Spin waves, the fundamental excitations in magnetic materials, are promising candidates for realizing low-dissipation information processing in spintronics. The ability to visualize and manipulate coherent spin-wave transport is crucial for the development of spin wave-based devices. We use a recently discovered method utilizing nitrogen vacancy (NV) centers, point defects in the diamond lattice, to measure spin waves in thin film magnetic insulators by detecting their magnetic stray field. We experimentally demonstrate enhanced contrast in the detected wavefront amplitudes by imaging spin waves underneath a reference stripline and phenomenologically model the results. By extracting the spin wave dispersion and comparing NV center based spin wave measurements to spin wave imaging conducted through the well-established time-resolved magneto-optical Kerr effect, we discuss the advantages and limitations of employing NV centers as spin wave sensors.

Spin waves represent collective excitations of the spins in a magnetic material.1–3 Their quanta are referred to as magnons. In insulating magnetic materials, these quasiparticles carry spin angular momentum and propagate through the material’s magnetic lattice, while avoiding heating associated with charge currents. Their unique properties, including long coherence times and low dissipation,4–8 as well as frequencies in the giga- to terahertz regime and wavelengths as small as several nanometers,9–11 render spin waves promising candidates for various applications, such as interference-based, ultrafast, nanoscale magnonic logic circuits, and miniaturized device technologies,6,8,12–17 particularly in the field of spintronics.12,18–20

Imaging spin waves with high spatial and temporal resolution is essential for exploiting their potential in technological advancements and understanding their intricate dynamics. Traditional imaging techniques, such as time-resolved magneto-optical Kerr effect (TR-MOKE) microscopy,21–26 Brillouin light scattering,27,28 and transmission x-ray microscopy,29,30 have provided valuable insights into spin wave behavior. However, these methods often have limitations in terms of spatial resolution and sensitivity, particularly when studying spin waves in nanoscale systems or beneath opaque materials, or they need large and expensive facilities for their implementation, as in the case of x-ray microscopy.

In recent years, the emergence of nitrogen-vacancy (NV) centers in diamond as sensitive magnetic field sensors has opened new avenues for spin wave imaging.31–34 NV centers are point-like defects in the diamond lattice, consisting of a substitutional nitrogen atom adjacent to a vacancy. They exhibit remarkable properties, including high sensitivity to static and fluctuating magnetic fields,35–38 nanoscale spatial resolution,39,40 even below opaque materials,41 and optical addressability.42 By using an ensemble of spins instead of a single NV center, the sensitivity can be further enhanced by a scaling of N, where N is the number of NV centers.32 However, this enhancement comes at the cost of reduced contrast, which can counteract the improvement in sensitivity to some extent.

In this work, we discuss the use of an ensemble of shallowly implanted NV centers to measure spin waves in a magnetic thin film, in this case the low damping ferrimagnetic insulator yttrium iron garnet (YIG),7,9 following the seminal work reported by Refs. 40 and 43. We demonstrate how the contrast in the imaged spin wave amplitude can be enhanced, allowing us to image spin waves over a long propagation distance. Furthermore, we phenomenologically model these measurements to explain the observed features in the NV spin wave images. By comparing the spin wave measurements obtained through NV center detection to spin wave imaging acquired with conventional TR-MOKE measurements, we discuss the advantages and limitations of utilizing NV centers for spin wave imaging, as well as the potential of integrating NV center measurements with TR-MOKE imaging into a unified setup.

We excite spin waves in a 200 nm thick YIG film grown by liquid phase epitaxy on a gadolinium gallium garnet substrate by a microwave current sent through a stripline (S1) fabricated onto the YIG surface. The diamond chip is placed on the YIG sample with the NV layer facing toward the magnetic surface, as shown in Fig. 1(a). Additional information on the diamond chip and the sample can be found in  Appendix A. As the spin waves propagate rightward (leftward), they generate a circularly polarized field with handedness that drives the ω(ω+) electron spin resonance (ESR) transition of the NVs, resulting in a spin-dependent photoluminescence (PL) signal.31,40,44 In Fig. 1(b), the NV photoluminescence signal, obtained via lock-in detection and measured ∼10 µm away from the center of S1, is shown in the dependence of the external magnetic field Bext and the microwave current frequency. The NV setup is described in detail in  Appendix B. Here, Bext is applied along S1 at an angle of ϕ = 35° relative to the sample plane, aligning it with one of the four possible orientations of the NV axes. The peaks in the PL signal belong to the ESR transitions in the ground (ω±, ω) and excited states (ωex), where only the PL signal of NV centers aligned with Bext results in a linear dependence on Bext. As the NVs are not only driven by the microwave field of S1 but also by the spin wave stray field Bsw, the ω transitions are enhanced for frequencies above the ferromagnetic resonance (FMR) limit. Due to the handedness of the circularly polarized spin wave stray field, the transitions involving ω (corresponding to spin waves with counterclockwise polarization) are preferentially enhanced when the excitation frequency exceeds the ferromagnetic resonance (FMR) frequency. This occurs because the circular polarization of the stray field aligns with the ω transitions, leading to a stronger coupling in this frequency range.

FIG. 1.

(a) Sample configuration: A diamond chip containing a layer of NVs is positioned on a 200 nm thick YIG film grown on a gadolinium gallium garnet substrate. Stripline S1 excites spin waves in the sample. Stripline S2, perpendicular to S1 and separated from S1 by a SiO2 layer, is used to generate an Oersted field Boe that oscillates at the same frequency as the spin wave stray field. An external magnetic field Bext is applied along S1 at an angle of ϕ = 35° relative to the sample plane. (b) Photoluminescence spectrum of the NV centers in the dependence of the external magnetic field Bext and the driving frequency f. Transitions in the ground (ω±, ω) and excited states (ωex) correspond to a peak in the signal. Due to the handedness of the circularly polarized spin wave stray field, the ω transitions are enhanced for frequencies above the ferromagnetic resonance (FMR) limit.

FIG. 1.

(a) Sample configuration: A diamond chip containing a layer of NVs is positioned on a 200 nm thick YIG film grown on a gadolinium gallium garnet substrate. Stripline S1 excites spin waves in the sample. Stripline S2, perpendicular to S1 and separated from S1 by a SiO2 layer, is used to generate an Oersted field Boe that oscillates at the same frequency as the spin wave stray field. An external magnetic field Bext is applied along S1 at an angle of ϕ = 35° relative to the sample plane. (b) Photoluminescence spectrum of the NV centers in the dependence of the external magnetic field Bext and the driving frequency f. Transitions in the ground (ω±, ω) and excited states (ωex) correspond to a peak in the signal. Due to the handedness of the circularly polarized spin wave stray field, the ω transitions are enhanced for frequencies above the ferromagnetic resonance (FMR) limit.

