1.
Introduction
Quantum sensing is a cutting-edge technology that leverages quantum properties, such as quantum coherence, superposition, and entanglement, to detect minuscule signals that are challenging to measure with classical counterparts. This technology enables the detection of physical parameters, such as magnetic fields, temperature, pressure, and gravity, with exceptional sensitivity and precision, surpassing classical limits.1 ,2
Among the various quantum sensing platforms, solid-state spin defects have emerged as a promising candidate due to their long spin coherence, compatibility with optical and microwave (MW) control, and ability to operate under ambient conditions.3 The nitrogen-vacancy (NV) center in diamond, one of the most extensively studied spin defects, was first observed in 1965 and further investigated in the late 1970s through electron paramagnetic resonance studies.4 ,5 Groundbreaking research in the early 2000s demonstrated optical initialization and readout of the NV’s spin states, highlighting its potential as both a defect-based qubit and a sensitive quantum sensor.6 Since then, a variety of spin defects in diverse host materials, such as silicon carbide (SiC), hexagonal boron nitride (hBN), gallium nitride (GaN), and silicon (Si), have been thoroughly explored,7 ,8 each offering unique advantages for quantum sensing applications.
Unlike cold atom and superconducting qubits, which often require cryogenic temperatures or vacuum environments,9 ,10 spin defects in wide-bandgap hosting materials, such as NV centers in diamond, operate robustly at room temperature and ambient conditions, facilitating practical and compact sensor implementations.11 ,12 The combination of high field sensitivity and atomic-scale sensor size makes them ideal for high-precision magnetometry13 ,14 and nanoscale imaging of temperature, magnetic, electric, and strain fields.15 ,16 Furthermore, the advancements in crystal growth, ion implantation, optical manipulation, and device integration technologies have significantly improved the performance of the spin defects, rapidly expanding their potential for practical applications.
Although previous reviews have introduced defect-based quantum materials, sensing principles, and performance metrics,3 ,17 –19 a comprehensive overview of quantum sensing based on spin defects from fundamental quantum sensing science to technologies for real-world applications based on spin defects remains lacking. In this review, we provide an up-to-date perspective spanning diverse spin defect platforms, recent advances in quantum control and photonic engineering, and application-driven quantum sensors, as depicted in Fig. 1. In Sec. 2, we begin with the basic working principles of solid-state spin defects, with a particular focus on diamond NV centers. We then discuss various quantum sensing protocols, including optically detected magnetic resonance (ODMR) in continuous and pulsed control techniques, as well as advanced protocols leveraging quantum coherence and entanglement. In addition, we present two imaging techniques using spin defects: scanning probe and wide-field optical imaging methods. In Sec. 3, we introduce novel methods that integrate spin defects with various optical platforms, such as nanostructured devices, micro-lenses, and photonic cavities, to enhance optical readout efficiency and light–matter interaction. Finally, in Sec. 4, we highlight recent progress in applying these techniques to practical quantum sensors, bridging the gap between laboratory research and real-world applications across a range of fields, including biomedical diagnostics, semiconductor device inspection, industrial sensing, and aerospace and military applications.
Fig. 1
Overview of this review: the quantum sensing landscape using spin defects, tracing the progression from fundamental research to technological developments aimed at enabling real-world applications. A directional bar is included to illustrate this trajectory—from foundational studies to advanced technological implementations—highlighting the increasing complexity and maturity of each stage.
2.
Quantum Sensing and Imaging Protocols Based on Spin Defects
Solid-state defects offer several advantages for quantum sensing applications. The defect energy levels are located well within the wide bandgap of crystals, effectively decoupling them from the electrical band structures of the host materials, which helps maintain high coherence properties even within solid-state environments. The defects can also operate under practical conditions at room temperature without the need for vacuum systems. When combined with imaging techniques, they enable high-sensitivity and high-resolution mapping of various physical quantities, such as magnetic, electric, and strain fields. In addition, the electronic spin states of these defects can be initialized, manipulated, and detected optically via fluorescence,20 making integration with optical platforms an attractive option for enhanced readout and versatile control of the spin states. The optical transitions manifest in the color of the host crystals, and the defects are often referred to as "color centers," as shown in Fig. 2.
Fig. 2
Example of a fluorescence image of the color centers. When ensemble NV centers in diamond are illuminated by a green laser at 532 nm, they emit red fluorescence around 620 to 800 nm.
In this section, we begin with a brief overview of the fundamental working principles of quantum sensors based on color centers, focusing on how their spin states interact with external optical and MW fields. We then explore various sensing protocols aimed at enhancing sensing performance. In addition, we discuss scanning and wide-field optical techniques that enable spatially resolved imaging implementations of these sensors. Finally, we compare a range of quantum materials that host different color centers, highlighting their distinct optical and spin properties.
2.1.
Basics of Spin Defects
Here, we discuss the basics of spin defects, with a particular focus on diamond NV centers. The NV center has been extensively studied in the field of quantum sensing due to its high sensitivity on a wide range of physical quantities, including temperature,21 –28 stress/strain,29 –31 electric field/charge,32 –34 magnetic field,35 –62 pressure,63 and rotation (using its intrinsic nuclear spin resource),64 –67 across a broad range of temperatures (mK to 1000 K)68 –71 and pressures (up to hundreds of GPa).63 ,72 –76 A detailed summary of the NV center Hamiltonian, coupling coefficients, and typical sensitivities for these physical properties is available in Ref. 77.
2.1.1.
Electronic structures of the diamond NV center
The NV center consists of three carbon atoms, a nitrogen substitutional atom, and a vacancy within a diamond lattice. A total of six electrons (three from the carbon atoms, two from the nitrogen atom, and one from the host materials’ electronic bands) occupy the ground energy levels of the NV center. Two unpaired electrons at the highest energy levels form an spin system, with sub-spin states of , , and . In the presence of an external magnetic field, , the NV Hamiltonian can be written as follows ():
Eq. (1)
In the ground state, the ZFS between and the degenerated states is 2.87 GHz at room temperature.27 The ZFS varies with temperature at a rate of 75.0(6) kHz/K,22 making the NV center suitable for sensitive thermometry. By contrast, the degenerated spin states can be split by the Zeeman effect in the presence of an external magnetic field. Monitoring the Zeeman splitting is the underlying mechanism of the NV magnetometry. Typically, one of the states is chosen to define a two-level system, and these states can act as a qubit. Therefore, developing techniques for properly reading out and manipulating these qubits is important in several quantum sensing protocols based on spin defects.
The energy levels and optical transitions associated with the NV center are illustrated in Fig. 3(a). Upon illumination with a green laser (), an electron in the ground electronic state is excited to the doublet state, followed by a rapid photon decay on the order of tens of nanoseconds.79 ,80 The resulting fluorescence spans ∼620–800 nm, with the zero-phonon line (ZPL) at 637 nm, and its intensity varies between the (bright) and (dark) states due to the spin-selective intersystem crossing (ISC) and transition via metastable singlet states. During the optical transition, mixed ground spin states can also be initialized into the state through the spin-flipping process. This unique property is the key to the spin-dependent optical readout and initialization of the spin qubit. Note that manipulation between the and states is achieved by applying MW fields with a frequency that matches the spin transition energies.
Fig. 3
Basics of NV center. (a) Schematic of NV’s energy levels and optical transitions. The energy levels of the NV center reside within the bandgap () of diamond and are well separated from the diamond’s conduction and valence bands. A green laser at excites the NV center from its electronic ground state () to the excited state (). Subsequent optical decay produces broad fluorescence in the range of to 800 nm (ZPL: 637 nm). During the optical transitions, the NV center experiences a spin-dependent decay process through the ISC and metastable singlet states. As a result, the fluorescence intensity for is up to 30% lower than that of , enabling optical readout of the spin states. Moreover, the spin-flipping process that occurs during the optical transition allows initialization of the spin states into . (b) Schematic of a confocal fluorescence microscope typically used for the NV experiment. In the optical excitation path, a green laser, an acousto-optic modulator (AOM), a scannable mirror (M), a dichroic mirror, and a high numerical aperture (NA) objective lens are used to focus the laser onto a single NV center in a diamond plate. The resulting fluorescence is collected by the objective lens, with eyepiece lenses (L1 and L2) and a pinhole, and detected by an avalanche photodiode (APD). MW circuitry near the diamond plate provides MW fields for spin transitions. A permanent magnet is occasionally used to break the degeneracy of prior to sensing experiments. (c) An example of an NV fluorescence image obtained by the confocal setup in (b). Individual NV centers can be clearly identified in the image. The scale bar is .
2.1.2.
Confocal fluorescence microscopy
As shown in Fig. 3(b), a confocal fluorescence microscope is commonly used to address single NV centers. The microscope consists of "confocal" optical components: an objective lens and an eyepiece lens (detector), with a pinhole acting as a spatial filter. In addition, a scanning system—either a scanning laser beam or a sample stage—is incorporated into the microscope to spatially map the fluorescence signals. The scanning confocal fluorescence microscope serves as a fundamental tool for characterizing the optical and spin properties of spin defects. An example confocal image of NV centers in Fig. 3(c) clearly shows individual NV centers. The identification of single NV centers can be further confirmed through measurements.81
To improve the spatial resolution and photon collection efficiency of individual NV centers, an objective lens with a high numerical aperture (NA) is typically used. However, the high refractive index of diamond (2.42) results in reduced collection efficiency, which decreases the number of readout photons and sensitivity. In general, the shot-noise-limited sensitivity () is related to the average photon number () as follows:82
Eq. (2)
2.1.3.
Optically detected magnetic resonance
Once a single NV center is identified from the confocal image, the next step is to determine the spin energy levels through electron spin resonance (ESR) spectroscopy. This is achieved by continuously applying the green laser and MW fields while collecting fluorescence signals as a function of the MW frequency [Fig. 4(a)]. This technique is referred to as continuous wave ESR (CW-ESR) spectroscopy. Continuous laser excitation and fluorescence detection enable spin initialization to and real-time readout of the spin states. When the MW frequency matches the NV center resonance, the NV center undergoes continuous Rabi oscillations. As the probability of the spin being in the states increases, the fluorescence intensity decreases, resulting in dark dips in the spectrum [Figs. 4(b) and 4(c)]. Without an applied DC magnetic field, the CW-ESR spectrum shows a single dip at the frequency of ZFS, i.e., 2.87 GHz at room temperature. However, when a DC magnetic field is applied, two split resonances appear, corresponding to the two transitions and , respectively. ESR spectroscopy through optical readout is referred to as ODMR.
Fig. 4
ODMR spectroscopy. (a) Continuous laser and MW excitations for ODMR spectroscopy. "Init." and "R.O." refer to the initialization and readout of the spin qubit. (b) ODMR spectra with varying magnitudes of the DC magnetic field. Normalized photoluminescence (PL) is plotted as a function of MW frequency. The plots are offset for clarity. (c) Two-dimensional (2D) intensity map of the ODMR spectra. The amount of frequency splitting, i.e., Zeeman splitting, is used to detect DC magnetic fields that are parallel to the NV center axis in the low field regime. Panel (c) is adapted with permission from Ref. 83; © 2020 The Authors.
2.2.
Quantum Sensing Protocols
As shown in Figs. 4(b) and 4(c), ODMR spectroscopy offers the simplest method for DC magnetometry. However, continuous driving of laser and MW fields can induce power broadening, which reduces the signal contrast and increases the linewidth of ESR resonances, ultimately degrading sensitivity.84 In this section, we discuss quantum sensing protocols based on both continuous wave and pulsed schemes of laser and MW fields. These protocols can be categorized into four different types: relaxometry, Rabi, Ramsey, and correlation protocols.
2.2.1.
Relaxometry
Quantum coherence is a valuable resource in quantum sensing,85 ,86 playing a crucial role in sensitivity.82 Quantum coherence time can be categorized into two types: longitudinal (spin-lattice) relaxation time and transverse (spin-spin) relaxation time . The time is governed by energy exchange processes between the spin and energy-matched levels in the surrounding environment, such as a thermal bath or spin bath.87 Through these interactions, the spin dissipates energy, and the corresponding relaxation rate is determined by the strength of the interaction.
relaxometry utilizes changes in the time due to interactions with environmental noise, especially when the relevant energy scale matches the spin two levels, typically in the GHz range for spin defects. As the electron’s precession frequency falls within this range, relaxometry has been used to probe GHz stochastic signals from ionic samples such as . Figure 5(a) illustrates the pulse sequences used for NV’s relaxometry. The NV center is initialized and prepared in a specific state (e.g., the excited state ) using laser and/or MW pulses. Note that we assume that after initialization, the qubit is prepared in its ground state, taking into account the depopulation from the metastable state. After a finite evolution time , qubit readout is performed with a laser pulse. By repeating the procedure with varying , the NV’s photoluminescence (or contrast) is plotted as a function of , and the time is obtained from the decay constant.
