Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

SERF-like magnetometry in room-temperature environment

Open Access Open Access

Abstract

We demonstrate an atomic magnetometry using amplitude-modulated pumping and hyperfine repumping techniques in a paraffin-coated cell. By exploiting the constructive interference between spins polarized by the pump beam and an additional repump beam, we observe a three-fold increase in the amplitude of magnetic resonance, along with a reduction in linewidth by approximately two times. The implementation of the repump beam effectively narrows the linewidth, demonstrating successful suppression of spin-exchange relaxation. This reduction in relaxation rate, combined with the enhanced signal, significantly improves the sensitivity of the magnetometer. Consequently, our technique offers a promising approach for achieving SERF-like magnetometry with sub-fT-level sensitivity in Earth-field range and room-temperature environment.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Atomic magnetometers with high sensitivity [18] are utilized in various of fields, including physics [914], biology [1520], chemistry [21,22], materials science [2327] and geology [28]. The sensitivity of these magnetometers is constrained by spin-projection noise $\delta \,B\simeq (\gamma \sqrt {NT\tau })^{-1},$ [4], which is determined by several factors, namely the gyromagnetic ratio ($\gamma$), the measurement time ($T$), the number of atoms ($N$), and the spin relaxation time ($\tau$). To improve the sensitivity, it is advantageous to maximize the atomic density and prolong the spin relaxation time.

Spin relaxation can be attributed to various processes, including spin-destructive collisions, spin-exchange (SE) collisions, spin-wall collisions, magnetic field gradients, and power broadening of the optical field [29]. Among these processes, spin-destructive collisions occur at a rate several orders of magnitude lower than SE collisions, while the influence of magnetic field gradients can be compensated by using gradient coils. To mitigate spin relaxation during spin-wall collisions, the collision rate can be effectively reduced by introducing a buffer gas in which the alkali atoms undergo a diffusive motion that slows down their wall collision rate. By operating such a cell at high temperatures and low magnetic fields, the atomic ensemble works in the spin-exchange-relaxation-free (SERF) regime, resulting in the most sensitive atomic magnetometer nowadays [3,68]. However, the requirement of high operating temperatures and low magnetic fields limits its application in the fields of biology and geology.

Outside the SERF regime, nonlinear-magnetic-optical-rotation (NMOR) magnetometry using paraffin coated cell is a promising candidate to achieve a sensitivity comparable to that of the SERF magnetometry in the Earth field range. To prevent spin relaxation during wall collisions, paraffin coatings are used supporting approximately $10^4$ spin-wall collisions before relaxation of the atomic spins [3032]. Anti-spin-relaxation coated cells typically do not contain any buffer gas which allows the atoms to explore the whole cell volume at their thermal velocity. To further improve the sensitivity, it is essential to develop technologies to suppress SE in paraffin coated cell.

In the presence of magnetic field, the atomic spins in the two hyperfine ground states undergo precession in opposite directions. SE collisions alter the population distribution between these two hyperfine ground states, leading to the relaxation of atomic spins. In this scenario, the suppression of SE relaxation can be achieved through a phenomenon known as light narrowing, which has been extensively studied in buffer-gas cells but has not been observed in the absence of buffer gas [29,3341]. By increasing the power of the pump beam, a majority of atoms can be efficiently pumped into the stretched state ($m_{F}=F$). This results in the suppression of SE relaxation due to the conservation of total angular momentum during SE collisions. In buffer-gas cells, only one laser is required to pump atoms in both hyperfine sublevels, because of pressure broadening. However, the presence of buffer gas renders the magnetometer more susceptible to magnetic-field gradients, which induces spin relaxation [29]. Without the buffer gas, an additional repump beam is necessary to transfer the atomic population to the desired hyperfine sublevel, while the pump beam polarizes the atoms into the stretched state. Nevertheless, the introduction of the repump beam disturbs the atomic states, leading to a significant broadening of the linewidth as the repump power increases.

In this work, we experimentally demonstrate the suppression of SE processes in a paraffin-coated cell using a pulsed repump beam. The pump and repump beams are utilized to create a near-zero population in the $F=1$ state while effectively pumping most of the atoms into the $F=2$, $m_{F}=2$ state. By manipulating the relative phase between the pump and repump beams, we are able to increase or eliminate the amplitude of the magnetic resonance signal. The constructive interference of the atomic spins generated by pump and repump beams leads to a reduction in the relaxation of atomic spins, resulting in the light narrowing effect. This effect is characterized by an enhanced signal and a narrowed linewidth, which consequently leads to an improvement in the sensitivity of the magnetometer. In addition, the reduction in transverse relaxation facilitates the increase in quantum-information storage time within the atomic vapor [42].

