Special Issue
Quantum Random Access Memory for Dummies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Commentary

Advances in Portable Atom Interferometry-Based Gravity Sensing

1
School of Physics and Astronomy, University of Birmingham, Birmingham B15 2TT, UK
2
QinetiQ, Malvern Technology Centre, St. Andrews Road, Malvern, Worcestershire WR14 3PS, UK
3
Aquark Technologies, Abbey Park Industrial Estate, Romsey SO51 9AQ, UK
4
School of Engineering, University of Birmingham, Birmingham B15 2TT, UK
*
Author to whom correspondence should be addressed.
Sensors 2023, 23(17), 7651; https://doi.org/10.3390/s23177651
Submission received: 7 July 2023 / Revised: 22 August 2023 / Accepted: 28 August 2023 / Published: 4 September 2023
(This article belongs to the Special Issue Quantum Sensors and Quantum Sensing)

Abstract

:
Gravity sensing is a valuable technique used for several applications, including fundamental physics, civil engineering, metrology, geology, and resource exploration. While classical gravimeters have proven useful, they face limitations, such as mechanical wear on the test masses, resulting in drift, and limited measurement speeds, hindering their use for long-term monitoring, as well as the need to average out microseismic vibrations, limiting their speed of data acquisition. Emerging sensors based on atom interferometry for gravity measurements could offer promising solutions to these limitations, and are currently advancing towards portable devices for real-world applications. This article provides a brief state-of-the-art review of portable atom interferometry-based quantum sensors and provides a perspective on routes towards improved sensors.
PACS:
03.75.Dg; 06.20.−f; 07.07.Df; 91.10.Op; 93.85.+q

1. Introduction

Gravity sensors are employed across various fields, including metrology [1], civil engineering [2,3], geology [4], archaeology [5], environmental monitoring [6,7,8], carbon capture and storage [9], and resource exploration [10]. Although traditional gravimeters are valuable sensors [11], they are limited by wear on their test masses, instrumental drift, microseismic vibrations, and variations in absolute values between different sensors, hindering their use for long-term observations and in arrays of multiple sensors [11,12,13]. A new generation of gravity sensors based on cold atoms, referred to as cold atom gravimeters (CAGs), could offer a solution to these limitations [14], providing consistency of measurements both spatially and temporally.
CAGs utilise cold atoms as test masses and measure gravity through atom interferometry. The first atom interferometer was at Stanford University in 1991, Kasevich et al. [15] and used to perform gravity measurements. Since this initial demonstration there has been an international wave of atom interferometry and related research [14,16], which has resulted in atom interferometers having become extremely sensitive devices achieving measurement sensitivities of up to 2.2 μGal/ H z (1 Gal = 1 cm/s 2 ) and resolutions of 0.08 μGal after integration times of 2000 s [17].
CAGs offer several advantages over classical gravimeters. They contain no macroscopic moving components, allowing for long-term continuous measurements and high sampling rates without degradation from thermal or wear effects [18]. The precision and stability of CAGs are also enhanced by the use of atom standards for reference enabling low drift. These benefits have sparked interest in using portable CAGs for various applications, such as civil engineering [19], atmospheric drag measurements [20], carbon capture and storage monitoring [21], climate change monitoring [22,23], navigation [18,23,24,25,26,27,28], geohazard monitoring [29], and military applications [30,31,32].
This article provides a brief review of the current advancements in developing portable atom interferometry-based gravity sensors designed for use in several applications, as well a providing a brief overview of different technological and physical techniques that could be used to produce improved CAGs and a more mature CAG market.

2. Operating Principle

An atom interferometer in its simplest form can be implemented by dropping a cloud of cold atoms in a vacuum [15]. These atoms are initially trapped, cooled, and prepared in a single atomic state before being allowed to fall under gravity, during which they are subjected to a sequence of three laser pulses, as shown in Figure 1.
The first laser pulse in this sequence puts the atoms into an equal superposition of two states; after a time, T, another laser pulse is applied which swaps the states, owing to atoms absorbing and emitting a photon, respectively. The associated momentum transfers make the atomic trajectories converge, such that they intersect at the same point in space after a further time T. A final laser pulse then closes the interferometer sequence, allowing for interference between the two different trajectories taken by the matter waves associated with the atoms. The gravitational acceleration, g, experienced by the atoms results in a phase shift Δ ϕ = 2 k e f f gT 2 , where k e f f is the effective wavevector of the light, which determines the momentum transfer. Δ ϕ can be read out from the atom interferometer by counting the number of excited-state atoms versus the number of ground-state atoms at the end of the sequence. The shot noise limited sensitivity for an atom interferometer is given by
Δ g g = 1 K e f f T 2 N
where N is the number of atoms that undergo the atom interferometry process.
It is possible to create a cold atom gravity gradiometer (CAGG), by utilising two atom interferometers spaced a distance apart, with the interferometry beams measuring on the same axis.
More detailed information about the underlying theory and operating principles can be found in references [18,33,34,35], while more information on the underpinning technologies can be found in references [36,37,38].

3. Progress to Date

Lab-based systems have been used in a variety of research areas including metrology and fundamental physics, where they have been used for measurements of the fine structure constant [39,40,41], the Newtonian gravitational constant [42,43], testing of general relativity [44,45], the isotropy of gravitational interaction [46,47], the equivalence principle [48,49,50,51] and in the search for new forces [52,53,54].
The success of the lab-based system has prompted the desire to transition these devices into portable sensors that can be used in a variety of applications [14,55,56]. To achieve this atom interferometers are evolving from complicated setups in research laboratories to practical, and in some cases transportable, instruments. As part of this process, the size, weight, and power (SWaP) of these sensors have had to be reduced along with improving robustness [57] to environmental and motional effects and usability, while also maintaining sufficient sensitivity to be useful in the desired application. Examples of signal sizes from different applications can be seen in Table 1.
Several portable CAGs have been developed to date, with earlier prototypes performing demonstrations outside [81,82,83] in lifts [84] or underground laboratories [85]. More recently, however, portable CAGs have been used in several real-world applications. A brief summary of the different demonstrations to date will follow.

3.1. Metrology

Within metrology, lab-based and portable CAGs have been used to better define the fundamental constants, with portable systems being used for applications where the instrument needs to be moved to work in tandem with another instrument which may be contained in a different laboratory far from the location of static CAGs. For example, in 2017, a portable CAG was used to determine the Planck constant using the LNE Kibble balance in air [86].
CAGs have participated in several comparisons of absolute gravimeters, comparing favourably with other technologies. The cold atom gravimeter (LNE-SYRTE) has been participating in international comparisons of absolute gravimeters since 2009 [87,88,89]. Another CAG participated in the first Asia-Pacific Comparison of Absolute Gravimeters, hosted by the National Institute of Metrology of China from December 2015 to March 2016 [90]. In 2017, six CAGs took part in the 10th International Comparison of Absolute Gravimeters, from institutions including the Zhejiang University of Technology, the Huazhong University of Science and Technology, the University of Science and Technology of China, the Changcheng Institute of Metrology and Measurement Beijing, the National Institute of Metrology and the Wuhan Institute of Physics and Mathematics. Several results from these sensors were accepted by the committee and showed performance comparable to classical corner cube gravimeters [91,92], showing the current generation of portable CAGs is suitable for use in metrological applications. Without the limitations of mechanical wear and with the potential for the sensitivity to improve further in the future, it is expected the use of CAGs will increase in metrological applications as they become more readily available commercially.

3.2. Environmental Monitoring

Portable CAGs have been demonstrated to be able to monitor tidal gravity variations [93,94]. For example, the CAG from the University of California [93] in 2017 was used to measure tidal gravity measurements and was capable of detecting both the effects of local and global tidal gravity variation.
CAGs have been used for seismic monitoring, for example, on the 28th of September 2013 a CAG from Zhejiang University detected a seismic wave, which originated in Pakistan from a 7.2 magnitude earthquake [95]. Other examples include the CAG from the University of California which was used to detect earthquakes in Berkeley originating from other parts of the world [93]. On the 6th of January 2019, the CAG detected a 6.6-magnitude earthquake that occurred in Indonesia at 17:27 UTC. From the measured variations in acceleration over time, the vertical component of the Rayleigh wave was determined to have a period of ∼30 s and a peak-to-peak amplitude of ∼90 μ m. These types of measurements are useful in the study and detection of earthquakes and suggest this could be a future use for CAGs.
Initial demonstrations of volcano monitoring were carried out on Mount Etna using the portable CAG from Exail Quantum Sensors, Paris, France (formerly iXblue) in 2022 [79,96]. The CAG was installed at an elevation of 800 m at the Pizzi Deneri Volcanological Observatory, approximately 2.5 km from the summit craters. During deployment, the CAG provided continuous gravity data allowing the tracking of volcano-related gravity changes with amplitudes ranging from tens to hundreds of nm s 2 over a wide range of time scales.
Environmental monitoring applications such as these have the potential to greatly improve resilience to geohazards through more accurate prediction and management of the risks. Furthermore, as performance improves, so does the potential to monitor other environmental signals such as water aquifers, allowing us to understand our changing environment and manage resources more effectively.

3.3. Small Scale Mapping for Engineering Applications

In 2022, a cold atom gravity gradiometer from the University of Birmingham was used to measure a utility tunnel under a road [97,98] with a signal-to-noise ratio of 8, locating the centre of the tunnel to within 20 cm. Such a result shows the technology is capable of having a transformative effect on reducing the risk of unforeseen ground conditions in the construction industry, as well as offering new mapping capabilities for archaeology, agriculture, natural resources, and defence capabilities.