Fig. 5
Basic quantum sensing protocols with diamond NV centers. Laser and MW pulse sequences and measurement results for (a) relaxometry, (b) Rabi oscillation, (c) Ramsey interferometry, and (d) spin Hahn echo. "Init." and "R.O." refer to qubit initialization and readout. We assume that after initialization, the qubit is prepared in its ground state, taking into account the depopulation from the metastable state. The evolution time corresponds to the time interval between the "Init." and "R.O." pulses. (b) A continuous MW field is applied, and the oscillation frequency corresponds to the Rabi frequency. (c) The oscillations in Ramsey measurement are related to the NV’s hyperfine triplet, which arises from the interaction with the nuclear spin. (d) The periodic contrast suppressions occur at the Larmor frequency of the nuclear spin bath and its higher-frequency replicas.
2.2.2.
Rabi-type sensing protocols
Although relaxometry measures changes in the time due to interactions with stochastic signals or noise at the spin resonance frequency, the Rabi protocol measures changes in the rate of population inversion between the ground state and excited state due to coherent driving at the spin resonance frequency. Figure 5(b) illustrates the pulse sequence and the resulting Rabi oscillations. After initialization, the qubit undergoes coherent transitions between and due to continuous MW. The magnitude of the MW field determines the oscillation period, known as the Rabi frequency. Changes in the Rabi frequency due to external coherent signals are used to detect GHz AC signals.88 ,89
The Rabi frequency , which varies according to the amplitude of the AC signal, can be calculated using the following equation:
Eq. (3)
Sometimes, pulsed and continuous protocols are distinguished by whether the qubit undergoes free evolution between initialization and readout. Although the Rabi protocol uses pulsed MWs, it can be classified as a continuous driving protocol because the qubit is continuously driven during the evolution period. Continuous MW driving, as used in the Rabi-type protocol, offers several advantages over pulse-based protocols. For instance, Rabi-type magnetometry has been demonstrated using continuous dynamical decoupling (CDD) methods.90 –94 CDD enables tunable control over the spin transition energy by transforming the bare spin states into dressed states, serving as an effective noise filter to suppress background noise outside the dressed state’s frequency. This approach allows for wide-bandwidth AC magnetometry with improved sensitivity. Note that typical CDD protocols resemble Rabi-type protocols when detecting coherent signals, whereas for randomly phased, incoherent signals, they become more like relaxometry techniques.
Recently, more advanced CDD methods have been developed. For example, concatenated CDD has successfully demonstrated prolonged coherence time, reduced noise floors, and extended spectral ranges toward both low-frequency (sub-MHz)95 and high-frequency (GHz) regimes.96 In addition, a phase-modulated CDD technique, known as phase-modulated Hartmann–Hahn double resonance, has demonstrated a four-order power reduction, resulting in spectral narrowing.97 Although pulse-based sensing protocols are still predominantly used in quantum sensing experiments, the Rabi-type protocol is emerging as a promising alternative methodology for quantum sensing.
2.2.3.
Ramsey-type sensing protocols
Ramsey interferometry is the most basic pulse-based sensing protocol for detecting DC and low-frequency magnetic signals (typically less than tens of kHz). Its sensitivity is fundamentally limited by the time; however, due to non-zero inhomogeneous low-frequency noise, the sensitivity is instead determined by the effective time, known as . Figure 5(c) illustrates the pulse sequence and an example result of Ramsey interferometry. After initialization, the spin qubit is transformed into a maximally superposed state of and , e.g., , with a pulse. During the free evolution period , the qubit acquires a spin-dependent phase, resulting in a phase difference given by
Eq. (4)
Although Ramsey interferometry is sensitive to detecting phase differences induced by DC signals, adding additional pulses in between the Ramsey sequence enables the detection of AC signals up to several tens of MHz. These pulse-based sensing protocols are known as pulsed dynamical decoupling (PDD) protocols. Figure 5(d) shows an example of the PDD protocol with a single pulse,98 known as the Hahn echo. The detection of AC magnetic signals from neighboring nuclear spins has been clearly demonstrated using the method [Fig. 5(d)]. Besides the Hahn echo, various PDD protocols have been developed by incorporating additional pulses, including Carr–Purcell–Meiboom–Gill (CPMG), the XY family (XY4, XY8, etc.), and Knill dynamical decoupling. Each protocol is designed to address specific challenges, such as mitigating phase errors or pulse imperfections.99 –101 PDD protocols have become fundamental building blocks both for conventional and NV-based nuclear magnetic resonance (NMR) spectroscopy.44 –62
The effects of pulses can be understood through filter functions in the toggling frame. Without pulses, as in Ramsey interferometry, the calculated filter function acts as a low-pass filter, allowing only DC and low-frequency signals to pass through. By contrast, pulses create a band-pass filter, where the central frequency shifts toward higher frequencies, and the filter’s bandwidth becomes narrower as the number of pulses, , increases, as depicted in Fig. 6(a). A higher central frequency and a narrower bandwidth enable the detection of AC signals, protect the qubit from band-mismatched noise, enhance coherence time, and improve both spectral resolution and sensitivity [Fig. 6(b)].13 ,35 ,102 –104 However, the performance is limited by the Rabi frequency (or the pulse width) and the time (which itself is limited to ). As a result, the highest AC frequency detectable using PDD protocols is typically limited to several tens of MHz. For AC sensing at higher frequencies ( to GHz), relaxometry and Rabi-type protocols can be used, as discussed previously.
Fig. 6
Performance of PDD with pulses. (a) Noise filter functions of the CPMG- protocol for a sensing time. denotes the number of pulses. As increases, the filter band shifts to higher frequencies, and the bandwidth becomes narrower. (b) Simulated magnetic field sensitivity to compare various PDD protocols. The calculations assume and an NV’s ESR contrast of 0.3. In addition, it is assumed that the time of an -pulse sequence is enhanced by a factor of .
Although PDD protocols are typically vulnerable to pulse errors, various mitigation methods have been developed over the decades. For instance, under the assumption of only 0th-order terms in average Hamiltonian theory,99 –101 the XY4 sequence is more robust to qubit rotation errors than the CPMG sequence. Pulse errors can also be mitigated by designing composite pulse sequences, such as the M. Levitt pulse and BB1, or by employing optimal control techniques.105 –107 To avoid inherent spurious effects caused by the symmetries of pulse sequences, appropriate pulse sequences can be selected, as each sequence exhibits a different spurious response. In addition, this issue can be fundamentally alleviated by utilizing correlation measurements or implementing pulse sequences with a correlated random phase scheme, which will be discussed in detail in Sec. 2.2.4.108 ,109
2.2.4.
Correlation-type sensing protocols
In addition to sensitivity, spectral resolution is also an important metric, particularly for NMR spectroscopy. Spectral resolution is typically limited by the spin coherence, which is on the order of a few milliseconds. In this section, we discuss how this challenge has been addressed, focusing on two sensing protocols: correlation measurement and quantum heterodyne measurement.
The spectral resolution of a sensing protocol is closely related to sampling theory: the longer the sampling time, the better the spectral resolution. In basic PDD protocols, the sampling time is limited by the coherence time, which for NV centers is typically less than 1 ms, resulting in a spectral resolution worse than 1 kHz. In nano-NMR spectroscopy, this resolution is insufficient to resolve chemical shifts (which depend on the applied magnetic field and are on the order of Hz for hydrogen NMR under a few teslas) or J-couplings (also on the order of Hz). The primary approach to overcoming this limitation is to utilize the correlation of sequential measurements, which effectively extends the sampling time beyond the limit, enabling significantly higher spectral resolution.
For instance, time correlations between two sequential phase measurements have demonstrated extended spectral resolution limited by ,54 which can be further enhanced with the quantum memory, as shown in Fig. 7. The phase information from the first sensing period is stored in the qubit population during the correlation time, which is related to , and then correlated with the phase acquired from the second sensing period. The spectral resolution can be improved further by a few orders of magnitude through the use of a nuclear memory qubit, with time that is much longer than that of the electron sensing qubit.54 ,110 ,111
Fig. 7
Time correlation protocol for NMR spectroscopy. NV center (electron spin) serves as a sensing qubit, while a nuclear spin serves as a memory qubit. The MREV-8 NMR pulse can be used to reduce the spectral broadening from the sample dynamics. Chemical shifts of both and nuclear spins on the order of Hz are resolved with the quantum memory assisted correlation measurements. Images are adapted and modified with permission from Ref. 110; © 2017 The Authors, some rights reserved; exclusive licensee American Association for the Advancement of Science.
The correlation measurement can be further improved by synchronizing the sequential measurements with an external clock, a technique known as quantum heterodyne (Qdyne) measurements, or, equivalently, coherently averaged synchronized readout (CASR). These methods have demonstrated sub-Hz spectral resolution, limited by stability of the clock and coherence time of the sample.52 ,112 ,113 Achieving sub-Hz resolution enables chemical shift-resolved NMR spectroscopy at relatively low magnetic fields ().
One of the challenges of conventional PDD techniques for NMR studies is their difficulty in operating under high magnetic fields (), as the nuclear Larmor frequency becomes too high for MW drives to track. This issue is critical in NMR spectroscopy because samples often require high magnetic fields to ensure long coherence time or high sample polarization. This challenge has been addressed by several methods, such as direct quantum memory access,110 ,111 ,114 electron nuclear double resonance quantum heterodyne,115 and amplitude-encoded radio-induced signal.116
Moreover, time correlation protocols, which have been used for a single NV (or a single group of ensemble NV centers), have recently been applied to spatial correlation measurements between more than two NV centers, demonstrating enhanced bandwidth and improved signal-to-noise ratio (SNR).37 –41
2.3.
Toward Quantum Advantage
One of the long-term goals of "quantum" sensing is to achieve quantum advantage over classical sensors by surpassing the standard quantum limit (SQL) of sensitivity or spectral resolution. Although many theoretical proposals exist to realize this advantage, only a few experimental demonstrations have been made in quantum sensing research based on solid-state spin defects. In this section, we focus on recent experimental demonstrations aimed at achieving this goal.
2.3.1.
Weak measurement
Investigating the fast-varying dynamics of a sample is an important objective of quantum spin sensors. However, conventional sensing protocols based on strong measurements inherently limit such studies by significantly perturbing the sample due to measurement backaction. By contrast, weak measurements harness the sequential detection of target dynamics using a weak probe. These relatively unperturbed measurements preserve the quantum properties of the target and allow for the tracking of nuclear spin dynamics, as illustrated in Fig. 8.117 Similar to the Qdyne measurement, however, the spectral resolution of the weak measurement is also limited by the sampling time and the target sample’s lifetime, resulting in a spectral linewidth on the order of Hz.
Fig. 8
Quantum sensing with the weak measurement. (a) Schematic of sequential weak measurements, which enable the tracking of nuclear spin dynamics. (b) An example NMR spectrum acquired using weak measurements, showing a single 13C nuclear spin with spectral resolution. Images are adapted with permission from Ref. 117; © 2019 The Authors, under exclusive license to Springer Nature Limited.
2.3.2.
Quantum Fisher information scaling
Increasing the number of sensing qubits () or expanding the dimension of the Hilbert space is one of the most straightforward approaches to increasing sensitivity or quantum Fisher information.118 –123 Entangled sensing qubits can coherently accumulate signals, allowing the quantum Fisher information to scale proportionally to , known as the Heisenberg scaling. By contrast, the Fisher information of classic sensors scales linearly with , which is eventually limited by the SQL. With robust quantum memory,124 it has been demonstrated that with an optimal phase mapping basis, the Fourier basis, the quantum Fisher information scales as (Fig. 9). This represents a state-of-the-art demonstration that combines quantum sensing and the quantum Fourier transform (QFT) technique, where the electron sensor detects the signal, and the nuclear memory spin stores and processes it. Note that additional energy levels in a larger spin system, known as a "qudit," can be advantageous, as demonstrated in Ref. 124. These extra levels enlarge the Hilbert space, which in principle leads to a more condensed quantum resource with a limited physical system. Furthermore, utilizing multiquantum coherence is expected to provide a quantum advantage,82 ,125 –130 benefiting from the large Hilbert space.
Fig. 9
Quantum phase estimation algorithm with QFT. A hybrid diamond quantum register is used to demonstrate beating the SQL. The register consists of an NV center as the sensor qubit and three nuclear spins as memory qubits. One of the nuclear spins is a spin triplet (), while the other two are spin doublet (), resulting in a total computational Hilbert space dimension of 12, corresponding to -qubits system. In this experiment, the electron spin qubit "reads" the signal from the target and "writes" it to the nuclear memory qubits, where the quantum information is processed. Images are adapted and modified with permission from Ref. 124; © 2021 The Authors.
2.3.3.
Entanglement-enabled noise filtering
A recent study131 demonstrated that quantum entanglement can be leveraged to distinguish pure quantum signals from classical noise (Fig. 10). As previously discussed, typical sensing information is acquired in the form of the accumulated phase. In this study, the base protocol is a simple correlation measurement, where the phase information is lost due to the decoherence process. However, if the signal originates from a quantum target, the entanglement backaction maps information in the qubit population, allowing it to survive against the decoherence process and effectively filtering out the classical noises. Another example of a quantum filter has been demonstrated with higher-order correlation measurements,132 which distinguishes quantum signals from classical ones by exploring the noncommuting nature of quantum signals. Similarly, other types of quantum correlations or quantum discord could serve as potential resources for quantum sensing.
Fig. 10
Quantum filter based on entanglement. (a) Classical noise is filtered out due to the decoherence process, while quantum signals remain in the qubit population through entanglement backaction. (b) Demonstration of classical noise-free detection of quantum signals. Images are adapted with permission from Ref. 131; © 2023 American Physical Society.
2.4.
Combination with Imaging Techniques
Solid-state spin defects can be integrated with imaging systems, extending the capabilities of quantum sensing to spatially resolved imaging. There are two main types of imaging based on spin defects: single-spin scanning magnetometry and wide-field quantum microscopy, as illustrated in Fig. 11. These imaging methods are applied to different target samples, depending on the required spatial resolution, imaging speed, and operating environments. For instance, single-spin scanning magnetometry has been used to study intriguing magnetic phenomena in condensed matter systems and to characterize current profiles in transport devices by providing high spatial resolution on the order of tens of nanometers.133 –136 This method offers high spatial resolution, but the overall slow scanning process limits the imaging time. By contrast, wide-field quantum microscopy has been used to image biological samples, magnetic materials, and current transport devices by providing imaging over a larger area with faster imaging speed but at the cost of reduced spatial resolution on the order of micrometers.85 ,137 ,138 Although this method allows relatively fast imaging, the spatial resolution is limited by the diffraction limit of the excitation light rather than the size of the sensing qubit, i.e., spin defect.
Fig. 11
Schematics of the two imaging techniques. (a) Schematic of single-spin scanning magnetometry. A single spin defect, e.g., NV center, is integrated at the tip of a fabricated diamond atomic force microscope (AFM). Confocal optics are employed to excite the NV center and collect the resulting fluorescence, while a sample beneath the AFM tip is scanned with the assistance of a three-dimensional scanning stage. Image is adapted with permission from Ref. 16; © 2012 Springer Nature Limited. (b) Schematic of wide-field quantum microscopy. Ensemble NV centers embedded in a diamond plate are used to detect magnetic fields from a sample on the diamond surface. Laser excitation occurs over the field-of-view of an objective lens, e.g., to , and the fluorescence signals from the NV ensemble are captured using a photosensitive camera, such as a complementary metal oxide semiconductor (CMOS) camera.
Figure 12 compares the spatial resolution and temporal resolution of the two imaging methods. Given a fixed ability to extract information, there is an inherent trade-off between spatial and temporal resolutions. For example, super-resolution imaging techniques have been implemented in wide-field quantum microscopy, achieving higher spatial resolution at the cost of reduced temporal resolution. Xu et al.139 used nine images to reconstruct a finer image for each pixel, resulting in a nine times slower imaging process. Conversely, pixel binning in an configuration can enhance sensitivity and, consequently, temporal resolution while sacrificing spatial resolution.140 Considering this relationship, the appropriate imaging technique should be selected based on the target sample and its dynamic properties.
Fig. 12
Comparison of spatial and temporal resolutions for the two imaging methods: single-spin scanning magnetometry (SM) and wide-field quantum microscopy (WF). WFSR refers to wide-field imaging using super-resolution techniques. Although single-spin scanning magnetometry offers higher spatial resolution but lower temporal resolution (defined as the inverse of the pixel rate), wide-field quantum microscopy provides the opposite—lower spatial resolution but higher temporal resolution. The numbers assigned in the figure correspond to Refs. 139 140. 141. 142. 143. 144. 145. 146. 147. 148. 149. 150. 151. 152. 153.–154.
2.4.1.
Single-spin scanning magnetometry
Single-spin scanning magnetometry has been implemented for various magnetic materials and current transport devices, operating in environments from room temperature down to cryogenic temperatures. Figure 13 presents imaging examples of magnetic domain walls,37 a single vortex in (YBCO) superconductor,148 2D ferromagnets,39 spin waves in yttrium iron garnet (YIG),135 multiferroic materials,150 current flows in graphene devices,155 and super current in Josephson junctions.156 This method has proven to be a powerful tool for investigating interesting magnetic phenomena in exotic materials, spintronics, magnonics, and transport devices.
Fig. 13
Imaging examples using single-spin scanning magnetometry. (a) Magnetic images of magnetic domains (left) and a domain wall (right). Adapted and modified with permission from Ref. 37; © 2014 American Association for the Advancement of Science. (b) 2D van der Waals ferromagnets. Adapted and modified with permission from Ref. 148; © 2019 The Authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. (c) Single vortex in YBCO superconductors. Adapted and modified with permission from Ref. 39. © 2016 Springer Nature Limited. (d) Spin waves propagating on YIG surface. Adapted and modified with permission from Ref. 135; © 2021 The Authors. (e) multiferroic materials. Adapted and modified with permission from Ref. 150; © 2017 Macmillan Publishers Limited, part of Springer Nature. (f) Current flows in graphene devices. Adapted and modified with permission from Ref. 155; © 2021 The Authors. (g) Super currents in Josephson junctions. Adapted and modified with permission from Ref. 156; © 2024 The Authors.
In addition to the image applications, ongoing efforts are focused on enhancing its performance. For example, the atomic force microscope (AFM) tip with a parabolically designed diamond nanopillar has demonstrated improved photon collection efficiency and directionality of emitted photons, achieving saturation photon count rate of 2.1 million counts per second (Mcps).157 Recently, the gradiometry technique has been introduced, which converts DC to AC magnetic fields by utilizing periodic driving of the AFM tip and the strong gradient of magnetic fields from the sample. This method enables the detection of DC signals using more sensitive AC sensing protocols.158 Furthermore, the use of multiple NV centers for imaging has been proposed to improve sensitivity through correlated signal measurements.159 These advancements continue to pave the way for further improvements and applications of defect-based scanning magnetometry.
2.4.2.
Wide-field quantum microscopy
By replacing a single-point photodetector with a multipixel array camera, wide-field quantum microscopy enables the simultaneous acquisition of fluorescence signals from spatially distributed NV centers within the field of view. This provides the advantage of short imaging times (seconds to minutes) and large imaging areas ( to ).105 ,160 ,161
Wide-field quantum microscopy has been applied to imaging magnetic materials, current transport devices, geological rocks, and biological samples. Particularly, the nontoxic and bio-friendly nature of diamond makes this imaging method a powerful tool for studying cells, tissues, neurons, etc. Figure 14 illustrates imaging examples of magnetotactic bacteria,162 Fe nanowires,138 spin waves in YIG,163 magnetized eucrite ALHA81001,160 current flows in graphene ribbons,149 and magnetic steganography.164
Fig. 14
Imaging examples using wide-field quantum microscopy. (a)–(f) Magnetic images. (a) Individual magnetotactic bacteria. Adapted and modified with permission from Ref. 162; © 2013 Springer Nature Limited. (b) Fe nanowires. Adapted and modified with permission from Ref. 138; © 2023 Wiley‐VCH GmbH. (c) Spin waves propagating on YIG surface. Adapted and modified with permission from Ref. 163. (d) Weakly magnetized eucrite ALHA81001. Adapted and modified with permission from Ref. 160; © 2017 American Geophysical Union. (e) Current flows in graphene ribbons. Adapted and modified with permission from Ref. 149; © 2017 The Authors. (f) Magnetic steganography (imaging hidden magnetic features). Adapted and modified with permission from Ref. 164; © 2025 The Authors.
To improve diffraction-limited spatial resolution, super-resolution imaging techniques have been introduced. These include stochastic reconstruction microscopy,165 structured light illumination microscopy,139 ,166 aberration correction through image deconvolution with an applied point spread function,167 and leveraging magnetic field phase changes under field gradients for super-resolution imaging.168 ,169
In addition, nanofabricated pillar arrays have been implemented to improve optical collection efficiency and increase photon signals.170 For example, an improvement in sensitivity by times with nano-pillar arrays has been demonstrated.170 Double-etched multicone structures have also been studied, narrowing the far-field angle and enabling coupling with low-NA objective lenses. These structures also improve spatial resolution due to the sharp apex of the pillar tips.171 These advancements in photon collection over a wide imaging area promote increased SNR and faster imaging acquisition rates.
2.5.
Emerging Materials and Defects for Quantum Sensing
Despite the advantages of NV centers in quantum sensing, their host material, diamond, presents several challenges, such as high cost, difficulties in growing large-sized single crystals, and challenges in nano-fabrication. In addition, the optical properties of NV centers, e.g., a small portion of ZPL and broad phonon sidebands, pose further barriers to their application and optical interfacing. To overcome these limitations, alternative types of defects and host materials have recently been explored through both theoretical and experimental research.
In diamonds, group IV defects with inversion symmetry, such as silicon vacancy (SiV), germanium vacancy (GeV), and tin vacancy (SnV) centers, have garnered increasing attention. These defects are notable for their large ZPL percentages and narrow emission linewidths. These properties enable all-optical coherent addressing of the electronic spin, showing potential for single-spin quantum sensing in tesla-range magnetic fields.172
Among other host materials, SiC has emerged as a promising alternative to diamond. With polytypes, SiC supports a diverse set of spin defects, such as silicon vacancies, divacancies, and nitrogen vacancies, with emissions spanning the near-infrared (NIR) to infrared (IR) range.173 –177 This makes SiC highly suitable for bio-sensing and imaging, as NIR and IR wavelengths minimize cellular damage and allow for deeper tissue penetration. Notably, the V2 defect among silicon vacancies ( to 1100 nm) has demonstrated magnetic fields sensing with a sensitivity of at room temperature, leveraging sharp photoluminescence variations near level anti-crossing without the need of an RF system.178 In addition, purified 4H-SiC hosts V2 defects with outstanding ODMR sensitivities,179 exhibiting a magnetic sensitivity of . On the other hand, divacancies ( to 1400 nm) have achieved thermal sensitivity of using the thermal CPMG (TCMPG) protocol.180 Similarly, nitrogen vacancies in 4H-SiC ( to 1400 nm) enable room temperature quantum sensing of AC fields with a spectral resolution of 10 kHz using XY8-2 correlation spectroscopy.181 The near-telecom wavelength fluorescence in SiC not only facilitates practical applications in medical, chemical, and biological analysis but also improves its compatibility with fiber optics. A fiber-integrated SiC magnetometer using a diameter sample achieved sensitivities of for silicon vacancies and for divacancies in 4H-SiC,182 ,183 highlighting their potential for portable and practical quantum sensors.
Hexagonal boron nitride has also drawn attention as a new candidate host material for defect-based quantum sensing. As a van der Waals 2D material, hBN possesses great potential in flexibility,184 integrability,185 and compatibility with existing fabrication techniques.186 Moreover, hBN hosts a variety of spin defects, including negatively charged boron vacancies ()187 –189 and carbon-related (C-related) defects.190 ,191 For instance, has a longitudinal (transverse) ZFS of 3.48 GHz () at room temperature. The magnetic response in the ODMR spectrum indicates as an spin system, with the calculated sensitivity expected to reach up to . Although this sensitivity is somewhat moderate compared with diamond NV centers, the low-dimensional characteristics of hBN significantly enhance its potential as a versatile quantum sensing platform. Leveraging the thin film nature of hBN, spin defects can be deterministically integrated onto photonic integrated circuits192 or fiber optics.193 They can also be transferred in close proximity to the sensing target. This has enabled magnetic imaging of van der Waals ferromagnets,194 ,195 temperature sensing via Joule heating from laser absorption,195 and strain sensing on nanostructured substrates.196 ,197 This versatility highlights the potential of hBN-based spin defects for a wide range of quantum sensing applications.
In addition to the host materials discussed so far, numerous other candidates are capable of hosting spin defects. For instance, recent studies have explored new defects in GaN, investigating their optical and spin properties. Room temperature ODMR measurements have been successfully carried out in the wavelength range from visible198 to telecom.199 Although the exact origin of these defects remains unclear, angle-resolved and magnetic field-dependent ODMR measurements have been used to deduce their spin quantization axis and spin quantum number. With GaN’s established presence in the light-emitting diode (LED) industry, it is positioned as a promising platform for spin defects.
Silicon, another representative semiconductor material, also shows great promise for quantum sensing. Spin defects such as T-centers and G-centers in Si are attracting attention due to their compatibility with complementary metal oxide semiconductor (CMOS) technology and emission at telecom wavelengths.200 ,201 This compatibility presents a unique opportunity to leverage established photonic platforms, such as photonic-integrated circuits and fiber optics. However, the primary challenge of using defects in Si lies in the host material’s relatively small bandgap, which restricts spin defect activation at cryogenic temperatures. Despite this limitation, Si remains a strong candidate for integrated quantum sensing platforms, as its solid-state architecture is easier to construct compared with other host materials.
In Table 1, we summarize and compare the optical properties (wavelength range, Debye–Waller factor, and lifetime) and spin properties (ZFS, spin quantum number, and spin coherence time) of the spin defects in the discussed host materials.
Table 1
Summary of optical and spin properties of spin defects in diamond, SiC, hBN, Si, and GaN.
Defect kinds | Optical properties | Spin properties | ||||
---|---|---|---|---|---|---|
Wavelength (nm) | FDebye−Waller | Lifetime | Zero-field splitting (DZFS, EZFS) | Spin quantum number (S) | Spin coherence (T1, T2*, T2) | |
Diamond | ||||||
to 800 nm202 ,203 (ZPL: 637 nm) | 0.044204 | : 12 ns81 | 71 ,202 | 205 | ||
206 | ||||||
: 7.8 ns81 | 207 | |||||
205 | ||||||
ZPL: 737 nm208 ,209 (weak phonon sideband) | 0.76 to 0.88208 | 1 to 2 ns210 ,211 | (ground state), 258.1 GHz (excited state)210 | 212 | ||
212 | ||||||
SiV0 | 946 nm213 | 0.9214 | 1.8 ns214 | 214 | 214 | |
SnV | ZPL: 619 nm215 ,216 | 0.57217 | 216 | (ground state), 3 THz (excited state)218 | 219 | |
218 | ||||||
GeV | ZPL: 602.7 nm220 | 0.6221 | 1.4 to 5.5 ns,220 | (ground state), 981 GHz (excited state)222 | 222 | 223 |
6.0 ns221 | ||||||
223 | ||||||
4H-SiC | ||||||
V1 | ZPL: 862 nm224 | 8%225 | 5.03 to 6.26 ns226 | (ground state), (excited state)225 | 224 | 227 |
861 nm227 ,228 | 40%228 | 227 | ||||
V2 | ZPL: 916 nm224 | 9%225 | 6.1 to 11.3 ns224 | (ground state), (excited state)225 | 226 | 179 |
229 | ||||||
229 | 229 | |||||
hBN | ||||||
to 900 nm (ZPL: 773 nm)230 | — | 1.2 ns187 | 189 | |||
231 ,232 | 232 | |||||
189 | ||||||
Carbon-related | 540 to 620 nm233 | 0.8233 | 4.4 ns233 | , () 190 | 234 | 191 |
, 1190 | 235 | |||||
, ()190 | ||||||
Silicon | ||||||
T-center201 | 935 meV (1326 nm) | 0.23 | ||||
G-center200 | ZPL: 1278 nm | 0.2 | ||||
GaN | ||||||
Ref. 198 | 667 nm | >0.5 | () | , 3/2 | ||
, () | ||||||
Ref. 199 | 1325 nm | , |
3.
Advancing Optical Interfaces for Spin Defects
Quantum sensing based on spin defects involves optical initialization, manipulation, and readout of spin states. Therefore, having an optimal optical interface between spin defects and optical systems is crucial for enhancing sensing capabilities, such as sensitivity, accuracy, and speed. Conventional confocal scanning microscopy with bulk host material limits these sensing capabilities in several points. The spatial resolution is mostly restricted by the diffraction limitation of the optical systems, which is defined as , referred to as the Abbe diffraction limit. In the case of sensitivity, following the equation ,19 the sensitivity of spin defects-based quantum sensing (: detectable frequency shift of ODMR spectrum) is strongly affected by the brightness () of spin defects as well as the linewidth () and contrast () of the ODMR spectrum. Furthermore, nonoptimal excitation and collection conditions of spin defects degrade fidelity and extend integration time, which restricts the implementation of advanced quantum sensing protocols and limits their practical use. The spatial resolution and brightness can be further improved through advancements in the optical interface. In this section, we introduce several strategies to enhance the optical performances of spin defects: (1) nanostructure fabrication for improving light extraction, (2) micro- and nano-scale optical lenses for improving light collection, and (3) cavity coupling for enhancing light–matter interaction. The strengths and potential of each approach are discussed in detail below.
3.1.
Low-Dimensional Nanostructure
As the high refractive index of host crystals hinders the extraction of fluorescence from spin defects, low-dimensional structures, such as nanoparticles, nanowires, and atomic-thin films, are getting great attention by serving significantly enhanced light extraction from defects.69 ,235 ,236 Moreover, these structures can efficiently interface with various sensing platforms, including scanning tips,16 ,145 ,146 ,237 fiber optics,238 –241 microfluid chips,242 and photonic chips.243 –245 There are numerous nanofabrication techniques available for creating such structures, broadly categorized into bottom-up growth and top-down fabrication.
3.1.1.
Zero-dimensional nanoparticle
Nanoparticles, as shown in Fig. 15(a), are a widely used form of nanostructures. These nanocrystals exhibit high optical brightness by minimizing total internal reflection,69 ,235 ,236 and they offer an excellent integrability with various platforms. Figure 15(b) describes how nanodiamonds can be easily integrated with sensing targets using a drop-casting method, a technique also applicable for practical platforms such as optical fibers.240 Moreover, their nanoscale size allows for precise positioning using a mechanical tip243 ,248 or optical trapping technique,246 ,249 as depicted in Fig. 15(c). These advantages of enhanced light extraction, integrability, and controllability make nanocrystals ideal for implementing efficient, deterministic, and potentially scalable quantum sensing through the arrangement of nanocrystal arrays.235
Fig. 15
Examples of low-dimensional nanostructures. (a) Schematic of spin defect in 0D nanocrystal and properties: (b) great integrability using a drop-casting method and (c) deterministic nanopositioning via mechanical tip and optical tweezer. (b) Adapted and modified with permission from Ref. 243; © Optica Publishing Group. (c) Adapted and modified with permission from Ref. 246; © 2021 American Chemical Society. (d) Schematic of spin defect in 1D nanowire and representative properties: (e) improved light extraction efficiency through optical guiding via one-dimensional nanostructure and (f) improved spatial resolution and spin addressing abilities via sharp-tip configuration. Adapted and modified with permission from Ref. 234 under CC-BY-NC-ND license. (g) Schematic of spin defect in 2D thin-film and its properties: (h) compatibility with nanofabrication. Adapted and modified with permission from Ref. 247 under CC-BY license. (i) Giant response to external mechanical deformation. Adapted and modified with permission from Ref. 196; © 2022 American Chemical Society.
One prominent example of such utility is a tip-attached quantum sensing probe capable of single-defect operation for nanoscale resolution. Various host materials support spin defect emission in nanoparticle structures, including diamond,69 ,250 ,251 SiC,252 and hBN.235 ,253 For high-resolution sensing, sub-10 nm nanocrystals are preferred due to their proximity to the target samples.236 However, shallow defects in such small crystals often suffer from poor optical stability, including brightness fluctuation, bleaching, or quenching caused by surface-bound states.254
To mitigate these issues, several approaches have been explored, including highly purified nanocrystals using chemical vapor deposition (CVD), slow-growth techniques at high-pressure high-temperature,255 –257 or mechanical milling of high-quality bulk crystals.258 ,259 Another approach involves post-growth treatments, such as surface modification via aerobic oxidation250 ,260 ,261 or post-annealing process at prolonged high temperature at 1600°C.262 In addition, scaling up the nanocrystal size, while keeping them below 100 nm, can help maintain optical and spin properties for nanoscale quantum sensing.69
3.1.2.
One-dimensional nanowire
Recently, one-dimensional (1D) nanostructures, such as nanotubes and nanowires [Fig. 15(d)], have gained significant attention for their utility and compatibility with scanning probe systems and integrated photonic platforms for wide-field imaging. Fabricated diamond nanowires, for instance, exhibit enhanced photon extraction through optical routing along the nanowire axis, demonstrating up to a 10-fold improvement compared with unstructured NV and SiV centers.263 ,264 Figure 15(e) displays a corresponding numerical simulation highlighting the enhanced light extraction via waveguiding effects in nanowire geometries.
Owing to these properties, ultra-bright wide-field imaging has been demonstrated using nanowire, nanopillar, and nanocone arrays fabricated on bulk diamond substrates.170 ,171 ,265 In addition, the directional light guiding enabled by these nanostructures allows for efficient collection of PL, particularly from sharp-tip scanning probes. For example, PL collection from a single NV center has reached up to 2.1 Mcps due to improved optical coupling to the objective lens.16 ,157 Beyond brightness enhancement, these 1D nanostructures also improve spatial resolution. Figure 15(f) illustrates a single C-related spin in a BN nanotube ( diameter) attached to an AFM cantilever, demonstrating high spatial resolution and strong photon collection for nanoscale sensing.234
Moreover, 1D elongated nanostructures can efficiently interface with mechanical forces, enabling ultra-sensitive force-sensing applications.266 ,267 Mechanical modes induced by external forces can couple to spin sublevels, forming hybrid spin-nanomechanical sensing platforms.29 ,268 ,269 Mechanical strain coupling, in particular, can modulate the spin qubit coherence, thereby enhancing spin coherence time and, in turn, improving sensitivity.270
Although the growth and synthesis of diamond-based nanowire architectures remain challenging, promising strategies involve the use of carbon nanotube271 or silicon nanowire templates.272 In addition, spin defects embedded in different stacking interfaces have attracted interest due to their unique optical properties, such as nontrivial optical performances in brightness and phonon dynamics. For example, recent studies on SiC nanowires have explored point defect-stacking fault complexes.273 These systems exhibit enhanced brightness and suppressed electron-phonon coupling due to the proximity of stacking faults to point defects. As a result, strong ZPL emission from is observed even at room temperature, with minimized phonon sidebands. Such features are highly advantageous for integration with photonic platforms, providing low-loss and low-dispersion optical interfaces.
3.1.3.
Two-dimensional thin film
The utilization of thin film structures [Fig. 15(g)] is also an attractive candidate for quantum sensing, offering excellent integrability with targets and engineerability through established fabrication techniques. Such thin films can be prepared by CVD growth,274 –276 epitaxial layer growth (especially for SiC or GaN),277 ,278 or by mechanical thinning and polishing of bulk crystals.279 –281 Among these, hBN, a van der Waals material, can be easily exfoliated down to atomic layers without requiring complex growth processes. This two-dimensional (2D) nature makes hBN compatible with nanofabrication, enabling precise engineering of the optical environment around spin defects.
These thin-film platforms are ideal for integrating photonic structures that tailor light–matter interactions. For example, Fig. 15(h) shows a photonic crystal cavity fabricated in an hBN membrane, where the coupling between spin defects and cavity photons enhances optical interfacing efficiency with external optical systems.247 In addition, the mechanical flexibility of thin film materials allows them to respond strongly to external strain, making them suitable for wide-area sensing applications. As illustrated in Fig. 15(i), hBN atomic layers have been used to probe nanobubble morphologies on deposited metal films through their strain sensitivity.196
Beyond hBN, emerging spin-hosting materials such as Si and GaN offer significant advantages due to their compatibility with existing industrial infrastructure. Thin film growth of these materials supports scalable and wide-area sensor fabrication, which is essential for translating quantum sensing technologies into real-world applications. Thus, the development and integration of 2D thin films in spin-hosting materials are critical steps toward commercialization and practical deployment of quantum sensing platforms.
3.2.
Micro- and Nano-Scale Optical Lenses
The aforementioned low-dimensional nanostructures offer excellent light extraction; however, their nanostructured surfaces often introduce optical instabilities, such as brightness fluctuations and blinking, as well as degraded spin coherence.254 ,282 –284 To mitigate these adverse effects, the use of micro- and nano-scale optical lenses for addressing spin defects has been extensively studied. These optical elements significantly enhance light extraction and photon collection efficiency without compromising the optical or spin properties of the emitters. By employing such lenses, the NA of the optical interface is effectively increased, and the far-field emission profile of spin defects can be optimized for improved signal collection.285 –287
Moreover, these modulated optical interfaces not only boost photon collection efficiency but also enable subwavelength spatial resolution.288 –291 Micro-optical lenses can be implemented either by fabricating the optical structures directly onto the host crystal, forming a monolithic platform,292 –294 or by integrating dielectric lenses onto the surface of spin-hosting materials.295 –299 More recently, the emerging field of meta-optics has been applied to quantum optics applications.300 –304 In particular, metalenses have demonstrated exceptional performance in photon collection, surpassing the limits of conventional diffractive optics. These improvements in brightness and optical resolution are critical for advancing the performance and reliability of quantum sensors.
3.2.1.
Solid immersion lens
A solid immersion lens (SIL) is a widely used microlens for defect characterization, whose schematic is illustrated in Fig. 16(a). The first reported SIL was fabricated using focused ion beam milling (FIB) to create a hemispherical solid immersion lens centered around single diamond NV centers.295 In this configuration, the NV center is precisely positioned at the focal point of the hemispherical SIL, enhancing the NA of the optical interface and significantly increasing the collimation angle of the emitted light. This improvement leads to a substantial boost in photon collection efficiency, with reported brightness enhancements exceeding an order of magnitude, as illustrated in Fig. 16(b). State-of-the-art implementations have achieved a saturated photon collection rate of up to 2.4 Mcps from a single NV center.305
Fig. 16
Examples of micro- and nano-scale optical lenses. (a) Schematic of a solid-immersion lens and (b) brightness enhancement showing collected PL counts as a function of excitation power, with and without the lens. Adapted and modified from Ref. 295, with the permission of AIP Publishing. (c) Schematic of microsphere-assisted microscopy and (d) super-resolution imaging capability presenting confocal scanning images with and without the microsphere. The scale bars are 500 nm. Adapted and modified with permission from Ref. 296 under CC BY-NC-ND license. (e) Schematic of metalens featuring near-unity numerical aperture and (f) simulated optical emission profile achieved using the flat metalens structure. Adapted and modified with permission from Ref. 300 under CC-BY license.
Recent advancements in SIL fabrication include the development of Fresnel-type SILs,306 which significantly reduce FIB milling time by more than 2/3 compared with hemispherical designs, thereby lowering fabrication costs. Beyond monolithic fabrication, alternative methods involve heterogeneous integration of commercially available SILs made from high-refractive-index materials such as gallium phosphide (GaP)285 and gallium nitride (GaN).298 These hybrid approaches preserve the structural integrity of the host crystal while still achieving considerable brightness enhancement without degrading the spin or optical properties of the defects. In addition, the lens material and dimensions can be tailored to optimize performance for specific types of spin defects and application requirements.
3.2.2.
Microsphere
Microsphere-type lenses, as depicted in Fig. 16(c), offer distinct advantages over hemispherical SILs, particularly due to their unique ability to generate a spatially localized excitation spot known as a photonic nanojet,289 ,307 and to efficiently convert evanescent waves into propagating waves.308 These advantages in excitation and collection enable highly magnified, super-resolution imaging that surpasses the diffraction limit, achieving spatial resolution as small as through a virtual image plane formed beneath the microsphere.288 ,309
When applied to spin defect characterization, microsphere-assisted confocal scanning microscopy facilitates the optical addressing of individual photons and spin states from closely spaced defects that are otherwise indistinguishable in conventional confocal microscopy.296 As shown in Fig. 16(d), NV centers that appear spatially unresolved in standard confocal imaging (labeled D and T) can be distinctly resolved into subcomponents (D1-2 and T1-3) using a microsphere lens. In addition, this virtual imaging configuration significantly enhances SNR, with reported improvements of up to fourfold compared with direct imaging at the sample surface.
Microspheres are readily available as off-the-shelf micro-optics and can be easily deployed onto host materials via simple drop-casting methods. For precise alignment, microspheres can be positioned deterministically using micro-manipulation tools such as microtips310 or optical tweezers.296 Importantly, this method requires no complex fabrication, making it a versatile and noninvasive strategy compatible with a wide range of bulk materials and standard optical microscopy setups. Therefore, microspheres provide a relocatable, low-cost, and effective optical interface for enhancing the performance of solid-state quantum emitters.
3.2.3.
Metalens
The introduction of meta-optics offers a powerful pathway to achieve near-unity NA in photon collection using thin membrane structures.311 As illustrated in Fig. 16(e), which shows a schematic of a metalens coupled to spin defects, flat meta-optical architectures such as monolithically fabricated diamond metalenses have been developed. These structures consist of nanopillar arrays engineered to mimic discretized Fresnel phase profiles.300 In this work, a brightness enhancement of 3.5 times was reported at a collimation NA of 0.75.
A key strength of this approach lies in the ability to fabricate these meta-optics architectures as ultra-thin membranes, as shown in Fig. 16(f). By harnessing this thin membrane design, hybrid integration of silicon dioxide () metalens membranes onto NV centers in bulk diamond has been demonstrated, achieving a 34-fold enhancement in photon collection compared to pristine bulk media.312 Moreover, their ultra-thin nature allows metalenses to interface with diverse photonic platforms, such as the facets of an optical fiber.313 ,314 In addition, metalenses offer tunable focusing depth315 and exhibit chiral functionality,316 ,317 enabling applications across a wide range of use cases—from addressing spin defects at varying depths to generating distinct optical responses to circularly polarized light. Therefore, meta-lenses represent a lightweight, customizable, and high-performance optical interface with transformative potential for quantum sensing.
3.3.
Cavity-Enhanced Light–Matter Interaction
Enhancing the brightness of spin defects with optical lenses is a useful way to improve sensing sensitivity. Beyond such passive optical interfaces, photonic cavities offer an active approach to tailoring light–matter interactions, enabling precise modification of the optical properties of coupled defects. These modifications include spectrally selective enhancement at a ZPL, increased spontaneous decay rates, and optimized optical propagation modes. To illustrate the effect of photonic cavities on promoting light–matter interaction between spin defects and cavity photons, we can refer to a quantum optical framework. In this context, the enhancement of spontaneous radiative emission is characterized by the Purcell factor ,318 where is the quality factor of the cavity mode and is the mode volume.
High-Q cavities, such as dielectric photonic crystals where is extremely small, exhibit ultra-narrow spectral enhancement and reduced dispersion. These features make them especially advantageous for integrated sensing platforms.319 By contrast, cavities with small , such as plasmonic antennas, are well suited for enhancing spatial resolution by confining spin defect emission within a tightly localized optical mode. These effects contribute to both improved SNR through enhanced radiative decay rates and optimized far-field emission patterns, and higher spatial resolution via confined mode volumes.
In the strong coupling regime, interactions between cavity photons and spin defects lead to the formation of new hybridized states. These states enable coherent transfer of information between spin and photonic degrees of freedom, forming the basis for spin-photon interfaces.320 Such modifications to light–matter interaction via photonic cavities have been explored across a variety of platforms and coupling regimes.
3.3.1.
Dielectric photonic cavity
The fabrication of dielectric photonic cavities on host media has been extensively developed across various fields of nanophotonics. Representative examples include photonic crystals [Fig. 17(a)]282 and circular Bragg gratings [Fig. 17(b)],325 which have been successfully integrated with NV centers in diamond. These monolithic dielectric photonic cavity structures are also compatible with a wide range of host materials, including the aforementioned crystals. By leveraging such dielectric photonic crystal cavities, enhanced spin sensitivity has been demonstrated through cavity-coupled spin defects, as in the case of the L3 photonic crystal cavity in SiC [Fig. 17(e)].321 The cavity-coupled emission enables enhanced optical signal readout, resulting in improved contrast in spin characterization.
Fig. 17
Examples of dielectric photonic cavities. Schematics of (a) photonic crystal cavity, (b) circular Bragg grating, (c) photonic crystal nanobeam, and (d) photonic crystal waveguide. Nanophotonic architectures offer several distinct advantages: (e) enhanced spin readout contrast in ODMR spectra, (f) selective cavity-enhanced PL emission with significant Purcell enhancement, (g) far-field engineering to narrow the emission angle for improving PL collection efficiency, and (h) the provision of an efficient nanophotonic interface with exceptionally high coupling efficiency to integrated photonic platforms. Images are adapted and modified with permission from Refs. 321 322. 323.–324.
Importantly, these dielectric photonic cavities can selectively enhance specific optical frequencies, allowing suppression of background fluorescence and the generation of low-dispersion modes in integrated photonic platforms. For instance, Fig. 17(f) shows selective enhancement of the NV center’s ZPL at 637 nm in diamond photonic crystal cavities.322 Beyond enhancing the radiative decay rate, the cavity’s diffractive structure can also modulate the far-field propagation profile. As shown in Fig. 17(g), a monolithic circular Bragg grating cavity in hBN exhibits a strong vertical emission pattern.323
For coupling photons in optical fibers or photonic integrated circuits, tapered nanobeam cavities324 and photonic crystal waveguides326 ,327 are employed, as shown in Figs. 17(c) and 17(d). These lateral photon-routing configurations enable efficient coupling, for example, a tapered photonic crystal nanobeam achieves coupling efficiency to the facet of a tapered silica fiber, as shown in Fig. 17(h).324 As a result, single-mode fibers can collect of single-photon emission from the SiV in diamond. In summary, the integration of dielectric micro- and nano-cavities enables simultaneous enhancement of radiative decay rates, reduction of optical dispersion, and improved photon collection along desired optical pathways. These advantages collectively facilitate the realization of highly sensitive and efficiently integrated quantum sensing platforms.
3.3.2.
Plasmonic antenna and cavity
Plasmonic cavities offer significant advantages due to their broadband optical coupling modes and ultra-small mode volumes, making them highly compatible with randomly distributed, phonon-broadened emission from spin defects. Their extremely small mode volumes enable very high Purcell enhancements. One of the simplest methods to induce plasmonic interactions between spin defects and localized plasmons involves positioning metal nanoparticles, such as gold or silver, around the spin defects.328 –330 Figure 18(a) shows schematics of metal nanoparticles placed close to diamond NV centers.328 Due to the broadband spectral response of plasmonic enhancement, the brightness of NV center emission is enhanced across a wide wavelength range, unlike high-Q dielectric photonic cavities, as shown in Fig. 18(d).
Fig. 18
Examples of plasmonic cavities and antennas. Schematics of (a) plasmonic nanoparticles, (b) surface plasmon resonance, and (c) plasmonic nanoantenna. Plasmonic cavities offer unique advantages including: (d) a broadband cavity mode spectrum of plasmonic nanoparticles interacting with diamond NV centers. Adapted and modified with permission from Ref. 328 under CC-BY license. (e) Large-area applicability with substantial brightness enhancement, as demonstrated in the confocal PL mapping of spin defects in SiC interacting with Au coplanar lines. Adapted and modified with permission from Ref. 331; © 2023, American Chemical Society. (f) Simulated PL enhancement as a function of the distance between the plasmonic antenna and spin defects. Adapted and modified with permission from Ref. 332; © 2023, American Chemical Society.
Another approach involves coupling shallow-embedded spin defects to surface plasmons on a metal thin film [Fig. 18(b)]. As shown in Fig. 18(e), the enhancement in brightness due to surface plasmonic interactions is evident in confocal PL images comparing SiC regions on Au coplanar waveguides with pristine regions.331 This configuration demonstrated a brightness enhancement of -fold and a -fold improvement in spin manipulation sensitivity, highlighting the strong performance of plasmonic cavity effects. Similarly, shallow centers in hBN, when coupled to surface plasmons, achieve magnetic field sensitivity of sub- even under continuous-wave ODMR measurement, a regime where typical sensitivities are around .333
Recent developments have focused on low-loss plasmonic nanocavity structures to further optimize plasmonic enhancement. For example, Fig. 18(c) illustrates an advanced nanopatch antenna consisting of a metallic naocube sandwiched with a metal substrate, yielding an extremely confined optical mode. Spin defects in hBN embedded between such a nanocube and a thin metallic film have demonstrated PL enhancement of up to -fold without compromising ODMR contrast. Figure 18(f) shows simulation results indicating the dependence of brightness enhancement on the distance between shallow spin defects and the metallic substrate, with optimal enhancement observed at distances below 10 nm.
Similar approaches have also been demonstrated using Au nanotrenches,334 metasurfaces,335 nanogaps,336 and other nanostructures.337 Importantly, plasmonic cavities are relatively simple to fabricate338 and can even be integrated with dielectric nanophotonic cavities,339 ,340 enabling hybrid metal-dielectric platforms. Such configurations offer promising pathways toward chip-integrated optical quantum sensing platforms with enhanced performance and scalability.
3.3.3.
External optical cavity
External optical cavities typically consist of two or more optical mirrors that trap light between them, enabling sustained photon confinement.341 A key advantage of such cavities is the tunability of their resonant modes, which can be readily adjusted to match the emission spectrum of target spin defects. By varying the distance between the mirrors, the cavity’s optical path length can be controlled, thereby tuning the steady-state photonic density and enhancing light–matter interactions when spin defects are positioned within the cavity.
Figure 19(a) illustrates the schematic of an optical setup for optical cavity-enhanced room temperature magnetometry. It consists of a 532 nm pump laser, a 1042 nm probe laser, and two spherical mirrors forming a cavity around diamond NV centers. The probing laser is resonant with the metastable singlet transition of the NV center and is used to acquire the ODMR transmission spectrum.335 ,341 To enable miniaturization, one study implemented a packaged cavity structure by bonding a bulk diamond with an integrated planar mirror to a piezoelectric transducer, paired with a spherical mirror.342 This configuration demonstrated magnetic field sensitivity near the photon shot-noise limit, confirming the sensitivity enhancement enabled by cavity-assisted detection [Fig. 19(b)].
Fig. 19
Examples of external optical cavities. (a) Schematic of the optical setup used for diamond NV centers embedded in an optical cavity, with excitation and probing lasers. Adapted and modified with permission from Ref. 341; © 2014 American Physical Society. (b) Measured magnetic field sensitivity of the cavity-coupled diamond NV center approaches the photon-shot noise limit, achieving . Adapted and modified with permission from Ref. 342; © 2017 American Physical Society. (c) Photograph of the optical cavity setup, consisting of a fiber-based micromirror aligned with a diamond membrane on a planar mirror. (d) Observed cavity mode spectrum corresponding to the setup in panel (c) shows significant PL enhancement at the ZPL of the NV center. (e) Second-order autocorrelation measurement revealing collective photon bunching characteristics for the cavity-coupled ZPL emission, with a pronounced bunching at zero-time delay on a sub-nanosecond timescale. By contrast, confocal imaging without spectral filtering at the same sample location exhibits photon anti-bunching, whereas filtering only the phonon sidebands shows neither bunching nor anti-bunching. (f) Enlarged view of from (e) near-zero time delay. Images in panels (c)–(f) are adapted and modified with permission from Ref. 343 under CC-BY license.
A similar approach has been used to explore collective effects in cavity-coupled NV centers. Figure 19(c) shows a photograph of a device comprising a spherical micromirror aligned with a diamond membrane on a planar mirror.343 This configuration significantly enhances ZPL emission from the NV centers, as observed in the cavity mode spectrum [Fig. 19(d)]. Second-order correlation measurements at the filtered ZPL reveal photon bunching behavior, indicative of superradiance among the coupled emitters [Figs. 19(e) and 19(f)]. These collective emissions enable access to the quantum-enhanced sensing regime.343 ,344
Importantly, cavity-mediated interactions among spin defects not only support enhanced sensing via superradiance but can also be used to explore subradiant states for potential quantum storage applications,345 which may further contribute to improving sensitivity.114 Therefore, photonic cavities provide a powerful platform for engineering photon-mediated interaction between spins, offering significant promise for advanced quantum sensing.
4.
Practical Application of Defect-Based Quantum Sensors
So far, we have discussed the fundamental principles and recent advancements in quantum sensing, particularly efforts to enhance sensitivity and optical properties. Alongside fundamental research, there is growing interest in utilizing quantum sensors for practical, real-world applications. Defect-based quantum sensors, with their unique advantages, are proving to be highly versatile and well-suited for use across a wide range of fields, including biomedical,52 ,57 ,85 ,346 –354 industrial,193 ,355 –363 and military sectors.29 ,65 ,364 –368 For these sensors to be practical in real-world applications, they must meet certain conditions, such as the ability to function at ambient room temperature and the development of miniaturized, integrated sensor systems.
In this context, defect-based quantum sensors offer several compelling benefits. Their qubits are highly sensitive to various physical properties, including magnetic,13 ,14 electric,32 and strain fields,369 as well as temperature,25 inertial,29 and rotational motion.368 They also can function across a wide range of temperatures, including room temperature,78 provide high-resolution imaging with both spatial38 and temporal precision,52 and can be integrated into compact, portable devices.193 ,359 –361 Although challenges remain, particularly in reducing costs and enhancing sensitivity to compete with existing classical sensor technologies, defect-based quantum sensors, especially those based on diamond NV centers, have made significant progress and show potential across numerous applications, as outlined in Tables 2 and 3.
Table 2
Diamond NV centers’ sensitivities to various physical properties.
Physical property | Sensitivity |
---|---|
Magnetic field | (single NV)13 |
(ensemble)14 | |
Electric field | [Ref. 33] |
Strain field | [Ref. 369] |
Temperature | [Ref. 370] |
Inertial motion | to [Ref. 29] |
Rotational motion | [Ref. 65] |
Table 3
Diamond NV centers’ unique advantages for quantum sensing applications.
Property | Performance |
---|---|
Spatial resolution | (single NV)38 |
Operating temperature | Sub-kelvin to over room temperature |
Diamond fabrication | Various nanostructures and devices can be fabricated |
Spectral resolution | 52 |
Detection bandwidth | DC − GHz |
Compact sensor design | [Ref. 359] |
Others | Optically stable (free from photobleaching); chemically stable, non-toxic, and bio-friendly; robust against harsh environments, e.g., temperature, pressure, high-voltage/current |
In this section, we highlight recent advancements in defect-based quantum sensors, focusing on their use in various fields, including biomedical diagnostics, semiconductor manufacturing device inspection, portable sensors for harsh industrial environments, and aerospace and military applications. Although other types of defects are still largely in the realm of basic research, diamond NV centers have already made substantial progress toward practical, real-world applications.
4.1.
Biomedical Diagnostics
Biomedical applications are among the fastest-growing fields leveraging NV-based quantum sensors. NV centers have been employed to probe local temperature and magnetic fields within individual cells, tissues, and neurons, offering several key advantages for biomedical use. They cover a wide frequency range, from DC to GHz, enabling the detection of static magnetic fields from magnetic nanoparticles (MNPs),85 ,162 ,349 ,351 ,371 sub-millisecond transient signals from action potentials,85 ,346 ,348 and GHz-frequency signals from ions used in magnetic resonance.85 ,353 ,354 ,372 ,373 NV centers also serve as highly effective fluorescence markers, especially when embedded in nanodiamonds.85 ,353 ,354 Compared with quantum dot-based markers, NV centers are more resistant to photobleaching, enhancing their durability and reliability in long-term imaging.77 In addition, the host material, diamond, is nontoxic and biocompatible, making it ideal for biological applications.77 When combined with imaging techniques such as scanning magnetometry and wide-field quantum microscopy, both single NV centers and ensembles provide powerful tools for biomedical diagnostics, providing resolution from the nanometer to micrometer scale.50 ,52 ,347 –349 ,351 ,373 –376 In this section, we explore various topics in biomedical applications, including nano- and micro-NMR, magnetic resonance imaging (MRI), cell detection, virus diagnostics, neuron activity monitoring, and magnetocardiography, highlighting the broad potential of NV centers in advancing biomedical science.
4.1.1.
NMR and MRI: nanometer, micrometer scales, and microfluidics
As discussed in Sec. 2, various sensing protocols have been developed to enhance sensitivity, spatial and spectral resolutions in NMR measurements. These protocols include pulsed and continuous dynamical decoupling, as well as coherently averaged synchronized readout, as discussed in Sec. 2. Individual proton spins have been detected within 1 s of integration using two-qubit quantum systems consisting of a single NV center and a neighboring nuclear spin.57 In addition, proton spins in a volume of ten picoliters have been detected using ensemble NV centers, with a sensitivity of .52 Alongside efforts to increase sensitivity, various methods to amplify the signal itself have been explored, such as polarizing nuclear spins beyond their thermal populations using techniques such as dynamic nuclear polarization and parahydrogen-based signal amplification by reversible exchange.377 –379 High spectral resolution is crucial for identifying different nuclear spins and chemical shifts resulting from interactions with the local chemical environment. Using nuclear memory qubits or synchronized readout techniques, high spectral resolution on the order of one part per million or sub-mHz has been demonstrated.52 ,57
For spatially resolved NMR, the imaging techniques described in Sec. 2.3 are combined with single or ensemble NV centers to achieve MRI at nanometer and micrometer scales. For nanometer-scale MRI, the NMR sample is mounted on a scanning probe microscope, e.g., an atomic force microscope, and scanned over a single NV center embedded in a diamond plate.50 ,374 This configuration is advantageous because the NV center is positioned in a fixed position, providing stability during the scanning process. As a result, sensitive NMR protocols, such as XY8-N, can be applied over prolonged periods. Early experiments have successfully demonstrated 2D NMR images of nuclear spins in poly-methyl methacrylate50 and nuclear spins in Teflon microspheres [Fig. 20(a)],374 with spatial resolution of and , respectively. As the NV’s sensing volume is fixed, the scanned image provides a laterally resolved NMR image, limiting it to 2D MRI. To extend this technique to full three-dimensional (3D) MRI, large magnetic field gradients are required to differentiate nuclear spins along the vertical direction. Recent experiments have demonstrated 3D NMR spectroscopy of neighboring nuclear spins in diamond by utilizing strong local magnetic field gradients, either from the intrinsic magnetic dipole field of the NV center380 ,381 or from externally fabricated microwire structures.382 Applying similar techniques to achieve 3D MRI on external samples remains an area for further research.
Fig. 20
NMR and MRI with diamond NV centers. Examples of 2D MRI images at nanometer (a) and micrometer (b) scales. (a) NMR signals of -rich Teflon microspheres, mounted at the tip apex of an atomic force microscope, are recorded and imaged by a single NV center located beneath the tip. Nano-indented grating patterns on the sample are captured in the scanning MRI experiment, demonstrating a spatial resolution of . Adapted and modified with permission from Ref. 374; © 2015 Springer Nature Limited. (b) 2D MRI image of nuclear spins in islands measured by wide-field diamond microscopy, achieving a spatial resolution of . Adapted and modified with permission from Ref. 175 under CC-BY license. Examples of 2D NMR spectroscopy (c) and imaging (d) based on microfluidic platforms. (c) The polarization and detection sites are separated using a microfluid channel, enabling NMR spectroscopy on water, trimethyl phosphate (TMP), and 1,4-difluorobenzene (DFB) samples, with a spectral resolution of . Adapted and modified with permission from Ref. 352; © 2019 The Authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. (d) Overhauser dynamic nuclear polarization (ODNP) and coherently averaged synchronized readout (CASR) techniques are employed to convert NMR signals in a microfluidic chip into optical signals, which are then imaged using wide-field diamond microscopy. Adapted and modified with permission from Ref. 376; © 2025 The Authors.
Although scanning probe-based MRI offers nanometer-scale resolution, its small imaging area and slow scanning process limit both the image size and acquisition time. This limitation can be overcome using wide-field quantum microscopy, which provides a larger field of view ( to ) and spatial resolution on the order of several hundred nanometers. Early experiments using ensemble NV centers with wide-field quantum microscopy demonstrated 2D MRI of nuclear spins in islands375 and fluorinated samples,49 achieving lateral resolutions of and , respectively. Recently, this technique has been integrated with microfluidic platforms.352 ,353 ,372 ,376 ,383 Microfluidic channels offer several advantages: biomolecular samples can be continuously supplied and maintained in a bio-friendly, in vivo environment with constant temperature and pH. In addition, samples can be polarized and detected at various points along the channels, providing flexibility in setup design and reducing the need for large-scale equipment.352 Furthermore, microfluidic systems enable precise control over sample flow, volume, biochemical reactions, and real-time monitoring.352 ,353 ,372 ,376 ,383 As a result, combining micro-MRI with microfluidic channels enhances the technique’s suitability for biomedical applications.
4.1.2.
Cell detections and virus diagnostics
NV centers are also utilized for the detection of individual cells and viruses, offering a novel diagnostic tool for the quantitative analysis of various biological samples. These samples range from naturally existing magnetic species, such as magnetic chains in magnetotactic bacteria162 and malarial hemozoin paramagnetic crystals,384 to magnetically labeled samples, including SKBR3 cancer cells,351 human lung tumor tissues,349 and mammalian cells and tissues.371
MNPs are commonly used as magnetic tags, which are incorporated into living cells and tissues to generate localized magnetic fields.349 ,351 ,371 Figure 21(a) illustrates the concept of magnetic tagging: MNPs are selectively bound to SKBR3 cancer cells, allowing magnetic imaging to differentiate between MNP-labeled cancer cells and nonlabeled healthy cells, which are not identifiable under optical microscopy.351 Figure 21(b) demonstrates immunomagnetic microscopy, a technique that quantitatively identifies the density of MNPs and reconstructs magnetic images using deep learning algorithms.349
Fig. 21
Cell detections and virus diagnostics with diamond NV centers. (a), (b) DC magnetic imaging of magnetic nanoparticles (MNPs). (a) SKBR3 cancer cells. Adapted and modified with permission from Ref. 351; © 2015 Springer Nature America, Inc. (b) Human lung tumor tissues. Adapted and modified with permission from Ref. 349; © 2022 The Authors. (c), (d) AC magnetic imaging of and ions. (c) At the nanometer scale using scanning probe microscopy. Adapted and modified with permission from Ref. 372; © 2024 The Authors. (d) At the micrometer scale using wide-field quantum microscopy. Adapted and modified with permission from Ref. 385; © 2017 The Authors. (e), (f) Diagnostic measurements using functionalized nanodiamonds. (e) SARS-CoV-2 viruses. Adapted and modified with permission from Ref. 353; © 2021 American Chemical Society. (f) HIV viruses. Adapted and modified with permission from Ref. 354; © 2020 The Authors.
Although MNPs serve as magnetic tags for DC field sensing, magnetic ions such as , , , and are employed as spin labels for AC field sensing.353 ,372 ,373 ,385 These ions are commonly used as contrast agents in magnetic resonance imaging or naturally occur in ferritin protein cells. Their ESR occurs at GHz frequencies, influencing the NV relaxation time. relaxometry measures this GHz magnetic noise from the ions by tracking changes in time. Imaging ions has been demonstrated at both nanometer and micrometer scales using scanning magnetometry and wide-field quantum microscopy. For instance, ferritin proteins in Hep G2 cells [Fig. 21(c)] and ions [Fig. 21(d)] have been imaged with spatial resolutions of and , respectively.372 ,385
Functionalized nanodiamonds have also been studied for rapid point-of-care diagnostic tests for contagious viruses, such as SARS-CoV-2 and HIV.353 ,354 Figure 21(e) illustrates the use of relaxometry to detect the SARS-CoV-2 virus with ions.353 In this setup, the surface of the nanodiamond is functionalized with -DNA linked to complex molecules. When viral RNA binds to the -DNA on the nanodiamond surface, the ions detach, reducing magnetic noise detected by the NV centers, which results in the recovery of the time. This method demonstrates sensitivity down to several hundred viral RNA copies, with a false-negative rate of less than 1%.353
In addition to relaxometry, a novel fluorescence detection method has been employed for in vitro biosensing.354 To separate the fluorescence signal from the NV center in nanodiamonds from background autofluorescence, an amplitude-modulated MW field is applied to the functionalized nanodiamonds, modulating the NV’s fluorescence emission, which is then detected by lock-in measurement. This technique provides a detection limit of mol for a biotin–avidin model and enables single-copy detection of HIV-1 RNA [Fig. 21(f)].354
4.1.3.
Toward clinical applications: MCG and MEG
Two potential clinical applications of NV-based magnetic sensors are magnetocardiography (MCG) and magnetoencephalography (MEG), which are used to detect cardiac and neural signals.346 ,347 ,386 ,387 MCG measures stray magnetic fields induced by electrical currents flowing through the heart, whereas MEG detects magnetic signals generated by action potentials in neural networks. MEG requires sensitivity at the level, but cardiac magnetic signals are typically several orders of magnitude stronger than neural signals, making them detectable by NV magnetic sensors. In addition, many heart diseases, such as atrial flutter and atrial fibrillation, are characterized by irregular current flow in the heart at the millimeter scale, necessitating miniaturized magnetic sensors. Diamond sensors based on ensemble NV centers can be compact () and offer sensitivities in the 1 to range, making them well suited for MCG applications in diagnosing cardiovascular diseases. Recent experiments with diamond magnetic sensors have successfully measured magnetic cardiac signals from anesthetized and thoracotomized rats. For instance, Fig. 22(a) demonstrates the detection of magnetic cardiac signals from a living rat with a sensitivity of using a 5 mm-sized sensor.347 The results of NV-based MCG are consistent with conventional electrocardiography (ECG) measurements, underscoring the potential of NV-based MCG sensors for clinical use.
Fig. 22
Magnetocardiography (MCG) and magnetoencephalography (MEG) sensors based on diamond NV centers. (a) Detection of magnetic cardiac signals from a living rat. Adapted and modified with permission from Ref. 347; © 2024 American Physical Society. (b) Measuring transient magnetic signals due to action potentials propagating through marine worm axons. Adapted and modified with permission from Ref. 346. (c) Simulated results of magnetic images in neural networks. Adapted and modified with permission from Ref. 387; © 2018 The Authors.
For MEG applications, NV centers still require further improvements in sensitivity. Early proof-of-principle experiments have demonstrated the ability to detect magnetic fields generated by ionic currents propagating through marine worm axons and mouse corpus callosum axons, as shown in Fig. 22(b). With a sensitivity range of 10 to and sub-microsecond temporal resolution, these experiments enable the measurement of magnetic field traces from transient action potentials occurring on the millisecond time scale.346
Recent theoretical calculations, shown in Fig. 22(c), suggest that neural network imaging will soon be possible with an improved wide-field quantum microscope setup. This setup could achieve an area-normalized sensitivity of , with spatial resolution and temporal resolution, averaged over to repetitions.387 However, single-shot imaging of neural activity still requires significant improvements in sensitivity by several orders of magnitude. This could be achieved by optimizing diamond treatment and enhancing .
4.2.
Semiconductor Device Inspection
The continuous advancement of semiconductor technology and the miniaturization of electronic components have led to the development of complex circuitries and sub-micrometer devices with high density and 3D compound structures. As these devices become more intricate and densely packed, there is an increasing demand for new methods to characterize device performance and detect manufacturing flaws at the microscopic level. With their high field sensitivity and spatial resolution, diamond NV centers provide powerful, noninvasive tools for investigating the magnetic and electronic properties of semiconductor devices. In this section, we explore recent efforts in semiconductor device inspection and diagnostics, focusing on applications such as individual bit detection in magnetic random access memory (MRAM) and current flow analysis in integrated circuit (IC) devices.356 –358
4.2.1.
Individual magnetic bits in MRAM
Diamond scanning magnetometry has proven to be an invaluable technique for characterizing MRAM devices at the nanometer scale.356 MRAM is a nonvolatile memory that stores information in magnetic bits. It relies on arrays of magnetic tunnel junctions (MTJs), which consist of two magnetic layers separated by a tunnel barrier. Magnetic information is stored based on the relative magnetization orientations of these layers: a parallel alignment represents "1," whereas an anti-parallel alignment represents "0." External magnetic fields can switch the magnetization direction of one of the layers, enabling data writing. As the MTJ pillars in MRAM devices are typically smaller than several tens of nanometers, analyzing device performance and identifying manufacturing flaws at the individual bit level requires both high field sensitivity and nanometer spatial resolution. Therefore, diamond scanning magnetometry serves as a novel tool for investigating MRAM devices.
For instance, Fig. 23 presents magnetic images of a spin transfer torque-MRAM (STT-MRAM) array consisting of bits.356 The NV center probes the distinct magnetic fields from the parallel and anti-parallel states of individual MTJ pillars, providing valuable insights into the switching behavior. These magnetic images offer a comprehensive overview of the bit switching process, allowing the characterization of key device properties such as magnetic characteristics, switching statistics, and uniformity across the magnetic bits. Figures 23(b) and 23(c) show the percentage of parallel bits as a function of the external magnetic field. By analyzing the bit-switching data, Borràs et al. were able to extract critical magnetic properties, including retention values () and anisotropy fields ().356 This novel nanoscale magnetic imaging technique holds significant potential for the MRAM industry, providing a method to characterize MRAM performance and screen device imperfections during the early stages of the manufacturing process.
Fig. 23
Nanometer-scale inspection of MRAM performance. (a) A magnetic image of STT-MRAM array of bits. Percentage of the parallel bits as a function of an external magnetic field (b) and corresponding magnetic images (c). Images are adapted and modified with permission from Ref. 356; © 2024 The Authors.
4.2.2.
IC devices
In addition to characterizing magnetic devices such as MRAM, NV centers can also be used to image current profiles in electronic devices by measuring the Oersted magnetic fields generated by currents in the DC to GHz frequency range. For instance, wide-field diamond microscopy has been successfully applied to image current flow in printed circuit board traces, MW devices,143 and photovoltaic devices.388 By measuring the Oersted stray field, a 2D current density map () can be reconstructed using the Biot–Savart law:
Eq. (5)
Recent research has also focused on expanding the capabilities of current flow imaging to more complex electronic circuits, such as ICs.357 ,358 Nanodiamond-based NV centers and wide-field quantum microscopy have been utilized to study IC microchips.357 ,358 ,389 For instance, nanodiamonds distributed over GaN HEMT junctions enable the simultaneous mapping of magnetic fields and temperature changes induced by current flows.389 On the other hand, wide-field quantum microscopy has been used to image magnetic fields generated by complex current paths in IC devices, facilitating the visualization of current profiles in multilayered devices and enabling the identification of malicious circuitry or Trojans.357
For example, Fig. 24(a) shows an experiment on a commercial FPGA designed to image current flow within ring oscillator clusters.357 These clusters, which consist of multiple ring oscillators, allow the control of current flow in different regions of the FPGA, switching between active and idle states. The magnetic field profiles measured by NV centers are overlaid on high-resolution optical images of the circuit die. Although most of the magnetic images align with the device’s designed structure, some features in the images reveal sub-surface structures or discontinuities in current flow. These hidden features, such as disconnected traces, are not easily detected in optical images, highlighting how magnetic imaging can offer a new approach to characterizing IC devices. In addition, quantitative analysis using machine learning techniques has been shown to identify the pre-programmed active states of the FPGA, paving the way for detecting hardware attacks such as malicious circuitry or Trojans and preventing counterfeiting during the manufacturing and distribution stages.
Fig. 24
Imaging current flows in integrated circuits (ICs). (a) Optical and magnetic images of an FPGA device. Magnetic images due to the current flows in the device are overlaid with optical images of the device. Adapted and modified with permission from Ref. 357; © 2020 American Physical Society. (b) 3D magnetic images due to the current flows in multilayered IC devices. Adapted and modified with permission from Ref. 358; © 2024 American Physical Society.
A key challenge in applying diamond microscopy at an industrial level is extending nearly 2D images into three dimensions. As the Oersted magnetic field induced by current flows in wires decays rapidly with distance (e.g., 1/distance), imaging sub-surface current flows at greater depths becomes challenging. Recent work, shown in Fig. 24(b), has demonstrated 3D current imaging up to in the vertical direction within multilayered IC devices.358 By modeling magnetic field profiles in three dimensions using the Biot–Savart law, Garsi et al.358 were able to isolate Oersted fields from different layers of the device. However, this method requires prior knowledge of the device’s structure and layout. Extending this technique to deeper layers of a chip will require further improvements in sensitivity and spatial resolution. In addition, the practical application of diamond microscopy in the semiconductor industry faces challenges due to the complex architectures of multilayer IC devices, which consist of billions of transistors, metal layers, and wire networks. This complexity makes image analysis challenging and necessitates further research to refine these techniques for more efficient and accurate device characterization.
4.3.
Industrial Applications
In addition to the high field sensitivity and broad dynamic range, diamond quantum sensors offer several other significant advantages for industrial applications. One of the key strengths is their ability to function across a wide range of temperature and pressure conditions, including room temperature and ambient environments. The host material, diamond, provides exceptional resilience to extreme temperatures, high radiation levels, and aggressive chemical conditions, ensuring the sensors remain durable and reliable even in harsh environments. Moreover, the sensor head, which houses a diamond plate with an ensemble of NV centers, can be integrated into compact, miniaturized sensor systems.11 ,12 ,359 ,360 ,390 –393 This is particularly beneficial in industries where space is limited or where multiple sensors need to be deployed in close proximity. This section highlights some of the pioneering efforts in the development of compact, portable diamond quantum sensors, as well as their emerging industrial applications, including magnetic detection of elevator motion, measurement of metal sheet thickness, real-time monitoring of current flows in electric vehicle (EV) batteries, and high-voltage current measurements in power network systems.193 ,361 –363
4.3.1.
Compact and portable sensors
In laboratory-scale experiments, diamond magnetometers based on ensembles of NV centers have demonstrated magnetic sensitivity, reaching as low as .14 However, there has been a growing effort to extend these advancements beyond the laboratory environment and develop compact magnetometers suitable for real-world industrial applications. Achieving this requires integrating essential sensor components, such as light sources, MW circuits, detectors, and the diamond sensor head, while ensuring that the high sensitivity observed in controlled lab environments is maintained.
Figure 25 showcases recent advancements in miniaturized diamond magnetometers, typically ranging in size from a few to several tens of .11 ,359 ,360 ,390 ,392 –394 The diamond plate itself is small, typically in the millimeter scale, and contains NV centers with densities ranging from to . The diamond plate is then integrated with various commercial or custom-designed components, such as LEDs, MW circuit boards, optical elements (e.g., lenses and filters), and photodetectors.11 ,360 ,392 ,393 To further reduce the sensor’s overall size and minimize potential spurious coupling between MW sources and the detector, some designs use external lasers and photodetectors connected to the sensor head via optical fibers.12 ,359 ,390 ,391 ,394 The reported sensitivities from these compact sensors range from a few hundred to , representing a significant step toward realizing practical, portable diamond magnetometers for industrial applications.
Fig. 25
Examples of miniaturized diamond quantum magnetometers with various configurations. (a)–(c) Photodetector-integrated designs using optical fibers for light delivery. (a) Adapted and modified with permission from Ref. 354; © 2020 The Authors. (b) Adapted and modified with permission from Ref. 11; © 2019 The Authors. (c) Adapted and modified with permission from Ref. 394; © 2020 IOP Publishing Ltd. (d), (f) Designs with both light source and detector integrated into the sensor head. (d) Adapted and modified with permission from Ref. 392; © 2024 The Authors. (e) Adapted and modified with permission from Ref. 360; © 2022 Elsevier B.V. (f) Adapted and modified with permission from Ref. 393; © 2022 IEEE. (g) A fiber-coupled design in which both the light source and detector are located outside the sensor head. Adapted and modified with permission from Ref. 390; © 2020 The Authors.
Figure 26 illustrates examples of portable diamond sensors for industrial applications, such as detecting elevator movement and measuring the thickness of steel sheets. The diamond sensor shown in Fig. 26(a) demonstrates real-time detection of magnetic fields induced by the movement of an elevator.361 The sensor successfully monitors changes in magnetic fields as the elevator approaches the second level of the building where the sensor is located. In addition, the sensor detects the magnetic field change in real time when the elevator doors open. On the other hand, Fig. 26(b) presents a fiber-integrated hBN sensor used to measure the thickness of metal sheets by assessing the amount of magnetic field shielding caused by the sheet.193 This compact fiber-based sensor is realized by integrating an hBN circular Bragg grating cavity, which contains spin defects, with an optical fiber. In this experiment, the steel sheet acts as a magnetic shield, and the sensor measures the suppression of the external magnetic field with varying sheet thicknesses, ranging from 0 to 9 mm.
Fig. 26
Examples of industrial applications using compact, portable defects-based magnetometers. (a) Real-time detection of magnetic fields induced by the movement of an elevator. Adapted and modified with permission from Ref. 361; © 2024 The Authors. (b) Measurements of the thickness of metal sheets by monitoring the amount of field shielding by the sheets. Adapted and modified with permission from Ref. 193; © 2024 The Authors.
For broader industrial applications of portable magnetometers, further development is necessary to make the entire sensor system even more compact, including the integration of external controllers and data analysis units, while maintaining high magnetic field sensitivity over a wide dynamic range. These advancements are crucial to enhancing the versatility and usability of portable sensors in diverse industrial environments.
4.3.2.
Battery monitor in electrical vehicles
Compact diamond sensors are also applied for monitoring EV battery systems. As the electric automobile market grows and concerns about battery safety become more prominent, accurate current sensors are critical for assessing the state of charge (SOC) of EV batteries. These sensors help predict remaining battery lifetime and drive range, and identify potential hazards in battery operations.
