This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

thanks: Present address: Department of Physics, National Tsing Hua University, Hsinchu 30013, Taiwanthanks: Present address: QuTech and Kavli Institute of Nanoscience, Delft University of Technology, Delft, 2600 GA, The Netherlandsthanks: Present address: Liquid Instruments, San Diego, CA 92130, USAthanks: Present address: School of Engineering, Brown University, Providence, RI 02912, USA

A near-quantum-limited diamond maser amplifier operating at millikelvin temperatures

Morihiro Ohta Experimental Quantum Information Physics Unit, Okinawa Institute of Science and Technology Graduate University, Onna, Okinawa 904-0495, Japan    Ching-Ping Lee The Science and Technology Group, Okinawa Institute of Science and Technology Graduate University, Onna, Okinawa 904-0495, Japan    Vincent P.M. Sietses Experimental Quantum Information Physics Unit, Okinawa Institute of Science and Technology Graduate University, Onna, Okinawa 904-0495, Japan    I. Kostylev The Science and Technology Group, Okinawa Institute of Science and Technology Graduate University, Onna, Okinawa 904-0495, Japan    J.R. Ball Quantum Dynamics Unit, Okinawa Institute of Science and Technology Graduate University, Onna, Okinawa 904-0495, Japan    P. Moroshkin Quantum Dynamics Unit, Okinawa Institute of Science and Technology Graduate University, Onna, Okinawa 904-0495, Japan    T. Hamamoto Experimental Quantum Information Physics Unit, Okinawa Institute of Science and Technology Graduate University, Onna, Okinawa 904-0495, Japan    Y. Kobayashi Sumitomo Electric Industries Ltd., Itami, Hyogo 664-0016, Japan    S. Onoda Quantum Materials and Applications Research Center, Takasaki Institute for Advanced Quantum Science, National Institutes for Quantum Science and Technology, Takasaki, Gunma 370-1292, Japan    T. Ohshima Quantum Materials and Applications Research Center, Takasaki Institute for Advanced Quantum Science, National Institutes for Quantum Science and Technology, Takasaki, Gunma 370-1292, Japan Department of Materials Science, Tohoku University, Sendai, Miyagi 980-8579, Japan    J. Isoya Graduate School of Pure and Applied Sciences, University of Tsukuba, Tsukuba, Ibaraki 305-8550, Japan    Hiroki Takahashi Experimental Quantum Information Physics Unit, Okinawa Institute of Science and Technology Graduate University, Onna, Okinawa 904-0495, Japan    Yuimaru Kubo yuimaru.kubo@oist.jp The Science and Technology Group, Okinawa Institute of Science and Technology Graduate University, Onna, Okinawa 904-0495, Japan
Abstract

Microwave quantum technologies require amplification of weak signals with minimal added noise at millikelvin temperatures. This stringent demand has been met with superconducting parametric amplifiers. While masers offer another fundamental approach, their dependence on cryogenic operation has historically posed challenges for classical communication technologies — a barrier that does not apply to microwave quantum technologies. In this work, we demonstrate an ultra-low-noise maser amplifier utilizing impurity spins in diamond at millikelvin temperatures. We achieve power gains exceeding 30dB\mathrm{30\,dB}, a minimum noise temperature of 0.86K\mathrm{0.86\,K} (corresponding to 2.22.2 noise photons), and a maximum 1dB\mathrm{1\,\text{dB}} output compression point of 63dBm\mathrm{-63\,dBm} at 6.595GHz\mathrm{6.595\,\text{GHz}}. Our results establish masers as viable components of microwave quantum technologies.

The tremendous developments in quantum information technologies that operate at millikelvin temperatures in the microwave frequency band over the past two decades have relied on ultra-low-noise amplification with near-quantum-limited noise [1, 2]. This demand has been met primarily by superconducting circuit technologies [3, 4, 5, 6]. Microwave amplification via masers (microwave amplification by stimulated emission of radiation) — the microwave-frequency counterpart of the laser — has been predicted to possess quantum-limited noise performance, particularly when thermal noise is absent[7] and sufficient population inversion is achieved [8, 9, 10].

Diamond masers, particularly those based on negatively charged nitrogen-vacancy (NV\text{NV}^{-}) centers, have emerged as both oscillators [9, 11] and amplifiers at room temperature [10] and cryogenic temperatures [12]. In addition, cavity cooling [13, 14, 10] and enhanced magnetometry [15] at room temperature have been demonstrated, suggesting the potential of diamond-based masers as resources for microwave quantum technology applications. However, the requirement for optical pumping to establish population inversion in NV\text{NV}^{-} centers presents a significant challenge for applications at millikelvin temperatures due to the high laser powers required and the limited cooling power of dilution refrigerators. In this work, we achieve population inversion in substitutional nitrogen defects (P1 centers) in diamond through microwave pumping at power levels of 100100 nanowatts or less. We demonstrate amplification with near-quantum-limited noise at millikelvin temperatures.

Maser amplifier device and setup

Refer to caption
Figure 1: Diamond maser amplifier device and setup. (A) Device schematic. A diamond crystal containing P1 centers is positioned within a loop-gap microwave resonator. Both pump and probe microwave tones are introduced via a coupling loop antenna. (B) Experimental setup at the cold plate (100mK100\,\text{mK}) and mixing chamber plate (13mK13\,\text{mK}) in a dilution refrigerator. The resonator is enclosed in a copper housing and thermalized to the mixing chamber plate. Amplified microwave signals are routed to the measurement line through a series of two circulators. Additionally, a temperature-variable noise source, weakly thermalized to 100mK100\,\text{mK}, generates input noise characterized by a temperature TinT_{\text{in}}.

Fig. 1A illustrates the maser amplifier device. It consists of a loop-gap microwave resonator [16] that houses a diamond crystal containing approximately 16ppm16\,\text{ppm} of P1 centers, together with about 2ppm2\,\text{ppm} of negatively-charged nitrogen-vacancy (NV-) centers [17, 16], with dimensions of 3×1.5×0.5mm33\times 1.5\times 0.5\,\text{mm}^{3}. The resonator’s frequency is ωr/2π6.595GHz\omega_{r}/2\pi\approx 6.595\,\text{GHz} and its internal quality factor is Qint3000Q_{\text{int}}\approx 3000. In this work, we operate the resonator with two different external quality factors, Qext250Q_{\text{ext}}\approx 250 and 730\approx 730, by adjusting the number of turns of the coupling loop antenna.

Fig. 1B illustrates a simplified measurement setup inside the refrigerator. The maser amplifier device is thermalized to the mixing chamber plate at 13mK13\,\text{mK} in a dilution refrigerator. Two microwave input lines, one heavily attenuated for the probe and another moderately attenuated for the pump, are combined through a directional coupler. After being combined, the pump and probe microwave tones go to the maser amplifier device through a circulator, which in turn directs the reflected signal to the measurement line and decouples the thermal noise from the following HEMT (high electron mobility transistor) amplifier placed at the 4K4\,\text{K} stage. An impedance-matched temperature-variable noise source [18] is integrated along the probe line, which is weakly thermalized to the cold plate at approximately 100mK100\,\text{mK}. We use this noise source to characterize the noise temperature of our diamond maser amplifier.

Operating principle

Refer to caption
Figure 2: Operating principle of the diamond maser amplifier. (A) Illustration of a P1 center in diamond and its energy levels under a static magnetic field of BDC232mTB_{\text{DC}}\approx 232\,\text{mT}, aligned parallel to the [001][001] crystallographic axis. (B) System dynamics with a pump: 1. A microwave pump tone at ω0\omega_{0} excites P10 spins at a rate of Γp\Gamma_{\text{p}}. 2. This pump induces the excitation of one P1+ and one P1- spin by exchanging the energy with two P10 spins at a rate of Γcr\Gamma_{\text{cr}} via four-spin cross-relaxation. 3. Most P1- spins quickly deplete to the lower energy state due to the “fast-relaxing spins” (detailed further in C and Supplementary Text). Consequently, only P1+ spins progressively accumulate in the higher energy state, culminating in population inversion in the P1+ transition. 4. The inverted P1+ state amplifies the input microwave probe tone through stimulated emission. (C) Asymmetry in energy relaxation times between P1+ and P1-, indicating faster decay in P1-. The inset shows the energy relaxation of the “fast-relaxing spins” on a timescale of approximately ten milliseconds. While the data in the main panel were extracted from reflection spectroscopy measurements, those in the inset were obtained using a saturation recovery pulse sequence (Supplementary Text).

The electron spins of the P1 centers are the active medium in our maser amplifier. These centers possess an electron spin S=1/2S=1/2 and a nuclear spin I=1I=1, with a hyperfine interaction tensor characterized by 𝒜/2π=81.33MHz\mathcal{A}_{\perp}/2\pi=81.33\,\text{MHz} and 𝒜||/2π=114.03MHz\mathcal{A}_{||}/2\pi=114.03\,\text{MHz}. Under a static magnetic field aligned parallel to the cubic crystallographic axis [001][001] of the diamond, the electron spin states (S=±12S=\pm\frac{1}{2}) split into three levels depending on the nuclear spin states, P1+, P10, and P1-, as shown in Fig. 2A. The corresponding transition frequencies, denoted as ω+\omega_{+}, ω0\omega_{0}, and ω\omega_{-}, satisfy an energy conserving relationship 2ω0=ω+ω+2\omega_{0}=\omega_{-}+\omega_{+}, which facilitates a four-spin cross-relaxation process between the central transition (mI=0m_{I}=0) and the two satellite transitions (mI=±1m_{I}=\pm 1[19, 20, 21].

The operating principle of our maser amplifier is described in Figs. 2B. First, we apply a static magnetic field BDCB_{\text{DC}} to align the frequency ω+\omega_{+} with the resonator frequency ωr\omega_{r}, and then pump the P10 transition at ω0\omega_{0} with a rate Γp\Gamma_{\text{p}}, as illustrated in Fig. 2B (1). This pumping triggers the excitation of one P1+ and one P1- spin by exchanging the energy with two P10 spins through the four-spin cross-relaxation process [19, 20] at a rate of Γcr\Gamma_{\text{cr}}, as depicted in Fig. 2B (2). Ordinarily, such pumping would lead to saturation across all transitions. However, the dynamics are modified by the difference between the relaxation times of the spins in P1- and P1+, as demonstrated in Fig. 2C, with respective values of T1-1030sT_{\text{1-}}\approx 1030\,\text{s} and T1+1650sT_{\text{1+}}\approx 1650\,\text{s}. The faster relaxation of P1- is induced by another “fast-relaxing” spin system, which coexists within the diamond crystal and exhibits relaxation times on the order of ten milliseconds (see inset of Fig. 2C and [22]). The transition frequencies of these “fast-relaxing” spins largely overlap with the P1- transition around ω\omega_{-}, thereby accelerating the relaxation of the P1- spins through magnetic dipole-dipole interactions, as indicated by ΓP1eff\Gamma_{\text{P1}_{-}}^{\text{eff}} in Fig. 2B (3). This imbalance in relaxation times prevents saturation and maintains continuous excitation of the P1+ state, resulting in population inversion. Although the exact nature of these fast-relaxing spins remains unidentified, they might be either neutrally-charged nitrogen-vacancy (NV0) centers [23, 24, 22] or nickel centers [25]. We refer to them as “fast-relaxing spins” throughout this paper.