3.4. Regional and Geological Scale Surveys

To perform larger scale regional and geological scale surveys, like those typically used in applications such as hydrology and oil prospecting, CAGs have been integrated into cars, trucks, ships, and planes.
The gravimeter GIRAFE from the Office National d’Etudes et de Recherches Aérospatiales (ONERA) has demonstrated dynamic measurements on ships, and planes. In 2018, the gravimeter achieved a shipboard measurement with a precision of 0.2–0.6 mGal under a 4 sea state condition and was used to map an area of the Meriadzec terrace located in the North Atlantic ocean [99]. In 2020, the GIRAFE system was used in an airborne campaign across Iceland, yielding gravity measurements with an estimated error of 1.7–3.9 mGal [100].
Truck-borne gravity mapping has been carried out using two different CAGs: the CAG from the University of California, Berkeley, and the CAG from Zhejiang University of Technology in Hangzhou. The former performed a gravity survey in the Berkeley Hills in 2019 achieving an uncertainty of 0.04 mGal, which allowed for the determination of subsurface rock density from the vertical gravity gradient [93]. In 2022 the CAG from Zhejiang University of Technology conducted a survey at the Xianlin Reservoir in Hangzhou. With internal and external coincidence accuracy of 35.4 μ Gal and 76.7 μ Gal, respectively, the results were verified by comparing the theoretical values obtained through forward modeling of a local high-resolution digital elevation model to the measured values, which showed good agreement [101].
In 2021, a car-mounted CAG from the Huazhong University of Science and Technology achieved the sensitivity of 1.9 mGal/ H z and the accuracy of more than 30 μ Gal in field measurements [102] while performing a survey on Yujia mountain.
In addition to these demonstrations, other CAGS are either under development [103] or have been developed, which could be used for large area mapping [104,105]. For example, the CAG from Huazhong University of Science and Technology in 2021 demonstrated operation in a moving vehicle. During this demonstration, it achieved a sensitivity of 60.88 mGal H z with T = 5 ms [104] and could be used in the future for large-area gravity mapping.
It is clear that portable CAGs are progressing to devices capable of precise measurements while in motion and will enable their use in applications such as hydrology, oil and mineral prospecting, as well as navigation.

3.5. Space Based Systems

CAGs for use in space are currently under development [106,107,108,109,110] for several applications such as global gravity field mapping and fundamental physics [26]. For example, in fundamental physics CAGs are being developed with the aim to test the Weak Equivalence Principle [111,112]. While there is a lot of work in developing CAGS for use in space, there have been some existing cold atom systems demonstrated in space [113]. The first interference experiments were performed in 2017 during a space flight on the MAIUS-1 rocket [114]. Later in May 2018, NASA’s Cold Atom Laboratory (CAL) was launched to the International Space Station and has been operating onboard since then [115]. CAL is a quantum facility for studying ultra-cold gases in a microgravity environment and is being used to perform research in a force-free environment inaccessible to terrestrial laboratories, allowing for greater T times to be realised than practical in ground-based experiments.

4. Towards Improved Sensors and a Mature Commercial Market

4.1. The Commercial Market for CAGs

The current generation of portable CAGs is already sensitive and robust enough to be used in some applications with several systems already commercially available. Most notably, the systems from Exail (formerly iXblue) [94,116] made a massive contribution towards defining and accelerating the global quantum sensing market. Other commercial sensors are under development at companies including M Squared lasers [117], AOSense, Inc. [118,119], Mugaltech [120] and CASColdAtom [121]. Valued at $474.06 million in 2022, the market is expected to grow to over $740 million by 2028, with gravity applications being one of the largest market segments next to timing [122]. Whilst the market advances in a stable way with a predicted compound annual growth rate of 7.8%, there is a significant potential for further technological development and with it, significant market growth.
For example, making this technology robust enough to survive in harsh environments will enable quantum gravimeters to access the borehole market for geothermal applications. This is a fast-growing market valued in 2021 at $5.3 billion and is expected to grow to over $7 billion by 2030 [123]. Enabling long time continuous use of these quantum gravimeters together with potential large-area scanning capabilities are another advancement that can enable additional markets such as carbon capture and storage (>$4 billion [124]) and monitoring of geohazards (>$680 million [125]). Ultimately, increasing the performance, robustness, and ease of use of these systems could have a major impact on the hydrocarbon market. This is one of the largest markets this technology can address, valued at over $71 trillion in 2022 [126].
Overall, the opportunities enabled by commercial advances of CAGs are significant and by tailoring subcomponents of the system to address the size, weight, power, cost, user accessibility, and performance the technology could see a significant acceleration in market adoption.

4.2. Technological Routes to Improved Sensors

To facilitate the current generation of portable CAGs, several simplifications compared to lab systems and innovations have been required to meet the required robustness and SWaP. For example, to simplify the laser system, some CAGs have simply dropped atoms after the laser cooling and trapping stage, rather than utilise a fountain launch, which is commonly used in lab systems. This simplifies the laser system by requiring a single frequency of cooling light rather than the two or more frequencies needed for a fountain launch. To simplify the laser and vacuum systems, novel atom cooling and trapping geometries have enabled a reduction in system SWAP through a reduction in optical complexity compared with traditional 6-beam magneto-optical traps (MOTs). For example, pyramid MOTs [94,127], prism MOTs [97], mirror MOTs [128], and grating MOTs [129,130,131] all enable magneto-optical trapping with fewer input beams than traditional six beam MOTs. Several systems have also reduced the vacuum system and optical system size by using the same optical path between the MOT beams and interferometry beams [94,97,127].
It is expected that improvements in sensitivity, measurement speed, robustness, usability, autonomy as well as reduced size, weight, and power requirements will be achieved in the next generation of portable GAGs through the use of the latest and future techniques [132,133,134], technological developments [37,135,136,137,138,139,140,141,142], improved sub-components [143], systems engineering [144,145], and optimisation of sensor design [146,147]. Some examples of technology developments, along with the expected benefits, are highlighted in Table 2.

4.3. Physics Routes to Improved Sensors

While technological developments are key to bringing CAGs to market, advancing in the understanding and ability to manipulate cold atoms is also key to reaching the full potential of CAGs. In particular, realising techniques to increase the sensitivity of CAGs will be important to realising as small a form factor as possible.
The minimum length of a quantum gravity sensor is limited by the T time of the instrument, requiring a drop distance to allow the atomic wavepackets to evolve in time. As instruments are designed for smaller applications, to maintain or further improve the instrument sensitivity, one key technique that will need to be demonstrated in portable systems is large momentum transfer (LMT), which by applying sequential light pulses augments the Mach–Zender type interferometer by transferring multiple photon momenta to the cloud, increasing the shot noise limit sensitivity proportionally to the photon momenta [148,149,150,151]. In principle, devices can operate at reduced T time to achieve the same sensitivity, which would result in more compact instruments. Demonstrations in laboratory systems have shown in excess of 400 k of photon momentum in an interferometry sequence, using optimised pulse schemes to achieve high fidelity atom optics [152], which if implemented in portable gravimeters brings the sensitivity of atom interferometers far higher than their classical counterparts. Large momentum transfer is a key requirement for future atom interferometry dark matter and gravitational wave detectors, where photon momenta of > 10 3 are required [153,154].
Techniques developed in achieving high fidelity atom optics from the fundamental physics applications should be applicable to future portable devices since the techniques are generally around light pulse engineering of the frequency, phase, and amplitude [155,156,157,158]. These schemes can introduce robustness to parameters such as detuning or intensity, reducing the effect of laser system noise on the atom interferometer. Some of these pulse schemes are applicable in LMT, and so may be implemented together in portable systems.
For high data rate applications, such as operation on moving platforms, single shot atom interferometry measurements will be beneficial to increase the spatial resolution of CAGs. Techniques such as phase shear have been used to read out the entire interferometer fringe in a single image by applying a tilt to the final interferometer pulse, creating a spatially varying interference pattern [133].
To reduce the power requirements of the laser system, cavity-enhanced atom interferometry may be useful in future portable CAGs, where atom interferometry occurs within an optical cavity [159]. Cavities are more stable and technologically simpler when smaller, which may lend themselves to compact sensor development. For larger and larger cavity systems, the cavity begins to hinder the atom interferometry pulses [159,160], requiring complicated schemes to overcome the cavity limits [161,162].
It is expected that several of the physics improvements described here will not be present in the first generation of portable systems, but in later generations of portable CAGs.
Table 2. Examples of technological developments which could be used to improve portable CAGs and their expected benefits.
Table 2. Examples of technological developments which could be used to improve portable CAGs and their expected benefits.
Technological DevelopmentSizeWeightPowerNoise and/or Bias ReductionRobustnessRefs.
Additive manufacturing
(e.g., 3D printing)
[135,136,137,163,164]
Beam shaping (e.g, Top hat) [139]
Clean atom sources [165,166,167]
Compact 2D MOTs [168,169]
Compact laser systems [170,171]
Micro-fabricated components [172]
Metasurfaces [173,174]
Compact spectroscopy cells [175,176,177,178]
Optimised Coil systems[140,179,180]
Optimised electronics[181,182]
Passive vacuum systems [132,183]
Vacuum compatible anti reflection coatings [138]
Vibration Compensation [104,142,184]

5. Conclusions

Portable atom interferometry-based gravity and gravity gradient sensors are becoming more prevalent, with some commercial devices entering the market [94,116,121] and many university systems under development [25,185]. While the current generation of portable CAGs is sensitive and robust enough for use in some applications, there is significant potential for further development. It is expected that improvements in sensitivity, robustness, and usability, as well as having reduced size, weight, and power requirements, will be achieved in future systems. These improvements will not only enhance the performance of the sensors but also enable their operation in new applications and on new platforms, such as unstaffed aerial vehicles [103,138], trains [25], cube satellites [186], and down boreholes [187].

Author Contributions

Conceptualization, J.V.; investigation, J.V., B.S., D.B. and A.D.; writing—original draft preparation, J.V., B.S., D.B. and A.D.; writing—review and editing, J.V., D.B. and A.D.; visualization, J.V.; supervision, J.V.; project administration, J.V.; funding acquisition, J.V. and A.D. All authors have read and agreed to the published version of the manuscript.