Close modal
The imaging resolution achieved with an ensemble of NV centers is ultimately limited by diffraction. In addition to diffraction, the distance z0 between the NV layer and the sample surface also imposes a constraint on the resolution.37 Furthermore, this parameter is needed to calculate the spin-wave amplitude from the detected field.31 To determine z0, we reconstruct the magnetic field generated by a DC current sent through S2. Simultaneously, an AC current is applied to S2 using a bias tee to detect the NV transitions. To lift the degeneracy of the NV transitions, we apply an external magnetic field Bext at a polar angle θ relative to the z-axis and an azimuthal angle ϕ relative to the x-axis, as indicated in Fig. 2(a). The NV1 and NV2 axes lie on the surface of a cone with an angle of θNV = 54.7° with respect to the z-axis, as indicated by the gray points in Fig. 2(a). Furthermore, NV1 lies in the xz-plane. The NV3 and NV4 axes, on the other hand, lie on the surface of a cone with an angle of θNV with respect to the negative z-axis, indicated by the gray squares in Fig. 2(a). In this coordinate system, the NV directions are given by
(1)
Furthermore, the magnetic field vector Bext is given by
(2)
FIG. 2.

(a) Illustration of the external field direction Bext and the NV center orientations NV1, NV2, NV3, and NV4. Bext is applied at a polar angle θ = (54.75 ± 0.04)° relative to the z-axis and an azimuthal angle ϕ = (14.77 ± 0.07)° relative to the x-axis. NV1 and NV2 lie on the surface of a cone with an angle of θNV = 54.7°, indicated by the gray points, while NV3 and NV4 lie on the surface of a cone with an angle of θNV with respect to the negative z-axis, indicated by the gray squares. (b) Optically detected magnetic resonance (ODMR) spectrum for this magnetic field direction. Fitting the data with eight Lorentzian functions (red line) yields the resonance frequencies of each NV family NVi, from which the external field magnitude Bmag = (14.81 ± 0.02) mT is obtained. (c) ODMR spectra plotted against the distance from the reference stripline S2, when a DC current of IDC = 80 mA is sent through S2. The total field sensed by the NVs is Btot = Bext + BDC. (d) Reconstructed DC magnetic field BDC. The y-component of the reconstructed DC field shows a constant offset of yoff = (0.25 ± 0.01) mT (gray dots). To fit the data, this offset was subtracted, resulting in the green dots. From the fit, the NV–S2 distance z0 = (1.4 ± 0.1) µm is obtained.

FIG. 2.

(a) Illustration of the external field direction Bext and the NV center orientations NV1, NV2, NV3, and NV4. Bext is applied at a polar angle θ = (54.75 ± 0.04)° relative to the z-axis and an azimuthal angle ϕ = (14.77 ± 0.07)° relative to the x-axis. NV1 and NV2 lie on the surface of a cone with an angle of θNV = 54.7°, indicated by the gray points, while NV3 and NV4 lie on the surface of a cone with an angle of θNV with respect to the negative z-axis, indicated by the gray squares. (b) Optically detected magnetic resonance (ODMR) spectrum for this magnetic field direction. Fitting the data with eight Lorentzian functions (red line) yields the resonance frequencies of each NV family NVi, from which the external field magnitude Bmag = (14.81 ± 0.02) mT is obtained. (c) ODMR spectra plotted against the distance from the reference stripline S2, when a DC current of IDC = 80 mA is sent through S2. The total field sensed by the NVs is Btot = Bext + BDC. (d) Reconstructed DC magnetic field BDC. The y-component of the reconstructed DC field shows a constant offset of yoff = (0.25 ± 0.01) mT (gray dots). To fit the data, this offset was subtracted, resulting in the green dots. From the fit, the NV–S2 distance z0 = (1.4 ± 0.1) µm is obtained.

Close modal

The optically detected magnetic resonance (ODMR) spectrum for this orientation of the external magnetic is shown in Fig. 2(b). By fitting the PL signal with eight Lorentzian functions, we obtain the resonance frequencies for each NV family. Using these values, we calculate the measured parallel Bi,m and perpendicular Bi,m projections of the external magnetic field along each NV axis using Eqs. (C1) and (C2) in  Appendix C. The magnitude of the external field vector is then obtained by Bmag,i=Bi,m2+Bi,m2. By averaging these values, we find Bmag = (14.81 ± 0.02) mT.

The theoretical parallel (Bexti,t) and perpendicular (Bexti,t) projections of Bext on the four NV axes are given by Eq. (C3) in  Appendix C. To determine the polar and azimuthal angles θ and ϕ at which the external field is applied, we minimize the error ΔBext between the measured and the theoretical parallel and perpendicular field components given by
(3)
with
(4)
This results in θ = (54.75 ± 0.04)° and ϕ = (14.77 ± 0.07)°. Using these values, we find Bext = (11.7, 3.1, 8.5) mT.

In addition to the external field Bext, we then apply a DC current of IDC = 80 mA through S2, which generates a magnetic DC-field that depends on the NV–S2 distance z0. The total magnetic field sensed by the NV centers is Btot = Bext + BDC, leading to spatially dependent shifts in the ODMR frequencies. By scanning the sample along the x-axis, we detect and fit the ODMR spectrum at each position x, as shown in Fig. 2(c), where the center of S2 is located at x ≈ 40 µm. Directly above S2, the AC field is strongest, resulting in an enhanced contrast in the ODMR spectrum. From the fits of the measured ODMR spectrum at each position x, we calculate the projections of the total field on the NV axes. The angles θ(x) and ϕ(x) are calculated by minimizing the error between the measured total field amplitudes and the reconstructed total field amplitudes. Thereby, we gain the total field vectors Btot(x).

