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

Tunable double split-ring resonator for quantum sensing using nitrogen-vacancy centers in diamond

Open Access Open Access

Abstract

For quantum sensing based on nitrogen-vacancies (NV) ensembles, microwave antennas can couple the microwave field to the NV center, which leads it to becoming the core of spin manipulation and can directly affect the sensitivity of quantum sensing. The double split-ring resonator is a widely used microwave device for NV ensembles due to the advantages of high radiation efficiency and uniform magnetic field in millimeter-scale areas. But the bandwidth (30 MHz) is quite narrow which limits the application in quantum sensing with NV ensembles. Here, we experimentally achieve continuous tuning of the resonant frequency of the double split-ring resonator by changing the copper sheet position on the edge of the outer ring. The frequency tuning range can reach 80 MHz, up to 2-3 times the bandwidth, which can cover the transition of the electron spin under different magnetic field conditions. The performance of the tunable antenna in the quantum operation of NV centers is verified by optically detected magnetic resonance and Rabi oscillation. This tunable antenna is promising in the fabrication of integrated and arrayed quantum sensors based on NV ensembles.

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

1. Introduction

Using the laws of quantum mechanics, microscopic quantum systems and their quantum states quantum sensing can realize high-precision and high-sensitivity measurement of specific physical parameters [1]. Nitrogen-vacancy (NV) center in diamond is an excellent solid-state quantum sensor for multi-physics parameters with high sensitivity and spatial resolution, such as magnetic field, electric field, temperature, and stress [2,3]. As a stable light-emitting defect, the NV center consists of a nitrogen atom replacing a carbon atom and a vacancy at the adjacent position as shown in Fig. 1(a). The NV center has a variety of charge states [4,5], and the negatively charged state NV (capturing an electron) is often used in quantum sensing. The NV center mentioned in this paper defaults to NV.

 figure: Fig. 1.

Fig. 1. (a) Crystal structure of NV center. (b) Level structure of NV center. (c) home-built confocal microscopy for ODMR experiment. The nanodiamonds are distributed on a 0.17-millimeter cover glass on the radiation structure.

Download Full Size | PDF

Figure 1(b) is the electron spin energy-level diagram of NV centers. For NV-based quantum sensing, the state-selective intersystem crossing (ISC) enables the electron spin state of the NV center to be optically initialized and readout at ambient temperatures [6,7]. The electron spin of NV centers can be excited from ground state 3A2 to excited state 3E by a 532nm laser. Then, the electron spin of excited state 3E can decay from the excited state 3E to ground state 3A2 by the non-radiative ISC transition and radiative transition. The electron spin in the ms = ±1 sublevels of excited state 3E has a significantly higher probability to undergo non-radiative ISC toward the singlet system and transit from excited state 3E to ms = 0 of ground state 3A2, which leads to being used to realize the initialization and readout of the NV center. The initialized electron spin in the ground state 3A2 can be manipulated by microwave (MW) and is measured by optically detected magnetic resonance (ODMR).

NV centers are along one of [1,1,1] orientations of the diamond lattice, and the Zeeman effect is sensitive to the projection of magnetic fields along the NV’s orientation. Therefore, under an external magnetic field, electron spin levels of an NV ensemble in a single crystal diamond may be non-degenerately divided into eight [8]. This property makes it an ideal vector magnetic field measurement carrier. In addition, these spin levels will also shift with the thermal expansion of the lattice, which allows the NV center to measure temperature. Based on the response of the energy levels of the NV center to the external environment, a large number of quantum measurement schemes have been proposed and realized [1,9].

Similar to most ODMR experimental setups, as shown in Fig. 1(c), the home-built confocal microscopy with a single photon detector is used for green light excitation (532 nm laser) and fluorescence detection (640 nm to 750 nm). During microwave scanning, the fluorescence intensity is weakened when the microwave resonates with the electron spin transition. This technique is called continuous wave optically detected magnetic resonance (CW-ODMR). Using microwave switches and acousto-optic modulators (AOM), we can generate microwave and light pulses on this setup to achieve pulsed ODMR, Rabi, and other pulsed experiments.