Funding

We acknowledge support from EPSRC through grant EP/T001046/1 and Innovate uk through grant 10032699 as part of the UK National Quantum Technologies Programme.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We would like to acknowledge Adam Seedat, Yu-hung Lien and Andrew Hinton for useful discussions during the preparation of this article.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
CAGCold atom gravimeter
CAGGCold atom gravity gradiometer
MOTMagneto-optical trap
UTCUniversal Time Coordinated

References

  1. Wood, B.M.; Sanchez, C.A.; Green, R.G.; Liard, J.O. A summary of the Planck constant determinations using the NRC Kibble balance. Metrologia 2017, 54, 399. [Google Scholar] [CrossRef]
  2. Tuckwell, G.; Grossey, T.; Owen, S.; Stearns, P. The use of microgravity to detect small distributed voids and low-density ground. Q. J. Eng. Geol. Hydrogeol. 2008, 41, 371–380. [Google Scholar] [CrossRef]
  3. Porzucek, S.; Loj, M. Microgravity survey to detect voids and loosening zones in the vicinity of the mine shaft. Energies 2021, 14, 3021. [Google Scholar] [CrossRef]
  4. Saadi, N.M.; Aboud, E.; Saibi, H.; Watanabe, K. Integrating data from remote sensing, geology and gravity for geological investigation in the Tarhunah area, Northwest Libya. Int. J. Digit. Earth 2008, 1, 347–366. [Google Scholar] [CrossRef]
  5. Bishop, I.; Styles, P.; Emsley, S.; Ferguson, N. The detection of cavities using the microgravity technique: Case histories from mining and karstic environments. Geol. Soc. Lond. Eng. Geol. Spec. Publ. 1997, 12, 153–166. [Google Scholar] [CrossRef]
  6. Montagner, J.P.; Juhel, K.; Barsuglia, M.; Ampuero, J.P.; Chassande-Mottin, E.; Harms, J.; Whiting, B.; Bernard, P.; Clévédé, E.; Lognonné, P. Prompt gravity signal induced by the 2011 Tohoku-Oki earthquake. Nat. Commun. 2016, 7, 13349. [Google Scholar] [CrossRef] [PubMed]
  7. Chen, J.; Li, J.; Zhang, Z.; Ni, S. Long-term groundwater variations in Northwest India from satellite gravity measurements. Glob. Planet. Chang. 2014, 116, 130–138. [Google Scholar] [CrossRef]
  8. Van Camp, M.; de Viron, O.; Watlet, A.; Meurers, B.; Francis, O.; Caudron, C. Geophysics from Terrestrial Time-Variable Gravity Measurements. Rev. Geophys. 2017, 55, 938–992. [Google Scholar] [CrossRef]
  9. Appriou, D.; Bonneville, A.; Zhou, Q.; Gasperikova, E. Time-lapse gravity monitoring of CO2 migration based on numerical modeling of a faulted storage complex. Int. J. Greenh. Gas Control 2020, 95, 102956. [Google Scholar] [CrossRef]
  10. Martinez, C.; Li, Y.; Krahenbuhl, R.; Braga, M.A. 3D inversion of airborne gravity gradiometry data in mineral exploration: A case study in the Quadrilátero Ferrífero, Brazil. Geophysics 2013, 78, B1–B11. [Google Scholar] [CrossRef]
  11. Marson, I. A short walk along the gravimeters path. Int. J. Geophys. 2012, 2012, 687813. [Google Scholar] [CrossRef]
  12. Francis, O. Performance assessment of the relative gravimeter Scintrex CG-6. J. Geod. 2021, 95, 116. [Google Scholar] [CrossRef]
  13. Middlemiss, R.; Samarelli, A.; Paul, D.; Hough, J.; Rowan, S.; Hammond, G. Measurement of the Earth tides with a MEMS gravimeter. Nature 2016, 531, 614–617. [Google Scholar] [CrossRef]
  14. Bongs, K.; Holynski, M.; Vovrosh, J.; Bouyer, P.; Condon, G.; Rasel, E.; Schubert, C.; Schleich, W.P.; Roura, A. Taking atom interferometric quantum sensors from the laboratory to real-world applications. Nat. Rev. Phys. 2019, 1, 731–739. [Google Scholar] [CrossRef]
  15. Kasevich, M.; Chu, S. Atomic interferometry using stimulated Raman transitions. Phys. Rev. Lett. 1991, 67, 181–184. [Google Scholar] [CrossRef] [PubMed]
  16. Dutta, P.; Maurya, S.S.; Patel, K.; Biswas, K.; Mangaonkar, J.; Sarkar, S.; D. Rapol, U. A decade of advancement of quantum sensing and metrology in India using cold atoms and ions. J. Indian Inst. Sci. 2022, 103, 609–632. [Google Scholar] [CrossRef]
  17. Zhang, T.; Chen, L.L.; Shu, Y.B.; Xu, W.J.; Cheng, Y.; Luo, Q.; Hu, Z.K.; Zhou, M.K. Ultrahigh-Sensitivity Bragg Atom Gravimeter and its Application in Testing Lorentz Violation. Phys. Rev. Appl. 2023, 20, 014067. [Google Scholar] [CrossRef]
  18. Narducci, F.A.; Black, A.T.; Burke, J.H. Advances toward fieldable atom interferometers. Adv. Phys. X 2022, 7, 1946426. [Google Scholar] [CrossRef]
  19. Boddice, D.; Metje, N.; Tuckwell, G. Capability assessment and challenges for quantum technology gravity sensors for near surface terrestrial geophysical surveying. J. Appl. Geophys. 2017, 146, 149–159. [Google Scholar] [CrossRef]
  20. Riou, I.; Chichet, L.; Edwards, I.; Kiss-Toth, M.; Reilly, J.; Renshaw, R. Embedded cold atom accelerometer for atmospheric drag measurement. Proc. SPIE 2022, 12093, 1209305. [Google Scholar]
  21. Ridley, K.; de Villiers, G.; Vovrosh, J.; Vincent, C.; Wilkinson, P.; Holynski, M. Quantum Technology Based Gravity and Gravity Gradiometry as a Tool for CCS Monitoring and Investigation; Elsevier: Amsterdam, The Netherlands, 2022. [Google Scholar]
  22. Berger, C.; Di Paolo, A.; Forrest, T.; Hadfield, S.; Sawaya, N.; Stęchły, M.; Thibault, K. Quantum technologies for climate change: Preliminary assessment. arXiv 2021, arXiv:2107.05362. [Google Scholar]
  23. Phillips, A.M.; Wright, M.J.; Kiss-Toth, M.; Read, I.; Riou, I.; Maddox, S.; Maskell, S.; Ralph, J.F. Augmented inertial navigation using cold atom sensing. Cold At. Quantum Technol. 2020, 11578, 115780C. [Google Scholar]
  24. Wu, L.; Wang, H.; Chai, H.; Zhang, L.; Hsu, H.; Wang, Y. Performance evaluation and analysis for gravity matching aided navigation. Sensors 2017, 17, 769. [Google Scholar] [CrossRef] [PubMed]
  25. Adams, B.; Macrae, C.; Entezami, M.; Ridley, K.; Kubba, A.; Lien, Y.H.; Kinge, S.; Bongs, K. The development of a High data rate atom interferometric gravimeter (HIDRAG) for gravity map matching navigation. In Proceedings of the 2021 IEEE International Symposium on Inertial Sensors and Systems (INERTIAL), Kailua-Kona, HI, USA, 22–25 March 2021; pp. 1–4. [Google Scholar]
  26. Phillips, A.M.; Wright, M.J.; Riou, I.; Maddox, S.; Maskell, S.; Ralph, J.F. Position fixing with cold atom gravity gradiometers. AVS Quantum Sci. 2022, 4, 024404. [Google Scholar] [CrossRef]
  27. Welker, T.C.; Pachter, M.; Huffman, R.E. Gravity gradiometer integrated inertial navigation. In Proceedings of the 2013 European Control Conference (ECC), Zurich, Switzerland, 17–19 July 2013; IEEE: Piscataway, NJ, USA, 2013; pp. 846–851. [Google Scholar]
  28. Wan, X.; Wu, Y.; Guo, H.; Li, M. Development Status and Influencing Factor Analysis of Underwater Matching Navigation Based on Gravity Field Products. 武汉大学学报 (信息科学版) 2023, 48, 879–890. [Google Scholar]
  29. Clare, M.A.; Vardy, M.E.; Cartigny, M.J.; Talling, P.J.; Himsworth, M.D.; Dix, J.K.; Harris, J.M.; Whitehouse, R.J.; Belal, M. Direct monitoring of active geohazards: Emerging geophysical tools for deep-water assessments. Near Surf. Geophys. 2017, 15, 427–444. [Google Scholar] [CrossRef]
  30. Krelina, M. Quantum technology for military applications. EPJ Quantum Technol. 2021, 8, 24. [Google Scholar] [CrossRef]
  31. Inglesant, P.; Jirotka, M.; Hartswood, M. Responsible Innovation in Quantum Technologies Applied to Defence and National Security; NQIT (Networked Quantum Information Technologies): Oxford, UK, 2018. [Google Scholar]
  32. Kubiak, K. Quantum Technology and Submarine Near-Invulnerability; Global Security Policy Brief; European Leadership Network: London, UK, 2020. [Google Scholar]
  33. Berman, P.R. Atom Interferometry; Academic Press: Cambridge, MA, USA, 1997. [Google Scholar]
  34. Foot, C.J. Atomic Physics; OUP Oxford: Oxford, UK, 2004; Volume 7. [Google Scholar]