Gain and bandwidth

Refer to caption
Figure 3: Maser amplifier’s gain and bandwidth. (A) Inverted spin population ΔNP1+\Delta N_{\mathrm{P1_{+}}} between the lower and upper P1+ states over time under various pump powers. Open symbols represent the experimental data, while solid curves represent simulations. The pump powers used here are consistent with those in (B). The inset shows the identical measurement with a higher external quality factor of Qext730Q_{\text{ext}}\approx 730, where the horizontal red dot-dashed line indicates the threshold of self-oscillation. The pump powers from bottom to top correspond to 54dBm-54\leavevmode\nobreak\ \mathrm{dBm}, 52dBm-52\leavevmode\nobreak\ \mathrm{dBm}, and 50dBm-50\leavevmode\nobreak\ \mathrm{dBm}, respectively. (B) Maser gain spectra under various pump powers. The horizontal axis represents the detuning from the central resonator frequency ωr/2π=6.595GHz\omega_{r}/2\pi=6.595\,\text{GHz}. The black dashed line represents the spectrum measured without the pump. (C) Gain-bandwidth product versus inverted spin population. The inset shows the gain, where the open symbols and dashed curves are the experimental data and theoretical curves, respectively (see [22]). (D) Maser gain spectrum versus magnetic field at a pump power of Ppump=44dBmP_{\text{pump}}=-44\,\text{dBm}, yielding a gain of 23dB23\,\text{dB}.

Fig. 3A shows the time evolution of the inverted spin population ΔNP1+=NP1+uNP1+l\Delta N_{\mathrm{P1_{+}}}=N_{\mathrm{P1_{+}}}^{u}-N_{\mathrm{P1_{+}}}^{l} between the upper (uu) and lower (ll) levels of P1+ transition, extracted from the measured reflection coefficients [22]. Above a pump power of 58dBm-58\,\text{dBm} at the device input, ΔNP1+\Delta N_{\mathrm{P1_{+}}} becomes positive, indicating population inversion. We also simulate the dynamics using rate equations that include all P1 states and their interactions [22], which are shown by the solid lines in Fig. 3A.

Fig. 3B displays the maser’s gain spectra under several different pump powers measured by a vector network analyzer (VNA) under a static magnetic field of BDC=232.15mTB_{\text{DC}}=232.15\,\text{mT}. Although the spectra show negative gains at weaker pump powers until 59dBm-59\,\text{dBm}, the gain becomes positive and increases above 58dBm-58\,\text{dBm}. With a pump of 39dBm-39\,\text{dBm}, the gain exceeds 30dB30\,\text{dB} with a bandwidth of approximately 70kHz70\,\text{kHz}, which results in the gain-bandwidth product of 2.8MHz2.8\,\text{MHz}. Fig. 3C plots the gain-bandwidth products under various pump powers for the two different external quality factors, Qext250Q_{\text{ext}}\approx 250 and 730730, of the microwave resonator, as a function of the inverted spin population ΔNP1+\Delta N_{\mathrm{P1_{+}}}. The inset presents the gain. For the maser amplifier device with Qext730Q_{\text{ext}}\approx 730, the system enters the self-oscillation regime at ΔNP1+2.4×1014\Delta N_{\mathrm{P1_{+}}}\approx 2.4\times 10^{14}, where coherent microwave radiation is spontaneously emitted. This is visible as the saturation of ΔNP1+\Delta N_{\mathrm{P1_{+}}} above the gain-bandwidth product of approximately 2.5MHz2.5\,\text{MHz}, as also indicated by the red dot-dashed lines in the insets of Fig. 3C and A.

Next, we demonstrate the frequency tunability of the diamond maser’s gain profile by varying the external static magnetic field BDCB_{\text{DC}}. As the transition frequencies of the P1 centers shift with BDCB_{\text{DC}}, the maser’s gain profile can also be tuned accordingly. Fig. 3D shows the measured maser gain versus BDCB_{\text{DC}}, demonstrating tunability over 10MHz10\,\text{MHz} in the central gain and about 6MHz6\,\text{MHz} in the full-width at half-maximum.

Noise temperature

Refer to caption
Figure 4: Maser amplifier’s noise temperature and saturation power. (A) Signal-to-noise ratio (SNR) in the presence of a weak coherent probe tone with the maser amplifier turned on and off. The ‘pump on’ data, shown in red, is vertically shifted downwards to align the peak values for a clearer comparison of the SNRs. (B) Noise power PoutP_{\text{out}}, detected by a spectrum analyzer, as a function of the input noise temperature TinT_{\text{in}} to the maser amplifier across various gains (2020, 17.517.5, 1515, 12.512.5, and 10dB10\,\text{dB}). Each TinT_{\text{in}} is corrected by taking the line attenuation and insertion losses into account [22]. The dashed lines are the linear fits with their intersection at Pout=0P_{\text{out}}=0 determining the system noise temperatures [18]. The inset shows the system’s background noise temperature, dominated by the following HEMT amplifier, when the maser amplifier is not in operation. (C) Gain saturation measurement showing the maser amplifier gain as a function of input signal power. The dashed curves are the fitting results, and the colored filled symbols denote the input 1dB1\,\text{dB} compression points. Inset: Schematic showing a switch to bypass the HEMT in order to prevent its saturation during these measurements.

To assess the noise performance of our maser amplifier, we measured the noise power spectrum in the presence of a weak microwave probe tone as a reference, with the maser turned on and off. The results displayed in Fig. 4A show a signal-to-noise ratio improvement of approximately 6dB6\,\text{dB}. This indicates that our maser amplifier achieves a noise level approximately four times lower than that of the following measurement line, which is dominated by the noise level of the HEMT amplifier.

To quantitatively evaluate our maser amplifier’s noise temperature, we use the impedance-matched temperature-variable noise source [18]. This device enables independent control of the input noise temperature TinT_{\text{in}} to the maser amplifier, resulting in accurate determination of noise temperature. It is important to note that we also account for the attenuation of the lines between the noise source and the maser amplifier [22], as well as the insertion losses at the directional coupler and the circulator. We have independently measured these attenuations to obtain the accurate value of TinT_{\text{in}} (Supplementary Text).

First, we characterized the base noise temperature of our measurement line with the maser amplifier turned off. In this configuration, the noise power after the HEMT amplifier as a function of TinT_{\text{in}}, detected within a bandwidth of δfBW\delta f_{\text{BW}}, is given by [18]

Pout(Tin)GHEMTkBδfBWTin+THEMT,\frac{P_{\text{out}}(T_{\text{in}})}{G_{\text{HEMT}}k_{\text{B}}\delta f_{\text{BW}}}\approx T_{\text{in}}+T_{\text{HEMT}}, (1)

where GHEMTG_{\text{HEMT}} and THEMTT_{\text{HEMT}} are the gain and noise temperature of the HEMT amplifier, respectively, and kBk_{\text{B}} is the Boltzmann constant. The results, shown in the inset of Fig. 4B, indicate that THEMT4.19(28)KT_{\text{HEMT}}\approx 4.19(28)\,\text{K}.

Subsequently, we turned on the diamond maser amplifier and repeated the noise temperature measurements at various maser amplifier gains. In this configuration, the output noise power after the HEMT amplifier is given by

Pout(Tin)GmaserGHEMTkBδfBWTin+Tmaser+THEMTGmaser.\frac{P_{\text{out}}(T_{\text{in}})}{G_{\text{maser}}G_{\text{HEMT}}k_{\text{B}}\delta f_{\text{BW}}}\approx T_{\text{in}}+T_{\text{maser}}+\frac{T_{\text{HEMT}}}{G_{\text{maser}}}. (2)

Here, GmaserG_{\text{maser}} and TmaserT_{\text{maser}} represent the gain and noise temperature of the maser amplifier, respectively. As presented in Fig. 4B, the minimum noise temperature of the entire system is approximately Tsys=Tmaser+(THEMT/Gmaser)0.902(3)KT_{\text{sys}}=T_{\text{maser}}+(T_{\text{HEMT}}/G_{\text{maser}})\approx 0.902(3)\,\text{K}. After subtracting the contribution from THEMTT_{\text{HEMT}}, the minimum noise temperature of the diamond maser amplifier is obtained to be Tmaser0.860(6)KT_{\text{maser}}\approx 0.860(6)\,\text{K}, corresponding to approximate noise photons of n¯maser=(eω/kBTmaser1)12.2\bar{n}_{\text{maser}}=\left(e^{\hbar\omega/k_{B}T_{\text{maser}}}-1\right)^{-1}\approx 2.2.

The measured noise temperature TmaserT_{\text{maser}}, which is nearly three times higher than the quantum limit of n¯QL=0.5\bar{n}_{\text{QL}}=0.5, corresponding to approximately 0.3K0.3\,\text{K} in our setup, primarily results from imperfect population inversion[8, 10]. While the spins in the higher energy level amplify the incoming signal through stimulated emission, those remaining in the lower energy level absorb photons. This counteraction effectively introduces additional noise photons, quantified by ns=(eω/kB|Ts|1)1n_{s}=\left(e^{\hbar\omega/k_{B}|T_{s}|}-1\right)^{-1}, where negative spin temperature TsT_{\text{s}} and the inversion ratio ρ=ΔNP1+/(NP1+u+NP1+l)\rho=\Delta N_{P1_{+}}/(N_{\mathrm{P1_{+}}}^{u}+N_{\mathrm{P1_{+}}}^{l}) are linked through

ρ=tanh(ω2kB|Ts|),\rho=\tanh\left(\frac{\hbar\omega}{2k_{\text{B}}|T_{\text{s}}|}\right), (3)

with \hbar denoting the reduced Planck constant. As shown in Fig. 3A, the measured ρ0.28\rho\approx 0.28 at 20dB20\,\text{dB} gain corresponds to a noise temperature of Tmaser0.79KT_{\text{maser}}\approx 0.79\,\text{K} [10], in good agreement with the measured TmaserT_{\text{maser}}. It is important to note that the measured TmaserT_{\text{maser}} values are by no means the fundamental limits in our system; they are constrained by the slow relaxation rate of the P1\text{P1}_{-} spins, ΓP1eff0.95mHz\Gamma_{\text{P1}_{-}}^{\text{eff}}\approx 0.95\,\text{mHz}, which in turn limits the effective pump rate on the P1+ spins. With a slightly faster ΓP1eff\Gamma_{\text{P1}_{-}}^{\text{eff}} of 1.2mHz1.2\,\text{mHz}, which could be realized by engineering a higher density of the fast-relaxing spins, the noise temperature could be reduced to Tmaser0.59KT_{\text{maser}}\approx 0.59\,\text{K}. Additionally, improving the internal quality factor QintQ_{\text{int}} would further lower the maser amplifier’s noise temperature. For instance, an order-of-magnitude improvement in QintQ_{\text{int}} to 104\sim 10^{4}, achievable without reducing the coupling constant g0g_{0} by employing a dielectric resonator [11, 26, 27], could lower the noise temperature by approximately 0.1K0.1\,\text{K}.