For NV-based quantum sensing schemes, the microwave is the key to manipulating the spins of NV centers. The magnetic field component in the microwave field will couple with multiple NV electron spins to realize the control of the quantum states. To improve microwave coupling and manipulation efficiency, a variety of microwave antennas are designed to adapt to specific environments and samples [1017]. Since our target is to manipulate ensembles or an array of NV centers within a certain area, the microwave device is also required to generate a relatively uniform microwave field on the millimeter-scale area in addition to the radiation requirements of the single spin experiment. In single NV center experiments, Omega ring and microstrip line [1821] waveguides are widely used due to the low requirements on microwave range and magnetic field uniformity. A microstrip line can provide an extremely high magnetic field if the NV center is close enough, but the intensity decreases with the square of the distance. Omega ring can generate a uniform field inside the ring, but the ring must be small to get a strong field. The other advantage of the Omega ring and microstrip line is their available bandwidth as broad as 1GHz or more.

However, the susceptibilities of electron spin energy levels with temperature and magnetic field are -74 kHz/K and 2.8MHz/Gauss respectively [2227], and the target we want to measure may only have a temperature change of tens of K or a magnetic field change of several Gauss. In this case, the measured energy level shift is only a few MHz usually. The inhomogeneous broadening linewidth of the NV ensemble electron spins in nanodiamonds is about 10MHz, so we may only need an antenna with a bandwidth of 30 to 40MHz in the experiment. Considering the randomness of nanodiamond lattice orientation, the uniformity of the magnetic field is less important, keeping it in the same order of magnitude is enough. Therefore, we put bandwidth on the back burner and adopted a double 89 split-ring resonator (DSRR) as the microwave antenna [2830], which can effectively couple 90 the microwave magnetic field to many NV centers with different orientations in the 91 millimeter-scale area. This kind of structure has been well-developed and used in 92 metamaterials [3133], microwave filters [34], and antennas. Generally speaking, the antenna 93 bandwidth is defined as S11 ≤ -10 dB [35]. However, due to the shortcoming of DSRR that the bandwidth is very narrow, only about 30MHz, which is only enough to cover one or two dips on ODMR spectrum. For example, in the quantum sensing based on NV ensembles in our laboratory, the frequency range of four levels (different orientations of the same electron spin) is about 100MHz, which is hard to cover by one DSRR antenna. So, in this article, we studied a way to tune the resonance point of the double-split ring by changing the capacitance between the outer ring and the ground. The frequency range of the DSRR can be shifted to meet the experimental requirements and cover the spin levels we care about by a piece of a moveable copper sheet, which will significantly benefit our quantum sensing experiment. In this article, we did simulations (CST Microwave Studio) and experiments (network analyzer, Rohde & Schwarz-ZNA26) on the tunable microwave device base on the analysis of the original DSRR, and the effect of our design is verified by the ODMR and Rabi experiment on nanodiamonds with an NV ensemble. This antenna is promising in the fabrication of integrated quantum detectors based on the ensemble of NV centers, nanodiamonds, or an array of nanodiamonds [36].

2. Model design and analysis

The design of DSRR starts from the closed metal ring whose electrical size at the first resonance point does not satisfy the subwavelength condition [36]. The resonant frequency of the entire unit can be gradually reduced by opening a slit on the edge of the simply closed metal ring, which is called the split resonant ring, and the structure with an approximately optimal electrical size that satisfies the subwavelength condition can be obtained. The resonant frequency can be further lowered by adding an inner ring to the closed metal ring and opening a slot at the symmetrical position of the second ring. The electromagnetic coupling between the inner and outer rings will also reduce the resonance frequency by bringing a capacitance. Then the structure with the lowest resonant frequency at this time is the DSRR unit. And the DSRR is usually etched on the dielectric substrate which is also helpful in reducing the electrical size. The resonance of DSRR mainly depends on the resonance of the outer ring because most of the energy of the electric field is concentrated in the outer ring [14,37]. The optimized DSRR structure satisfies the sub-wavelength condition and has a high symmetry. It has been so perfect and is widely used in the microwave and radio frequency area. As shown in Fig. 2(a), w is the width of the rings and wire, r1 represents the inner radius of the inner ring, r2 represents the inner radius of the outer ring, r0 = (r1 + r2 + w)/2 is the mean radius between two rings, d0 is the splits, and d2 is the distance between the microstrip line and outer ring. These parameters determine the resonance performance of the DSRR. The equivalent circuit model of DSRR is an LC resonant circuit, as shown in Fig. 2. (b) [37], C1 represents the equivalent capacitance of the ring, C1 = 2πr0C0, C0 represents the capacitance per unit length, CS is the series capacitance, CS = C1/4, LS represents the equivalent inductance of the ring, then the resonant frequency of the entire resonant tank is as follows:

$$\begin{array}{{c}} {{{\boldsymbol f}_{\boldsymbol r}} = \frac{\mathbf 1}{{\mathbf 2{\boldsymbol \pi }\sqrt {{{\boldsymbol L}_{\boldsymbol s}}{{\boldsymbol C}_{\boldsymbol s}}} }}} \end{array}$$

 figure: Fig. 2.

Fig. 2. (a) Model design of the DSRR. (b) Equivalent Circuit of the DSRR. (c) Simulated and experimental S-parameters of loading sample and unloaded (without sample) resonator. (d) The simulated and experimental resonant frequency of these DSRRs with different radii of rings.

Download Full Size | PDF

Based on this simplified model, we did a series of simulations and experiment to optimize the structure. Metal copper is used as the material of the ring, and the intermediate dielectric substrate is Rogers 6010, whose dielectric constant is 10.2. The backside is covered with copper as the ground wire, and the holes at the corners are connected to the ground on the backside. The microwave is provided through a microstrip line that meets the impedance matching of 50 Ohms and is coupled to the double ring through the gap. The gap between the microstrip line and the ring determines the S11 by impendence matching. Theoretically, the smaller the gap, the higher the coupled microwave power. The optimized size is 0.05 mm, but because the precision of the copper sinking process is hard to achieve, the minimum size that can be well achieved by the process is selected to be 0.1 mm. The S11 of simulation and experimental measurement with or without samples is shown in Fig. 2(c). And the deviation between simulation and experimental measurement is less than 8 MHz. In addition, compared to unloaded, a small piece of glass (0.1 mm) with nanodiamonds was glued on the structure, which cause a change in permittivity and shifted the resonant frequency of the structure in the experimental measurement loading sample. So we can find some sample-related deviation from the experiment with simulation.

The quality factor Q expresses the magnitude of the resonant frequency relative to the bandwidth, and also means the ratio of the electromagnetic energy in the resonator to the energy lost in one cycle. A larger Q results in a larger magnetic field amplitude at the center of the resonator, but usually at the expense of narrow bandwidth. The resonant frequency of the resonant ring can be tuned within a range by changing the mean radius r0. In the simulation and experiment, we keep widths and gaps constant, then tune r1 and r2 together. The resonant frequency corresponding to r1 (r1 = r2-1.2 mm) is plotted in Fig. 2(d). As the radius increases, the frequency gradually decreases, so we chose three radius sizes with frequencies close to 2.87 GHz for casting, to test the performance experimentally. Finally, after the network analyzer test, the optimized inner diameter of the size is 1.1 mm, that is, the double split-ring structure with an average radius of 2.2 mm, resonant frequency f = 2.878 GHz, S11=-31 dB, bandwidth is 30 MHz, Q = 96, it can be estimated that 99% of the microwave power at the center frequency can be directly coupled to the resonator.

The energy stored in the resonator is (1-S112)QPin/wr, Pin represents the input microwave power delivered to the device, and wr represents the resonant frequency. The magnetic field energy is the strongest at the resonant point [33].

Optimizing the structural parameters of the antenna can tune the resonant frequency to meet the requirement of different experiments. But repeated antenna replacements are time-consuming and risky to damage samples, especially nanodiamond arrays and fragile ultrathin diamonds. Therefore, as theoretically suggested in the literature [33], we choose to paste a small piece of conductor on the surface of a microwave antenna to fine-tune the resonant frequency. As discussed earlier, a small piece of copper sheet is used to slightly change the capacitance of the DSRR structure, lowering the resonant frequency of the antenna.

3. Simulation and experiment of the tunable radiation structure

Frequency tuning using stubs and strips is a common practice in microwave circuits [38]. In general, stub tuning is simply the process of adding a length of the transmission line to the existing length in either series or shunt circuit configuration to match the line to the load. This is achieved by placing a specific length of stub a specific distance away from the load. stub tuners are simple to implement and inexpensive to make, they only require the same material as making transmission lines. The stub is designed to counteract the reactance component of the load to be matched, and special stubs help increase the bandwidth of the circuit. In our case, the copper tape affects the capacitive coupling between the outer ring and the ground plane (backside of the PCB) and hence shifts the resonance frequency.