  35. Foot, C. Laser cooling and trapping of atoms. Contemp. Phys. 1991, 32, 369–381. [Google Scholar] [CrossRef]
  36. Geiger, R.; Landragin, A.; Merlet, S.; Pereira Dos Santos, F. High-accuracy inertial measurements with cold-atom sensors. AVS Quantum Sci. 2020, 2, 024702. [Google Scholar] [CrossRef]
  37. Rushton, J.; Aldous, M.; Himsworth, M. Contributed review: The feasibility of a fully miniaturized magneto-optical trap for portable ultracold quantum technology. Rev. Sci. Instrum. 2014, 85, 121501. [Google Scholar] [CrossRef]
  38. de Angelis, M.; Angonin, M.; Beaufils, Q.; Becker, C.; Bertoldi, A.; Bongs, K.; Bourdel, T.; Bouyer, P.; Boyer, V.; Dörscher, S.; et al. iSense: A Portable Ultracold-Atom-Based Gravimeter. Procedia Comput. Sci. 2011, 7, 334–336. [Google Scholar] [CrossRef]
  39. Morel, L.; Yao, Z.; Cladé, P.; Guellati-Khélifa, S. Determination of the fine-structure constant with an accuracy of 81 parts per trillion. Nature 2020, 588, 61–65. [Google Scholar] [CrossRef]
  40. Bouchendira, R.; Cladé, P.; Guellati-Khélifa, S.; Nez, F.; Biraben, F. New Determination of the Fine Structure Constant and Test of the Quantum Electrodynamics. Phys. Rev. Lett. 2011, 106, 080801. [Google Scholar] [CrossRef]
  41. Parker, R.H.; Yu, C.; Zhong, W.; Estey, B.; Müller, H. Measurement of the fine-structure constant as a test of the Standard Model. Science 2018, 360, 191–195. [Google Scholar] [CrossRef] [PubMed]
  42. Fixler, J.B.; Foster, G.T.; McGuirk, J.M.; Kasevich, M.A. Atom Interferometer Measurement of the Newtonian Constant of Gravity. Science 2007, 315, 74–77. [Google Scholar] [CrossRef] [PubMed]
  43. Rosi, G.; Sorrentino, F.; Cacciapuoti, L.; Prevedelli, M.; Tino, G. Precision measurement of the Newtonian gravitational constant using cold atoms. Nature 2014, 510, 518–521. [Google Scholar] [CrossRef]
  44. Dimopoulos, S.; Graham, P.W.; Hogan, J.M.; Kasevich, M.A. Testing general relativity with atom interferometry. Phys. Rev. Lett. 2007, 98, 111102. [Google Scholar] [CrossRef]
  45. Asenbaum, P.; Overstreet, C.; Kovachy, T.; Brown, D.D.; Hogan, J.M.; Kasevich, M.A. Phase shift in an atom interferometer due to spacetime curvature across its wave function. Phys. Rev. Lett. 2017, 118, 183602. [Google Scholar] [CrossRef]
  46. Müller, H.; Chiow, S.w.; Herrmann, S.; Chu, S.; Chung, K.Y. Atom-Interferometry Tests of the Isotropy of Post–Newtonian Gravity. Phys. Rev. Lett. 2008, 100, 031101. [Google Scholar] [CrossRef] [PubMed]
  47. Duan, X.C.; Deng, X.B.; Zhou, M.K.; Zhang, K.; Xu, W.J.; Xiong, F.; Xu, Y.Y.; Shao, C.G.; Luo, J.; Hu, Z.K. Test of the Universality of Free Fall with Atoms in Different Spin Orientations. Phys. Rev. Lett. 2016, 117, 023001. [Google Scholar] [CrossRef] [PubMed]
  48. Tarallo, M.G.; Mazzoni, T.; Poli, N.; Sutyrin, D.; Zhang, X.; Tino, G. Test of Einstein equivalence principle for 0-spin and half-integer-spin atoms: Search for spin-gravity coupling effects. Phys. Rev. Lett. 2014, 113, 023005. [Google Scholar] [CrossRef] [PubMed]
  49. Barrett, B.; Antoni-Micollier, L.; Chichet, L.; Battelier, B.; Gominet, P.; Bertoldi, A.; Bouyer, P.; Landragin, A. Correlative methods for dual-species quantum tests of the weak equivalence principle. New J. Phys. 2015, 17, 085010. [Google Scholar] [CrossRef]
  50. Geiger, R.; Trupke, M. Proposal for a quantum test of the weak equivalence principle with entangled atomic species. Phys. Rev. Lett. 2018, 120, 043602. [Google Scholar] [CrossRef] [PubMed]
  51. Asenbaum, P.; Overstreet, C.; Kim, M.; Curti, J.; Kasevich, M.A. Atom-Interferometric Test of the Equivalence Principle at the 10−12 Level. Phys. Rev. Lett. 2020, 125, 191101. [Google Scholar] [CrossRef]
  52. Jaffe, M.; Haslinger, P.; Xu, V.; Hamilton, P.; Upadhye, A.; Elder, B.; Khoury, J.; Müller, H. Testing sub-gravitational forces on atoms from a miniature in-vacuum source mass. Nat. Phys. 2017, 13, 938–942. [Google Scholar] [CrossRef]
  53. Haslinger, P.; Jaffe, M.; Xu, V.; Schwartz, O.; Sonnleitner, M.; Ritsch-Marte, M.; Ritsch, H.; Müller, H. Attractive force on atoms due to blackbody radiation. Nat. Phys. 2018, 14, 257–260. [Google Scholar] [CrossRef]
  54. Sabulsky, D.O.; Dutta, I.; Hinds, E.A.; Elder, B.; Burrage, C.; Copeland, E.J. Experiment to Detect Dark Energy Forces Using Atom Interferometry. Phys. Rev. Lett. 2019, 123, 061102. [Google Scholar] [CrossRef]
  55. Vovrosh, J.; Lien, Y.H. Applications of Cold-Atom-Based Quantum Technology. Atoms 2022, 10, 30. [Google Scholar] [CrossRef]
  56. Zhong, J.; Tang, B.; Chen, X.; Zhou, L. Quantum gravimetry going toward real applications. Innovation 2022, 3, 100230. [Google Scholar] [CrossRef] [PubMed]
  57. Lellouch, S.; Bongs, K.; Holynski, M. Using atom interferometry to measure gravity. Contemp. Phys. 2023, 63, 138–155. [Google Scholar] [CrossRef]
  58. Pašteka, R.; Pánisová, J.; Zahorec, P.; Papčo, J.; Mrlina, J.; Fraštia, M.; Vargemezis, G.; Kušnirák, D.; Zvara, I. Microgravity method in archaeological prospection: Methodical comments on selected case studies from crypt and tomb detection. Archaeol. Prospect. 2020, 27, 415–431. [Google Scholar] [CrossRef]
  59. PÁNISOVÁ, J.; PAŠTEKA, R. The use of microgravity technique in archaeology: A case study from the St. Nicolas Church in Pukanec, Slovakia. Contrib. Geophys. Geod. 2009, 39, 237–254. [Google Scholar] [CrossRef]
  60. Gasperikova, E.; Hoversten, G. Gravity monitoring of CO2 movement during sequestration: Model studies. Geophysics 2008, 73, WA105–WA112. [Google Scholar] [CrossRef]
  61. Chadwick, R. Measurement and monitoring technologies for verification of carbon dioxide (CO2) storage in underground reservoirs. In Developments and Innovation in Carbon Dioxide (CO2) Capture and Storage Technology; Woodhead Publishing: Sawston, UK, 2010; Volume 2, pp. 203–239. [Google Scholar]
  62. Nooner, S.L.; Eiken, O.; Hermanrud, C.; Sasagawa, G.S.; Stenvold, T.; Zumberge, M.A. Constraints on the in situ density of CO2 within the Utsira formation from time-lapse seafloor gravity measurements. Int. J. Greenh. Gas Control 2007, 1, 198–214. [Google Scholar] [CrossRef]
  63. Braitenberg, C.; Sampietro, D.; Pivetta, T.; Zuliani, D.; Barbagallo, A.; Fabris, P.; Rossi, L.; Fabbri, J.; Mansi, A.H. Gravity for detecting caves: Airborne and terrestrial simulations based on a comprehensive karstic cave benchmark. Pure Appl. Geophys. 2016, 173, 1243–1264. [Google Scholar] [CrossRef]
  64. Detecting gypsum caves with microgravity and ERT under soil water content variations (Sorbas, SE Spain). Eng. Geol. 2015, 193, 38–48. [CrossRef]
  65. Styles, P.; McGrath, R.; Thomas, E.; Cassidy, N. The use of microgravity for cavity characterization in karstic terrains. Q. J. Eng. Geol. Hydrogeol. 2005, 38, 155–169. [Google Scholar] [CrossRef]
  66. Prasad, A.; Middlemiss, R.P.; Noack, A.; Anastasiou, K.; Bramsiepe, S.G.; Toland, K.; Utting, P.R.; Paul, D.J.; Hammond, G.D. A 19 day earth tide measurement with a MEMS gravimeter. Sci. Rep. 2022, 12, 13091. [Google Scholar] [CrossRef]
  67. Boddice, D.; Atkins, P.; Rodgers, A.; Metje, N.; Goncharenko, Y.; Chapman, D. A novel approach to reduce environmental noise in microgravity measurements using a Scintrex CG5. J. Appl. Geophys. 2018, 152, 221–235. [Google Scholar] [CrossRef]
  68. Zhang, M.; Liu, Z.; Wu, Q.; Teng, Y.; Zhang, X.; Du, F.; Jiang, Y. Hydrologic changes of in situ gravimetry. Geophysics 2022, 87, B117–B127. [Google Scholar] [CrossRef]
  69. Creutzfeldt, B.; Güntner, A.; Vorogushyn, S.; Merz, B. The benefits of gravimeter observations for modelling water storage changes at the field scale. Hydrol. Earth Syst. Sci. 2010, 14, 1715–1730. [Google Scholar] [CrossRef]