Compression point

To assess the dynamic range of our maser amplifier, we measured the 1dB1\,\text{dB} compression points at various maser amplifier gains. Fig. 4C shows the amplifier gains as a function of input probe power. The output (input) 1dB1\,\text{dB} compression points (P1dBP_{\text{1dB}}) are measured to be approximately 65dBm-65\,\text{dBm} (85dBm-85\,\text{dBm}) at a gain of 20dB20\,\text{dB}, and approximately 63dBm-63\,\text{dBm} (73dBm-73\,\text{dBm}) at a gain of 10dB10\,\text{dB}. These measured compression points are 2020 to 30dB30\,\text{dB} higher than those typically reported in Josephson traveling-wave parametric amplifiers [4, 2], but are two or three orders of magnitude lower than those achieved by state-of-the-art kinetic inductance traveling-wave parametric amplifiers [28, 29].

Conclusion

We demonstrated a diamond maser amplifier operating at millikelvin temperatures and achieved low noise photon levels of n¯maser2.2\bar{n}_{\text{maser}}\approx 2.2. The noise level could be further reduced by increasing the effective pump rate through a higher density of fast-relaxing spins and by employing a resonator with a higher internal quality factor. By revisiting masers at millikelvin temperatures, our work reestablishes maser technology as a robust and low-noise alternative for microwave amplification in quantum technologies, such as quantum computing, magnetic resonance spectroscopy, and particle physics.

Acknowledgements.
We thank J. Pla, Ç.O. Girit, P. Bertet, A. Bienfait, D. Xiao, R.-B. Liu, J.J.L. Morton, K. Mølmer, E. Flurin, M. Hatifi, F. Quijandría, J. Teufel, Y. Ota, S. Iwamoto, and S. Kono for their discussions, and M. Ishida for assistance with device illustration. We are also grateful to the members of the Hybrid Quantum Device Team within the Science and Technology Group and the Experimental Quantum Information Physics Unit at Okinawa Institute of Science and Technology (OIST) Graduate University for their insights. Additionally, we acknowledge the support from S. Ikemiyagi, M. Kuroda, and P. Kennedy at the Engineering Section of OIST.
Funding:

This work was supported by the Proof-of-Concept (PoC) program by OIST Innovation, the JST Moonshot R&D Program (Grant No. JPMJMS2066), JST-PRESTO (Grant No. JPMJPR15P7), JSPS KAKENHI Grant-in-Aid for Scientific Research (B) (Grant No. 18H01817), Grant-in-Aid for Scientific Research on Innovative Areas (Grant No. 18H04295), the Science Research Promotion Fund from the Promotion and Mutual Aid Corporation for Private Schools of Japan (PMAC), and the Research Encouragement Grant from the AGC Inc. Research Foundation. M.O. and T.H. acknowledge financial support from the JSPS Grant-in-Aid for Fellows (Grant Nos. 23KJ2135 and 24KJ2178), and J.R.B. acknowledges support from OIST Graduate University.

Author contributions:

M.O. and C.-P.L. contributed equally to this work and were primarily responsible for data collection with partial assistance from V.P.M.S., T.H., I.K., and Y.Ku. M.O. and C.-P.L. analyzed the data with assistance from T.H., I.K., V.P.M.S., and Y.Ku. Y.Ku. conceived the project and initiated the preliminary experiments with J.R.B. and P.M. S.O., T.O., Y.Ko., and J.I. were responsible for the preparation of the diamond sample. Y.Ku. and H.T. jointly supervised the project. All authors contributed to the manuscript.

Competing interests:

There are no competing interests to declare.

Data and materials availability:

The data that support the findings of this study are available from the corresponding author upon reasonable request.

References

  • Clerk et al. [2010] A. A. Clerk, M. H. Devoret, S. M. Girvin, F. Marquardt, and R. J. Schoelkopf, Introduction to quantum noise, measurement, and amplification, Reviews of Modern Physics 82, 1155 (2010).
  • Esposito et al. [2021] M. Esposito, A. Ranadive, L. Planat, and N. Roch, Perspective on traveling wave microwave parametric amplifiers, Applied Physics Letters 119, 120501 (2021).
  • Yamamoto et al. [2008] T. Yamamoto, K. Inomata, M. Watanabe, K. Matsuba, T. Miyazaki, W. D. Oliver, Y. Nakamura, and J. S. Tsai, Flux-driven Josephson parametric amplifier, Applied Physics Letters 93, 042510 (2008).
  • Macklin et al. [2015] C. Macklin, K. O’Brien, D. Hover, M. E. Schwartz, V. Bolkhovsky, X. Zhang, W. D. Oliver, and I. Siddiqi, A near–quantum-limited Josephson traveling-wave parametric amplifier, Science 350, 307 (2015).
  • Eom et al. [2012] B. H. Eom, P. K. Day, H. G. LeDuc, and J. Zmuidzinas, A wideband, low-noise superconducting amplifier with high dynamic range, Nature Physics 8, 623 (2012).
  • Malnou et al. [2021] M. Malnou, M. Vissers, J. Wheeler, J. Aumentado, J. Hubmayr, J. Ullom, and J. Gao, Three-Wave Mixing Kinetic Inductance Traveling-Wave Amplifier with Near-Quantum-Limited Noise Performance, PRX Quantum 2, 010302 (2021).
  • Weber [1957] J. Weber, Maser Noise Considerations, Physical Review 108, 537 (1957).
  • Siegman [1964] A. E. Siegman, Microwave solid-state masers (McGraw-Hill, 1964).
  • Jin et al. [2015] L. Jin, M. Pfender, N. Aslam, P. Neumann, S. Yang, J. Wrachtrup, and R.-B. Liu, Proposal for a room-temperature diamond maser, Nature Communications 6, 1 (2015).
  • Day et al. [2024] T. Day, M. Isarov, W. J. Pappas, B. C. Johnson, H. Abe, T. Ohshima, D. R. McCamey, A. Laucht, and J. J. Pla, Room-Temperature Solid-State Maser Amplifier, Physical Review X 14, 041066 (2024), publisher: American Physical Society.
  • Breeze et al. [2018] J. D. Breeze, E. Salvadori, J. Sathian, N. M. Alford, and C. W. M. Kay, Continuous-wave room-temperature diamond maser, Nature 555, 493 (2018).
  • Sherman et al. [2022] A. Sherman, O. Zgadzai, B. Koren, I. Peretz, E. Laster, and A. Blank, Diamond-based microwave quantum amplifier, Science Advances 8, eade6527 (2022), publisher: American Association for the Advancement of Science.
  • Ng et al. [2021] W. Ng, H. Wu, and M. Oxborrow, Quasi-continuous cooling of a microwave mode on a benchtop using hyperpolarized NV- diamond, Applied Physics Letters 119, 234001 (2021), publisher: American Institute of Physics.
  • Fahey et al. [2023] D. P. Fahey, K. Jacobs, M. J. Turner, H. Choi, J. E. Hoffman, D. Englund, and M. E. Trusheim, Steady-State Microwave Mode Cooling with a Diamond N-$V$ Ensemble, Physical Review Applied 20, 014033 (2023), publisher: American Physical Society.
  • Wu et al. [2022] H. Wu, S. Yang, M. Oxborrow, M. Jiang, Q. Zhao, D. Budker, B. Zhang, and J. Du, Enhanced quantum sensing with room-temperature solid-state masers, Science Advances 8, eade1613 (2022), publisher: American Association for the Advancement of Science.
  • Ball et al. [2018] J. R. Ball, Y. Yamashiro, H. Sumiya, S. Onoda, T. Ohshima, J. Isoya, D. Konstantinov, and Y. Kubo, Loop-gap microwave resonator for hybrid quantum systems, Applied Physics Letters 112, 204102 (2018).
  • Grezes et al. [2014] C. Grezes, B. Julsgaard, Y. Kubo, M. Stern, T. Umeda, J. Isoya, H. Sumiya, H. Abe, S. Onoda, T. Ohshima, V. Jacques, J. Esteve, D. Vion, D. Esteve, K. Mølmer, and P. Bertet, Multimode Storage and Retrieval of Microwave Fields in a Spin Ensemble, Physical Review X 4, 021049 (2014).
  • Simbierowicz et al. [2021] S. Simbierowicz, V. Vesterinen, J. Milem, A. Lintunen, M. Oksanen, L. Roschier, L. Grönberg, J. Hassel, D. Gunnarsson, and R. E. Lake, Characterizing cryogenic amplifiers with a matched temperature-variable noise source, Review of Scientific Instruments 92, 034708 (2021), tex.ids= simbierowiczCharacterizingCryogenicAmplifiers2021a publisher: American Institute of Physics.
  • Sorokin et al. [1960] P. P. Sorokin, G. J. Lasher, and I. L. Gelles, Cross Relaxation Studies in Diamond, Physical Review 118, 939 (1960).
  • Ma et al. [2019] C. Ma, Z. Wang, Y. Wu, Z. Bao, Y. Song, H. Zhang, and L. Duan, Four-spin cross relaxation in a hybrid quantum device, Physical Review A 100, 012322 (2019).
  • Zhang et al. [2021] G.-Q. Zhang, Z. Chen, D. Xu, N. Shammah, M. Liao, T.-F. Li, L. Tong, S.-Y. Zhu, F. Nori, and J. Q. You, Exceptional Point and Cross-Relaxation Effect in a Hybrid Quantum System, PRX Quantum 2, 020307 (2021), publisher: American Physical Society.
  • [22] Materials and methods are available as supplementary materials.
  • Baier et al. [2020] S. Baier, C. Bradley, T. Middelburg, V. Dobrovitski, T. Taminiau, and R. Hanson, Orbital and Spin Dynamics of Single Neutrally-Charged Nitrogen-Vacancy Centers in Diamond, Physical Review Letters 125, 193601 (2020), publisher: American Physical Society.
  • Barson et al. [2019] M. S. J. Barson, E. Krausz, N. B. Manson, and M. W. Doherty, The fine structure of the neutral nitrogen-vacancy center in diamond, Nanophotonics 8, 1985 (2019), publisher: De Gruyter.
  • Isoya et al. [1990] J. Isoya, H. Kanda, and Y. Uchida, EPR studies of interstitial Ni centers in synthetic diamond crystals, Physical Review B 42, 9843 (1990).
  • Kato et al. [2023] K. Kato, R. Sasaki, K. Matsuura, K. Usami, and Y. Nakamura, High-cooperativity cavity magnon-polariton using a high-Q dielectric resonator, Journal of Applied Physics 134, 083901 (2023).
  • Hamamoto et al. [2024] T. Hamamoto, A. Bhunia, R. K. Bhattacharya, H. Takahashi, and Y. Kubo, Dielectric microwave resonator with large optical apertures for spin-based quantum devices, Applied Physics Letters 124, 234001 (2024).
  • Faramarzi et al. [2024] F. Faramarzi, R. Stephenson, S. Sypkens, B. H. Eom, H. LeDuc, and P. Day, A 4–8 GHz kinetic inductance traveling-wave parametric amplifier using four-wave mixing with near quantum-limited noise performance, APL Quantum 1, 036107 (2024).
  • Parker et al. [2022] D. J. Parker, M. Savytskyi, W. Vine, A. Laucht, T. Duty, A. Morello, A. L. Grimsmo, and J. J. Pla, Degenerate Parametric Amplification via Three-Wave Mixing Using Kinetic Inductance, Physical Review Applied 17, 034064 (2022), publisher: American Physical Society.
  • Probst et al. [2015] S. Probst, F. B. Song, P. A. Bushev, A. V. Ustinov, and M. Weides, Efficient and robust analysis of complex scattering data under noise in microwave resonators, Review of Scientific Instruments 86, 024706 (2015).
  • Walls and Milburn [2008] D. Walls and G. J. Milburn, eds., Quantum Optics (Springer, Berlin, Heidelberg, 2008).
  • Kubo et al. [2010] Y. Kubo, F. R. Ong, P. Bertet, D. Vion, V. Jacques, D. Zheng, A. Dréau, J.-F. Roch, A. Auffeves, F. Jelezko, J. Wrachtrup, M. F. Barthe, P. Bergonzo, and D. Esteve, Strong Coupling of a Spin Ensemble to a Superconducting Resonator, Physical Review Letters 105, 140502 (2010).
  • Choi and Englund [2023] H. Choi and D. Englund, Ultrastrong magnetic light-matter interaction with cavity mode engineering, Communications Physics 6, 1 (2023), number: 1 Publisher: Nature Publishing Group.
  • Collett and Gardiner [1984] M. J. Collett and C. W. Gardiner, Squeezing of intracavity and traveling-wave light fields produced in parametric amplification, Physical Review A 30, 1386 (1984), publisher: American Physical Society.
  • Diniz et al. [2011] I. Diniz, S. Portolan, R. Ferreira, J. M. Gérard, P. Bertet, and A. Auffèves, Strongly coupling a cavity to inhomogeneous ensembles of emitters: Potential for long-lived solid-state quantum memories, Physical Review A 8410.1103/PhysRevA.84.063810 (2011).
  • Kurucz et al. [2011] Z. Kurucz, J. H. Wesenberg, and K. Mølmer, Spectroscopic properties of inhomogeneously broadened spin ensembles in a cavity, Physical Review A 8310.1103/PhysRevA.83.053852 (2011).
  • Julsgaard and Mølmer [2012] B. Julsgaard and K. Mølmer, Dynamical evolution of an inverted spin ensemble in a cavity: Inhomogeneous broadening as a stabilizing mechanism, Physical Review A 86, 063810 (2012).
  • Carmichael [2013] H. J. Carmichael, Statistical Methods in Quantum Optics 1: Master Equations and Fokker-Planck Equations (Springer Science & Business Media, 2013) google-Books-ID: 5hv2CAAAQBAJ.
  • Schweiger and Jeschke [2001] A. Schweiger and G. Jeschke, Principles of Pulse Electron Paramagnetic Resonance (Oxford University Press, Oxford, New York, 2001).
  • Iakoubovskii and Stesmans [2002] K. Iakoubovskii and A. Stesmans, Dominant paramagnetic centers in 17 O -implanted diamond, Physical Review B 66, 045406 (2002).
  • Meesala et al. [2018] S. Meesala, Y.-I. Sohn, B. Pingault, L. Shao, H. A. Atikian, J. Holzgrafe, M. Gündoğan, C. Stavrakas, A. Sipahigil, C. Chia, R. Evans, M. J. Burek, M. Zhang, L. Wu, J. L. Pacheco, J. Abraham, E. Bielejec, M. D. Lukin, M. Atatüre, and M. Lončar, Strain engineering of the silicon-vacancy center in diamond, Physical Review B 97, 205444 (2018), publisher: American Physical Society.
  • Kurokawa et al. [2024] H. Kurokawa, K. Wakamatsu, S. Nakazato, T. Makino, H. Kato, Y. Sekiguchi, and H. Kosaka, Coherent electric field control of orbital state of a neutral nitrogen-vacancy center, Nature Communications 15, 4039 (2024), publisher: Nature Publishing Group.
  • Liu et al. [2021] A. Liu, S. T. Cundiff, D. B. Almeida, and R. Ulbricht, Spectral broadening and ultrafast dynamics of a nitrogen-vacancy center ensemble in diamond, Materials for Quantum Technology 1, 025002 (2021), publisher: IOP Publishing.
  • Acosta et al. [2013] V. M. Acosta, K. Jensen, C. Santori, D. Budker, and R. G. Beausoleil, Electromagnetically Induced Transparency in a Diamond Spin Ensemble Enables All-Optical Electromagnetic Field Sensing, Physical Review Letters 110, 213605 (2013).