The copper sheet not only changes the area of the outer ring but also changes the shape of the outer ring. The definition of the copper sheet positions on the microwave device is shown in Fig. 3(a). The materials are the same as mentioned in Sec. 2, and parameters of the structure are listed in Table 1. To study the effect of the shape of the outer ring on the resonant frequency, we simulated and measured the resonant frequency of the antenna with a 1.5mm × 1.5 mm copper sheet at different positions, shown in Fig. 3(b). During the experiment, we keep the coincidence distance dc = 0.2 mm to balance the stability and tuning range of the antenna. Smaller coincidence distance dc provides a larger tuning range if the copper sheet contacts the outer ring well. The relationships between coincidence distance dc and resonant frequency at the different positions are simulated in Supplement 1.

Tables Icon

Table 1. Optimized parameters of the radiation structure

 figure: Fig. 3.

Fig. 3. (a) Position of the copper sheet in tuning the resonance frequency. The overlapping of the copper sheet and the outer ring is represented by the coincidence distance dc. In the experiment of resonance turning, we tried to keep dc = 0.2 mm (±0.1 mm). (b) simulation and measurement of the tuned resonance frequency of the structure by moving the copper sheet. The data of S11 is plotted in Fig. 2(c). (c) Resonant Rabi frequency for different microwave frequencies on the same spin level with and without tuning copper sheet on the antenna. The spin level is scanned by a magnetic field. (d) Resonant Rabi frequency of the fixed spin levels with the tuning of the microwave antenna.

Download Full Size | PDF

For a specific energy level of NV centers in a nanodiamond, the transition magnetic moment $\vec{\mu }$ is fixed, so the Rabi frequency can be expressed as:

$$\begin{array}{{c}} {\boldsymbol \Omega = \frac{{\vec{{\boldsymbol \mu }} \cdot \vec{{\boldsymbol B}}}}{\hbar }} \end{array}$$

The direction of the alternating magnetic field generated by the antenna is perpendicular to the plane of the device, which is also fixed. So, the resonant Rabi frequency of the specific transition is proportional to the amplitude of the alternating magnetic field $\vec{B}$. In this case, we can use the Rabi frequency of a specific energy level to characterize the performance of the antenna.

In order to reflect the amplitude of the magnetic field from the antenna, we did Rabi oscillation before and after tuning the antenna with the copper sheet. In the two cases, we choose the same energy level of NV centers in the same nanodiamond, fine-tuning the external magnetic field to scan the frequency of the electron spin levels and measure the resonant Rabi frequency at each frequency. Then the Rabi frequency, shown in Fig. 3(c), can reflect the amplitude of the AC magnetic field from the microwave antenna. The variations of Rabi frequencies show the full width at half maximum of the antenna as about 40 MHz. Tuning the antenna frequency with a copper sheet, it is observed that the peak of Rabi frequencies shifted clearly, which is consistent with the previous simulation and network analyzer test results.

Under a fixed static magnetic field (fixed energy level), the antenna frequency is tuned by changing the position of the copper sheet. Within the tunable range of about 80MHz, the Rabi frequency changes by about two times, which means the magnetic field amplitude at the transition frequency is doubled by tuning the antenna by the copper sheet. We also did the Rabi test with two different microwave powers by scan the resonance frequency of a NV level with fine-tuning the external static magnetic field, and the results are consistent with our expectations (Fig. 3(d)). The Rabi frequency changed from 6MHz to 13.3MHz when the input microwave power is 10W. Since the lattice orientation of nanodiamonds is completely random, we did not calculate the magnetic field precisely. The Rabi frequency can be further increased by using bulk diamonds with the [1,1,1] direction perpendicular to the microwave device.

4. ODMR with the tunable double split-ring resonator

In quantum sensing based on NV centers, the amplitude of the microwave magnetic field at the sample directly affects the signal contrast, which determines the measurement sensitivity. We used ODMR and Rabi experiments to test and verify the performance of the microwave device.