  70. Gettings, P.; Hurlow, H.; Chapman, D.; Harris, R. Gravity Monitoring of the Weber River Aquifer Storage Project. In AGU Fall Meeting Abstracts; American Geophysical Union: Washington, DC, USA, 2004. [Google Scholar]
  71. Pringle, J.; Stimpson, I.; Toon, S.; Caunt, S.; Lane, V.; Husband, C.; Jones, G.; Cassidy, N.; Styles, P. Geophysical characterization of derelict coalmine workings and mineshaft detection: A case study from Shrewsbury, United Kingdom. Near Surf. Geophys. 2008, 6, 185–194. [Google Scholar] [CrossRef]
  72. Benson, R.; Kaufmann, R.; Yuhr, L.; Hopkins, R. Locating and characterizing abandoned mines using microgravity. In Proceedings of the Geophysical Technologies for Detecting Underground Coal Mine Voids Forum, Lexington, KY, USA, 28–30 July 2003. [Google Scholar]
  73. Pringle, J.K.; Styles, P.; Howell, C.P.; Branston, M.W.; Furner, R.; Toon, S.M. Long-term time-lapse microgravity and geotechnical monitoring of relict salt mines, Marston, Cheshire, UK. Geophysics 2012, 77, B287–B294. [Google Scholar] [CrossRef]
  74. Pazzi, V.; Di Filippo, M.; Di Nezza, M.; Carlà, T.; Bardi, F.; Marini, F.; Fontanelli, K.; Intrieri, E.; Fanti, R. Integrated geophysical survey in a sinkhole-prone area: Microgravity, electrical resistivity tomographies, and seismic noise measurements to delimit its extension. Eng. Geol. 2018, 243, 282–293. [Google Scholar] [CrossRef]
  75. Kobe, M.; Gabriel, G.; Weise, A.; Vogel, D. Time-lapse gravity and levelling surveys reveal mass loss and ongoing subsidence in the urban subrosion-prone area of Bad Frankenhausen, Germany. Solid Earth 2019, 10, 599–619. [Google Scholar] [CrossRef]
  76. Cuss, R.J.; Styles, P. The application of microgravity in industrial archaeology: An example from the Williamson tunnels, Edge Hill, Liverpool. Geol. Soc. Lond. Spec. Publ. 1999, 165, 41–59. [Google Scholar] [CrossRef]
  77. Wilson, S.S.; Crawford, N.C.; Croft, L.A.; Howard, M.; Miller, S.; Rippy, T. Autonomous robot for detecting subsurface voids and tunnels using microgravity. In Sensors, and Command, Control, Communications, and Intelligence (C3I) Technologies for Homeland Security and Homeland Defense V; Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series; SPIE: Bellingham, WA, USA, 2006; Volume 6021, p. 62011. [Google Scholar]
  78. Rymer, H.; Brown, G. Gravity fields and the interpretation of volcanic structures: Geological discrimination and temporal evolution. J. Volcanol. Geotherm. Res. 1986, 27, 229–254. [Google Scholar] [CrossRef]
  79. Carbone, D.; Antoni-Micollier, L.; Hammond, G.; de Zeeuw van Dalfsen, E.; Rivalta, E.; Bonadonna, C.; Messina, A.; Lautier-Gaud, J.; Toland, K.; Koymans, M.; et al. The NEWTON-g Gravity Imager: Toward New Paradigms for Terrain Gravimetry. Front. Earth Sci. 2020, 8, 573396. [Google Scholar] [CrossRef]
  80. Sainz-Maza Aparicio, S.; Arnoso Sampedro, J.; González Montesinos, F.; Martí Molist, J. Volcanic signatures in time gravity variations during the volcanic unrest on El Hierro (Canary Islands). J. Geophys. Res. Solid Earth 2014, 119, 5033–5051. [Google Scholar] [CrossRef]
  81. Schmidt, M.; Senger, A.; Hauth, M.; Freier, C.; Schkolnik, V.; Peters, A. A mobile high-precision absolute gravimeter based on atom interferometry. Gyroscopy Navig. 2011, 2, 170–177. [Google Scholar] [CrossRef]
  82. Hauth, M.; Freier, C.; Schkolnik, V.; Senger, A.; Schmidt, M.; Peters, A. First gravity measurements using the mobile atom interferometer GAIN. Appl. Phys. B 2013, 113, 49–55. [Google Scholar] [CrossRef]
  83. Hinton, A.; Perea-Ortiz, M.; Winch, J.; Briggs, J.; Freer, S.; Moustoukas, D.; Powell-Gill, S.; Squire, C.; Lamb, A.; Rammeloo, C.; et al. A portable magneto-optical trap with prospects for atom interferometry in civil engineering. Philos. Trans. R. Soc. Math. Phys. Eng. Sci. 2017, 375, 20160238. [Google Scholar] [CrossRef]
  84. Bidel, Y.; Carraz, O.; Charrière, R.; Cadoret, M.; Zahzam, N.; Bresson, A. Compact cold atom gravimeter for field applications. Appl. Phys. Lett. 2013, 102, 144107. [Google Scholar] [CrossRef]
  85. Farah, T.; Guerlin, C.; Landragin, A.; Bouyer, P.; Gaffet, S.; Pereira Dos Santos, F.; Merlet, S. Underground operation at best sensitivity of the mobile LNE-SYRTE cold atom gravimeter. Gyroscopy Navig. 2014, 5, 266–274. [Google Scholar] [CrossRef]
  86. Thomas, M.; Ziane, D.; Pinot, P.; Karcher, R.; Imanaliev, A.; Dos Santos, F.P.; Merlet, S.; Piquemal, F.; Espel, P. A determination of the Planck constant using the LNE Kibble balance in air. Metrologia 2017, 54, 468. [Google Scholar] [CrossRef]
  87. Jiang, Z.; Pálinkáš, V.; Arias, F.; Liard, J.; Merlet, S.; Wilmes, H.; Vitushkin, L.; Robertsson, L.; Tisserand, L.; Dos Santos, F.P.; et al. The 8th International Comparison of Absolute Gravimeters 2009: The first Key Comparison (CCM. G-K1) in the field of absolute gravimetry. Metrologia 2012, 49, 666. [Google Scholar] [CrossRef]
  88. Francis, O.; Baumann, H.; Volarik, T.; Rothleitner, C.; Klein, G.; Seil, M.; Dando, N.; Tracey, R.; Ullrich, C.; Castelein, S.; et al. The European comparison of absolute gravimeters 2011 (ECAG-2011) in Walferdange, Luxembourg: Results and recommendations. Metrologia 2013, 50, 257. [Google Scholar] [CrossRef]
  89. Gillot, P.; Francis, O.; Landragin, A.; Dos Santos, F.P.; Merlet, S. Stability comparison of two absolute gravimeters: Optical versus atomic interferometers. Metrologia 2014, 51, L15. [Google Scholar] [CrossRef]
  90. Fu, Z.; Wang, Q.; Wang, Z.; Wu, B.; Cheng, B.; Lin, Q. Participation in the absolute gravity comparison with a compact cold atom gravimeter. Chin. Opt. Lett. 2019, 17, 011204. [Google Scholar]
  91. Wu, S.; Feng, J.; Li, C.; Su, D.; Wang, Q.; Hu, R.; Hu, L.; Xu, J.; Ji, W.; Ullrich, C.; et al. The results of CCM.G-K2.2017 key comparison. Metrologia 2020, 57, 07002. [Google Scholar]
  92. Wu, S.; Feng, J.; Li, C.; Su, D.; Wang, Q.; Hu, R.; Mou, L. The results of 10th International Comparison of Absolute Gravimeters (ICAG-2017). J. Geod. 2021, 95, 63. [Google Scholar] [CrossRef]
  93. Wu, X.; Pagel, Z.; Malek, B.S.; Nguyen, T.H.; Zi, F.; Scheirer, D.S.; Müller, H. Gravity surveys using a mobile atom interferometer. Sci. Adv. 2019, 5, eaax0800. [Google Scholar] [CrossRef] [PubMed]
  94. Ménoret, V.; Vermeulen, P.; Le Moigne, N.; Bonvalot, S.; Bouyer, P.; Landragin, A.; Desruelle, B. Gravity measurements below 10−9 g with a transportable absolute quantum gravimeter. Sci. Rep. 2018, 8, 12300. [Google Scholar] [CrossRef] [PubMed]
  95. Wu, B.; Wang, Z.; Cheng, B.; Wang, Q.; Xu, A.; Lin, Q. The investigation of a μGal-level cold atom gravimeter for field applications. Metrologia 2014, 51, 452. [Google Scholar] [CrossRef]
  96. Antoni-Micollier, L.; Carbone, D.; Ménoret, V.; Lautier-Gaud, J.; King, T.; Greco, F.; Messina, A.; Contrafatto, D.; Desruelle, B. Detecting Volcano-Related Underground Mass Changes with a Quantum Gravimeter. Geophys. Res. Lett. 2022, 49, e2022GL097814. [Google Scholar] [CrossRef]
  97. Stray, B.; Lamb, A.; Kaushik, A.; Vovrosh, J.; Rodgers, A.; Winch, J.; Hayati, F.; Boddice, D.; Stabrawa, A.; Niggebaum, A.; et al. Quantum sensing for gravity cartography. Nature 2022, 602, 590–594. [Google Scholar] [CrossRef]
  98. Vovrosh, J.; Boddice, D.; Holynski, M. Using the quantum properties of atoms to reveal what is underground. TheScienceBreaker 2023, 9. [Google Scholar] [CrossRef]
  99. Bidel, Y.; Zahzam, N.; Blanchard, C.; Bonnin, A.; Cadoret, M.; Bresson, A.; Rouxel, D.; Lequentrec-Lalancette, M.F. Absolute marine gravimetry with matter-wave interferometry. Nat. Commun. 2018, 9, 627. [Google Scholar] [CrossRef]
  100. Bidel, Y.; Zahzam, N.; Bresson, A.; Blanchard, C.; Cadoret, M.; Olesen, A.V.; Forsberg, R. Absolute airborne gravimetry with a cold atom sensor. J. Geod. 2020, 94, 20. [Google Scholar] [CrossRef]