Supplementary Materials for
A near-quantum-limited diamond maser amplifier operating at millikelvin temperatures

I Materials and Methods

I.1 Sample and device

In this section, we describe the diamond sample and the microwave resonator device used in this work.

Refer to caption
Figure S1: Quality factors of the bare resonator. Magnitude (upper panels) and phase (lower panels) of the resonator reflection coefficient r(ω)r(\omega) for (A) Qext250Q_{\mathrm{ext}}\approx 250 and (B) Qext730Q_{\mathrm{ext}}\approx 730. The black dashed lines represent the fits.

I.1.1 Diamond sample

The diamond sample used in this work is identical to those described in previous works (Refs. [17, 16]). It was synthesized by the temperature gradient method under high-pressure and high-temperature (HPHT) conditions of 5.5GPa5.5\,\text{GPa} and 1350C1350\,^{\circ}\text{C} using an Fe-Ni-Co alloy as the metal solvent and natural diamond grains as the carbon source. A (100)(100) plate obtained by laser-cutting and polishing was irradiated with 3MeV3\leavevmode\nobreak\ \text{MeV} electrons at a fluence of 4×1017e/cm24\times 10^{17}\,\mathrm{e/cm^{2}} at room temperature, followed by annealing at 850C850\leavevmode\nobreak\ ^{\circ}\text{C} for 2 hours in vacuum. Subsequently, it was further irradiated with 2MeV2\,\text{MeV} electrons at 700C700\,^{\circ}\text{C} to a fluence of 5×1017e/cm25\times 10^{17}\,\mathrm{e/cm^{2}} and annealed at 1000C1000\,^{\circ}\text{C} for 2 hours in vacuum. The final dimensions of the diamond (3×1.5×0.5mm33\times 1.5\times 0.5\,\mathrm{mm}^{3}) used in this study were obtained from the central part of the plate via laser-cutting and polishing. Following electron irradiation and annealing, the concentrations of P1 centers and negatively-charged nitrogen-vacancy (NV-) centers were measured to be approximately 16ppm16\,\mathrm{ppm} and 2ppm2\,\mathrm{ppm}, respectively.

I.1.2 Loop-gap resonator

The microwave resonator used in this work is a loop-gap type [16]. We determined its internal and external quality factors using the circle-fit method [30] after subtracting background amplitude oscillations to achieve a flat baseline. The results are shown in Fig. S1 for both Qext250Q_{\text{ext}}\approx 250 and Qext730Q_{\text{ext}}\approx 730. The internal quality factor QintQ_{\text{int}} consistently stays around 30003000. The resonator frequency, ωr\omega_{r}, is very sensitive to the distances of the two gaps. Consequently, after each disassembly and reassembly of the resonator and its enclosure to adjust QextQ_{\text{ext}}, ωr\omega_{r} varied over 6.59GHz6.59\,\text{GHz} to 6.86GHz6.86\,\text{GHz} upon each cooldown.

I.2 Cryogenic and room temperature setups

In this section, we describe the wiring and setup both inside the dilution refrigerator and at room temperature.

Refer to caption
Figure S2: Cryogenic wiring and configurations. “LPF” and “IRF” denote a 7.5GHz7.5\,\mathrm{GHz} low pass filter and an infrarad filter, respectively. The microwave switch located below the high-mobility transistor amplifier (HEMT) at 4K4\,\mathrm{K} is used only for pulse-ESR (electron spin resonance) measurements. The decibel numbers indicate the values of attenuators (gray round rectangular) and directional couplers. (A) Setup at 13mK13\,\mathrm{mK} for maser gain and noise temperature measurements. (B) Setup at 13mK13\,\mathrm{mK} for the saturation measurements. The right-side switch at 13mK13\,\mathrm{mK} is used to bypass the HEMT during these measurements.
Refer to caption
Figure S3: Room temperature measurement setup. “BPF” and “DCB” denote a 94108MHz94-108\,\mathrm{MHz} band pass filter and a DC block, respectively. (A) Setup for masing measurement. (B) Setup for pulse ESR measurement.

I.2.1 Wiring configuration in the dilution refrigerator

The detailed wiring and configuration inside the dilution refrigerator (Bluefors, LD-400) are illustrated in Fig. S2. Cryogenic attenuators (XMA, 2082-624X-CRYO series), represented as gray round rectangles, are inserted in each input line at every temperature stage to attenuate thermal noise from the microwave tone injected from room temperature. The heavily attenuated “Probe” line and moderately attenuated “Pump” line are combined at a directional coupler (Marki Microwave, C20-0226) on the mixing chamber stage. The labels “LPF” and “IRF” represent low-pass filters (Marki Microwave, FLP-0750) and homemade infrared filters using iron-loaded epoxy Eccosorb CR110, respectively. A HEMT (High Electron Mobility Transistor) amplifier (Low Noise Factory, LNA-LNC4_8C) was installed on the 4K4\,\text{K} plate.

The calibrated noise source [18] is used for the noise temperature measurements. While the noise temperature is heated maximally up to 4.5K\approx 4.5\,\text{K}, the base temperature of the mixing chamber plate slightly increases up to 19mK\approx 19\,\text{mK}.

Two microwave switches (Radiall, R573423600) were used to check the total attenuation of the coaxial cables to and from the noise source, which turned out to be negligible. The microwave switch (Analog Devices Inc., HMC547ALP3E), depicted inside the dashed box in Fig. S2, was installed and used only for pulse electron spin resonance (pulse-ESR) measurements to protect the HEMT. Further details are described in the Supplementary Text.

I.2.2 Pump cancellation

The probe tone was fed into the fridge through the line labeled “Probe”, while the pump tone was sent through the line “Pump”. To minimize the impact of the pump tone on the HEMT and the following receiver circuits, a cancellation microwave tone was introduced through a dedicated microwave input line labeled “Cancellation” (see also Fig. S3). When operating the maser amplifier, a microwave tone was generated by a signal generator (PSG; Keysight Technologies, E8267D) for both pump and cancellation. This tone was then split by a power splitter (Mini-Circuits, ZC2PD-01263-S+), as shown in Fig. S3A. On one branch, the cancellation tone was adjusted using a voltage-controlled phase shifter (Qorvo, CMD297P34) and a voltage-controlled variable attenuator (Pulsar Microwave, AAT-22-479A/7S). The amplitude and phase of this cancellation tone were adjusted to destructively interfere with the reflected pump tone through a directional coupler placed after the two circulators (Quinstar, QCY-G040082AS), resulting in suppression of the pump power by at least 50dB50\,\text{dB}.

I.2.3 Setup for maser gain and noise temperature measurement

The reflection coefficient r(ω)r(\omega) of the resonator was measured using a vector network analyzer (VNA; Keysight, N5230C). The data presented in Figs. 2C and 3A, B, and D in the main text were measured at a power of 110dBm-110\,\mathrm{dBm}, corresponding to a mean intra-resonator photon number of approximately 10610^{6}, which is sufficiently lower than the number of inverted spins, ΔNP1+4×1014\Delta N_{P1+}\approx 4\times 10^{14}, ensuring minimal impact on the spin saturation. Although this power level almost reaches the typical saturation power of traveling-wave parametric amplifiers [2], it remains well below the saturation power of our maser amplifier as discussed in Fig. 4C in the main text.