Figure 27 demonstrates how compact diamond sensors are used to monitor current flows in the busbar of an EV battery module.362 ,395 These sensors can detect magnetic fields generated by a busbar current of 20 mA, even when external magnetic noise is as high as . The system employs two separate diamond sensors placed on opposite sides of the battery busbar, measuring magnetic fields from opposing currents. By performing a differential measurement between the two sensors, the system effectively cancels out external magnetic noise, which is common in real driving situations. This approach significantly improves the accuracy of SOC estimation, suggesting that the EV’s driving range could be increased by 10%.362 ,395 Despite these advancements, further research and optimization are necessary to improve the sensor’s versatility for different charging situations. With ongoing development, these sensors will be better equipped to monitor EV battery performance reliably under real driving conditions.
Fig. 27
Battery monitoring in electric vehicles (EVs) using compact diamond sensors. (a), (b) Schematics of compact diamond sensors mounted in the busbar of an EV battery module. (a) Adapted and modified with permission from Ref. 395; © 2022 The Authors. (b) Adapted and modified with permission from Ref. 362; © 2023 The Authors. (c) Busbar current and differential measurements of magnetic fields detected by two diamond sensors. Adapted and modified with permission from Ref. 362; © 2023 The Authors.
4.3.3.
Harsh environment: high-voltage grids in power network systems
In high-voltage power grid systems, accurate current measurement is critical for minimizing energy losses and preventing hazards such as overheating of cables and equipment. As power network systems require current sensors capable of measuring large currents (several kA) within confined spaces, diamond sensors present a significant solution for monitoring currents in these environments.363
Figure 28 illustrates an example of how diamond sensors can be implemented in electrical power systems. Liu et al.363 proposed a design that separates the diamond sensor head from the control units for practical reasons, such as limited space near the high-voltage side and challenges associated with delivering MW signals through coaxial cables in high-voltage areas. The compact diamond sensor is placed on the high-voltage side, whereas the control units (including laser, MW circuits, detectors, and power supplies) are positioned on the low-voltage side. Optical excitation and readout of the NV centers are achieved using optical fibers, and MW drives are transmitted with directional antennas. In proof-of-principle experiments, this remote sensor system, positioned 11 m apart, successfully measured currents with high accuracy, showing only a 0.4% error over 1000 A.363 This demonstrates the potential of diamond sensors for high-voltage power grid applications. However, for practical implementation in the power grid, further improvements are needed to enhance the sensor’s sensitivity and the efficiency of the directional antenna system.
Fig. 28
Current sensors in the high-voltage grids of power network systems. (a) Schematic of diamond sensors for high-voltage grids in power network systems. (b) Photographs of remote diamond sensor system. Directional antennas are used to transmit microwave drive for the sensor. (c) Current measurements with controlling the cable current. Images are adapted and modified with permission from Ref. 363; © 2024 The Authors.
4.4.
Military and Navigation Applications
Diamond sensors are becoming increasingly valuable in the aerospace and military sectors due to their unique capabilities in providing self-reliant navigation systems that operate without dependence on Global Navigation Satellite Systems (GNSS) such as GPS.29 ,65 ,364 –368 These systems are crucial in environments where GNSS signals are unavailable, unreliable, or deliberately compromised, such as underground tunnels, underwater, or during wartime. Navigation sensors, or inertial measurement units (IMUs), typically combine accelerometers, gyroscopes, and magnetometers to track motion, orientation, and position relative to Earth’s magnetic field. Diamond NV centers can serve as a quantum platform for these applications, offering extraordinary sensitivity to a wide range of physical properties, including magnetic fields, temperature, inertial, and rotational motion. The exceptional precision of these sensors allows for the detection of minute variations in the environment, such as slight changes in magnetic fields or tiny shifts in rotational motion. Therefore, diamond sensors can also be applied in mine detection, submarine navigation, and other security-related applications.
This section highlights recent advancements in the integration of diamond quantum sensors into aerospace and military technologies. These innovations encompass inertial sensors that measure time-varying magnetic or strain fields induced by mechanical motion, gyroscopes capable of detecting rotational pseudo-magnetic fields generated by rotating nuclear spins, and magnetometers used for mapping magnetic fields in complex outdoor environments or detecting hidden magnetic objects.
4.4.1.
Inertial sensors
Inertial sensing involves the measurement of force, acceleration, or displacement. Two primary methods for sensing inertial motion with diamond NV centers are detecting external magnetic fields or internal strain fields induced by the motion of mechanical oscillators. For example, the oscillatory motion of a magnetized AFM tip generates a time-varying magnetic field.396 This magnetic field drives spin transitions in the NV centers, which can be detected using NV’s AC sensing protocols. In this approach, both driven and Brownian motions of the AFM resonator have been detected with a precision of less than 6 picometers under ambient room temperature conditions.396
On the other hand, NV centers are also sensitive to strain fields through spin-lattice interaction. For instance, Ovartchaiyapong et al.31 demonstrated strain-mediated coupling between a diamond cantilever and an embedded NV center, quantitatively characterizing both axial and transverse strain sensitivities. With a strain sensitivity of , they were able to detect motion within a 1-s measurement. These measurements highlight the potential of inertial sensing, whether through spin coupling to time-varying external magnetic fields or by detecting internal strain fields.
Figure 29(a) illustrates a concept for force sensing using diamond nanopillars with NV centers positioned at two different locations: one at the tip of the nanopillar and the other at the base.29 The NV center at the tip measures the spin-magnetic response caused by the nanopillar’s motion under a strong transverse magnetic field gradient, whereas the NV center at the base measures the stress and strain induced by the same mechanical motion. Each method offers distinct advantages and limitations. Magnetic sensitivity is significantly higher than strain sensitivity as evidenced by the measured displacements in the picometer and nanometer range.31 However, calculations from Ref. 29 suggest that spin-magnetic coupling requires large field gradients (e.g., ), presenting practical challenges, such as the need to position nanoferromagnetic structures close to the tip, which limits the range of mechanical motion. Consequently, both methods may need to be employed together to obtain complementary inertial data, depending on the scale of the motion.
Fig. 29
Inertial sensors based on diamond NV centers. (a) Schematic of a diamond nanopillar with two NV centers located at the tip and base of the nanopillar. The former NV is used to detect inertial force via magnetic field gradient, while the latter probes strain field induced by the internal force. (b) Schematic of inertial force sensor using an array of diamond nanopillars. Images are adapted and modified with permission from Ref. 29; © 2017 American Chemical Society.
Figure 29(b) presents a concept for force sensing based on arrays of diamond nanopillars, capable of detecting forces induced by biological cells through spin–strain coupling.29 This approach can also be adapted for navigation applications. Recent advancements in diamond nanofabrication have enabled the development of various diamond nanomechanical structures with high mechanical quality factors. By combining nanofabrication techniques with advanced sensing protocols, the inertial sensitivity of these devices can be significantly enhanced.
4.4.2.
Rotation sensors: gyroscope
Alongside inertial sensing, rotation sensing or gyroscopes are essential components for navigation and IMU systems. The fundamental mechanisms of rotation sensing based on NV centers can be categorized into two main approaches: directly monitoring the rotational motion of the NV electron spin or sensing the effective magnetic field generated by rotating nuclear spins.64 –67 ,397 For instance, Fig. 30(a) demonstrates stroboscopic imaging of NV centers in a rotating diamond, where the laser and MW fields are relatively stationary.367 The fluorescence from a rotating NV center is collected by applying synchronized laser and MW pulses that match the rotational period of a spindle driven by an electric motor. This technique enables the detection of rapid rotations, e.g., 200,000 r/min, using a single NV center. Further experiments show that the phase shift of the NV center, induced by the relative motion between a fixed external MW field and the rotating NV spin, can be precisely measured.397 The external MW field, in combination with the rotating NV center, generates geometric or Berry phases in the NV’s Bloch sphere. The accumulated phase can then be detected using spin Hahn echo protocols.
Fig. 30
Gyroscopes using electron and nuclear spins in diamond. (a) Stroboscopic imaging of NV centers in a rotating diamond. Adapted and modified with permission from Ref. 367; © 2018 The Authors. (b) Rotation detection with NV centers using rotational pseudo-magnetic fields generated by rotating nuclear spins. Adapted and modified with permission from Ref. 65; © 2021 The Authors.
Another approach leverages nuclear spins near the NV center. This method exploits nuclear spins as effective generators of magnetic fields due to rotation, known as rotational pseudo-fields.368 The NV center serves as a sensing qubit to probe these magnetic fields using AC sensing protocols. In a rotating frame, a precessing nuclear spin generates an effective magnetic field, , which is determined by the precession frequency, , and the spin gyromagnetic ratio, , such that . As the gyromagnetic ratio of nuclear spins is times smaller than that of electron spins, nuclear spins can generate significantly larger rotational pseudo-fields.368 The and nuclear spins associated with NV centers are utilized for nuclear spin gyroscopes, with an example system illustrated in Fig. 30(b).65 In this system, both the diamond sensor and control components (including laser, MW source, and detector) are mounted on the spindle and rotate together. Double quantum (DQ) pulse protocols, such as DQ Ramsey and DQ echo, are employed to monitor nuclear spin precession while minimizing the effects of temperature fluctuations. Recent experiments have demonstrated sensitivities in the range of 1 to , with the best sensitivity being and a bias stability of .65
4.4.3.
Mapping magnetic fields and detecting hidden targets
One important military application of diamond quantum sensors is the mapping of Earth’s magnetic fields, which is vital for navigation, geophysical surveillance, and detecting concealed targets.364 Mapping the Earth’s magnetic field or identifying magnetic anomalies caused by underground magnetic compositions or hidden objects holds significant potential for military use. Miniaturized, portable sensors with high magnetic field sensitivity can be integrated into unmanned aerial vehicles for remote detection of magnetic fields from both the Earth’s surface and hidden magnetic objects. This provides a highly effective tool for surveillance and reconnaissance.
Figure 31 illustrates examples of using diamond magnetometers to map magnetic fields from road surfaces and detect concealed magnetic objects.365 ,366 The sensor systems are mounted on a vehicle to probe changes in magnetic fields in both indoor and outdoor environments [Fig. 31(a)].365 In outdoor environments, several factors contribute to the total magnetic field, including Earth’s magnetic field, as well as permanent or induced fields from the vehicle and its control units. Using two magnetometers fixed to the vehicle and measuring the vector magnetic fields along 3D coordinates, Graham et al. effectively canceled out the permanent magnetic fields from the vehicle itself, enabling them to measure variations in the Earth’s core magnetic field along the vehicle’s track.365
Fig. 31
Mapping magnetic fields from road surfaces and concealed objects. (a) Mapping magnetic fields from road surfaces using diamond sensors mounted in a vehicle. Adapted and modified with permission from Ref. 365; © 2025 Elsevier B.V. (b) Detecting hidden magnetic objects from magnetic images measured by miniaturized diamond sensors mounted in a 2D scanning stage. Adapted and modified with permission from Ref. 366; © 2025 The Authors.
On the other hand, miniaturized diamond sensors, integrated with a 2D scanning stage, are used to image stray magnetic fields from various sizes of magnets embedded in a toy diorama, as shown in Fig. 31(b).366 This setup emulates a scenario in which remote detection of landmines on a battlefield or concealed objects at construction sites is required. By comparing the measured data with simulations, Choi et al.366 were able to accurately estimate the location and size of the magnets. In addition, vector magnetometry was proposed to compensate for any relative tilt between the sensor and the target, reducing the uncertainty in determining the target’s location from several centimeters to tenths of millimeters.366
5.
Conclusion
In this review, we provide a comprehensive overview of quantum sensing based on solid-state spin defects, covering the fundamental principles as well as recent progress toward realizing practical use cases. We discussed both basic and advanced sensing protocols, imaging applications at two different length scales, the integration with optical devices to enhance optical properties, and ongoing efforts toward practical applications for real-world use.
Solid-state spin defects have emerged as a promising quantum platform for enhanced metrology tools in various fields, offering several advantages, including sensitivity, robustness, and compactness. The unique ability to engineer spin defects significantly enhances their performance and versatility. The challenges posed by spin defects from the surrounding solid-state environment can be tailored through advanced quantum control techniques that decouple environmental noises from spin defects, as well as by interfacing with micro- and nano-photonic structures that improve optical properties. These efforts to optimize sensitivity, coherence time, and integration with optics further extend the utility of spin defects as quantum sensors.
Although natural atoms or superconductor-based sensors also provide high quantum sensing performances, the ability of spin defects to operate under ambient conditions at room temperature offers an incomparable level of control and scalability for practical sensor applications. Along with NV centers in diamond, which lead defect-based quantum sensing research, alternative materials such as SiC, hBN, GaN, and Si show great promise in expanding quantum sensing capabilities. These materials open new opportunities for applications in CMOS, bio-compatible platforms, and integrated photonic circuits. However, further exploration is necessary to fully characterize their spin properties and optimize quantum coherence to unlock their full potential.
Efforts to advance the practical implementation of quantum sensing in industrial, military, and biomedical applications continue to improve sensing performance, device robustness, and real-time data processing techniques. In the future, the development of miniaturized and portable quantum sensors, fiber-based devices, and chip-integrated systems will be crucial in bridging the gap between experimental research and field-ready systems, facilitated by close interdisciplinary collaboration.395 ,398
Given their maturity and near-term applicability, quantum sensors based on spin defects are expected to be the first widely adopted quantum technology, benefiting our daily lives before large-scale quantum computing and communication technologies arrive.