  101. Wang, H.; Wang, K.; Xu, Y.; Tang, Y.; Wu, B.; Cheng, B.; Wu, L.; Zhou, Y.; Weng, K.; Zhu, D.; et al. A Truck-Borne System Based on Cold Atom Gravimeter for Measuring the Absolute Gravity in the Field. Sensors 2022, 22, 6172. [Google Scholar] [CrossRef]
  102. Zhang, J.Y.; Xu, W.J.; Sun, S.D.; Shu, Y.B.; Luo, Q.; Cheng, Y.; Hu, Z.K.; Zhou, M.K. A car-based portable atom gravimeter and its application in field gravity survey. AIP Adv. 2021, 11, 115223. [Google Scholar] [CrossRef]
  103. Weiner, S.; Wu, X.; Pagel, Z.; Li, D.; Sleczkowski, J.; Ketcham, F.; Mueller, H. A Flight Capable Atomic Gravity Gradiometer With a Single Laser. In Proceedings of the 2020 IEEE International Symposium on Inertial Sensors and Systems (INERTIAL), Hiroshima, Japan, 23–26 March 2020. [Google Scholar]
  104. Guo, J.; Ma, S.; Zhou, C.; Liu, J.; Wang, B.; Pan, D.; Mao, H. Vibration Compensation for a Vehicle-Mounted Atom Gravimeter. IEEE Sens. J. 2022, 22, 12939–12946. [Google Scholar] [CrossRef]
  105. Zhou, Y.; Zhang, C.; Chen, P.; Cheng, B.; Zhu, D.; Wang, K.; Wang, X.; Wu, B.; Qiao, Z.; Lin, Q.; et al. A Testing Method for Shipborne Atomic Gravimeter Based on the Modulated Coriolis Effect. Sensors 2023, 23, 881. [Google Scholar] [CrossRef] [PubMed]
  106. Yu, N.; Kohel, J.; Kellogg, J.; Maleki, L. Development of an atom-interferometer gravity gradiometer for gravity measurement from space. Appl. Phys. B 2006, 84, 647–652. [Google Scholar] [CrossRef]
  107. Sorrentino, F.; Bongs, K.; Bouyer, P.; Cacciapuoti, L.; De Angelis, M.; Dittus, H.; Ertmer, W.; Giorgini, A.; Hartwig, J.; Hauth, M.; et al. A compact atom interferometer for future space missions. Microgravity Sci. Technol. 2010, 22, 551–561. [Google Scholar] [CrossRef]
  108. Trimeche, A.; Battelier, B.; Becker, D.; Bertoldi, A.; Bouyer, P.; Braxmaier, C.; Charron, E.; Corgier, R.; Cornelius, M.; Douch, K.; et al. Concept study and preliminary design of a cold atom interferometer for space gravity gradiometry. Class. Quantum Gravity 2019, 36, 215004. [Google Scholar] [CrossRef]
  109. Wang, G.; Gao, D.; Ni, W.T.; Wang, J.; Zhan, M. Orbit design for space atom-interferometer AIGSO. Int. J. Mod. Phys. D 2020, 29, 1940004. [Google Scholar] [CrossRef]
  110. Zhu, Z.; Liao, H.; Tu, H.; Duan, X.; Zhao, Y. Spaceborne Atom-Interferometry Gravity Gradiometry Design towards Future Satellite Gradiometric Missions. Aerospace 2022, 9, 253. [Google Scholar] [CrossRef]
  111. Tino, G.; Sorrentino, F.; Aguilera, D.; Battelier, B.; Bertoldi, A.; Bodart, Q.; Bongs, K.; Bouyer, P.; Braxmaier, C.; Cacciapuoti, L.; et al. Precision Gravity Tests with Atom Interferometry in Space. Nucl. Phys. B Proc. Suppl. 2013, 243–244, 203–217. [Google Scholar] [CrossRef]
  112. Aguilera, D.N.; Ahlers, H.; Battelier, B.; Bawamia, A.; Bertoldi, A.; Bondarescu, R.; Bongs, K.; Bouyer, P.; Braxmaier, C.; Cacciapuoti, L.; et al. STE-QUEST—Test of the universality of free fall using cold atom interferometry. Class. Quantum Gravity 2014, 31, 115010. [Google Scholar] [CrossRef]
  113. Becker, D.; Lachmann, M.D.; Seidel, S.T.; Ahlers, H.; Dinkelaker, A.N.; Grosse, J.; Hellmig, O.; Müntinga, H.; Schkolnik, V.; Wendrich, T.; et al. Space-borne Bose–Einstein condensation for precision interferometry. Nature 2018, 562, 391–395. [Google Scholar] [CrossRef] [PubMed]
  114. Lachmann, M.D.; Ahlers, H.; Becker, D.; Dinkelaker, A.N.; Grosse, J.; Hellmig, O.; Müntinga, H.; Schkolnik, V.; Seidel, S.T.; Wendrich, T.; et al. Ultracold atom interferometry in space. Nat. Commun. 2021, 12, 1317. [Google Scholar] [CrossRef]
  115. Oudrhiri, K.; Kohel, J.M.; Harvey, N.; Kellogg, J.R.; Aveline, D.C.; Butler, R.L.; Bosch-Lluis, J.; Callas, J.L.; Cheng, L.Y.; Croonquist, A.P.; et al. NASA’s Cold Atom Laboratory: Four Years of Quantum Science Operations in Space. arXiv 2023, arXiv:2305.13285. [Google Scholar]
  116. Janvier, C.; Ménoret, V.; Desruelle, B.; Merlet, S.; Landragin, A.; dos Santos, F.P. Compact differential gravimeter at the quantum projection-noise limit. Phys. Rev. A 2022, 105, 022801. [Google Scholar] [CrossRef]
  117. Thom, J.; Picken, C.; Malcolm, J.; Kelly, A.; Cheng, X.; Hinton, A.; Bunting, A.; Maker, G.; Hempler, N.; Malcolm, G. Commercial quantum sensors using atom interferometry and laser sources for precision metrology. Opt. Quantum Sens. Precis. Metrol. 2021, 11700, 1170007. [Google Scholar]
  118. Young, B.; Black, A.; Boyd, M. Cold atom inertial sensors for precision navigation. In Proceedings of the 2011 Joint Navigation Conference, Colorado Springs, CO, USA, 28–30 June 2011. [Google Scholar]
  119. Libby, S.; Sonnad, V.; Kreek, S.; Brady, K.; Matthews, M.; Dubetsky, B.; Vitouchkine, A.; Young, B. Feasibility Study of a Passive, Standoff Detector of High Density Masses with a Gravity Gradiometer Based on Atom Interferometry; Technical Report; Lawrence Livermore National Lab. (LLNL): Livermore, CA, USA, 2011. [Google Scholar]
  120. Mugaltech. Available online: https://www.mugaltech.com (accessed on 12 December 2022).
  121. CASColdatom. Available online: https://www.cascoldatom.com/en/sys-pd/1.html?fromColId=104 (accessed on 12 December 2022).
  122. Market Data Forecast. Available online: https://www.marketdataforecast.com/market-reports/quantum-sensing-market (accessed on 13 April 2023).
  123. Precedence Research. Available online: https://www.precedenceresearch.com/geothermal-power-market (accessed on 13 April 2023).
  124. Future Market Insights. Available online: https://www.futuremarketinsights.com/reports/carbon-capture-storage-market#:~:text=The%20global%20carbon%20capture%20and,US%24%209%20Billion%20by%202032 (accessed on 13 April 2023).
  125. Fortune Business Insights. Available online: https://www.fortunebusinessinsights.com/geohazard-market-105033 (accessed on 13 April 2023).
  126. Data Bridge Market Research. Available online: https://www.databridgemarketresearch.com/reports/global-hydrocarbons-market (accessed on 13 April 2023).
  127. Bodart, Q.; Merlet, S.; Malossi, N.; Dos Santos, F.P.; Bouyer, P.; Landragin, A. A cold atom pyramidal gravimeter with a single laser beam. Appl. Phys. Lett. 2010, 96, 134101. [Google Scholar] [CrossRef]
  128. Heine, N.; Matthias, J.; Sahelgozin, M.; Herr, W.; Abend, S.; Timmen, L.; Müller, J.; Rasel, E.M. A transportable quantum gravimeter employing delta-kick collimated Bose–Einstein condensates. Eur. Phys. J. D 2020, 74, 174. [Google Scholar] [CrossRef]
  129. Lee, J.; Grover, J.; Orozco, L.; Rolston, S. Sub-Doppler cooling of neutral atoms in a grating magneto-optical trap. JOSA B 2013, 30, 2869–2874. [Google Scholar] [CrossRef]
  130. McGilligan, J.P.; Griffin, P.F.; Elvin, R.; Ingleby, S.J.; Riis, E.; Arnold, A.S. Grating chips for quantum technologies. Sci. Rep. 2017, 7, 384. [Google Scholar] [CrossRef] [PubMed]
  131. Sitaram, A.; Elgee, P.; Campbell, G.K.; Klimov, N.; Eckel, S.; Barker, D. Confinement of an alkaline-earth element in a grating magneto-optical trap. Rev. Sci. Instrum. 2020, 91, 103202. [Google Scholar] [CrossRef] [PubMed]
  132. Little, B.J.; Hoth, G.W.; Christensen, J.; Walker, C.; De Smet, D.J.; Biedermann, G.W.; Lee, J.; Schwindt, P.D.D. A passively pumped vacuum package sustaining cold atoms for more than 200 days. AVS Quantum Sci. 2021, 3, 035001. [Google Scholar] [CrossRef]
  133. Sugarbaker, A.; Dickerson, S.M.; Hogan, J.M.; Johnson, D.M.S.; Kasevich, M.A. Enhanced Atom Interferometer Readout through the Application of Phase Shear. Phys. Rev. Lett. 2013, 111, 113002. [Google Scholar] [CrossRef] [PubMed]