Using CW-ODMR, as shown in Fig. 4(a), we can find that the four pairs energy levels of the NV ensembles in different orientations are separated by the Zeeman effect from the magnetic field, and the levels located near the resonant frequency of the antenna (${f_r}$=2878 MHz) has better signal contrast. The eight inversed peaks also mean the nanodiamond is an ideal single crystal. After the 1.5mm × 1.5 mm copper sheet is pasted at 30° outside the outer ring to adjust the microwave device, as shown in Fig. 3(a), the resonance frequency of the antenna is tuned down to 2830 MHz. With this copper sheet, the resonant frequency of the antenna can completely cover one of the electron spin energy levels, and at the same time, it is relatively close to other energy levels with the spin ms = -1at the lower frequency side. The CW-ODMR result with the tuned antenna is plotted in Fig. 4(b), the contrasts of the four levels at left (ms = -1) are much better than the right four (ms=+1). Compared with the ODMR spectrum of the untuned antenna in Fig. 4(a), the contrast of the three energy levels on the left is significantly improved, and the fourth energy level at 2844 MHz doesn’t change much because the detuning magnitude is close to the previous one. In this case, by measuring the one-sided CW-ODMR spectrum at a specific temperature, we have been able to carry out the vector measurement of a small-scale variation magnetic field [39,40].

 figure: Fig. 4.

Fig. 4. (a) CW-ODMR spectrum of a nanodiamond with a magnetic field when the antenna is untuned, ${f_r}$=2878 MHz. The red crive is the fitted line. (b) CW-ODMR spectrum when the resonance of the antenna ${f_r}$ is tuned to 2830 MHz by the copper sheet. The red curve is the fitted line. (c) Rabi oscillation of the third spin level (2819 MHz) with an untuned antenna, Ωuntuned = 2π×7.1 MHz. (d) Rabi oscillation of the same spin level close to the resonance of the antenna tuned by the copper sheet, Ωtuned = 2π×13.3 MHz.

Download Full Size | PDF

5. Conclusion

In this work, we analyzed and optimized the classical DSRR antenna for a millimeter-scale area first. based on the DSRR structure, we tune the antenna via a moveable copper sheet. Compared with the work of K. Bayat et al. [33], the bandwidth is increased. Besides simulation and experiment, the tunable antenna is applied to an NV ensemble in a nanodiamond. The performance of the microwave device is verified by CW-ODMR and Rabi on the NV ensemble. In more than 50 MHz range, we can tune the antenna to cover the resonance of the level and reach 13.3 MHz Rabi frequency at 10 W microwave input. For different experimental purposes and environments, when the spin level is shifted by a magnetic field, we can quickly change the resonant frequency of the antenna without moving the sample, which provides great convenience for many experiments, for example the detection of nano magnetic particles or two-dimensional magnetic materials. This tunable antenna has considerable application prospects in nanodiamond arrays and wide-field experiments of NV ensembles.

Funding

National Key Research and Development Program of China (2020YFA0309400); National Natural Science Foundation of China (11904070, 12174081); Fundamental Research Funds for the Central Universities (JZ2021HGQA0260).

Acknowledgments

We thanks the National Natural Science Foundation for help identifying collaborators for this work.

Disclosures

The authors declare no conflicts of interest.

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.

Supplemental document

See Supplement 1 for supporting content.

References

1. C. L. Degen, F. Reinhard, and P. Cappellaro, “Quantum sensing,” Rev. Mod. Phys. 89(3), 035002 (2017). [CrossRef]  

2. M. W. Doherty, N. B. Manson, P. Delaney, F. Jelezko, J. Wrachtrup, and L. C. L. Hollenberg, “The nitrogen-vacancy colour centre in diamond,” Phys. Rep. 528(1), 1–45 (2013). [CrossRef]  

3. J. F. Barry, J. M. Schloss, E. Bauch, M. J. Turner, C. A. Hart, L. M. Pham, and R. L. Walsworth, “Sensitivity optimization for NV-diamond magnetometry,” Rev. Mod. Phys. 92(1), 015004 (2020). [CrossRef]  

4. P. Qian, Y. Zhai, J. Hu, R. Zheng, B. Chen, and N. Xu, “Multicolor-illuminated charge-state dynamics of the nitrogen-vacancy center in diamond,” Phys. Rev. A 106(3), 033506 (2022). [CrossRef]  

5. S. Dhomkar, J. Henshaw, H. Jayakumar, and C. A. Meriles, “Long-term data storage in diamond,” Sci. Adv. 2, (2016).

6. N. B. Manson, J. P. Harrison, and M. J. Sellars, “Nitrogen-vacancy center in diamond: Model of the electronic structure and associated dynamics,” Phys. Rev. B 74(10), 104303 (2006). [CrossRef]  

7. P. Delaney, J. C. Greer, and J. A. Larsson, “Spin-polarization mechanisms of the nitrogen-vacancy center in diamond,” Nano Lett. 10(2), 610–614 (2010). [CrossRef]  

8. B. Chen, B. Chen, X. Zhu, J. Fan, Z. Yu, P. Qian, and N. Xu, “Sensitivity-enhanced magnetometry using nitrogen-vacancy ensembles via adaptively complete transitions overlapping,” Rev. Sci. Instrum. 93(12), 125105 (2022). [CrossRef]  

9. W. W. Hsiao, Y. Y. Hui, P. C. Tsai, and H. C. Chang, “Fluorescent Nanodiamond: A Versatile Tool for Long-Term Cell Tracking, Super-Resolution Imaging, and Nanoscale Temperature Sensing,” Acc. Chem. Res. 49(3), 400–407 (2016). [CrossRef]  

10. O. R. Opaluch, N. Oshnik, R. Nelz, and E. Neu, “Optimized Planar Microwave Antenna for Nitrogen Vacancy Center Based Sensing Applications,” Nanomaterials 11(8), 2108 (2021). [CrossRef]  

11. A. Savitsky, J. Zhang, and D. Suter, “Variable bandwidth, high efficiency microwave resonator for control of spin-qubits in nitrogen-vacancy centers,” arXiv, arXiv preprint arXiv:2301.03911 (2023).

12. V. V. Soshenko, O. R. Rubinas, V. V. Vorobyov, S. V. Bolshedvorskii, P. V. Kapitanova, V. N. Sorokin, and A. V. Akimov, “Microwave Antenna for Exciting Optically Detected Magnetic Resonance in Diamond NV Centers,” Bull. Lebedev Phys. Inst. 45(8), 237–240 (2018). [CrossRef]  

13. V. Yaroshenko, V. Soshenko, V. Vorobyov, S. Bolshedvorskii, E. Nenasheva, I. Kotel’nikov, A. Akimov, and P. Kapitanova, “Circularly polarized microwave antenna for nitrogen vacancy centers in diamond,” Rev. Sci. Instrum. 91(3), 035003 (2020). [CrossRef]  

14. M. Shamonin, E. Shamonina, V. Kalinin, and L. Solymar, “Resonant frequencies of a split-ring resonator: Analytical solutions and numerical simulations,” Microw. Opt. Technol. Lett. 44(2), 133–136 (2005). [CrossRef]  

15. N. Zhang, C. Zhang, L. Xu, M. Ding, W. Quan, Z. Tang, and H. Yuan, “Microwave Magnetic Field Coupling with Nitrogen-Vacancy Center Ensembles in Diamond with High Homogeneity,” Appl. Magn. Reson. 47(6), 589–599 (2016). [CrossRef]  

16. Y. Chen, H. Guo, W. Li, D. Wu, Q. Zhu, B. Zhao, L. Wang, Y. Zhang, R. Zhao, W. Liu, F. Du, J. Tang, and J. Liu, “Large-area, tridimensional uniform microwave antenna for quantum sensing based on nitrogen-vacancy centers in diamond,” Appl. Phys. Express 11(12), 123001 (2018). [CrossRef]  

17. J. Herrmann, M. A. Appleton, K. Sasaki, Y. Monnai, T. Teraji, K. M. Itoh, and E. Abe, “Polarization- and frequency-tunable microwave circuit for selective excitation of nitrogen-vacancy spins in diamond,” Appl. Phys. Lett. 109(18), 183111 (2016). [CrossRef]  

18. X. Yang, N. Zhang, H. Yuan, G. Bian, P. Fan, and M. Li, “Microstrip-line resonator with broadband, circularly polarized, uniform microwave field for nitrogen vacancy center ensembles in diamond,” AIP Adv. 9(7), 075213 (2019). [CrossRef]  

19. L. Qin, Y. Fu, S. Zhang, J. Zhao, J. Gao, H. Yuan, Z. Ma, Y. Shi, and J. Liu, “Near-field microwave radiation function on spin assembly of nitrogen vacancy centers in diamond with copper wire and ring microstrip antennas,” Jpn. J. Appl. Phys. 57(7), 072201 (2018). [CrossRef]  

20. K. Sasaki, Y. Monnai, S. Saijo, R. Fujita, H. Watanabe, J. Ishi-Hayase, K. M. Itoh, and E. Abe, “Broadband, large-area microwave antenna for optically detected magnetic resonance of nitrogen-vacancy centers in diamond,” Rev. Sci. Instrum. 87(5), 053904 (2016). [CrossRef]  

21. F. Ge, B. Chen, Y. Wang, F. Zhou, R. Zheng, X. Yang, P. Qian, and N. Xu, “A Wideband Balun-Based Microwave Device for Quantum Information Processing With Nitrogen–Vacancy Centers in Diamond,” J. Lightwave Technol. 40(23), 7572–7577 (2022). [CrossRef]  

22. N. Wang, G.-Q. Liu, W.-H. Leong, H. Zeng, X. Feng, S.-H. Li, F. Dolde, H. Fedder, J. Wrachtrup, X.-D. Cui, S. Yang, Q. Li, and R.-B. Liu, “Magnetic Criticality Enhanced Hybrid Nanodiamond Thermometer under Ambient Conditions,” Phys. Rev. X 8(1), 011042 (2018). [CrossRef]  

23. G.-Q. Liu, X. Feng, N. Wang, Q. Li, and R.-B. Liu, “Coherent quantum control of nitrogen-vacancy center spins near 1000 kelvin,” Nat. Commun. 10(1), 1344 (2019). [CrossRef]  

24. D. M. Toyli, D. J. Christle, A. Alkauskas, B. B. Buckley, C. G. Van de Walle, and D. D. Awschalom, “Measurement and Control of Single Nitrogen-Vacancy Center Spins above 600 K,” Phys. Rev. X 2(3), 031001 (2012). [CrossRef]  

25. V. M. Acosta, E. Bauch, M. P. Ledbetter, A. Waxman, L. S. Bouchard, and D. Budker, “Temperature dependence of the nitrogen-vacancy magnetic resonance in diamond,” Phys. Rev. Lett. 104(7), 070801 (2010). [CrossRef]  

26. V. M. Acosta, E. Bauch, M. P. Ledbetter, C. Santori, K. M. C. Fu, P. E. Barclay, R. G. Beausoleil, H. Linget, J. F. Roch, F. Treussart, S. Chemerisov, W. Gawlik, and D. Budker, “Diamonds with a high density of nitrogen-vacancy centers for magnetometry applications,” Phys. Rev. B 80(11), 115202 (2009). [CrossRef]  

27. M. W. Doherty, V. M. Acosta, A. Jarmola, M. S. J. Barson, N. B. Manson, D. Budker, and L. C. L. Hollenberg, “Temperature shifts of the resonances of the NV−center in diamond,” Phys. Rev. B 90(4), 041201 (2014). [CrossRef]  

28. J. B. Pendry, A. J. Holden, D. J. Robbins, and W. J. Stewart, “Magnetism from conductors and enhanced nonlinear phenomena,” IEEE Trans. Microwave Theory Techn. 47(11), 2075–2084 (1999). [CrossRef]  

29. J. B. Pendry, “Negative refraction makes a perfect lens,” Phys. Rev. Lett. 85(18), 3966–3969 (2000). [CrossRef]  

30. J.-W. Fan, J. Xu, M.-T. Cheng, and Y. Yang, “Vacuum induced transparency in metamaterials,” Opt. Express 26(15), 19498–19512 (2018). [CrossRef]  

31. J. Martel, R. Marques, F. Falcone, J. D. Baena, F. Medina, F. Martin, and M. Sorolla, “A new LC series element for compact bandpass filter design,” IEEE Microw. Wireless Compon. Lett. 14(5), 210–212 (2004). [CrossRef]  

32. W. N. Hardy and L. A. Whitehead, “Split-ring resonator for use in magnetic resonance from 200–2000MHz,” Rev. Sci. Instrum. 52(2), 213–216 (1981). [CrossRef]  

33. K. Bayat, J. Choy, M. Farrokh Baroughi, S. Meesala, and M. Loncar, “Efficient, Uniform, and Large Area Microwave Magnetic Coupling to NV Centers in Diamond Using Double Split-Ring Resonators,” Nano Lett. 14(3), 1208–1213 (2014). [CrossRef]  

34. F. C. Fehsenfeld, K. M. Evenson, and H. P. Broida, “Microwave Discharge Cavities Operating at 2450 MHz,” Rev. Sci. Instrum. 36(3), 294–298 (1965). [CrossRef]  

35. N. Atanasov, G. Atanasova, and B. Atanasov, “Wearable textile antennas with high body-antenna isolation: design, fabrication, and characterization aspects,” IntechOpen, 2020.

36. Z. Xu, L. Wang, X. Huan, H. Lee, J. Yang, Z. Zhou, M. Chen, S. Hu, Y. Liu, S.-P. Feng, T. Zhang, F. Xu, Z. Chu, and J. T. Kim, “On-Demand, Direct Printing of Nanodiamonds at the Quantum Level,” Adv. Sci. 9(5), 2103598 (2022). [CrossRef]  

37. J. D. Baena, J. Bonache, F. Martín, R. M. Sillero, F. Falcone, T. Lopetegi, M. A. G. Laso, J. García-García, I. Gil, M. F. Portillo, and M. Sorolla, “Equivalent-circuit models for split-ring resonators and complementary split-ring resonators coupled to planar transmission lines,” IEEE Trans. Microwave Theory Techn. 53(4), 1451–1461 (2005). [CrossRef]  

38. L. Wong K and F. Lin Y, “Circularly polarised microstrip antenna with a tuning stub[J],” Electron. Lett. 34(9), 831–832 (1998). [CrossRef]  

39. T. Weggler, C. Ganslmayer, F. Frank, T. Eilert, F. Jelezko, and J. Michaelis, “Determination of the Three-Dimensional Magnetic Field Vector Orientation with Nitrogen Vacany Centers in Diamond,” Nano Lett. 20(5), 2980–2985 (2020). [CrossRef]  

40. S. Kitazawa, Y. Matsuzaki, S. Saijo, K. Kakuyanagi, S. Saito, and J. Ishi-Hayase, “Vector-magnetic-field sensing via multifrequency control of nitrogen-vacancy centers in diamond,” Phys. Rev. A 96(4), 042115 (2017). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       Supplementary materials

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 (4)

Fig. 1.
Fig. 1. (a) Crystal structure of NV center. (b) Level structure of NV center. (c) home-built confocal microscopy for ODMR experiment. The nanodiamonds are distributed on a 0.17-millimeter cover glass on the radiation structure.
Fig. 2.
Fig. 2. (a) Model design of the DSRR. (b) Equivalent Circuit of the DSRR. (c) Simulated and experimental S-parameters of loading sample and unloaded (without sample) resonator. (d) The simulated and experimental resonant frequency of these DSRRs with different radii of rings.
Fig. 3.
Fig. 3. (a) Position of the copper sheet in tuning the resonance frequency. The overlapping of the copper sheet and the outer ring is represented by the coincidence distance dc. In the experiment of resonance turning, we tried to keep dc = 0.2 mm (±0.1 mm). (b) simulation and measurement of the tuned resonance frequency of the structure by moving the copper sheet. The data of S11 is plotted in Fig. 2(c). (c) Resonant Rabi frequency for different microwave frequencies on the same spin level with and without tuning copper sheet on the antenna. The spin level is scanned by a magnetic field. (d) Resonant Rabi frequency of the fixed spin levels with the tuning of the microwave antenna.
Fig. 4.
Fig. 4. (a) CW-ODMR spectrum of a nanodiamond with a magnetic field when the antenna is untuned, ${f_r}$=2878 MHz. The red crive is the fitted line. (b) CW-ODMR spectrum when the resonance of the antenna ${f_r}$ is tuned to 2830 MHz by the copper sheet. The red curve is the fitted line. (c) Rabi oscillation of the third spin level (2819 MHz) with an untuned antenna, Ωuntuned = 2π×7.1 MHz. (d) Rabi oscillation of the same spin level close to the resonance of the antenna tuned by the copper sheet, Ωtuned = 2π×13.3 MHz.

Tables (1)

Tables Icon

Table 1. Optimized parameters of the radiation structure

Equations (2)

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

f r = 1 2 π L s C s
Ω = μ B
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.