The position dependent magnetic DC field is finally obtained by BDC(x) = Btot(x) − Bext. In Fig. 2(d), the x-, y-, and z-components of the obtained DC magnetic field (blue, gray, and red dots) are plotted against the coordinate x. The y-component (gray dots) shows an offset yoff = (0.25 ± 0.01) mT. The green dots show the y-component, when this offset was subtracted. It is possible that the magnet generating the external field has drifted slightly over time, resulting in this field offset. The deviation of By from zero in the vicinity of the edges of S2 results from the error of the fitted resonance frequencies in this region. To determine z0, we subtract the offset from the y-component of the DC field and fit the DC-field components using Eq. (D3) from the supplements. This yields an NV–S2 distance of z0 = (1.4 ± 0.1) µm. Additional fitting parameters are listed in  Appendix C.

To obtain the phase sensitivity to image individual wavefronts of the spin waves, a second perpendicular reference stripline, S2, isolated from S1 by an insulating SiO2 layer, is used, as shown in Fig. 1(a). A microwave current that generates an Oersted field Boe oscillating at the same frequency as Bsw is sent through S2. The superposition of Boe with the spin wave stray field Bsw results in a standing wave pattern in the total magnetic field that drives the NV ESR with a spatial periodicity equal to the spin wave wavelength.31 In Fig. 3(a), the PL signal for a spatial scan over the sample is shown. The x-axis, aligned with S1, is located ∼10 µm away from the center of S1. Furthermore, we scanned over S2, which is aligned with the y-axis, with its center located at x = 40 µm. We find an enhanced contrast in the imaged spin wave amplitude directly above S2. In addition, we observe a 180° phase shift of the wavefronts at the edges of S2 and a light bending of the wavefronts toward the center of S2.

FIG. 3.

(a) Spatially resolved PL signal measured above the YIG film, when a spin wave with frequency fsw = 2.24 GHz, corresponding to the external field B = 22.5 mT, is excited by S1. S1 is located at y = −10 µm, and the center of S2 is located at x = 40 µm. Bext is applied parallel to the x-axis at an angle of ϕ = 35° relative to the sample plane. The microwave current is split between S1 and S2 to generate an additional Oersted field Boe oscillating at fsw, which results, together with the stray field of the spin wave, in a standing wave of the total magnetic field above the YIG film. (b) Simulated PL signal corresponding to the same spatial position as in (a). The phase shift of the simulated wavefronts at the edges of S2 occurs due to the change in sign of Boe from the left side to the right side of S2. (c) Normalized TR-MOKE signal of a spin wave with frequency fsw = 2.24 GHz, excited by S1. The external field is applied parallel to the x-axis.

FIG. 3.

(a) Spatially resolved PL signal measured above the YIG film, when a spin wave with frequency fsw = 2.24 GHz, corresponding to the external field B = 22.5 mT, is excited by S1. S1 is located at y = −10 µm, and the center of S2 is located at x = 40 µm. Bext is applied parallel to the x-axis at an angle of ϕ = 35° relative to the sample plane. The microwave current is split between S1 and S2 to generate an additional Oersted field Boe oscillating at fsw, which results, together with the stray field of the spin wave, in a standing wave of the total magnetic field above the YIG film. (b) Simulated PL signal corresponding to the same spatial position as in (a). The phase shift of the simulated wavefronts at the edges of S2 occurs due to the change in sign of Boe from the left side to the right side of S2. (c) Normalized TR-MOKE signal of a spin wave with frequency fsw = 2.24 GHz, excited by S1. The external field is applied parallel to the x-axis.

Close modal
To explain these characteristics, we phenomenologically model our PL signal by calculating the time-averaged magnitude of the perpendicular time-dependent magnetic field Btot relative to the NV axis. Thereby, we take into account that ESR transitions can only be driven by rf-fields perpendicular to their quantization axis. Furthermore, we average over time, as we measure a time independent PL signal at a fixed position on the sample, which scales with the magnitude of the standing wave in the total perpendicular magnetic field. We use a coordinate system where the x-axis is aligned with S1, the y-axis is aligned with S2, and the z-axis is aligned with the normal vector of the sample plane, consistent with Fig. 3. In Btot, we consider the spin wave stray field above the film Bsw as well as the Oersted field Boe of S2,
(5)
The time-dependent magnetic fields Bsw(t) and Boe(t), as well as further details about the simulation, are given in  Appendix D. In Fig. 3(b), the simulated PL signal corresponding to the same spatial position on the sample as the measurement depicted in Fig. 3(a) is shown. The observed phase shift at the edges of S2 is attributed to the change in sign of Boe from the left side to the right side of S2. Moreover, we find an enhanced contrast in the PL signal above S2 due to the comparable magnitudes of Bsw(t) and Boe(t) in the vicinity of S2. However, reproducing the bending of the wavefronts toward the center of S2 requires a more advanced model, which is beyond the scope of this study.

In the following, we compare our NV measurements to TR-MOKE measurements in the Damon–Eshbach (DE) configuration,45 where the external field is applied in the YIG thin film plane along S1. TR-MOKE measurements involve synchronizing the rf excitation and optical probing pulses. A fixed phase relationship between excitation and probing is established to capture a stationary rastered image of the dynamic out-of-plane component. To achieve this synchronization, typically a pulsed Ti:Sa-laser with a repetition rate frep = 80 MHz is employed, while the excitation frequency fex of the spin waves fulfills fex = n·frep, with n being an integer. However, for a direct comparison between NV and TR-MOKE imaging, this frequency limitation presents a disadvantage. NV centers permit the detection of spin waves within a narrow frequency range, limiting the number of comparative images that can be captured. To overcome this constraint in TR-MOKE, we employ super-Nyquist sampling MOKE.46 In that case, we can tune the rf excitation to any intermediate frequency fex = n·frep + ɛ, where ɛ is a rational number. Demodulating the Kerr signal at the frequency ɛ enables direct extraction of the real and imaginary components of the magnetic rf-susceptibility, thereby providing phase-resolved measurements of the spin precession.46 See  Appendix B for further information concerning the MOKE setup. The lock-in detected Kerr signal SKerr is then normalized to the topographic signal STopo to gain an enhanced contrast in the measured spin wavefronts. Figure 3(c) shows a TR-MOKE measurement of propagating spin waves. As it is not possible to image spin waves propagating beneath S2 with TR-MOKE, the measurement is taken far away from S2. In  Appendix F, we additionally show spin wave measurements with different wavelengths.

To extract the spin wave dispersion from both measurement techniques, we adjust the external static magnetic field Bext such that the spin wave frequency coincides with the ω NV ESR frequencies. The mathematical equations to calculate the dispersion are given in  Appendix E. This allows us to probe spin waves with varying wavelength, detectable by both measurement techniques. In the NV measurements, the field is applied along S1 and one of the NV quantization axes. The NV measurements were only taken above 23 mT because the external magnetic field was applied using a permanent magnet mounted on a linear translation stage. Due to the limited movement range of the stage, fields below 23 mT could not be achieved without employing a different magnet setup. Given that the external field is always less than 30 mT, significantly smaller than the saturation magnetization of YIG μ0MS ≈ 185 mT, the static magnetization of the YIG film tilts only slightly out of plane, at an angle B sin(ϕ)/(μ0MS) ≤ 5.3°.31 Hence, a direct comparison between the NV measurements and the in-plane MOKE measurements is feasible. In addition, in  Appendix G, we demonstrate that within this range of magnetic fields, MOKE measurements in the DE configuration and with the field applied along the NV axis yield consistent wavelengths within the error margin. To facilitate a direct comparison of the measured wavelengths, we transpose the in-plane fields of the TR-MOKE measurements into the direction of the NV center axis. In Fig. 4, the theoretical dispersion curve (red line), alongside the fitted wavelength of the NV measurement (blue dots) and the MOKE measurements (green dots), is shown, which are in good agreement with each other.  Appendix H provides information about the fit model. This consistency demonstrates the reliability and equivalence of the two measurement techniques.

FIG. 4.

Dependence of the wavelength of the excited spin waves on the external magnetic field. The theoretical dispersion curve (red line) is in good agreement with the fitted wavelength from the NV measurements (blue dots) and the TR-MOKE measurements (green dots).

FIG. 4.

Dependence of the wavelength of the excited spin waves on the external magnetic field. The theoretical dispersion curve (red line) is in good agreement with the fitted wavelength from the NV measurements (blue dots) and the TR-MOKE measurements (green dots).

Close modal

In conclusion, the utilization of NV centers presents several advantages in the measurement of spin waves. Their high resolution capabilities can extend beyond the diffraction limit,37,39,47–52 potentially enabling more precise characterization of spin wave phenomena. Moreover, their ability to measure through opaque materials offers unique opportunities for non-invasive investigations.41,53 In addition, NV centers exhibit remarkable sensitivity to magnetic fields,37,54,55 enhancing their utility in detecting subtle magnetic variations. However, it is worth noting some limitations of NV centers, including the lack of time resolution and the requirement of a standing wave for effectively imaging wavefronts. In addition, NV centers are limited to probing only spin waves at the NV ESR frequencies. In comparison, Kerr microscopy offers high time resolution and the flexibility to probe a wide range of frequencies. However, TR-MOKE cannot image spin waves through opaque materials thicker than the skin depth, and the spatial resolution is ultimately limited by diffraction. To overcome these limitations and harness the complementary strengths of both techniques, we propose a combined approach. By integrating NV center measurements with Kerr microscopy, one can benefit from high-resolution imaging, non-invasive probing through opaque materials, high sensitivity to magnetic fields, high time resolution, and flexibility in probing frequencies. This integrated approach promises to advance the capability to probe spin wave dynamics and enable new insights into the field.

This work was supported by the Bayerisches Staatsministerium für Wissenschaft und Kunst through project IQ-Sense via the Munich Quantum Valley (MQV) and by the DFG via the Munich Center for Quantum Science and Technology (MCQST) under Germany’s Excellence Strategy EXC-2111 (Project No. 390814868).

The authors have no conflicts to disclose.

Carolina Lüthi: Data curation (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Lukas Colombo: Validation (equal); Writing – review & editing (equal). Franz Vilsmeier: Validation (equal); Writing – review & editing (equal). Christian Back: Conceptualization (equal); Formal analysis (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Project administration (equal); Resources (equal); Supervision (equal); Validation (equal); Writing – review & editing (equal).

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

1. Diamond substrate

We use a type IIa single-crystal diamond of size 0.5 mm × 0.5 mm × 100 µm grown by chemical vapor deposition, with nitrogen content <1 ppm, surface roughness <2 nm Ra, and (100)-orientation (Applied Diamond). It was implanted with15 N+ ions at an energy of 4 keV, a dose of 1.2 × 1013 ions /cm2, and a tilt angle of 7° (performed by CuttingEdge Ions), resulting in an implantation depth of 6 nm. Following this, it underwent annealing in a vacuum environment of ∼10−6 mbar. The annealing process involved a 45-min ramp to 900 °C, followed by 3 h at 900 ± 10 °C and a 2 h ramp to room temperature. To remove the graphitic layer that formed during the annealing, it was acid cleaned for 4 h by a boiling three-acid mix consisting of 3 ml sulfuric acid (H2SO4), 3 ml nitric acid (HNO3), and 3 ml perchloric acid (HClO4). After thoroughly cleaning both the YIG substrate and the diamond membrane in isopropanol and acetone using an ultrasonic bath, the diamond membrane was carefully placed on the YIG sample using tweezers. To ensure proper alignment of the NV center orientation, which lies in the plane parallel to the sides of the diamond, the placement process was conducted under a microscope. A small droplet of isopropanol was applied to the YIG surface to aid adhesion. The diamond membrane was gently positioned on the droplet and softly pressed onto the surface with tweezers to ensure close contact. The tweezers were released once the isopropanol had fully evaporated, as observed through the microscope. This evaporation step improved the adhesion between the diamond membrane and the YIG substrate, ensuring a stable placement.

2. YIG sample

We use a 200 nm thick YIG thin film grown on a gadolinium gallium garnet substrate via liquid phase epitaxy. The thin film possesses a saturation magnetization of μ0MS = 0.185 mT and an exchange stiffness of Aex=3.71012Jm, as determined from a fit to the dispersion measured by the NV centers. To excite spin waves, stripline S1 with a width of 5 µm was fabricated on top of the YIG film by optical lithography and electron beam evaporation of Ti(5 nm)/Au(100 nm). A second stripline, S2, of Ti(5 nm)/Au(200 nm) with width 30 µm as well as a 150 µm thick SiO2 layer were fabricated on top of S1 by optical lithography and electron beam evaporation.

1. NV center setup

The setup used for the NV center measurements is an in-house built confocal microscope. The NV centers were optically excited by a 515 nm laser (iBeam smart-S 515-S, Toptica), which was focused to a diffraction-limited spot by an objective with a numerical aperture of 0.9 and a magnification of 63. Before the objective, a polarizer in combination with a λ/2-plate was used to rotate the polarization of the laser light, maximizing the contrast in the ω±. The NV PL was gathered by the same objective, separated from the excitation light by a dichroic short pass mirror (cutoff wavelength 600 nm, Edmund Optics) and a long-pass filter (cut-on wavelength 590 nm). Microwaves for driving the NVs and spin waves were generated using a microwave generator (N5183A MXG, Agilent) at 0 dBm output power. To simultaneously send a microwave current through the pair of striplines S1 and S2 depicted in Fig. 1(a), the microwave excitation was divided using a power combiner (ZFRSC-123-S+, Mini-Circuits). To measure the PL signal with a lock-in amplifier, the microwave signal was modulated by a coaxial switch (139-ZASWA-2-50DRA, Mini-Circuits) at a frequency of 1.742 kHz. The PL photons were collected by an avalanche photodiode and, after passing a voltage amplifier, measured by the lock-in amplifier. Thereby, the microwave excitation of the NV centers and the spin waves was modulated, allowing the lock-in amplifier to isolate the dynamic component of the photoluminescence (PL) signal from the background signal. As the lock-in amplifier is designed to detect only variations caused by the modulated microwave signal, the static background PL signal (PL0) is not detected. Consequently, we do not normalize the signal by PL0. Each data point was measured for 1 s at the third filter order. To gain an improved signal, each column in Fig. 3(a) was measured five times and averaged. All measurements were conducted at room temperature.

2. TR-MOKE setup

The TR-MOKE measurements utilized a custom-built confocal microscope. A pulsed 800 nm Ti:Sa-laser was employed, which was first attenuated and polarized using a half-wave plate and a polarizer capable of handling the high optical power densities produced by the laser. The beam then passed through a pellicle beam splitter with 92% transmission. A Wollaston prism splits the returning beam into two linearly polarized components, each directed toward a differential detector with two photodiodes. To keep the repetition rate of the laser constant, the laser’s cavity length was actively adjusted to match the repetition rate to a global frequency reference. This adjustment was facilitated by a control loop incorporating a photodiode detecting generated pulses and mixed with the reference signal to produce a down-converted signal for use by a phase-locked loop and a proportional-integral-derivative controller. The reference signal, derived from a 10 MHz rubidium atomic clock, underwent frequency multiplication to achieve the desired 80 MHz value. The 10 MHz reference signal was also utilized for an arbitrary waveform generator producing MW signals at 20 dBm output power sent to the sample’s stripline S1 [Fig. 1(a)]. The output of the lock-in, which is the differential signal generated by the rotation in polarization and obtained by taking the difference of the outputs of the detectors, is referred to as the Kerr signal SKerr, while the summed signal of the two photodiodes is referred to as the topographic signal STopo. Each data point was measured for 1 s at the third filter order.

The measured parallel and perpendicular projections of an external magnetic field along an NV axis are given by40 
(C1)
(C2)
where D = 2.87 GHz is the zero-field splitting and γ = 2π· 2.8025 MHz/G is the NV gyromagnetic ratio. Furthermore, the theoretical parallel (Bexti,t) and perpendicular (Bexti,t) projections of Bext on the four NV axes are given by
(C3)

In Table I, the expected and fitted parameters for the DC magnetic field, obtained using Eq. (D3), are presented. Here, IDC denotes the DC current through S2, x0 represents the offset of the center of S2 from x = 0, w is the width of the microstrip line, and z0 is the unknown distance between the NV centers and the YIG surface. We find a good agreement between these values.

TABLE I.

Comparison of the expected and fitted values used to determine the NV–S2 distance by fitting the reconstructed DC magnetic field using Eq. (D3).

IDC (mA)x0 (μm)w (μm)z0 (μm)
Expected value 80 40 30 ⋯ 
Fitted value 78.1 ± 0.7 37.2 ± 0.7 34.2 ± 0.2 1.4 ± 0.1 
IDC (mA)x0 (μm)w (μm)z0 (μm)
Expected value 80 40 30 ⋯ 
Fitted value 78.1 ± 0.7 37.2 ± 0.7 34.2 ± 0.2 1.4 ± 0.1 
To simulate the PL signal resulting from the standing wave pattern of the total magnetic field above the YIG thin film, the time-averaged magnitude of total magnetic stray field perpendicular to the NV axis |Btot|t was calculated, where Btot(t) is given by
(D1)
The time-dependent spin wave stray field Bsw(t) is given by31 
(D2)
Bsw=μ0m0(1+sgn(ky)η)|k|dekyz0, where k is the wave vector, ω = 2πf is the frequency, η is the degree of ellipticity, and t is the time. For the parameters, we used m0 = 0.1MS for the magnetization, z0 = 1.4 µm for the distance between the NV layer and the YIG surface, and η = 1 for the ellipticity. Furthermore, the time-dependent Oersted field of S2 is given by31 
(D3)
where we used the parameters I = 0.8 A for the current, w = 30 µm for the width of S2, and z0 = 1.4 µm for the distance between the NV layer and the YIG film surface.
The spin wave dispersion for free magnetization boundary conditions is given by31 
(E1)
with ΩH = B cos(ϕ)/μ0MS, the exchange stiffness αex, the saturation magnetization MS, the angle ϕ = 35°, and
(E2)
where d = 200 nm is the YIG thickness. To compare the NV measurements with the TR-MOKE measurements, the frequency of the ω transition of the NV centers aligned with the external magnetic field Bext is determined via ESR measurements. This frequency is subsequently used to excite spin waves in the TR-MOKE measurements.

Figures 5 and 6 provide additional images of the observed spin wave wavefronts, captured by the NV setup and the TR-MOKE setup respectively, under various external fields, resulting in different excitation frequencies and wavelengths.

FIG. 5.

Spatial PL signal measured above the YIG film, when spin waves excited by the frequency depending on the external applied field strength Bext ranging from 22.33 to 25.19 mT (a) to (d), travel through the YIG thin film.

FIG. 5.

Spatial PL signal measured above the YIG film, when spin waves excited by the frequency depending on the external applied field strength Bext ranging from 22.33 to 25.19 mT (a) to (d), travel through the YIG thin film.

Close modal
FIG. 6.

Spatial TR-MOKE signal, when spin waves excited by the frequency depending on the external applied field strength Bext ranging from 13.74 to 20.8 mT (a) to (d), travel through the YIG thin film.

FIG. 6.

Spatial TR-MOKE signal, when spin waves excited by the frequency depending on the external applied field strength Bext ranging from 13.74 to 20.8 mT (a) to (d), travel through the YIG thin film.

Close modal

In Fig. 7, the spin wave dispersion (for spin waves excited at the ω transition frequency of the NVs) extracted for the case where the external magnetic field is applied in the DE configuration (green dots) and along S1 at an angle of ϕ = 35° relative to the sample plane, aligning it with one of the four possible NV center axis directions (blue dots), is plotted against the theory curve (red line). The deviation in the fitted wavelengths arises from errors in the magnetic field calibration, which was done separately for the DE and NV configurations. However, within the bounds of this error, the measurements closely match the expected theoretical curve, demonstrating the negligible effect of the perpendicular component of the external field with respect to the thin film plane.

FIG. 7.

Dependence of the wavelength of the excited spin waves if the external magnetic field is applied in the DE configuration (green dots) and along S1 at an angle of ϕ = 35° relative to the sample plane. The theoretical dispersion curve (red line) is in good agreement with the fitted wavelengths.

FIG. 7.

Dependence of the wavelength of the excited spin waves if the external magnetic field is applied in the DE configuration (green dots) and along S1 at an angle of ϕ = 35° relative to the sample plane. The theoretical dispersion curve (red line) is in good agreement with the fitted wavelengths.

Close modal
The wavefronts A(y) of the NV- and TR-MOKE measurements were both fitted by
(H1)
where A is the wavefront amplitude, y0 is the distance of the x-axis from S1, λ is the wavelength of the spin wave, ϕ0 is a phase constant, aoff is a constant for the signal offset, and a0 is the slope of the signal.
1.
F.
Bloch
, “
Zur Theorie des Ferromagnetismus
,”
Z. Phys.
61
,
206
219
(
1930
).
2.
A. G.
Gurevich
and
G. A.
Melkov
,
Magnetization Oscillations and Waves
(
CRC Press
,
Boca Raton, FL
,
1996
).
3.
D. D.
Stancil
and
A.
Prabhakar
,
Spin Waves: Theory and Applications
, 1st ed. (
Springer
,
Berlin
,
2008
).
4.
H.
Wang
,
R.
Yuan
,
Y.
Zhou
,
Y.
Zhang
,
J.
Chen
,
S.
Liu
,
H.
Jia
,
D.
Yu
,
J.-P.
Ansermet
,
C.
Song
, and
H.
Yu
, “
Long-distance coherent propagation of high-velocity antiferromagnetic spin waves
,”
Phys. Rev. Lett.
130
,
096701
(
2023
).
5.
V. V.
Kruglyak
,
S. O.
Demokritov
, and
D.
Grundler
, “
Magnonics
,”
J. Phys. D: Appl. Phys.
43
,
264001
(
2010
).
6.
J. R.
Hortensius
,
D.
Afanasiev
,
M.
Matthiesen
,
R.
Leenders
,
R.
Citro
,
A. V.
Kimel
,
R. V.
Mikhaylovskiy
,
B. A.
Ivanov
, and
A. D.
Caviglia
, “
Coherent spin-wave transport in an antiferromagnet
,”
Nat. Phys.
17
,
1001
1006
(
2021
).
7.
A. A.
Serga
,
A. V.
Chumak
, and
B.
Hillebrands
, “
Yig magnonics
,”
J. Phys. D: Appl. Phys.
43
,
264002
(
2010
).
8.
A. V.
Chumak
,
A. A.
Serga
, and
B.
Hillebrands
, “
Magnon transistor for all-magnon data processing
,”
Nat. Commun.
5
,
4700
(
2014
).
9.
V.
Cherepanov
,
I.
Kolokolov
, and
V.
L’vov
, “
The saga of YIG: Spectra, thermodynamics, interaction and relaxation of magnons in a complex magnet
,”
Phys. Rep.
229
,
81
144
(
1993
).
10.
T.
Balashov
,
P.
Buczek
,
L.
Sandratskii
,
A.
Ernst
, and
W.
Wulfhekel
, “
Magnon dispersion in thin magnetic films
,”
J. Phys.: Condens. Matter
26
,
394007
(
2014
).
11.
T.-H.
Chuang
,
K.
Zakeri
,
A.
Ernst
,
Y.
Zhang
,
H. J.
Qin
,
Y.
Meng
,
Y.-J.
Chen
, and
J.
Kirschner
, “
Magnetic properties and magnon excitations in Fe(001) films grown on Ir(001)
,”
Phys. Rev. B
89
,
174404
(
2014
).
12.
A. V.
Chumak
,
V. I.
Vasyuchka
,
A. A.
Serga
, and
B.
Hillebrands
, “
Magnon spintronics
,”
Nat. Phys.
11
,
453
461
(
2015
).
13.
C.
Hahn
,
V. V.
Naletov
,
G.
de Loubens
,
O.
Klein
,
O.
d’Allivy Kelly
,
A.
Anane
,
R.
Bernard
,
E.
Jacquet
,
P.
Bortolotti
,
V.
Cros
,
J. L.
Prieto
, and
M.
Muñoz
, “
Measurement of the intrinsic damping constant in individual nanodisks of Y3Fe5O12 and Y3Fe5O12|Pt
,”
Appl. Phys. Lett.
104
,
152410
(
2014
).
14.
A.
Khitun
, “
Multi-frequency magnonic logic circuits for parallel data processing
,”
J. Appl. Phys.
111
,
054307
(
2012
).
15.
A.
Khitun
,
M.
Bao
, and
K. L.
Wang
, “
Magnonic logic circuits
,”
J. Phys. D: Appl. Phys.
43
,
264005
(
2010
).
16.
T.
Schneider
,
A. A.
Serga
,
T.
Neumann
,
B.
Hillebrands
, and
M. P.
Kostylev
, “
Phase reciprocity of spin-wave excitation by a microstrip antenna
,”
Phys. Rev. B
77
,
214411
(
2008
).
17.
J. M.
Owens
,
J. H.
Collins
, and
R. L.
Carter
, “
System applications of magnetostatic wave devices
,”
Circuits Syst. Signal Process.
4
,
317
334
(
1985
).
18.
J.
Han
,
R.
Cheng
,
L.
Liu
,
H.
Ohno
, and
S.
Fukami
, “
Coherent antiferromagnetic spintronics
,”
Nat. Mater.
22
,
684
695
(
2023
).
19.
S.
Singh
,
V.
Kumar
,
S.
Tyagi
,
N.
Saxena
,
Z. H.
Khan
, and
P.
Kumar
, “
Room temperature ferromagnetism in metal oxides for spintronics: A comprehensive review
,”
Opt. Quantum Electron.
55
,
123
(
2023
).
20.
Y.
Zhang
,
X.
Feng
,
Z.
Zheng
,
Z.
Zhang
,
K.
Lin
,
X.
Sun
,
G.
Wang
,
J.
Wang
,
J.
Wei
,
P.
Vallobra
,
Y.
He
,
Z.
Wang
,
L.
Chen
,
K.
Zhang
,
Y.
Xu
, and
W.
Zhao
, “
Ferrimagnets for spintronic devices: From materials to applications
,”
Appl. Phys. Rev.
10
,
011301
(
2023
).
21.
F.
Vilsmeier
,
C.
Riedel
, and
C. H.
Back
, “
Spatial control of hybridization-induced spin-wave transmission stop band
,”
Appl. Phys. Lett.
124
(
13
),
132407
(
2024
).
22.
K.
Perzlmaier
,
G.
Woltersdorf
, and
C. H.
Back
, “
Observation of the propagation and interference of spin waves in ferromagnetic thin films
,”
Phys. Rev. B
77
,
054425
(
2008
).
23.
M.
Farle
,
T.
Silva
, and
G.
Woltersdorf
, “
Spin dynamics in the time and frequency domain
,”
Springer Tracts Mod. Phys.
246
,
37
83
(
2013
).
24.
Y.
Au
,
T.
Davison
,
E.
Ahmad
,
P. S.
Keatley
,
R. J.
Hicken
, and
V. V.
Kruglyak
, “
Excitation of propagating spin waves with global uniform microwave fields
,”
Appl. Phys. Lett.
98
,
122506
(
2011
).
25.
H. G.
Bauer
,
J.-Y.
Chauleau
,
G.
Woltersdorf
, and
C. H.
Back
, “
Coupling of spinwave modes in wire structures
,”
Appl. Phys. Lett.
104
,
102404
(
2014
).
26.
J.
Stigloher
,
T.
Taniguchi
,
H. S.
Körner
,
M.
Decker
,
T.
Moriyama
,
T.
Ono
, and
C. H.
Back
, “
Observation of a goos-hänchen-like phase shift for magnetostatic spin waves
,”
Phys. Rev. Lett.
121
,
137201
(
2018
).
27.
B.
Hillebrands
, “
Progress in multipass tandem Fabry–Perot interferometry: I. A fully automated, easy to use, self-aligning spectrometer with increased stability and flexibility
,”
Rev. Sci. Instrum.
70
,
1589
1598
(
1999
).
28.
S.
Demokritov
, “
Brillouin light scattering studies of confined spin waves: Linear and nonlinear confinement
,”
Phys. Rep.
348
,
441
489
(
2001
).
29.
T.
Warwick
,
K.
Franck
,
J. B.
Kortright
,
G.
Meigs
,
M.
Moronne
,
S.
Myneni
,
E.
Rotenberg
,
S.
Seal
,
W. F.
Steele
,
H.
Ade
,
A.
Garcia
,
S.
Cerasari
,
J.
Denlinger
,
S.
Hayakawa
,
A. P.
Hitchcock
,
T.
Tyliszczak
,
J.
Kikuma
,
E. G.
Rightor
,
H.-J.
Shin
, and
B. P.
Tonner
, “
A scanning transmission x-ray microscope for materials science spectromicroscopy at the advanced light source
,”
Rev. Sci. Instrum.
69
,
2964
2973
(
1998
).
30.
V.
Sluka
,
T.
Schneider
,
R. A.
Gallardo
,
A.
Kákay
,
M.
Weigand
,
T.
Warnatz
,
R.
Mattheis
,
A.
Roldán-Molina
,
P.
Landeros
,
V.
Tiberkevich
,
A.
Slavin
,
G.
Schütz
,
A.
Erbe
,
A.
Deac
,
J.
Lindner
,
J.
Raabe
,
J.
Fassbender
, and
S.
Wintz
, “
Emission and propagation of 1D and 2D spin waves with nanoscale wavelengths in anisotropic spin textures
,”
Nat. Nanotechnol.
14
,
328
333
(
2019
).
31.
I.
Bertelli
,
J. J.
Carmiggelt
,
T.
Yu
,
B. G.
Simon
,
C. C.
Pothoven
,
G. E. W.
Bauer
,
Y. M.
Blanter
,
J.
Aarts
, and
T.
van der Sar
, “
Magnetic resonance imaging of spin-wave transport and interference in a magnetic insulator
,”
Sci. Adv.
6
,
eabd3556
(
2020
).
32.
C. M.
Purser
,
V. P.
Bhallamudi
,
F.
Guo
,
M. R.
Page
,
Q.
Guo
,
G. D.
Fuchs
, and
P. C.
Hammel
, “
Spinwave detection by nitrogen-vacancy centers in diamond as a function of probe–sample separation
,”
Appl. Phys. Lett.
116
,
202401
(
2020
).
33.
C.
Koerner
,
R.
Dreyer
,
M.
Wagener
,
N.
Liebing
,
H. G.
Bauer
, and
G.
Woltersdorf
, “
Frequency multiplication by collective nanoscale spin-wave dynamics
,”
Science
375
,
1165
1169
(
2022
).
34.
J.-P.
Tetienne
,
T.
Hingant
,
L.
Rondin
,
A.
Cavaillès
,
L.
Mayer
,
G.
Dantelle
,
T.
Gacoin
,
J.
Wrachtrup
,
J.-F.
Roch
, and
V.
Jacques
, “
Spin relaxometry of single nitrogen-vacancy defects in diamond nanocrystals for magnetic noise sensing
,”
Phys. Rev. B
87
,
235436
(
2013
).
35.
P.
Maletinsky
,
S.
Hong
,
M. S.
Grinolds
,
B.
Hausmann
,
M. D.
Lukin
,
R. L.
Walsworth
,
M.
Loncar
, and
A.
Yacoby
, “
A robust scanning diamond sensor for nanoscale imaging with single nitrogen-vacancy centres
,”
Nat. Nanotechnol.
7
,
320
324
(
2012
).
36.
P.
Appel
,
M.
Ganzhorn
,
E.
Neu
, and
P.
Maletinsky
, “
Nanoscale microwave imaging with a single electron spin in diamond
,”
New J. Phys.
17
,
112001
(
2015
).
37.
L.
Rondin
,
J.-P.
Tetienne
,
T.
Hingant
,
J.-F.
Roch
,
P.
Maletinsky
, and
V.
Jacques
, “
Magnetometry with nitrogen-vacancy defects in diamond
,”
Rep. Prog. Phys.
77
,
056503
(
2014
).
38.
T.
Weggler
,
C.
Ganslmayer
,
F.
Frank
,
T.
Eilert
,
F.
Jelezko
, and
J.
Michaelis
, “
Determination of the three-dimensional magnetic field vector orientation with nitrogen vacancy centers in diamond
,”
Nano Lett.
20
,
2980
2985
(
2020
).
39.
G.
Balasubramanian
,
I. Y.
Chan
,
R.
Kolesov
,
M.
Al-Hmoud
,
J.
Tisler
,
C.
Shin
,
C.
Kim
,
A.
Wojcik
,
P. R.
Hemmer
,
A.
Krueger
,
T.
Hanke
,
A.
Leitenstorfer
,
R.
Bratschitsch
,
F.
Jelezko
, and
J.
Wrachtrup
, “
Nanoscale imaging magnetometry with diamond spins under ambient conditions
,”
Nature
455
,
648
651
(
2008
).
40.
T.
van der Sar
,
F.
Casola
,
R.
Walsworth
, and
A.
Yacoby
, “
Nanometre-scale probing of spin waves using single electron spins
,”
Nat. Commun.
6
,
7886
(
2015
).
41.
I.
Bertelli
,
B. G.
Simon
,
T.
Yu
,
J.
Aarts
,
G. E. W.
Bauer
,
Y. M.
Blanter
, and
T.
van der Sar
, “
Imaging spin-wave damping underneath metals using electron spins in diamond
,”
Adv. Quantum Technol.
4
,
2100094
(
2021
).
42.
M. W.
Doherty
,
N. B.
Manson
,
P.
Delaney
,
F.
Jelezko
,
J.
Wrachtrup
, and
L. C. L.
Hollenberg
, “
The nitrogen-vacancy colour centre in diamond
,”
Phys. Rep.
528
,
1
45
(
2013
).
43.
T. X.
Zhou
,
J. J.
Carmiggelt
,
L. M.
Gächter
,
I.
Esterlis
,
D.
Sels
,
R. J.
Stöhr
,
C.
Du
,
D.
Fernandez
,
J. F.
Rodriguez-Nieva
,
F.
Büttner
,
E.
Demler
, and
A.
Yacoby
, “
A magnon scattering platform
,”
Proc. Natl. Acad. Sci. U. S. A.
118
,
e2019473118
(
2021
).
44.
P.
Andrich
,
C. F.
de las Casas
,
X.
Liu
,
H. L.
Bretscher
,
J. R.
Berman
,
F. J.
Heremans
,
P. F.
Nealey
, and
D. D.
Awschalom
, “
Long-range spin wave mediated control of defect qubits in nanodiamonds
,”
npj Quantum Inf.
3
,
28
(
2017
).
45.
R. W.
Damon
and
J. R.
Eshbach
, “
Magnetostatic modes of a ferromagnet slab
,”
J. Phys. Chem. Solids
19
,
308
320
(
1961
).
46.
R.
Dreyer
,
N.
Liebing
,
E. R. J.
Edwards
,
A.
Müller
, and
G.
Woltersdorf
, “
Spin-wave localization and guiding by magnon band structure engineering in yttrium iron garnet
,”
Phys. Rev. Mater.
5
,
064411
(
2021
).
47.
C. L.
Degen
, “
Scanning magnetic field microscope with a diamond single-spin sensor
,”
Appl. Phys. Lett.
92
,
243111
(
2008
).
48.
J. M.
Taylor
,
P.
Cappellaro
,
L.
Childress
,
L.
Jiang
,
D.
Budker
,
P. R.
Hemmer
,
A.
Yacoby
,
R.
Walsworth
, and
M. D.
Lukin
, “
Erratum: High-sensitivity diamond magnetometer with nanoscale resolution
,”
Nat. Phys.
7
,
270
(
2011
).
49.
B. J.
Maertz
, “
Vector magnetic field microscopy using nitrogen vacancy centers in diamond
,”
Appl. Phys. Lett.
96
,
092504
(
2010
).
50.
C. A.
Meriles
,
L.
Jiang
,
G.
Goldstein
,
J. S.
Hodges
,
J.
Maze
,
M. D.
Lukin
, and
P.
Cappellaro
, “
Imaging mesoscopic nuclear spin noise with a diamond magnetometer
,”
J. Chem. Phys.
133
,
124105
(
2010
).
51.
A.
Laraoui
,
J. S.
Hodges
, and
C. A.
Meriles
, “
Magnetometry of random AC magnetic fields using a single nitrogen-vacancy center
,”
Appl. Phys. Lett.
97
,
143104
(
2010
).
52.
N.
Zhao
,
J.-L.
Hu
,
S.-W.
Ho
,
J. T. K.
Wan
, and
R. B.
Liu
, “
Atomic-scale magnetometry of distant nuclear spin clusters via nitrogen-vacancy spin in diamond
,”
Nat. Nanotechnol.
6
,
242
246
(
2011
).
53.
M.
Borst
,
P. H.
Vree
,
A.
Lowther
,
A.
Teepe
,
S.
Kurdi
,
I.
Bertelli
,
B. G.
Simon
,
Y. M.
Blanter
, and
T.
van der Sar
, “
Observation and control of hybrid spin-wave–Meissner-current transport modes
,”
Science
382
,
430
434
(
2023
).
54.
J. R.
Maze
,
P. L.
Stanwix
,
J. S.
Hodges
,
S.
Hong
,
J. M.
Taylor
,
P.
Cappellaro
,
L.
Jiang
,
M. V. G.
Dutt
,
E.
Togan
,
A. S.
Zibrov
,
A.
Yacoby
,
R. L.
Walsworth
, and
M. D.
Lukin
, “
Nanoscale magnetic sensing with an individual electronic spin in diamond
,”
Nature
455
,
644
647
(
2008
).
55.
S.
Steinert
,
F.
Dolde
,
P.
Neumann
,
A.
Aird
,
B.
Naydenov
,
G.
Balasubramanian
,
F.
Jelezko
, and
J.
Wrachtrup
, “
High sensitivity magnetic imaging using an array of spins in diamond
,”
Rev. Sci. Instrum.
81
,
043705
(
2010
).