  134. Louchet-Chauvet, A.; Farah, T.; Bodart, Q.; Clairon, A.; Landragin, A.; Merlet, S.; Santos, F.P.D. The influence of transverse motion within an atomic gravimeter. New J. Phys. 2011, 13, 065025. [Google Scholar] [CrossRef]
  135. Vovrosh, J.; Voulazeris, G.; Petrov, P.G.; Zou, J.; Gaber, Y.; Benn, L.; Woolger, D.; Attallah, M.M.; Boyer, V.; Bongs, K.; et al. Additive manufacturing of magnetic shielding and ultra-high vacuum flange for cold atom sensors. Sci. Rep. 2018, 8, 2023. [Google Scholar] [CrossRef] [PubMed]
  136. Cooper, N.; Coles, L.; Everton, S.; Maskery, I.; Campion, R.; Madkhaly, S.; Morley, C.; O’Shea, J.; Evans, W.; Saint, R.; et al. Additively manufactured ultra-high vacuum chamber for portable quantum technologies. Addit. Manuf. 2021, 40, 101898. [Google Scholar] [CrossRef]
  137. Madkhaly, S.; Coles, L.; Morley, C.; Colquhoun, C.; Fromhold, T.; Cooper, N.; Hackermüller, L. Performance-Optimized Components for Quantum Technologies via Additive Manufacturing. PRX Quantum 2021, 2, 030326. [Google Scholar] [CrossRef]
  138. Vovrosh, J.; Earl, L.; Thomas, H.; Winch, J.; Stray, B.; Ridley, K.; Langlois, M.; Bongs, K.; Holynski, M. Reduction of background scattered light in vacuum systems for cold atoms experiments. AIP Adv. 2020, 10, 105125. [Google Scholar] [CrossRef]
  139. Mielec, N.; Altorio, M.; Sapam, R.; Horville, D.; Holleville, D.; Sidorenkov, L.A.; Landragin, A.; Geiger, R. Atom interferometry with top-hat laser beams. Appl. Phys. Lett. 2018, 113, 161108. [Google Scholar] [CrossRef]
  140. Hobson, P.J.; Vovrosh, J.; Stray, B.; Packer, M.; Winch, J.; Holmes, N.; Hayati, F.; McGovern, K.; Bowtell, R.; Brookes, M.J.; et al. Bespoke magnetic feld design for a magnetically shielded cold atom interferometer. Sci. Rep. 2022, 12, 10520. [Google Scholar] [CrossRef] [PubMed]
  141. López-Vázquez, A.; Maldonado, M.A.; Gomez, E.; Corzo, N.V.; de Carlos-López, E.; Villafañe, J.F.; Jiménez-García, K.; Jiménez-Mier, J.; López-González, J.L.; López-Monjaraz, C.J.; et al. Compact laser modulation system for a transportable atomic gravimeter. Opt. Express 2023, 31, 3504–3519. [Google Scholar] [CrossRef]
  142. Gong, W.; Li, A.; Luo, J.; Che, H.; Ma, J.; Qin, F. A Vibration Compensation Approach for Atom Gravimeter Based on Improved Sparrow Search Algorithm. IEEE Sens. J. 2023, 23, 5911–5919. [Google Scholar] [CrossRef]
  143. Cannon, R.; Dyer, S.; Griffin, P.; McGilligan, J.; Bremner, D.; Riis, E. Miniaturized high-reliability lasers for quantum technologies. In Photonics Quantum; SPIE: Bellingham, WA, USA, 2022; p. 1224308. [Google Scholar]
  144. Everitt, M.J.; Michael, J.D.C.; Dwyer, V.M. Quantum Systems Engineering: A structured approach to accelerating the development of a quantum technology industry. In Proceedings of the 18th International Conference on Transparent Optical Networks (ICTON), Trento, Italy, 10–14 July 2016; IEEE: Piscataway, NJ, USA, 2016; pp. 1–4. [Google Scholar]
  145. Bjergstrom, K.N.; Huish, W.G.B.; de C. Henshaw, M.J.; Dwyer, V.M.; Everitt, M.J. Transformational Effects of Applying Systems Engineering in Laboratory Scientific Research. IEEE Syst. J. 2019, 13, 1924–1935. [Google Scholar] [CrossRef]
  146. Cervantes, J.M.; Maldonado, M.A.; Franco-Villafañe, J.A.; Roach, T.; Valenzuela, V.M.; Gomez, E. Selection of a Raman beam waist in atomic gravimetry. OSA Contin. 2021, 4, 1996–2007. [Google Scholar] [CrossRef]
  147. Hu, Q.Q.; Zhou, H.; Luo, Y.K.; Luo, Q.; Kuang, W.J.; Wan, F.B.; Zhong, Y.Y.; Xu, F.F. Improving the fringe contrast in an atomic gravimeter by optimizing the Raman laser intensity. Optik 2023, 276, 170637. [Google Scholar] [CrossRef]
  148. McGuirk, J.; Snadden, M.; Kasevich, M. Large area light-pulse atom interferometry. Phys. Rev. Lett. 2000, 85, 4498. [Google Scholar] [CrossRef]
  149. Müller, H.; Chiow, S.w.; Long, Q.; Herrmann, S.; Chu, S. Atom interferometry with up to 24-photon-momentum-transfer beam splitters. Phys. Rev. Lett. 2008, 100, 180405. [Google Scholar] [CrossRef] [PubMed]
  150. Chiow, S.w.; Kovachy, T.; Chien, H.C.; Kasevich, M.A. 102 ℏk large area atom interferometers. Phys. Rev. Lett. 2011, 107, 130403. [Google Scholar] [CrossRef]
  151. Rudolph, J.; Wilkason, T.; Nantel, M.; Swan, H.; Holland, C.M.; Jiang, Y.; Garber, B.E.; Carman, S.P.; Hogan, J.M. Large Momentum Transfer Clock Atom Interferometry on the 689 nm Intercombination Line of Strontium. Phys. Rev. Lett. 2020, 124, 083604. [Google Scholar] [CrossRef]
  152. Wilkason, T.; Nantel, M.; Rudolph, J.; Jiang, Y.; Garber, B.E.; Swan, H.; Carman, S.P.; Abe, M.; Hogan, J.M. Atom Interferometry with Floquet Atom Optics. Phys. Rev. Lett. 2022, 129, 183202. [Google Scholar] [CrossRef] [PubMed]
  153. Badurina, L.; Bentine, E.; Blas, D.; Bongs, K.; Bortoletto, D.; Bowcock, T.; Bridges, K.; Bowden, W.; Buchmueller, O.; Burrage, C.; et al. AION: An atom interferometer observatory and network. J. Cosmol. Astropart. Phys. 2020, 2020, 011. [Google Scholar] [CrossRef]
  154. Graham, P.W.; Hogan, J.M.; Kasevich, M.A.; Rajendran, S.; Romani, R.W. Mid-band gravitational wave detection with precision atomic sensors. arXiv 2017, arXiv:1711.02225. [Google Scholar]
  155. Berg, P.; Abend, S.; Tackmann, G.; Schubert, C.; Giese, E.; Schleich, W.P.; Narducci, F.A.; Ertmer, W.; Rasel, E.M. Composite-Light-Pulse Technique for High-Precision Atom Interferometry. Phys. Rev. Lett. 2015, 114, 063002. [Google Scholar] [CrossRef] [PubMed]
  156. Saywell, J.; Carey, M.; Belal, M.; Kuprov, I.; Freegarde, T. Optimal control of Raman pulse sequences for atom interferometry. J. Phys. B At. Mol. Opt. Phys. 2020, 53, 085006. [Google Scholar] [CrossRef]
  157. Saywell, J.; Carey, M.; Kuprov, I.; Freegarde, T. Biselective pulses for large-area atom interferometry. Phys. Rev. A 2020, 101, 063625. [Google Scholar] [CrossRef]
  158. Lellouch, S.; Ennis, O.; Haditalab, R.; Langlois, M.; Holynski, M. Polychromatic atom optics for atom interferometry. EPJ Quantum Technol. 2023, 10, 9. [Google Scholar] [CrossRef]
  159. Hamilton, P.; Jaffe, M.; Brown, J.M.; Maisenbacher, L.; Estey, B.; Müller, H. Atom Interferometry in an Optical Cavity. Phys. Rev. Lett. 2015, 114, 100405. [Google Scholar] [CrossRef]
  160. Dovale-Álvarez, M.; Brown, D.D.; Jones, A.W.; Mow-Lowry, C.M.; Miao, H.; Freise, A. Fundamental limitations of cavity-assisted atom interferometry. Phys. Rev. A 2017, 96, 053820. [Google Scholar] [CrossRef]
  161. Nourshargh, R.; Hedges, S.; Langlois, M.; Bongs, K.; Holynski, M. Doppler compensation for cavity-based atom interferometry. Opt. Express 2022, 30, 30001–30011. [Google Scholar] [CrossRef]
  162. Nourshargh, R.; Lellouch, S.; Hedges, S.; Langlois, M.; Bongs, K.; Holynski, M. Circulating pulse cavity enhancement as a method for extreme momentum transfer atom interferometry. Commun. Phys. 2021, 4, 257. [Google Scholar] [CrossRef]
  163. Saint, R.; Evans, W.; Zhou, Y.; Barrett, T.; Fromhold, T.; Saleh, E.; Maskery, I.; Tuck, C.; Wildman, R.; Oručević, F.; et al. 3D-printed components for quantum devices. Sci. Rep. 2018, 8, 8368. [Google Scholar] [CrossRef]
  164. Norrgard, E.B.; Barker, D.S.; Fedchak, J.A.; Klimov, N.; Scherschligt, J.; Eckel, S. Note: A 3D-printed alkali metal dispenser. Rev. Sci. Instrum. 2018, 89, 056101. [Google Scholar] [CrossRef] [PubMed]
  165. Kohn, R.N.; Bigelow, M.S.; Spanjers, M.; Stuhl, B.K.; Kasch, B.L.; Olson, S.E.; Imhof, E.A.; Hostutler, D.A.; Squires, M.B. Clean, robust alkali sources by intercalation within highly oriented pyrolytic graphite. Rev. Sci. Instrum. 2020, 91, 035108. [Google Scholar] [CrossRef] [PubMed]
  166. McGilligan, J.P.; Moore, K.R.; Kang, S.; Mott, R.; Mis, A.; Roper, C.; Donley, E.A.; Kitching, J. Dynamic Characterization of an Alkali-Ion Battery as a Source for Laser-Cooled Atoms. Phys. Rev. Appl. 2020, 13, 044038. [Google Scholar] [CrossRef]