2. Experimental setup

The experimental setup is illustrated in Fig. 1. A paraffin-coated cylindrical cell [3032], with dimensions of 5 cm in length and 5 cm in diameter, containing isotopically enriched $^{87}$Rb atoms is placed within a four-layer $\mu$-metal magnetic shield and operated at a temperature of 42 $^\circ$ C.

 figure: Fig. 1.

Fig. 1. Experimental setup. HWP: half-wave plate. QWP: quarter-wave plate. POL: polarizer. BPD: balanced photo-detector. LIA: lock-in amplifier. FG: function generator.

Download Full Size | PDF

The atoms are stretched along the $\hat {x}$-axis through optical pumping using two circularly polarized laser beams (pump and repump) directed along $\hat {x}$. The pump beam is resonant with the $^{87}$Rb D1 $F=2$ to $F^{'}=2$ transition, while the repump beam is resonant with the $^{87}$Rb D1 $F=1$ to $F^{'}=1$ or $2$ transitions. A dichroic atomic vapor laser lock (DAVLL) technique is employed to stabilize the laser frequencies of the pump and repump beams [43]. The intensities of the pump and repump beams are modulated using two acousto-optic modulators (AOMs) controlled by two channels of a function generator. The optimal duty cycle for the pump and repump beams is determined through parameter optimization. Theoretically, an infinitely large amplitude delta function would be ideal. In experiment, for a certain power of pump and repump beams, expanding the duty cycle increases both the linewidth and amplitude of magnetic resonance. Taking into account the power and relaxation rate, a value of $3{\% }$ causes the ratio of amplitude to linewidth to be maximize, so it is chosen as the optimal duty cycle. The modulation frequency of the pump and repump beams is set to match the Larmor frequency of the applied magnetic field (aligned along the $\hat {z}$-axis), allowing for strengthening of the atomic polarization with each pumping pulse in every cycle. A linearly polarized probe beam, with a power of 10 $\mu$W, is used to detect the Larmor precession. After passing through the cell, the polarization of the probe beam is altered due to the atomic polarization, which is subsequently detected by a balanced polarimeter as the optical rotation signal. This optical rotation signal is fed into a lock-in amplifier and demodulated at the modulation frequency.

3. Experimental results

3.1 Magnetic resonance

To illustrate the magnetic resonance with and without the repump beam, we keep the modulation frequency fixed at a specific value and scan the leading magnetic field. Figure 2 presents the magnetic resonance data for three different pumping and probing schemes. In all schemes, the pump power was set at 150 $\mu$W, and the optical field is resonant with the D1 $F=2$ to $F^{'}=2$ transition. The probe power is fixed at 10 $\mu$W. Two probe schemes are employed to detect the polarized atoms populated in the $F=2$ and $F=1$ states, as depicted in Fig. 2 (a) and (b), respectively. in Fig. 2 (a), the probe light is red-detuned by 1.5 GHz from the D2 $F=2$ to $F^{'}=2$ transitions, enabling the detection of the Larmor precession of atomic spins in the $F=2$ state. in Fig. 2 (b), the probe light is red-detuned by 500 MHz from the D2 $F=1$ to $F^{'}=1$ transitions, detecting spins in the $F=1$ state. The amplitudes of the magnetic resonances in Figs. 2 (a) and (b) are approximately the same. The difference in the center frequency of the two resonances is due to the variation in the gyromagnetic ratio between the $F=1$ and $F=2$ states. in Fig. 2 (c), an additional repump beam with the power of 150 $\mu$W and resonating with the D1 $F=1$ to $F^{'}=2$ transition is introduced. Nearly all the atoms contribute to this magnetic resonance signal. It is evident from this figure that the amplitude of the resonance in the presence of the repump beam is larger than that of the traditional scheme.

 figure: Fig. 2.

Fig. 2. Magnetic resonance data for a pump-modulation frequency of 4932 Hz as a function of the leading magnetic field along the $\hat {z}$-axis. In all scheme, pump beam is on resonance with D1 $F=2$ to $F^{'}=2$ transition. (b) The probe light is 500 MHz red detunned from D2 $F=1$ to $F^{'}=1$ transitions. (c) The probe light has same frequency with scheme (a), The repump beam is on resonance with D1 $F=1$ to $F^{'}=2$ transition.

Download Full Size | PDF

3.2 Spin interference

We notice that the repump beam not only pumps atoms back to the $F=2$ state but also contributes to spin polarization on its own. The polarized spins generated by the pump beam with amplitude $S_1$ and the repump beam with amplitude $S_2$ both oscillate at the Larmor frequency and can interfere with each other. The total spin $S$ can be described by the following equation:

$$S=\sqrt{S_1^2+S_2^2+2S_1S_2\cos\phi},$$
where $\phi$ represents the relative phase between $S_1$ and $S_2$. Figure 3 illustrates the variation in the amplitude of the magnetic resonance as a function of the phase $\phi$, with a repump power of 150 $\mu$W (a) and 100 $\mu$W (b,c) respectively. When a repump beam power of 150 $\mu$W is applied, the repump beam generates more spins than the pump beam. In this case, the visibility of the magnetic resonance is approximately 0.6. On the other hand, when a repump beam power of 100 $\mu$W is applied, the magnetic resonance is completely eliminated, indicating that the spins generated by the pump and repump beams have the same amplitude. In Figs. 3 (a,b), the polarization of the repump beam is the same as that of the pump beam. The signal is maximized when the relative phase is zero. in Fig. 3 (c), where the polarization of the repump beam is $\sigma _{-}$, the signal is completely suppressed when the relative phase is zero.

 figure: Fig. 3.

Fig. 3. The amplitude of magnetic resonance change with relative phase between pump and repump. The polarization of pump beam is $\sigma _+$. The pump power is 150 $\mu$W. (a) The polarization of repump beam is $\sigma _+$. The repump power is 150 $\mu$W. (b) The polarization of repump beam is $\sigma _+$. The repump power is 100 $\mu$W. (c) The polarization of repump beam is $\sigma _{-}$. The repump power is 100 $\mu$W.

Download Full Size | PDF

Figure 4 depicts the measured Faraday rotation signal, which is proportional to the population of the ground state, as it changes with the optical detuning $\Delta$ of the probe beam at constructive interference. In the presence of the pump beam, a dispersion signal emerges when the optical field is resonant with the F=2 to excited states transition. Conversely, an absorption signal arises when the optical field is resonant with the F=1 to excited states transition. With the repump beam, the signal associated with the F=2 to excited states transition is enhanced, while the signal related to the F=1 to excited states transition is eliminated. This observation serves as evidence that the atoms originally populated in the F=1 state are all pumped to the F=2 state, satisfying the necessary condition for the suppression of spin-exchange processes.

 figure: Fig. 4.

Fig. 4. The optical rotation signal with (a) and without (b) repump beam. The Fano profile shown in (b) is caused by imbalance population of $F=1, m_{F}=1$ and $F=1, m_{F}=-1$ states. The absorption spectrum of the D2 line of $^{87}$Rb (c).

Download Full Size | PDF

3.3 Spin relaxation time

In order to demonstrate the suppression of spin-exchange (SE) processes, it is important to maintain SE collisions as the dominant mechanism for spin relaxation. By employing a paraffin-coated cell, the magnetic sensor becomes insensitive to magnetic field gradients, and relaxation resulting from spin-wall collisions is significantly suppressed.

To further avoid power broadening effects resulting from the pump beam, we analyze the spin relaxation rate through the measurement of the FID signal. The transverse coherence in F=2 state is generated by applying a sequence of short pump pulses in the $\hat {x}$ direction, while the leading field is present. The repetition frequency of the pump beam is set to match the Larmor frequency of the leading field, maximizing the transverse coherence. After the pump beam is turned off, the optical rotation signal, demodulated by a lock-in amplifier, is observed as shown in Fig. 5 (i). In comparison to conventional FID detection, Fig. 5 (ii) demonstrates the application of an additional repump beam during the pumping period. With the presence of both the pump and repump beams, atoms initially located in the F=1 and F=2 states are now fully polarized to the F=2, $m_{F}=2$ state. This leads to an increase in the signal amplitude and the suppression of spin-exchange (SE) processes between the F=2 and F=1 hyperfine manifolds.

 figure: Fig. 5.

Fig. 5. Relaxation of atomic spins change with different repumping scheme. (i,ii) FID without and with per-repumping. (iii,iv) The polarized spins decay with in phase and out of phase repumping. Inset: the normalized relaxation signal.

Download Full Size | PDF

However, despite the utilization of the sequence shown in Fig. 5 (ii), several decay effects can still cause a gradual transfer of some atoms back to the F=1 state during the probing period. Consequently, the decay induced by SE collisions becomes active once again. in Fig. 5 (iii), a synchronized repump beam is applied during both the pumping and detection periods to further eliminate the decoherence effect during the probing period. Atoms that have decayed to the F=1 state during the probing period can be repopulated back to the F=2 state by the repump beam. The repump beams correct the phase of atomic spins at every cycle. As a result, the relaxation time is further improved. The relative phase between the pump and repump beams plays a crucial role in suppressing SE processes. When the relative phase is non-zero, the medium polarization is washed out, resulting in a shorter relaxation time, as depicted in Fig. 5 (iv).

The variation of the experimental measured and theoretical simulated spin relaxation rate with repump power is demonstrated in Fig. 6 using in-phase repumping. The relaxation rate is measured by exponential Fitting of the FID signal. Here, the relaxation primarily occurs due to SE processes and power broadening. As the repump power increases, the population of the $F=1$ state decreases, leading to a suppression of SE relaxation, while the power broadening effect becomes more prominent. By balancing the influence of these two effects, the minimum relaxation rate is achieved at 2.2 Hz for a repump power of 3 mW.

 figure: Fig. 6.

Fig. 6. Experimental measured and theoretical simulated relaxation rate of atomic spin change with the repump power while the pump power is fixed at 150 $\mu$W.

Download Full Size | PDF

In theoretical simulation, the relaxation rate is determined by adding the SE rate and optical power broadening. The SE rate is modeled as $\,n\rho _{11}\,\sigma _{SE}\sqrt {8k_{B}T/(\pi \,M)}$, where $n=6\times \,10^{16}$ m$^{-3}$ is the atomic density of $^{87}$Rb at 42 $^\circ$ C, $\rho _{11}$ is the population in the $F=1$ state which is calculated by Liouville equations, $\sigma _{SE}=1.8\times \,10^{-19}$ m$^{2}$ is the SE rate, $k_{B}$ is the Boltzmann constant, and $M=7.2\times \,10^{-26}$ kg is the reduced mass of the $^{87}$Rb-$^{87}$Rb collision [4446]. The power broadening is modeled as linearly increasing with the power of the repump beam. To determine the population of atomic spin in the presence of the pump and repump beams, we solve the Liouville equation for the rotating-frame density matrix [47]

$$i\frac{d}{dt}\hat{\rho}=[\hat{H},\hat{\rho}]-i\frac{1}{2}(\hat{\Gamma}\hat{\rho}+\hat{\rho}\hat{\Gamma})+i\hat{\Lambda},$$
where $\hat {H}=\hbar \sum _{F,F^{'},m_{F},m_{F}^{'}}\Omega _{F,F^{'}}\hat {Q}_{F,F'}\delta _{m_F,m_{F}^{'}-1} |F,m_F \rangle \langle F',m_{F}^{'} |$ is the atom-light interaction, $\Omega _{F,F^{'}}=\epsilon \langle \,F|\hat {d}|F'\rangle /\hbar$ is Rabi frequencies of pump and repump beams, $\epsilon$ is the amplitude of electric field, $\hat {d}$ is the atomic electric dipole operator. The standard spherical components of the operator $\hat {Q}_{F,F'}$ are $\hat {Q}^{q}_{F,F'}=(-1)^{F-m}\frac {1}{\sqrt {2F'+1}} \langle F,-m,1,q|F',-m' \rangle$ with ($q=-1,0,1$). $\hat {\rho }$ is the density matrix in the rotating frame, $\hat {\Gamma }$ is the relaxation matrix, and $\hat {\Lambda }$ is repopulation matrix. The excited states undergo spontaneous decay at a rate $\Gamma$ and the ground states undergo relaxation at a rate $\gamma$ due to SE and optical power broadening. The repopulation matrix ensures that the atomic density is normalized to unity. Figure. 7 (a,b) show the protocols of optical pumping and repumping, while (c,d) show the populations of $|F=1, m_{F}=-1\rangle$ state, $|F=1, m_{F}=1\rangle$ state, $|F=2, m_{F}=-2\rangle$ state, and $|F=2, m_{F}=2\rangle$ state during the optical pumping process with and without repump beam. When the atomic spin state reaches a steady state, the population for $F=2, m_{F}=2$ with the repump beam is approximately four times larger than that without the repump beam, indicating a stronger signal for magnetic measurement. Furthermore, as the repump power increases, the population in the F=1 states gradually diminishes, leading to the suppression of SE.

 figure: Fig. 7.

Fig. 7. The protocols of optical pumping without and with repump beam (a,b). The populations of $|F=1, m_{F}=-1\rangle$ state, $|F=1, m_{F}=1\rangle$ state, $|F=2, m_{F}=-2\rangle$ state, and $|F=2, m_{F}=2\rangle$ state during the optical pumping process without (c) and with repump beam (d).

Download Full Size | PDF

4. Conclusion

In conclusion, we have demonstrated the suppression of spin-exchange (SE) relaxation in a paraffin-coated cell operating at 42 $^\circ$ C. By introducing an additional repump beam at spin constructive interference, we observe an increase in the amplitude of the magnetic resonance signal. In the case of FID, clear light narrowing is found, indicating the effective suppression of SE in the presence of a strong magnetic field. This suppression of SE relaxation results in a three-fold enhancement of the magnetic resonance signal and approximately two-fold reduction in the relaxation rate, leading to a significant improvement in sensitivity. This work has demonstrated the potential for achieving SERF-like magnetometry in a paraffin-coated cell with sub-fT/$\sqrt {\text {Hz}}$-level sensitivity in the Earth-field range. It should be noted that in theory, the amplitude of magnetic resonance with repump beam is anticipated to increase fourfold. However, the measured signal only exhibits a threefold increase. Further enhancement and suppression of the amplitude and linewidth can be achieved through improving the polarization of the spin state. Further studies are also needed to investigate light narrowing in paraffin-coated cells with high atomic density. This method has potential to be applied in the magnetic field measurement of biological samples that require a room-temperature environment, as well as geological applications in Earth field range. Our work opens up new possibilities for applications of SERF atomic system in quantum metrology achieving high sensitivity and quantum information with long coherence time.

Funding

Innovation Program for Quantum Science and Technology (2021ZD0303200); National Natural Science Foundation of China (12234014, 11654005, 12274132, 12204304, 91536114, 11874152); Fundamental Research Funds for the Central Universities; the Shanghai Municipal Science and Technology Major Project (2019SHZDZX01); Innovation Program of the Shanghai Municipal Education Commission (202101070008E00099); National Key Research and Development Program of China (2016YFA0302001); Fellowship of China Postdoctoral Science Foundation (2020TQ0193, 2021M702146).

Acknowledgments

W. Z. also acknowledges additional support from the Shanghai talent program.

Disclosures

The authors declare that there are no conflicts of interest related to this article.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. D. Budker, D. F. Kimball, S. M. Rochester, and V. V. Yashchuk, “Nonlinear magneto-optical rotation via alignment-to-orientation conversion,” Phys. Rev. Lett. 85(10), 2088–2091 (2000). [CrossRef]  

2. D. Budker, W. Gawlik, D. Kimball, S. Rochester, V. Yashchuk, and A. Weis, “Resonant nonlinear magneto-optical effects in atoms,” Rev. Mod. Phys. 74(4), 1153–1201 (2002). [CrossRef]  

3. I. K. Kominis, T. W. Kornack, J. C. Allred, and M. V. Romalis, “A subfemtotesla multichannel atomic magnetometer,” Nature 422(6932), 596–599 (2003). [CrossRef]  

4. D. Budker and M. Romalis, “Optical magnetometry,” Nat. Phys. 3(4), 227–234 (2007). [CrossRef]  

5. M. P. Ledbetter, I. M. Savukov, V. M. Acosta, D. Budker, and M. V. Romalis, “Spin-exchange-relaxation-free magnetometry with Cs vapor,” Phys. Rev. A 77(3), 033408 (2008). [CrossRef]  

6. H. B. Dang, A. C. Maloof, and M. V. Romalis, “Ultrahigh sensitivity magnetic field and magnetization measurements with an atomic magnetometer,” Appl. Phys. Lett. 97(15), 151110 (2010). [CrossRef]  

7. D. Sheng, S. Li, N. Dural, and M. V. Romalis, “Subfemtotesla scalar atomic magnetometry using multipass cells,” Phys. Rev. Lett. 110(16), 160802 (2013). [CrossRef]  

8. D. Sheng, A. R. Perry, S. P. Krzyzewski, S. Gellerl, J. Kitching, and S. Knappe, “A microfabricated optically-pumped magnetic gradiometer,” Appl. Phys. Lett. 110(3), 031106 (2017). [CrossRef]  

9. S. K. Lee, K. L. Sauer, S. J. Seltzer, O. Alem, and M. V. Romalis, “Subfemtotesla radio-frequency atomic magnetometer for detection of nuclear quadrupole resonance,” Appl. Phys. Lett. 89(21), 214106 (2006). [CrossRef]  

10. S. Groeger, G. Bison, P. E. Knowles, R. Wynands, and A. Weis, “Laser-pumped cesium magnetometers for high-resolution medical and fundamental research,” Sens. Actuators, A 129(1-2), 1–5 (2006). [CrossRef]  

11. G. Vasilakis, J. M. Brown, T. W. Kornack, and M. V. Romalis, “Limits on New Long Range Nuclear Spin-Dependent Forces Set with a K- 3He Comagnetometer,” Phys. Rev. Lett. 103(26), 261801 (2009). [CrossRef]  

12. I. Altarev, C. A. Baker, G. Ban, et al., “Test of Lorentz invariance with spin precession of ultracold neutrons,” Phys. Rev. Lett. 103(8), 081602 (2009). [CrossRef]  

13. M. S. Safronova, D. Budker, D. DeMille, Derek F. Jackson Kimball, A. Derevianko, and Charles W. Clark, “Search for new physics with atoms and molecules,” Rev. Mod. Phys. 90(2), 025008 (2018). [CrossRef]  

14. M. Jiang, H. Su, A. Garcon, X. Peng, and D. Budker, “Search for axion-like dark matter with spin-based amplifiers,” Nat. Phys. 17(12), 1402–1407 (2021). [CrossRef]  

15. I. M. Savukov, V. S. Zotev, P. L. Volegov, M. A. Espy, A. N. Matlashov, J. J. Gomez, and R. H. KrausJr, “MRI with an atomic magnetometer suitable for practical imaging applications,” J. Magn. Reson. 199(2), 188–191 (2009). [CrossRef]  

16. E. Boto, N. Holmes, J. Leggett, G. Roberts, V. Shah, S. S. Meyer, L. D. Mu noz, K. J. Mullinger, T. M. Tierney, S. Bestmann, G. R. Barnes, R. Bowtell, and M. J. Brookes, “Moving magnetoencephalography towards real-world applications with a wearable system,” Nature 555(7698), 657–661 (2018). [CrossRef]  

17. R. M. Hill, E. Boto, N. Holmes, C. Hartley, Z. A. Seedat, J. Leggett, G. Roberts, V. Shah, T. M. Tierney, M. W. Woolrich, C. J. Stagg, G. R. Barnes, R. Bowtell, R. Slater, and M. J. Brookes, “A tool for functional brain imaging with lifespan compliance,” Nat. Commun. 10(1), 4785 (2019). [CrossRef]  

18. K. Jensen, M. A. Skarsfeldt, H. Stærkind, J. Arnbak, Mikhail V. Balabas, S-P. Olesen, B. H. Bentzen, and E. S. Polzik, “Magnetocardiography on an isolated animal heart with a room-temperature optically pumped magnetomete,” Sci. Rep. 8(1), 16218 (2019). [CrossRef]  

19. K Zhu, A. M. Shah, and J Berkow, “Miniature Coil Array for Passive Magnetocardiography in Non-Shielded Environments,” IEEE J. Electromagn. RF Microw. Med. Biol. 5(2), 124–131 (2020). [CrossRef]  

20. E. Elzenheimer and H. Laufs, “Magnetic measurement of electrically evoked muscle responses with optically pumped magnetometers,” IEEE Trans. Neural Syst. Rehabil. Eng. 28(3), 756–765 (2020). [CrossRef]  

21. T. Theis, P. Ganssle, G. Kervern, S. Knappe, J. Kitching, M. P. Ledbetter, D. Budker, and A. Pines, “Parahydrogen-enhanced zero-field nuclear magnetic resonance,” Nat. Phys. 7(7), 571–575 (2011). [CrossRef]  

22. J. W. Blanchard, D. Budker, and A. Trabesinger, “Lower than low: Perspectives on zero-to ultralow-field nuclear magnetic resonance,” J. Magn. Reson. 323, 106886 (2021). [CrossRef]  

23. Y. J Kim, I. Savukov, J. H. Huang, and P. Nath, “Magnetic microscopic imaging with an optically pumped magnetometer and flux guides,” Appl. Phys. Lett. 110(4), 043702 (2017). [CrossRef]  

24. C. Deans, L. Marmugi, S. Hussain, and F. Renzoni, “Electromagnetic induction imaging with a radio-frequency atomic magnetometer,” Appl. Phys. Lett. 108(10), 103503 (2016). [CrossRef]  

25. C. Deans, L. Marmugi, and F. Renzoni, “Through-barrier electromagnetic imaging with an atomic magnetometer,” Opt. Express 25(15), 17911 (2017). [CrossRef]  

26. M. C. D. Tayler, J. Ward-Williams, and L. F. Gladden, “Ultralow-field nuclear magnetic resonance of liquids confined in ferromagnetic and paramagnetic materials,” Appl. Phys. Lett. 115(7), 072409 (2019). [CrossRef]  

27. Y. Hu, G. Z. Iwata, M. Mohammadi, E. V. Silletta, A. Wickenbrock, J. W. Blanchard, D. Budker, and A. Jerschow, “Sensitive magnetometry reveals inhomogeneities in charge storage and weak transient internal currents in Li-ion cells,” Proc. Natl. Acad. Sci. U. S. A. 117(20), 10667–10672 (2020). [CrossRef]  

28. F. P. Bustos, D. B. Calia, D. Budker, M. Centrone, J. Hellemeier, P. Hickson, R. Holzlöhner, and S. Rochester, “Remote sensing of geomagnetic fields and atomic collisions in the mesosphere,” Nat. Commun. 9(1), 3981 (2018). [CrossRef]  

29. R. Han, M. Balabas, C. Hovde, W. Li, H. M. Roig, T. Wang, A. Wickenbrock, E. Zhivun, Z. You, and D. Budker, “Is light narrowing possible with dense-vapor paraffin coated cells for atomic magnetometers?” AIP Adv. 7(12), 125224 (2017). [CrossRef]  

30. W. Happer, “Optical pumping,” Rev. Mod. Phys. 44(2), 169–249 (1972). [CrossRef]  

31. M. V. Balabas, T. Karaulanov, M. P. Ledbetter, and D. Budker, “Polarized alkali-metal vapor with minute-long transverse spin-relaxation time,” Phys. Rev. Lett. 105(7), 070801 (2010). [CrossRef]  

32. S. Pustelny, D. J. Kimball, S. Rochester, V. Yashchuk, and D. Budker, “Influence of magnetic-field inhomogeneity on nonlinear magneto-optical resonances,” Phys. Rev. A. 74(6), 063406 (2006). [CrossRef]  

33. S. Appelt, A. Ben-Amar Baranga, A. R. Young, and W. Happer, “Light narrowing of rubidium magnetic-resonance lines in high-pressure optical-pumping cells,” Phys. Rev. A 59(3), 2078–2084 (1999). [CrossRef]  

34. T. Scholtes, V. Schultze, R. IJsselsteijn, S. Woetzel, and H.-G. Meyer, “Light-narrowed optically pumped Mx magnetometer with a miniaturized Cs cell,” Phys. Rev. A 84(4), 043416 (2011). [CrossRef]  

35. A. Yang, Guo. Yang, X. Cai, Y. Xu, and Q. Lin, “A laser pump-re-pump atomic magnetometer,” Chin. Phys. B 22(12), 120702 (2013). [CrossRef]  

36. V. Schultze, T. Scholtes, R. IJsselsteijn, and H. G. Meyer, “Improving the sensitivity of optically pumped magnetometers by hyperfine repumping,” J. Opt. Soc. Am. B 32(5), 730–736 (2015). [CrossRef]  

37. Z. Ding, J. Yuan, and X. Long, “Influence of optical pumping on the transverse spin relaxation of Cs atoms in different ground-state hyperfine levels,” EPL 120(4), 43001 (2017). [CrossRef]  

38. Y. Fu, X. Liu, and J. Yuan, “Light narrowing of cesium magnetic-resonance lines in a radio-frequency atomic magnetometer,” AIP Adv. 9(1), 015304 (2019). [CrossRef]  

39. Y. Guo, S Wan, X Sun, and J. Qin, “Compact, high-sensitivity atomic magnetometer utilizing the light-narrowing effect and in-phase excitation,” Appl. Opt. 58(4), 734–738 (2019). [CrossRef]  

40. T. Scholtes, V. Schultze, R. IJsselsteijn, S. Woetzel, and H.-G. Meyer, “Light-shift suppression in a miniaturized Mx optically pumped Cs magnetometer array with enhanced resonance signal using off-resonant laser pumping,” Opt. Express 20(28), 29217 (2012). [CrossRef]  

41. A. Korver, R. Wyllie, B. Lancor, and T. Walker, “Suppression of spin-exchange relaxation using pulsed parametric resonance,” Phys. Rev. Lett. 111(4), 043002 (2013). [CrossRef]  

42. O. Katz and O. Firstenberg, “Light storage for one second in room-temperature alkali vapor,” Nat. Commun. 9(1), 2074 (2018). [CrossRef]  

43. V. V. Yashchuk, D. Budker, and J. R. Davis, “Laser frequency stabilization using linear magneto-optics,” Rev. Sci. Instrum. 71(2), 341–346 (2000). [CrossRef]  

44. N. W. Resslen, R. H. Sands, and T. E. Stark, “Measurement of Spin-Exchange Cross Sections for Cs133, Rb87, Rb85, K39, and Na23,” Phys. Rev. 184(1), 102 118 (1969). [CrossRef]  

45. W. Happer, Yuan-Yu Jau, and T. Walker, Optically pumped atoms (John Wiley and Sons, 2010).

46. A. Grosz, M. J. Haji-Sheikh, and S. C. Mukhopadhyay, High sensitivity magnetometers (Springer, Switzerland, 2017).

47. M. Auzinsh, D. Budker, and S. Rochester, Optically polarized atoms: understanding light-atom interactions (Oxford University, New York, 2010).

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (7)

Fig. 1.
Fig. 1. Experimental setup. HWP: half-wave plate. QWP: quarter-wave plate. POL: polarizer. BPD: balanced photo-detector. LIA: lock-in amplifier. FG: function generator.
Fig. 2.
Fig. 2. Magnetic resonance data for a pump-modulation frequency of 4932 Hz as a function of the leading magnetic field along the $\hat {z}$-axis. In all scheme, pump beam is on resonance with D1 $F=2$ to $F^{'}=2$ transition. (b) The probe light is 500 MHz red detunned from D2 $F=1$ to $F^{'}=1$ transitions. (c) The probe light has same frequency with scheme (a), The repump beam is on resonance with D1 $F=1$ to $F^{'}=2$ transition.
Fig. 3.
Fig. 3. The amplitude of magnetic resonance change with relative phase between pump and repump. The polarization of pump beam is $\sigma _+$. The pump power is 150 $\mu$W. (a) The polarization of repump beam is $\sigma _+$. The repump power is 150 $\mu$W. (b) The polarization of repump beam is $\sigma _+$. The repump power is 100 $\mu$W. (c) The polarization of repump beam is $\sigma _{-}$. The repump power is 100 $\mu$W.
Fig. 4.
Fig. 4. The optical rotation signal with (a) and without (b) repump beam. The Fano profile shown in (b) is caused by imbalance population of $F=1, m_{F}=1$ and $F=1, m_{F}=-1$ states. The absorption spectrum of the D2 line of $^{87}$Rb (c).
Fig. 5.
Fig. 5. Relaxation of atomic spins change with different repumping scheme. (i,ii) FID without and with per-repumping. (iii,iv) The polarized spins decay with in phase and out of phase repumping. Inset: the normalized relaxation signal.
Fig. 6.
Fig. 6. Experimental measured and theoretical simulated relaxation rate of atomic spin change with the repump power while the pump power is fixed at 150 $\mu$W.
Fig. 7.
Fig. 7. The protocols of optical pumping without and with repump beam (a,b). The populations of $|F=1, m_{F}=-1\rangle$ state, $|F=1, m_{F}=1\rangle$ state, $|F=2, m_{F}=-2\rangle$ state, and $|F=2, m_{F}=2\rangle$ state during the optical pumping process without (c) and with repump beam (d).

Equations (2)

Equations on this page are rendered with MathJax. Learn more.

S = S 1 2 + S 2 2 + 2 S 1 S 2 cos ϕ ,
i d d t ρ ^ = [ H ^ , ρ ^ ] i 1 2 ( Γ ^ ρ ^ + ρ ^ Γ ^ ) + i Λ ^ ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.