Noise power spectra were measured by a spectrum analyzer (SPA; Keysight Technologies; N9010A). The output and input ports of the VNA are connected to the Probe and Output ports of the fridge via directional couplers (input: Pulsar Microwave, CS20-09-436/9; output: Marki, C10-0226). A signal generator (SG, Keysight Technologies, E8247C) generates a microwave probe tone during the signal-to-noise ratio (SNR) measurements, as presented in Fig. 4A.

I.2.4 Setup for compression measurement

Figure S2B displays the wiring configuration on the mixing chamber plate used to measure the 1dB1\,\text{dB} compression points, as presented in Fig. 4C in the main text. To ensure the HEMT and subsequent components were not saturated, we bypassed the HEMT using a microwave switch.

I.2.5 Setup for pulse electron spin resonance measurements

In the pulse ESR measurement detailed in the Supplementary Text, we use an FPGA-based integrated measurement system (Quantum Machines, OPX+) to generate pulse envelopes, control microwave switches, and detect spin echo signals. Gaussian-shaped pulse envelope signals, modulated at 100MHz100\,\mathrm{MHz}, are generated by the OPX arbitrary waveform generator (AWG). We use the single sideband modulation technique, where a baseband microwave tone (“LO” in Fig. S3B) is mixed with the pulse envelopes with an internal IQ mixer of the PSG. These pulses are subsequently amplified by a high-power amplifier (Cree, CMPA601C025F) and fed to the resonator through the “Pump” line. To minimize microwave leakage noise from the amplifier, a semiconductor-based microwave switch (Analog Devices Inc., ADRF5019-EVALZ) is installed immediately after the amplifier’s output. Similarly, another semiconductor-based microwave switch (Analog Devices Inc., HMC547ALP3E) is installed to protect the HEMT amplifier from the pulses.

For echo signal detection, the signals from the output port of the refrigerator are amplified by a series of room-temperature amplifiers (B&Z Technologies, BZ-02000800-090826-182020, and Mini-Circuits, ZVA-183+) and then down-converted using an IQ mixer (Marki, IQ-0307L). The local oscillator (LO) tones are extracted from a portion of the PSG’s locking signal (option HCC). After demodulation, the signals are passed through band-pass filters (Mini-Circuits, SBP-101+), and then amplified by a preamplifier (SRS, SR445A). Finally, they are detected by the OPX’s digitizer, which processes the signals digitally for heterodyne detection.

I.3 Noise temperature measurement

In this section, we detail the methodology used to measure the noise temperature of our maser amplifier using an impedance-matched temperature-variable noise source.

I.3.1 Y-factor method using a noise source

We characterized the noise temperature of the maser amplifier by using a procedure based on the Y-factor method with a calibrated noise source [18]. The noise source is weakly thermalized to the cold plate at about 100mK100\,\mathrm{mK}, which allows for quasi-independent temperature control using an attached heater and thermometer. This setup enables precise control over the input noise temperature TinT_{\text{in}} into the maser amplifier device.

Thermal noise power is given by Pnoise=kBTnoiseδfBWP_{\text{noise}}=k_{\text{B}}T_{\text{noise}}\delta f_{\mathrm{BW}}, where kBk_{\text{B}}, TnoiseT_{\text{noise}}, and δfBW\delta f_{\mathrm{BW}} are the Boltzmann constant, the noise temperature, and the measurement bandwidth, respectively. When a signal with a noise temperature of TinT_{\text{in}} is fed into an amplifier, the output power PoutP_{\text{out}} is given by

Pout(Tin)GkBδfBW=Tin+Tamp,\displaystyle\frac{P_{\text{out}}(T_{\text{in}})}{Gk_{\text{B}}\delta f_{\mathrm{BW}}}=T_{\text{in}}+T_{\text{amp}}, (S1)

where GG and TampT_{\mathrm{amp}} are the power gain and noise temperature of the amplifier, respectively. In our experiment, without the maser amplifier, PoutP_{\text{out}} can thus be expressed as:

Pout(Tin)GHEMTkBδfBW=Tin+THEMT+TbkgGHEMT,\displaystyle\frac{P_{\text{out}}(T_{\mathrm{in}})}{G_{\text{HEMT}}k_{\text{B}}\delta f_{\mathrm{BW}}}=T_{\text{in}}+T_{\text{HEMT}}+\frac{T_{\text{bkg}}}{G_{\text{HEMT}}}, (S2)

where GHEMTG_{\text{HEMT}} and THEMTT_{\text{HEMT}} represent the gain and noise temperature of the HEMT amplifier, respectively. The additional noise contribution from the final room temperature amplifier, including cable attenuation from the HEMT to the room temperature setup, is denoted as TbkgT_{\text{bkg}}. Equation (S2) can be approximated to Eq. (1) in the main text, considering that GHEMT>104G_{\text{HEMT}}>10^{4} and Tbkg300KT_{\text{bkg}}\approx 300\,\text{K}.

Similarly, the noise power with the maser amplifier on can be expressed as:

Pout(Tin)GmaserGHEMTkBδfBW=Tin+Tmaser+THEMTGmaser+TbkgGmaserGHEMT,\displaystyle\frac{P_{\text{out}}(T_{\text{in}})}{G_{\text{maser}}G_{\text{HEMT}}k_{\text{B}}\delta f_{\mathrm{BW}}}=T_{\text{in}}+T_{\text{maser}}+\frac{T_{\text{HEMT}}}{G_{\text{maser}}}+\frac{T_{\text{bkg}}}{G_{\text{maser}}G_{\text{HEMT}}}, (S3)

where GmaserG_{\text{maser}} and TmaserT_{\text{maser}} represent the gain and noise temperature of the maser amplifier, respectively. Therefore, Eq. (S3) can be approximated to Eq. (2) in the main text under the same considerations.

I.3.2 Impact of insertion losses and attenuation on noise temperature

In our setup, a directional coupler and a circulator, as well as lossy non-superconducting cables, are placed between the noise source and the maser amplifier. Their insertion losses and attenuations must, therefore, be taken into account in the evaluation of the input noise temperature TinT_{\text{in}}, as they attenuate the noise photons from the noise source. To model this effect, we employ the beam-splitter model [31]. If the jj-th component in the transmission line has an insertion loss βj\beta_{j}, the number of noise photons nj+1n_{j+1} can be expressed as

nj+1=βj(nj+12)+(1βj)(nth+12),\displaystyle n_{j+1}=\beta_{j}\left(n_{j}+\frac{1}{2}\right)+\left(1-\beta_{j}\right)\left(n_{\mathrm{th}}+\frac{1}{2}\right), (S4)

where njn_{j} is the number of input noise photons of the jj-th component, and nthn_{\mathrm{th}} is the thermal noise photons corresponding to the physical temperature of the component. All the presented values of TinT_{\text{in}} are corrected using Eq. (S4) based on the measurements detailed in the Supplementary Texts.

I.4 Extracting spin information through resonator

In this section, we describe how we extracted the spectral information of the spin ensemble through the resonator’s reflection coefficient r(ω)r(\omega) using the input-output theory. This enables us to determine the corresponding number of spins, as shown in Fig. 3A in the main text.

I.4.1 Single Spin Coupling Constant

The single-spin coupling constant g0g_{0}, which represents the interaction strength between a single P1 center and a single microwave photon in a resonator, is expressed as [32]

g0=γeδB01/2|𝐒|+1/2,\displaystyle g_{0}=-\gamma_{e}\delta B_{0}\langle{-1/2}|\mathbf{S}|{+1/2}\rangle, (S5)

where γe\gamma_{e} and δB0\delta B_{0} denotes the gyromagnetic ratio of the electron spin and the magnetic vacuum fluctuation of the resonator, respectively, and 𝐒\mathbf{S} is the spin vector operator with |1/2|{-1/2}\rangle and |+1/2|{+1/2}\rangle are the P1 center’s electron spin eigenstates. The magnetic vacuum fluctuation is given by

δB0=μ0ωr2𝒱eff,\displaystyle\delta B_{0}=\sqrt{\frac{\mu_{0}\hbar\omega_{r}}{2\mathcal{V}_{\mathrm{eff}}}}, (S6)

where μ0\mu_{0}, ωr\omega_{r}, and 𝒱eff\mathcal{V}_{\mathrm{eff}} are the vacuum permeability, the resonator frequency, and the effective mode volume of the resonator, respectively. 𝒱eff\mathcal{V}_{\mathrm{eff}} is defined as the ratio of the integral of the magnetic field energy density over the whole resonator mode volume VV to the maximum magnetic energy density [33]

𝒱eff=V|𝐁(𝐫)|2/μ(𝐫)dV|𝐁(𝐫0)|2/μ(𝐫0),\displaystyle\mathcal{V}_{\mathrm{eff}}=\frac{\int_{V}|\mathbf{B}(\mathbf{r})|^{2}/\mu(\mathbf{r})\mathrm{d}V}{|\mathbf{B}(\mathbf{r}_{0})|^{2}/\mu(\mathbf{r}_{0})}, (S7)

where 𝐁(𝐫)\mathbf{B}(\mathbf{r}) is the magnetic field in the resonator, maximized at the position 𝐫0\mathbf{r}_{0}. Using a finite element simulation software, COMSOL Multiphysics, the effective resonator mode volume 𝒱eff\mathcal{V}_{\mathrm{eff}} of our loop-gap resonator is estimated to be approximately 𝒱eff2.5×108m3\mathcal{V}_{\mathrm{eff}}\approx 2.5\times 10^{-8}\,\mathrm{m^{3}}. Therefore, the single-spin coupling constant of the P1 center with S=1/2S=1/2 is obtained to be g02π×0.15Hzg_{0}\approx 2\pi\times 0.15\,\mathrm{Hz}.

I.4.2 Input-output theory

To determine the number of inverted or non-inverted spins through the resonator reflection r(ω)r(\omega), we employed the input-output theory [34] for a coupled resonator-spin ensemble system with a coherent probe tone αin=α0eiωt\alpha_{\text{in}}=\alpha_{0}e^{-i\omega t} [17, 10]. The Hamiltonian can be expressed using the Tavis-Cummings model with Holstein-Primakoff approximation [35, 36, 37] as follows:

H/=\displaystyle H/\hbar= ωraa+12j=1Nωjsjsj\displaystyle\omega_{r}a^{\dagger}a+\frac{1}{2}\sum_{j=1}^{N}\omega_{j}s^{\dagger}_{j}s_{j}
+j=1Ngj(sja+sja)iκext(αinaαina),\displaystyle+\sum_{j=1}^{N}g_{j}\left(s^{\dagger}_{j}a+s_{j}a^{\dagger}\right)-i\sqrt{\kappa_{\mathrm{ext}}}(\alpha_{\text{in}}a^{\dagger}-\alpha_{\text{in}}^{\ast}a), (S8)

where aa (sjs_{j}) and aa^{\dagger} (sjs_{j}^{\dagger}) denote the annihilation and creation operators for the resonator mode (jj-th spin), respectively. In addition, NN represents the number of spins, and gjg_{j} is the single spin coupling constant between the resonator and the jj-th spin. κext=ωr/Qext\kappa_{\mathrm{ext}}=\omega_{r}/Q_{\mathrm{ext}} and κint=ωr/Qint\kappa_{\mathrm{int}}={\omega_{r}}/{Q_{\mathrm{int}}} are the external and internal loss rates of the resonator, with QextQ_{\mathrm{ext}} (QintQ_{\mathrm{int}}) representing the external (internal) quality factor.

I.4.3 The spin function K(ω)K(\omega)

Following the standard method for deriving the reflection coefficient [34], we obtain

r(ω)=iκext(ωωr)+iκext+κint2±K(ω)1,\displaystyle r(\omega)=\frac{i\kappa_{\mathrm{ext}}}{(\omega-\omega_{r})+i\frac{\kappa_{\mathrm{ext}}+\kappa_{\mathrm{int}}}{2}\pm K(\omega)}-1, (S9)

where K(ω)K(\omega), called the “spin function”, includes all information about the spin ensemble. As will be discussed in the following section, K(ω)K(\omega) is linked to the spin susceptibility [17] and is defined as shown in Refs. [35, 36, 37],

K(ω)=j=1Ngj2(ωωj)+iγj/2=gens2(ωωs)+iΓ/2,\displaystyle K(\omega)=\sum_{j=1}^{N}\frac{g_{j}^{2}}{(\omega-\omega_{j})+i\gamma_{j}/2}=\frac{g_{\text{ens}}^{2}}{(\omega-\omega_{s})+i\Gamma/2}, (S10)

with gens=jgj=g0Ng_{\text{ens}}=\sum_{j}g_{j}=g_{0}\sqrt{N} representing the ensemble coupling constant between the resonator and the spin ensemble. The decay rate of individual spins is denoted by γj\gamma_{j}, while ωs\omega_{s} and Γ/2\Gamma/2 represent the center frequency and the half-width at half-maximum (HWHM) of the spin ensemble, respectively, assumed to follow a Lorentzian distribution here. The sign in front of K(ω)K(\omega) in Eq. (S9) represents the state of the spin ensemble, i.e., a positive sign (++) corresponds to an inverted spin ensemble, whereas a negative sign (-) represents a non-inverted spin ensemble.

We modify K(ω)K(\omega) to incorporate the spin polarization ρ¯=ΔN/Ntot\bar{\rho}=\Delta N/N_{\mathrm{tot}}, where ΔN=NlNu\Delta N=N_{l}-N_{u} represents the population difference between the lower and upper levels of the P1+\text{P1}_{+} state, and Ntot=Nl+NuN_{\mathrm{tot}}=N_{l}+N_{u} is the total number of spins in the P1+\text{P1}_{+} transition. In a system with a population difference ΔN\Delta N, the ensemble coupling constant gensg_{\text{ens}} is replaced by the effective coupling constant gens,effg_{\text{ens,eff}}, defined as gens,eff2=g02ΔNg_{\text{ens,eff}}^{2}=g_{0}^{2}\Delta N. Consequently, K(ω)K(\omega) is reformulated in terms of ρ¯\bar{\rho} as

K(ω)\displaystyle K(\omega) =gens,eff2(ωωs)+iΓ/2\displaystyle=\frac{g_{\text{ens,eff}}^{2}}{(\omega-\omega_{s})+i\Gamma/2}
=g02ΔN(ωωs)+iΓ/2=ρ¯gens,total2(ωωs)+iΓ/2,\displaystyle=\frac{g_{0}^{2}\Delta N}{(\omega-\omega_{s})+i\Gamma/2}=\frac{\bar{\rho}g_{\text{ens,total}}^{2}}{(\omega-\omega_{s})+i\Gamma/2}, (S11)

where gens,total2=g02Ntotalg_{\mathrm{ens,total}}^{2}=g_{0}^{2}N_{\mathrm{total}} represents the “conventional” collective coupling constant, assuming all spins are polarized into the lower energy state. The sign of K(ω)K(\omega) reflects the sign of ΔN\Delta N; it becomes negative in the case of population inversion and positive in a non-inverted state. To ensure consistency with Eq. (S9), we rewrite the reflection coefficient r(ω)r(\omega) as follows

r(ω)=iκext(ωωr)+iκext+κint2K(ω)1.\displaystyle r(\omega)=\frac{i\kappa_{\mathrm{ext}}}{(\omega-\omega_{r})+i\frac{\kappa_{\mathrm{ext}}+\kappa_{\mathrm{int}}}{2}-K(\omega)}-1. (S12)

I.4.4 Measurement of population differences ΔN\Delta N

The function K(ω)K(\omega) can be expressed in terms of the reflection coefficient r(ω)r(\omega) by rearranging Eq. (S12):

K(ω)=(ωωr)+i2(κint+κextr(ω)1r(ω)+1),\displaystyle K(\omega)=(\omega-\omega_{r})+\frac{i}{2}\left(\kappa_{\mathrm{int}}+\kappa_{\mathrm{ext}}\frac{r(\omega)-1}{r(\omega)+1}\right), (S13)

which allows K(ω)K(\omega) to be directly determined from the experimentally obtained r(ω)r(\omega) data. Fig. S4 shows such examples converted from experimentally obtained data and fits using the imaginary parts of Eq. (S11).

In an ideal spin ensemble, the imaginary part of K(ω)K(\omega) exhibits a Lorentzian profile, enabling straightforward determination of the population difference ΔN\Delta N by fitting the measured spectrum using the function:

y=y0+A(xx0)2+B2.\displaystyle y=y_{0}+\frac{A}{(x-x_{0})^{2}+B^{2}}. (S14)

From this function and the imaginary part of Eq. (S11), the population difference ΔN\Delta N can be extracted as ΔN=A/g02B\Delta N=A/g_{0}^{2}B.

Refer to caption
Figure S4: Measured spin function K(ω)K(\omega). Imaginary parts of K(ω)K(\omega) obtained from the measured maser gain spectra, with fits (black dashed lines) with Eq. (S15). The data shown in (A) correspond to selected time points from the time dynamics presented in Fig. 3A in the main text at a pump power of 54dBm-54\,\mathrm{dBm}, while those in (B) represent gain spectra under the three pump powers after the gain was stabilized.

However, significant deviations from the ideal Lorentzian profile are observed under certain experimental conditions. Immediately after initiating the pump, and before reaching the steady state, the spectrum becomes distorted due to the rapid inversion of the spins near the center of the distribution, while those farther from the center remain non-inverted. Furthermore, even in the steady state, intermediate pump powers lead to partial inversion of the spin ensemble, resulting in non-uniform inversion and persistent spectral distortion. This behavior is illustrated in Fig. S4. Panel A shows spectra at selected time points extracted from the time evolution in Fig. 3A of the main text. Among these, the spectra at 146.6 and 326.7 seconds clearly exhibit transient distortions following pump initiation. Panel B shows spectra at various pump powers extracted from the gain profiles in Fig. 3B. Notably, the spectrum at a pump power of 59dBm-59\,\mathrm{dBm} exhibits pronounced steady-state distortion due to partial inversion.

To accurately account for these spectral conditions, we employ a two-Lorentzian fitting model in which the spectrum is represented as a superposition of two separate Lorentzian contributions; one representing the inverted upper-level spins and the other the lower-level spins:

y=y0+A1(xx1)2+B12A2(xx2)2+B22.\displaystyle y=y_{0}+\frac{A_{1}}{(x-x_{1})^{2}+B_{1}^{2}}-\frac{A_{2}}{(x-x_{2})^{2}+B_{2}^{2}}. (S15)

The results of fits using this model are illustrated in Fig. S4. From this fitting model, the population difference ΔN\Delta N can be described as

ΔN=1g02(A2B2A1B1).\displaystyle\Delta N=\frac{1}{g_{0}^{2}}\left(\frac{A_{2}}{B_{2}}-\frac{A_{1}}{B_{1}}\right). (S16)

I.4.5 Maser gain

The peak maser gain is defined by the absolute square of the reflection coefficient r(ω)r(\omega) at the resonance, where ω=ωr=ωs\omega=\omega_{r}=\omega_{s}. According to Eq. (S12), this relationship can be expressed as

G=|r(ω)|2=(κextκint+|κs|)2(κext+κint|κs|)2,\displaystyle G=|r(\omega)|^{2}=\frac{(\kappa_{\mathrm{ext}}-\kappa_{\mathrm{int}}+|\kappa_{\mathrm{s}}|)^{2}}{(\kappa_{\mathrm{ext}}+\kappa_{\mathrm{int}}-|\kappa_{\mathrm{s}}|)^{2}}, (S17)

where we defined a new parameter, κs=4gens2/Γ=Im[2K(ωr=ωs)]\kappa_{s}=4g_{\mathrm{ens}}^{2}/\Gamma=\mathrm{Im}[2K(\omega_{r}=\omega_{s})] [10, 37], as the “spin transition rate”. This parameter quantifies the effective transition rate of the spin ensemble to the resonator mode, i.e., the rate at which spins undergo upwards (downwards) transitions per unit time through the stimulated emission (absorption) with a rate Γs=4g02/Γ\Gamma_{\text{s}}=4g_{0}^{2}/\Gamma. Moreover, κs\kappa_{\text{s}} is directly correlated with the “magnetic quality factor” QmagQ_{\text{mag}} [8, 12], which characterizes a spin ensemble system as a maser gain medium:

Qmag=1ηχ′′=1ηIm[2K(ωr=ωs)/(ωrη)]=ωrΓ4gens2=ωrκs,\displaystyle Q_{\text{mag}}=\frac{1}{\eta\chi^{\prime\prime}}=\frac{1}{\eta\cdot\textrm{Im}[2K(\omega_{r}=\omega_{s})/(\omega_{r}\eta)]}=\frac{\omega_{r}\Gamma}{4g_{\mathrm{ens}}^{2}}=\frac{\omega_{r}}{\kappa_{s}}, (S18)

where η\eta is the magnetic field filling factor and χ′′\chi^{\prime\prime} is the imaginary part of the spin susceptibility at resonance, defined as χ(ω)=2K(ω)/(ωrη)\chi(\omega)=-2K(\omega)/(\omega_{r}\eta) [17].

Using Eq. (S17), we calculated the gains G(ωr=ωs)G(\omega_{r}=\omega_{s}) for the two external quality factors, Qext250Q_{\text{ext}}\approx 250 and 730730, used in the experiments. The results are shown as dashed curves in the inset of Fig. 3C.

I.5 Modeling spin dynamics

In this section, we describe the semi-classical rate equations we used to analyze the spin system’s dynamics.

I.5.1 Semi-classical Rate Equations

We modeled the dynamics of our diamond maser amplifier system as detailed below and solved them using semi-classical rate equations [19]. Although quantum Langevin equations would yield equivalent results [9, 10], our lack of expertise with that numerical treatment led us to opt for a semi-classical approach. The system under consideration, as depicted in Fig. S5A, consists of eight energy levels, which include two levels of the “fast-relaxing spins” and the three two-level systems associated with the P1 centers nuclear spin states, P1+, P10, and P1-. This configuration leads to eight rate equations, in addition to a photon number rate equation, to describe the population dynamics and photon numbers within the system. However, due to the computational complexity of solving nine equations, we simplified the model to seven equations by combining the two levels of P1\text{P}1_{-} with the fast-relaxing spins. The resulting set of seven rate equations, including the photon number rate equation, are represented as follows [19, 38]:

N1˙\displaystyle\dot{N_{1}} =Γ41N4+ΓcrNtot3(N22N4N6N52N1N3)Γsn(N1N4),\displaystyle=\Gamma_{41}N_{4}+\frac{\Gamma_{\mathrm{cr}}}{N_{\text{tot}}^{3}}(N_{2}^{2}N_{4}N_{6}-N_{5}^{2}N_{1}N_{3})-\Gamma_{s}n(N_{1}-N_{4}), (S19)
N2˙\displaystyle\dot{N_{2}} =Γ52N52ΓcrNtot3(N22N4N6N52N1N3)Γp(N2N5),\displaystyle=\Gamma_{52}N_{5}-2\frac{\Gamma_{\mathrm{cr}}}{N_{\text{tot}}^{3}}(N_{2}^{2}N_{4}N_{6}-N_{5}^{2}N_{1}N_{3})-\Gamma_{\text{p}}(N_{2}-N_{5}),
N3˙\displaystyle\dot{N_{3}} =Γ63effN6+ΓcrNtot3(N22N4N6N52N1N3),\displaystyle=\Gamma_{63}^{\mathrm{eff}}N_{6}+\frac{\Gamma_{\mathrm{cr}}}{N_{\text{tot}}^{3}}(N_{2}^{2}N_{4}N_{6}-N_{5}^{2}N_{1}N_{3}),
N4˙\displaystyle\dot{N_{4}} =Γ41N4ΓcrNtot3(N22N4N6N52N1N3)+Γsn(N1N4),\displaystyle=-\Gamma_{41}N_{4}-\frac{\Gamma_{\mathrm{cr}}}{N_{\text{tot}}^{3}}(N_{2}^{2}N_{4}N_{6}-N_{5}^{2}N_{1}N_{3})+\Gamma_{s}n(N_{1}-N_{4}),
N5˙\displaystyle\dot{N_{5}} =Γ52N5+2ΓcrNtot3(N22N4N6N52N1N3)+Γp(N2N5),\displaystyle=-\Gamma_{52}N_{5}+2\frac{\Gamma_{\mathrm{cr}}}{N_{\text{tot}}^{3}}(N_{2}^{2}N_{4}N_{6}-N_{5}^{2}N_{1}N_{3})+\Gamma_{\text{p}}(N_{2}-N_{5}),
N6˙\displaystyle\dot{N_{6}} =Γ63effN6ΓcrNtot3(N22N4N6N52N1N3),\displaystyle=-\Gamma_{63}^{\mathrm{eff}}N_{6}-\frac{\Gamma_{\mathrm{cr}}}{N_{\text{tot}}^{3}}(N_{2}^{2}N_{4}N_{6}-N_{5}^{2}N_{1}N_{3}),

where NiN_{i} represents the number of spins of each level, Γ41\Gamma_{41}, Γ52\Gamma_{52}, and Γ63\Gamma_{63} are the relaxation rates of each transition, Γcr\Gamma_{\mathrm{cr}} is the four-spin cross-relaxation rate, and Γs=4g02/Γ\Gamma_{s}=4g_{0}^{2}/\Gamma is the stimulated emission rate. The rate equation for the intra-resonator photon number nn is:

n˙\displaystyle\dot{n} =κtotnΓsn(N1N4)+ΓsN4.\displaystyle=-\kappa_{\mathrm{tot}}n-\Gamma_{s}n(N_{1}-N_{4})+\Gamma_{s}N_{4}. (S20)
Refer to caption
Figure S5: Rate equation model and comparison to measurements. (A) Energy level diagram of P1 center and the relaxation rates used in the rate-equation model. (B) Steady-state population difference ΔNP1+\Delta N_{\mathrm{P1_{+}}} between the lower and upper P1+ states, obtained from measured r(ω)r(\omega) using Eq. (S13) in the Materials and Methods, plotted versus the pump rate Γp\Gamma_{\mathrm{p}} calculated from Eq. (S23) based on the applied pump power PinP_{\mathrm{in}}. The dashed lines indicate the results of rate-equation simulations, and the red dot-dashed lines represent the threshold of both QextQ_{\mathrm{ext}} values.

I.5.2 Cross-relaxation rate

Table S1: Relaxation rates used in the rate-equation simulation. The values of Γ41\Gamma_{41} and Γ63\Gamma_{63} are determined from T1T_{1} measurements of the P1+ and P1- transitions shown in Fig. S7.
Parameter Γ41/2π\Gamma_{41}/2\pi (Hz) Γ52/2π\Gamma_{52}/2\pi (Hz) Γ63/2π\Gamma_{63}/2\pi (Hz) Γcr/2π\Gamma_{\mathrm{cr}}/2\pi (Hz)
Value 9.6×1059.6\times 10^{-5} 9.6×1059.6\times 10^{-5} 1.5×1041.5\times 10^{-4} 6.0×1026.0\times 10^{-2}
Table S2: Resonator and spin parameters used in the rate-equation simulations.
QextQ_{\mathrm{ext}} ωr/2π\omega_{r}/2\pi (Hz) κext/2π\kappa_{\mathrm{ext}}/2\pi (Hz) κint/2π\kappa_{\mathrm{int}}/2\pi (Hz) Γ/2/2π\Gamma/2/2\pi (Hz)
250250 6.59×1096.59\times 10^{9} 25×10625\times 10^{6} 2.6×1062.6\times 10^{6} 0.75×1060.75\times 10^{6}
730730 6.86×1096.86\times 10^{9} 9.4×1069.4\times 10^{6} 2.2×1062.2\times 10^{6} 0.90×1060.90\times 10^{6}

The cross-relaxation rate Γcr\Gamma_{\mathrm{cr}} in our system cannot be directly measured due to the presence of the “fast-relaxing spins”, contrasting with the previous reports [19, 20, 21]. To estimate this rate indirectly, we compared computational results from the rate equations with experimental results obtained under various pump powers, as depicted in Fig. S5B and explained in the next subsection. We determined Γcr\Gamma_{\mathrm{cr}} to be within the range of 1\sim 1 to 10mHz10\,\text{mHz}, with the parameters listed in Tables S1 and S2. While previously reported cross-relaxation rates for highly nitrogen-doped diamonds typically range from 10Hz\sim 10\,\text{Hz} to 100Hz\sim 100\,\text{Hz} [19, 20], theoretical models of the four-spin interactions imply that Γcrd6\Gamma_{\mathrm{cr}}\propto d^{6} [20], where dd denotes the spin density. This supports our findings of ΓcrmHz\Gamma_{\mathrm{cr}}\sim\text{mHz}, as the P1 center density in our diamond sample of 16ppm\approx 16\,\text{ppm} is considerably lower than the >100ppm>100\,\text{ppm} in those previous studies.

I.5.3 Calculation of pump rate Γp\Gamma_{\text{p}}

Spins absorb incoming microwave pump tones and are excited to the upper state, a process governed by the same underlying physics as stimulated emission. Therefore, the photon absorption rate, i.e., the pump rate Γp\Gamma_{\text{p}}, is calculated as a product of the stimulated emission rate and the number of incoming photons, as described in Eqs. (7.7) and (7.8) on p.260 in Ref. [38]:

Γp=g02Γ(ωpωs)2+Γ2/4nin,\displaystyle\Gamma_{\text{p}}=\frac{g_{0}^{2}\Gamma}{(\omega_{p}-\omega_{s})^{2}+\Gamma^{2}/4}n_{\mathrm{in}}, (S21)

where ωp\omega_{p} is the pump frequency, ωs\omega_{s} is the average spin resonance frequency of P10, and ninn_{\mathrm{in}} is the number of input photons. The expression g02Γ/[(ωpωs)2+Γ2/4]{g_{0}^{2}\Gamma}/\left[{(\omega_{p}-\omega_{s})^{2}+\Gamma^{2}/4}\right] on the right-hand side is the rigorous stimulated emission rate with detuning between ωp\omega_{p} and ωs\omega_{s}.

From the input-output theory, ninn_{\mathrm{in}} with a coherent pump tone βin=β0eiωt\beta_{\mathrm{in}}=\beta_{0}e^{-i\omega t} applied to a single-port resonator is expressed as

nin\displaystyle n_{\mathrm{in}} =4κextκtot2+4(ωpωr)2|β0|2=4κextκtot2+4(ωpωr)2Pinω,\displaystyle=\frac{4\kappa_{\mathrm{ext}}}{\kappa_{\mathrm{tot}}^{2}+4(\omega_{p}-\omega_{r})^{2}}|\beta_{0}|^{2}=\frac{4\kappa_{\mathrm{ext}}}{\kappa_{\mathrm{tot}}^{2}+4(\omega_{p}-\omega_{r})^{2}}\frac{P_{\mathrm{in}}}{\hbar\omega}, (S22)

where Pin=ω|β0|2P_{\mathrm{in}}=\hbar\omega|\beta_{0}|^{2} denotes the input pump power, and κtot=κext+κint\kappa_{\mathrm{tot}}=\kappa_{\mathrm{ext}}+\kappa_{\mathrm{int}} is the total loss rate of the resonator. Therefore, the pump rate Γp\Gamma_{\text{p}} can be expressed as a function of PinP_{\mathrm{in}} as follows;

Γp(Pin)=g02Γ(ωpωs)2+Γ2/44κextκtot2+4(ωpωr)2Pinωp.\displaystyle\Gamma_{\text{p}}(P_{\mathrm{in}})=\frac{g_{0}^{2}\Gamma}{(\omega_{p}-\omega_{s})^{2}+\Gamma^{2}/4}\cdot\frac{4\kappa_{\mathrm{ext}}}{\kappa_{\mathrm{tot}}^{2}+4(\omega_{p}-\omega_{r})^{2}}\frac{P_{\mathrm{in}}}{\hbar\omega_{p}}. (S23)

During maser operation, the pump frequency ωp\omega_{p} is tuned to match the resonance frequency of the P10 spins, i.e., ωp=ωs=ω0\omega_{p}=\omega_{s}=\omega_{0}, but it is off-resonance with the resonator, thus ωpωr\omega_{p}\neq\omega_{r}. Figure S5B displays the population difference in the P1+ transition, ΔNP1+\Delta N_{\text{P1+}}, plotted against the calculated pump rates Γp\Gamma_{\text{p}}. This population difference is extracted using the spin function K(ω)K(\omega), as discussed in the above section. The plot also includes overlayed simulation results from the rate equations Eqs. (S19) and (S20).

I.5.4 Self-oscillation threshold

If the photons produced by stimulated emission exceed the resonator losses, the system enters the self-oscillating or free-running maser regime. In this regime, the maser is capable of emitting microwave photons into the resonator without the need for an externally applied probe tone. The threshold population difference ΔNthr\Delta N_{\mathrm{thr}} can be calculated from the photon number rate equation given in Eq. (S20[38]. At the threshold, the two terms in the equation are balanced, hence,

0=κtotnssΓsnssΔNthr+ΓsN4thr,\displaystyle 0=-\kappa_{\mathrm{tot}}n_{\mathrm{ss}}-\Gamma_{s}n_{\mathrm{ss}}\Delta N_{\mathrm{thr}}+\Gamma_{s}N_{4}^{\mathrm{thr}}, (S24)

where N4thrN_{4}^{\mathrm{thr}} is the number of spins in the state 4 at the threshold and nssn_{\mathrm{ss}} is the photon number in the steady state in the free-running maser regime. Rearranging this equation gives

ΔNthr=κtotΓs+N4thrnss.\displaystyle\Delta N_{\mathrm{thr}}=-\frac{\kappa_{\mathrm{tot}}}{\Gamma_{s}}+\frac{N_{4}^{\mathrm{thr}}}{n_{\mathrm{ss}}}. (S25)

The second term can often be neglected since typically nss1n_{\mathrm{ss}}\gg 1. Therefore,

|ΔNthr|κtotΓs=κtotΓ4g02.\displaystyle|\Delta N_{\mathrm{thr}}|\approx\frac{\kappa_{\mathrm{tot}}}{\Gamma_{s}}=\frac{\kappa_{\mathrm{tot}}\Gamma}{4g_{0}^{2}}. (S26)

In the case of Qext250Q_{\mathrm{ext}}\approx 250, |ΔNthr||\Delta N_{\mathrm{thr}}| is approximately 4.81×10144.81\times 10^{14}, while for Qext730Q_{\mathrm{ext}}\approx 730, it is approximately 2.42×10142.42\times 10^{14}. These thresholds are shown as the red dot-dashed lines in the insets of Fig. 3A and C in the main text.

II Supplementary Text

Refer to caption
Figure S6: Electron spin resonance spectroscopy and relaxation time measurements. (A) Pulse sequence used for the echo-detected field-sweep (EDFS) spectroscopy and saturation recovery measurements. The wait time TWT_{\text{W}} was fixed at 100 ms for EDFS spectroscopy and was varied for T1T_{1} measurements. (B) EDFS signals as a function of the magnetic field. Black points depict experimental data, while the blue solid curve indicates the fit using nine Lorentzian functions. Orange dashed curves correspond to fitting results for known defects in diamond, i.e., P1 and NV- centers, and their hyperfine transitions with 13C centers at the nearest neighbors. The red solid curve represents fast-relaxing spins. Calculated resonant magnetic field values for known defects are indicated by short arrows at the top. (C) Results of saturation recovery measurements. Echo area as a function of wait time TWT_{W} at a magnetic field of B0=250mTB_{0}=250\,\mathrm{mT} (upper panel) and B0=255mTB_{0}=255\,\mathrm{mT} (lower panel) are plotted. Orange dashed lines show the result of bi-exponential fitting in the upper panel and single-exponential fitting in the lower panel. (D) T1T_{1} as a function of the magnetic field, with the EDFS spectroscopy data overlaid. Circular and square markers represent the shorter and longer T1T_{1} components obtained from bi-exponential fits. Data obtained above 253mT253\,\text{mT} can be fitted with a single exponential.

II.1 Echo-detected field-sweep spectroscopy

To characterize the impurity spins in our diamond, specifically the “fast-relaxing spins,” we conducted an echo-detected field-sweep (EDFS) measurement using the setup depicted in Fig. S3B, with the pulse sequence illustrated in Fig. S6A. A practical challenge arises from the “long relaxing spins”, namely the P1 and NV- centers, which exhibit very long relaxation times ranging from approximately 103\sim 10^{3} to 104\sim 10^{4} seconds. Additionally, the substantial quantity of these impurity spins may hinder the signals from the “fast-relaxing spins”, which turned out to be about two orders of magnitude less abundant. To mitigate this, we first saturated the P1 and NV- centers using a 200-ms-long saturation pulse with frequency modulation of ±10MHz\pm 10\,\mathrm{MHz} centered at the resonator frequency of 6.788GHz\mathrm{6.788\,GHz} in this cooldown. This allowed us to primarily detect the spin echo signals from the “fast-relaxing spins”. This approach also enabled optimal adjustment of the room temperature amplifier chain to utilize the full dynamic range of the digitizer. After a waiting time of TW=100msT_{W}=100\,\mathrm{ms}, the echo signal was detected using a Hahn-echo sequence consisting of π/2\pi/2 and π\pi pulses.

The results of the EDFS measurement are shown in Fig. S6B. Despite the saturation pulse, several peaks associated with the P1 and NV- centers remain visible. This is attributed to imperfections in the saturation pulse and partial population recovery during the 100ms100\,\mathrm{ms} wait time. Nevertheless, a distinct “broad shoulder” feature is observed, extending from around the P1- peak at approximately 245mT245\,\text{mT} to around 260mT260\,\text{mT}. We fitted the data with a sum of nine Lorentzian functions. These components are also indicated by short arrows at the top of Fig. S6B. The broad signal stemming from the “fast-relaxing spins” is represented by a solid red curve.

II.2 Characterization of “fast-relaxing spins”

Refer to caption
Figure S7: Spin relaxation measurements of two different diamond samples. (A) Results of the sample used in this work. The data were fitted using a bi-exponential function, and the two relaxation times, i.e., the shorter T1,ST_{\text{1,S}} and longer T1,LT_{\text{1,L}}, are displayed. (B) Results of a yellow diamond sample.

We measured the longitudinal relaxation times T1T_{1} of the “fast-relaxing spins” under constant magnetic fields ranging from 247mT247\,\mathrm{mT} to 260mT260\,\mathrm{mT} using the saturation recovery pulse sequence [39]. The results are shown in Fig. S6C and D. Near the transition of P1-, the relaxation curves exhibit bi-exponential decays, as presented in the upper panel of Fig. S6C. In contrast, in the fields higher than 253mT253\,\mathrm{mT}, the relaxation curves can be fitted by a single exponential decay (bottom panel of Fig. S6C). This is attributed to the cross-relaxation effects with the P1- transition and their 13C hyperfine-coupled transition.

As noted in the main text, we have not been able to rigorously identify the specific spin species constituting the “fast-relaxing spins.” Nevertheless, the evidence presented below suggests that these may be neutrally-charged nitrogen-vacancy (NV0) centers. First, we measured T1T_{1} of P1 centers in a type Ib yellow diamond, an as-grown HPHT sample purchased from Sumitomo Electric Industries, Ltd., containing approximately 5050 to 100ppm100\,\mathrm{ppm} of P1 centers, at 10mK10\,\mathrm{mK}. The results, displayed in Fig. S7B, show spin relaxation times exceeding 4×1034\times 10^{3} seconds in the short T1T_{1}, about four or five times longer than those measured in the sample used for the maser experiments as presented in Fig. S7A. Similarly, the longer T1T_{1} also show 5.2×1045.2\times 10^{4} and 4×1044\times 10^{4} seconds. These imply that the electron irradiation, i.e., the introduction of vacancies into our diamond crystal, likely created the “fast-relaxing spins” that exhibit the broad spin distribution (Fig. S6B), subsequently impacting the transitions of P1 centers.

Vacancy clusters [40] initially appeared to be possible candidates. However, these defects typically exhibit spin transitions with much narrower linewidths (significantly less than 1mT1\,\text{mT}) at temperatures below 100K100\,\text{K} [40], which stands in stark contrast to the broad transition observed for the fast-relaxing spins in our diamond crystal. On the other hand, similar to negatively-charged silicon-vacancy (SiV-) centers [41], the spins of NV0 centers are sensitive to crystal strain [23, 42]. Such strain becomes significant in diamond crystals containing defect concentrations exceeding 1ppm\approx 1\,\mathrm{ppm} [43, 44], which is indeed the case for the diamond crystal used in our experiments.

II.3 Noise temperature of baseline

Refer to caption
Figure S8: Noise temperature measurements over wide frequency range. Noise temperature as a function of frequency detuning, measured with the pump on (red squares), and with the pump off and an off-resonant magnetic field for P1+ (blue circles). (A) Wide range: ±200MHz\pm 200\,\mathrm{MHz}. (B) Zoomed-in view: ±10MHz\pm 10\,\mathrm{MHz}, corresponding to the dashed window in panel A.

We measured the noise temperature at several frequencies around the maser peak gain, both with the pump on and off, as shown in Fig. S8. First, we measured the noise temperature with the pump off, namely, THEMTT_{\text{HEMT}}. To minimize the absorption effect by the spins, which is equivalent to having additional attenuation along the line, the magnetic field was detuned from the resonance of P1+.

Under these conditions, as seen in Fig. S8, the noise temperatures are measured to be between 22 and 4K4\,\mathrm{K}, in reasonable agreement with the values reported in the HEMT’s datasheet. The oscillations are mostly attributed to small impedance mismatches among the components along the measurement chain from 10mK10\,\text{mK} up to the HEMT amplifier. At the resonance frequency ωr6.6GHz\omega_{r}\approx 6.6\,\text{GHz}, it reaches 4.2K\approx 4.2\,\text{K}, which is used as the HEMT’s noise temperature THEMTT_{\text{HEMT}} in the main text. During this measurement with the pump off, we disconnected both the pump and cancellation lines and terminated at room temperature to minimize the noise from the room-temperature setup, especially from the amplifier in the cancellation line (Fig. S3).

When the pump and cancellation tones are turned on and tuned for maser amplifier’s gain of 20dB20\,\mathrm{dB}, the noise baseline increases by approximately 1K1\,\text{K} over the entire frequency range (except near ωr\omega_{r}, i.e., the maser amplifier frequency), most likely due to noise coming from the room-temperature amplifier connected to the “Cancellation” line. The significant increase of the noise around 100MHz\sim-100\,\mathrm{MHz} is due to the pump.

II.4 Detailed line attenuation characterization

Refer to caption
Figure S9: Cryogenic wiring used to characterize the line attenuations. The vertical arrows next to each line denote the materials used for the inner and outer conductors, with CN, B, and SB representing cupronickel, beryllium copper, and silver-plated beryllium copper, respectively.

In order to accurately estimate the power of the pump and probe tones arriving at the input of the maser amplifier device, we independently characterized the attenuation of both the probe and pump lines in a separate cooldown. To this end, we measured the transmission coefficients over the “symmetric lines”, i.e, pairs of lines with identical configurations, including coaxial cable materials, lengths, and attenuators at each temperature stage, as illustrated in Fig. S9. These symmetric lines were interconnected by a short, low-loss coaxial cable at the mixing chamber stage.

We also characterized the attenuation of the cables and components in the Probe line at the mixing chamber stage, specifically from the infrared filter through the circulator, noise source, directional coupler, and the final circulator to the input of the maser device (see also Fig. S2). This was performed using two mechanical microwave switches (Radiall, R573423600), as depicted in Fig. S9. Four measurement paths (Ch. 1–Ch. 4) were established for this purpose. Channels 1 and 2 correspond to the paths from the noise source to the maser amplifier device and from the infrared filter to the noise source, respectively. Channel 3 was used to isolate the attenuation introduced by the components measured in Ch. 1 and Ch. 2 by using a low-loss flexible coaxial cable. Channel 4 was used to estimate the attenuation of the low-loss flexible coaxial cables used to connect the components by comparing with Ch. 3. The results of these measurements are summarized in Table S3. Note that the insertion loss of the mechanical switches must also be taken into account in the calculation of Eq. (S4) for the noise temperature measurement; here we relied on the value on the datasheet of 0.3dB0.3\,\text{dB}.

Table S3: Line attenuations
Line Attenuation value (dB)
Probe 53.3
Pump, Cancellation 27.6
Channel 1 1.26
Channel 2 1.50