  167. Li, C.; Chai, X.; Wei, B.; Yang, J.; Daruwalla, A.; Ayazi, F.; Raman, C. Cascaded collimator for atomic beams traveling in planar silicon devices. Nat. Commun. 2019, 10, 1831. [Google Scholar] [CrossRef] [PubMed]
  168. Kellogg, J.R.; Schlippert, D.; Kohel, J.M.; Thompson, R.J.; Aveline, D.C.; Yu, N. A compact high-efficiency cold atom beam source. Appl. Phys. B 2012, 109, 61–64. [Google Scholar] [CrossRef]
  169. Imhof, E.; Stuhl, B.K.; Kasch, B.; Kroese, B.; Olson, S.E.; Squires, M.B. Two-dimensional grating magneto-optical trap. Phys. Rev. A 2017, 96, 033636. [Google Scholar] [CrossRef]
  170. Wang, Q.; Wang, Z.; Fu, Z.; Liu, W.; Lin, Q. A compact laser system for the cold atom gravimeter. Opt. Commun. 2016, 358, 82–87. [Google Scholar] [CrossRef]
  171. Lee, J.; Ding, R.; Christensen, J.; Rosenthal, R.R.; Ison, A.; Gillund, D.P.; Bossert, D.; Fuerschbach, K.H.; Kindel, W.; Finnegan, P.S.; et al. A compact cold-atom interferometer with a high data-rate grating magneto-optical trap and a photonic-integrated-circuit-compatible laser system. Nat. Commun. 2022, 13, 5131. [Google Scholar] [CrossRef]
  172. McGilligan, J.; Gallacher, K.; Griffin, P.; Paul, D.; Arnold, A.; Riis, E. Micro-fabricated components for cold atom sensors. Rev. Sci. Instrum. 2022, 93, 091101. [Google Scholar] [CrossRef]
  173. Zhu, L.; Liu, X.; Sain, B.; Wang, M.; Schlickriede, C.; Tang, Y.; Deng, J.; Li, K.; Yang, J.; Holynski, M.; et al. A dielectric metasurface optical chip for the generation of cold atoms. Sci. Adv. 2020, 6, eabb6667. [Google Scholar] [CrossRef]
  174. Jin, M.; Zhang, X.; Liu, X.; Liang, C.; Liu, J.; Hu, Z.; Li, K.; Wang, G.; Yang, J.; Zhu, L.; et al. A Centimeter-Scale Dielectric Metasurface for the Generation of Cold Atoms. Nano Lett. 2023, 23, 4008–4013. [Google Scholar] [CrossRef] [PubMed]
  175. Pétremand, Y.; Affolderbach, C.; Straessle, R.; Pellaton, M.; Briand, D.; Mileti, G.; de Rooij, N.F. Microfabricated rubidium vapour cell with a thick glass core for small-scale atomic clock applications. J. Micromech. Microeng. 2012, 22, 025013. [Google Scholar] [CrossRef]
  176. Tuzson, B.; Mangold, M.; Looser, H.; Manninen, A.; Emmenegger, L. Compact multipass optical cell for laser spectroscopy. Opt. Lett. 2013, 38, 257–259. [Google Scholar] [CrossRef] [PubMed]
  177. Gharavipour, M.; Affolderbach, C.; Kang, S.; Mileti, G. Double-resonance spectroscopy in Rubidium vapour-cells for high performance and miniature atomic clocks. J. Phys. Conf. Ser. 2017, 793, 012007. [Google Scholar] [CrossRef]
  178. Graf, M.; Emmenegger, L.; Tuzson, B. Compact, circular, and optically stable multipass cell for mobile laser absorption spectroscopy. Opt. Lett. 2018, 43, 2434–2437. [Google Scholar] [CrossRef] [PubMed]
  179. Packer, M.; Hobson, P.; Holmes, N.; Leggett, J.; Glover, P.; Brookes, M.; Bowtell, R.; Fromhold, T. Optimal inverse design of magnetic field profiles in a magnetically shielded cylinder. Phys. Rev. Appl. 2020, 14, 054004. [Google Scholar] [CrossRef]
  180. Packer, M.; Hobson, P.; Holmes, N.; Leggett, J.; Glover, P.; Brookes, M.; Bowtell, R.; Fromhold, T. Planar coil optimization in a magnetically shielded cylinder. Phys. Rev. Appl. 2021, 15, 064006. [Google Scholar] [CrossRef]
  181. Malcolm, J.I. Construction of a Portable Platform for Cold Atom Interferometry. Ph.D. Thesis, University of Birmingham, Birmingham, UK, 2016. [Google Scholar]
  182. Schkolnik, V.; Döringshoff, K.; Gutsch, F.B.; Oswald, M.; Schuldt, T.; Braxmaier, C.; Lezius, M.; Holzwarth, R.; Kürbis, C.; Bawamia, A.; et al. JOKARUS-design of a compact optical iodine frequency reference for a sounding rocket mission. EPJ Quantum Technol. 2017, 4, 9. [Google Scholar] [CrossRef]
  183. Burrow, O.S.; Osborn, P.F.; Boughton, E.; Mirando, F.; Burt, D.P.; Griffin, P.F.; Arnold, A.S.; Riis, E. Stand-alone vacuum cell for compact ultracold quantum technologies. Appl. Phys. Lett. 2021, 119, 124002. [Google Scholar] [CrossRef]
  184. Che, H.; Li, A.; Zhou, Z.; Gong, W.; Ma, J.; Qin, F. An Approach of Vibration Compensation for Atomic Gravimeter under Complex Vibration Environment. Sensors 2023, 23, 3535. [Google Scholar] [CrossRef]
  185. Lyu, W.; Zhong, J.Q.; Zhang, X.W.; Liu, W.; Zhu, L.; Xu, W.H.; Chen, X.; Tang, B.; Wang, J.; Zhan, M.S. Compact High-Resolution Absolute-Gravity Gradiometer Based on Atom Interferometers. Phys. Rev. Appl. 2022, 18, 054091. [Google Scholar] [CrossRef]
  186. Devani, D.; Maddox, S.; Renshaw, R.; Cox, N.; Sweeney, H.; Cross, T.; Holynski, M.; Nolli, R.; Winch, J.; Bongs, K.; et al. Gravity sensing: Cold atom trap onboard a 6U CubeSat. CEAS Space J. 2020, 12, 539–549. [Google Scholar] [CrossRef]
  187. Vovrosh, J.; Wilkinson, K.; Hedges, S.; McGovern, K.; Hayati, F.; Carson, C.; Selyem, A.; Winch, J.; Stray, B.; Earl, L.; et al. Magneto-optical trapping in a near-surface borehole. arXiv 2022, arXiv:2211.11415. [Google Scholar]
Figure 1. The atom interferometry sequence: As the atoms fall under the effect of gravity, a laser pulse is used to put them into a quantum superposition of the ground and excited states. The action of this pulse can be thought of as sending each atom along two simultaneous but different paths in the gravity field (as a consequence of wave-particle duality). After a time, T, a second laser pulse is then used to cause the paths to converge again at time 2T. At this time, the atoms are recombined using a final laser pulse; the final state of the atoms encodes the value of gravity. Dotted and solid lines show the sequence with and without gravity, respectively, where the deflection due to gravity introduces a phase shift Δ ϕ to the interferometer.
Figure 1. The atom interferometry sequence: As the atoms fall under the effect of gravity, a laser pulse is used to put them into a quantum superposition of the ground and excited states. The action of this pulse can be thought of as sending each atom along two simultaneous but different paths in the gravity field (as a consequence of wave-particle duality). After a time, T, a second laser pulse is then used to cause the paths to converge again at time 2T. At this time, the atoms are recombined using a final laser pulse; the final state of the atoms encodes the value of gravity. Dotted and solid lines show the sequence with and without gravity, respectively, where the deflection due to gravity introduces a phase shift Δ ϕ to the interferometer.
Sensors 23 07651 g001
Table 1. Examples of the signal sizes across different applications.
Table 1. Examples of the signal sizes across different applications.
TargetSignal Size (μGal)References
Archaeology15–40[5,58,59]
Carbon storage monitoring0–16[60,61,62]
Cave detection and mapping0–1500[63,64,65]
Earth tide measurements100–300[66]
Earthquake detection0–16,000[6,67]
Hydrology0–100[68,69,70]
Mine shafts0–100[71,72]
Sinkholes0–40[73,74,75]
Tunnels0–300[19,76,77]
Volcano monitoring0–60,000[78,79,80]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Vovrosh, J.; Dragomir, A.; Stray, B.; Boddice, D. Advances in Portable Atom Interferometry-Based Gravity Sensing. Sensors 2023, 23, 7651. https://doi.org/10.3390/s23177651

AMA Style

Vovrosh J, Dragomir A, Stray B, Boddice D. Advances in Portable Atom Interferometry-Based Gravity Sensing. Sensors. 2023; 23(17):7651. https://doi.org/10.3390/s23177651

Chicago/Turabian Style

Vovrosh, Jamie, Andrei Dragomir, Ben Stray, and Daniel Boddice. 2023. "Advances in Portable Atom Interferometry-Based Gravity Sensing" Sensors 23, no. 17: 7651. https://doi.org/10.3390/s23177651

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop