Direct Nanoscale Sensing of the Internal Electric Field in Operating

Jan 23, 2017 - Here, we demonstrate the direct sensing of the internal electric field in diamond power devices using single nitrogen–vacancy (NV) ce...
36 downloads 13 Views 3MB Size
Direct Nanoscale Sensing of the Internal Electric Field in Operating Semiconductor Devices Using Single Electron Spins Takayuki Iwasaki,*,†,‡ Wataru Naruki,† Kosuke Tahara,†,‡ Toshiharu Makino,‡,§ Hiromitsu Kato,‡,§ Masahiko Ogura,‡,§ Daisuke Takeuchi,‡,§ Satoshi Yamasaki,‡,§ and Mutsuko Hatano†,‡ †

Department of Electrical and Electronic Engineering, Tokyo Institute of Technology, Meguro, Tokyo 152-8552, Japan CREST, Japan Science and Technology Agency, Chiyoda, Tokyo 102-0076, Japan § Advanced Power Electronics Research Center, National Institute of Advanced Industrial Science and Technology, Tsukuba, Ibaraki 305-8568, Japan ‡

S Supporting Information *

ABSTRACT: The electric field inside semiconductor devices is a key physical parameter that determines the properties of the devices. However, techniques based on scanning probe microscopy are limited to sensing at the surface only. Here, we demonstrate the direct sensing of the internal electric field in diamond power devices using single nitrogen−vacancy (NV) centers. The NV center embedded inside the device acts as a nanoscale electric field sensor. We fabricated vertical diamond p-i-n diodes containing the single NV centers. By performing optically detected magnetic resonance measurements under reverse-biased conditions with an applied voltage of up to 150 V, we found a large splitting in the magnetic resonance frequencies. This indicated that the NV center senses the transverse electric field in the space-charge region formed in the i-layer. The experimentally obtained electric field values are in good agreement with those calculated by a device simulator. Furthermore, we demonstrate the sensing of the electric field in different directions by utilizing NV centers with different N−V axes. This direct and quantitative sensing method using an electron spin in a wide-band-gap material provides a way to monitor the electric field in operating semiconductor devices. KEYWORDS: nanoscale electric field sensing, semiconductor device, diamond, nitrogen−vacancy center, single electron spin

T

this study, we propose a method for quantitatively sensing the internal electric field using a single electron spin embedded in the devices. The atomic-level structures with the single electron spin, which are sensitive to the electric field, can be formed in the wide-band-gap materials, such as a nitrogen−vacancy (NV) center in diamond4 and vacancies in 4H-SiC.5−7 Here, the single NV centers were fabricated in a diamond device, and the direct nanoscale measurement of the internal electric field was performed. The NV center consists of a nitrogen atom and a vacancy formed in the diamond lattice.4 Its energy level is confined in the band gap of diamond and is sensitive to various perturbations, such as magnetic8−10 and electric11−13 fields, pressure,14 and temperature.15−20 Because the NV center is an atomic-level structure, sensing can be performed at a nanoscale

he electric properties of semiconductor devices are controlled by their internal electric field, which enables their operation as rectifiers and switching devices. Strong electric fields are induced inside the devices under the reverse bias condition. Electronic devices based on wide-bandgap materials, such as 4H-SiC, GaN, and diamond, exhibit particularly strong electric fields,1 which is a key property for low-loss electronic applications. When the internal electric field values exceed the breakdown electric field, the system no longer operates correctly and safely; this situation is difficult to predict by device simulation. Therefore, in situ monitoring and direct sensing of the internal electric field in the semiconductor devices with high spatial resolution is needed to monitor the unusual behavior and understand the device performance. Several techniques have been proposed for measuring the electric properties based on scanning probe microscopy, such as Kelvin force microscopy2 and scanning capacitance microscopy.3 However, these methods can detect only the surface of the materials and cannot obtain the information regarding the interior of the devices under the application of high voltage. In © 2017 American Chemical Society

Received: July 6, 2016 Accepted: January 10, 2017 Published: January 23, 2017 1238

DOI: 10.1021/acsnano.6b04460 ACS Nano 2017, 11, 1238−1245

Article

www.acsnano.org

Article

ACS Nano

Figure 1. Device structure and setup for electric field sensing. (a) Photograph of the sample configuration. The sample was fixed on the anode contact metal with an electrically conducting adhesive. The devices were fabricated on the 2 × 2 mm2 diamond substrate. Inset: Top-view optical microscope image of a diamond vertical p-i-n diode. (b) Measurement setup. The device was placed in immersion oil. Microwave (MW) for magnetic resonance was radiated from a thin Cu wire. (c) AFM image (top panel) and line profile of the step structure along the dashed line in the image (bottom curve). “C” in the top panel refers to the cathode electrode. (d) I−V curve measured in immersion oil. The voltage was applied to the n+-type diamond at the top of the device.

result is evidence that step-flow growth occurred to form the high-quality diamond film. This surface morphology is not expected to affect the electric field at NV centers several hundred nanometers below the surface. Figure 1d shows the I− V curve of the diamond p-i-n diode, measured in immersion oil. The device turns on with a threshold voltage of −4 V, and its on current increases to 1.6 mA at −10 V. The reverse bias (positive bias) expands the SCR in the i-layer. Although the leakage current gradually increases as the voltage is increased, it remains low and breakdown is not observed up to 200 V; this trait is important for sensing the strong electric field in the device. An on/off ratio of 105 was obtained at ±10 V, indicating that the device functions as a diode. For electric field sensing in semiconductor devices, a region where the electric field is concentrated must be measured. Because the n+-type diamond is patterned in the p-i-n diode used here, the electric field concentrates and achieves the maximum value at the edge of the n+-i interface. Hence, we performed confocal fluorescence microscope observations to locate the NV centers in the strong electric field region. Figure 2a shows a large-area top-view scanning confocal image near the n+-i interface. The components of the diode structure can be well-resolved by the difference in the fluorescence intensity. The i-layer has the lowest background intensity, enabling the observation of single NV centers. In the magnified image (Figure 2b), we observed an NV center 540 nm from the n+-i interface (denoted as NV in Figure 2b) at a depth of 410 nm (Supporting Information, Figure S1). A diffraction-limited spatial resolution of approximately 300 nm was obtained for the sensing. The selected spot was confirmed to correspond to a single NV center based on the Hanbury Brown and Twiss

resolution. Unique applications for magnetometers, such as nanoscale NMR, have been intensively studied.21−26 Regarding the electric field, the sensing principle and benefits of using the NV center, such as the sensing capability at room temperature and quantitative measurement of the electric field, have been reported.11,12 Furthermore, the detection of a single charge has been demonstrated as an application of electric field sensing.27 In the context of combined work on electronic devices and fluorescent centers, charge state control, 28−31 electrical excitation,32−34 and electrical readout35−37 have been reported. For the charge state control in the proximity of the surface, band bending with an electric field plays a key role.38,39 For the electrical readout process, the electric field was observed on a diamond substrate containing an ensemble of NV centers.35 In this study, we focus on the high electric field generated in the space-charge region (SCR) of an electronic device structure and report the direct sensing of this internal electric field under reverse-biased conditions by observing the nuclear hyperfine structure of NV centers.

RESULTS AND DISCUSSION We fabricated vertical diamond p-i-n diodes31 for the electric field measurements (Figure 1). A photograph and an illustration of the measurement setup are shown in Figure 1a,b, respectively. The NV centers were fabricated inside the intrinsic layer (i-layer) of the device by ion implantation with a projected depth of approximately 350 nm. Figure 1c shows an atomic force microscopy (AFM) image of the diamond p-i-n diode, which clearly indicates a periodic structure in the i-layer with a terrace width of 1.5 μm and a step height of 20 nm. This 1239

DOI: 10.1021/acsnano.6b04460 ACS Nano 2017, 11, 1238−1245

Article

ACS Nano

study, we utilized NV centers with different axis directions to sense the internal electric field in different directions. The axis of the single NV center in Figure 2 was confirmed to be oriented along the [111] direction normal to the sample surface. Figure 3a illustrates the cross-sectional device structure with the SCR at different voltages. Even without an external bias (0 V), the SCR expands with a depth of 3.5 μm and a width of 5 μm owing to the built-in potential. Thus, in the SCR, the NV center is exposed to the electric field in both the axial and transverse directions with respect to the N−V axis (Figure 3a). The SCR becomes larger with increasing voltage and reaches the p-type substrate at V1 ( V1 > 0 V) and a single NV center exposed to the axial and transverse electric fields. (b) ODMR spectra of the single NV center under reverse-biased conditions up to 150 V. Gray lines are experimental data. Black and red lines indicate fitting for mI = 0 and mI = ±1, respectively. Blue lines are the envelopes of the fitting curves. (c) Relationship of transverse electric fields and strain fields (Π⊥) and ODMR splitting width. Solid lines are drawn using the theoretical equations in the main text. 1240

DOI: 10.1021/acsnano.6b04460 ACS Nano 2017, 11, 1238−1245

Article

ACS Nano

Figure 4. Estimation of the internal transverse electric field. (a) Simulation of the transverse electric field at 150 V with a relative permittivity of 3 for the immersion oil. The NV center is located at (x,z) = (0.54 μm, 0.41 μm). (b) Dependence of the transverse electric field on the applied reverse voltage, simulated at z = 0.41 μm. The dotted line denotes the experimental NV position, x = 0.54 μm. (c) Transverse electric field inside the diamond p-i-n diode measured with the NV center. Simulated results at the same NV position as the experimental observation (marked as a dot in panel a) are also shown. The uncertainty in the simulation results is caused by the uncertainty in the value of the relative permittivity of immersion oil and the detection limit of the boron concentration in the i-layer.

of the dips for mI = ±1 was observed even upon the application of the bias, indicating that the magnetic field generated by the leakage current did not affect the measurements (Supporting Information, Figure S3). As mentioned above, because of the small electric dipole moment corresponding to Ez (dgs∥/h = 0.35 Hz·cm·V−1),11 no clear tendency regarding a shift in the center position of the ODMR lines was observed (Supporting Information, Figures S4 and S5). Under the condition of the zero magnetic field, the relationship between the transverse electric field and ODMR splitting width is described considering the nuclear hyperfine interaction41 and is given by (also see Supporting Information, Figure S6) dgs⊥Π⊥ h

dgs⊥Π⊥ h

=

1 W0 for mI = 0 2

=

1 [W±12 − (2A /h)2 ]1/2 for mI = ±1 2

relationship in Figure 3c. At 50 V, we obtain the similar dgs⊥Π⊥/h values for each nuclear hyperfine state, while the value for mI = ±1 becomes much lower at 30 V. This difference is due to the nonlinearity in the small electric field regime for mI = ±1. Although the nuclear hyperfine structure for mI = 0 is unresolved at 100 and 150 V, probably because of the small difference in the magnetic resonance frequencies, we can obtain the dgs⊥Π⊥/h values for mI = ±1. The transverse electric field in the p-i-n diode was simulated using a device simulator (Figure 4a). Figure 4b shows the dependence of the transverse electric field on the applied voltage and the distance x from the n+-edge at a fixed depth of 0.41 μm. As expected, the electric field increases with approaching the n+-edge (x = 0 μm) and with increasing voltage. Figure 4c shows the estimated experimental electric field plotted versus the applied voltage for mI = 0 and mI = ±1. The vertical error bars originate from the unknown direction of the strain field. For mI = ±1, the large difference from the simulation is observed at 30 V. This is because of the aforementioned nonlinear relationship. Thus, more correct electric field can be obtained with mI = 0 for the small voltage regime. At a voltage of 150 V, the field magnitude reaches approximately 350 kV/cm, the highest electric field detected by the NV center. The experimental values obtained by the NV center were in good agreement with the simulation results, thus confirming the capability of the NV center to quantitatively measure the internal electric field in diamond devices with a high spatial resolution. To sense the electric field in the device in a different direction, we utilized an NV center with an inclined axial

where Π⊥ is given by Π⊥ = E⊥ + σ⊥. W0 and W±1 are the splitting width between the dips in the ODMR spectrum for mI = 0 and mI = ±1, respectively. A∥ is the axial magnetic hyperfine parameter, provided as A∥/h = 2.2 MHz. These theoretical relationships are illustrated in Figure 3c. Although different behaviors were observed (i.e., linear for mI = 0 and nonlinear for mI = ±1), the transverse electric field could be obtained from either curve. Each component for mI = 0 and mI = ±1 can be resolved at 30 and 50 V in the bias-applied ODMR spectra (Figure 3b). Then, dgs⊥Π⊥/h can be obtained from the theoretical 1241

DOI: 10.1021/acsnano.6b04460 ACS Nano 2017, 11, 1238−1245

Article

ACS Nano

Figure 5. Sensing the internal electric field in a different direction. (a) Schematic of the four possible NV orientations in the device. (b) Determination of the N−V axis direction by measuring the ODMR spectra under an applied magnetic field (∼2.5 mT) in different directions parallel to [111] and [1̅1̅1]. The gray lines are the experimental data, and the colored lines are the fit curves. (c) ODMR spectra of the NV center with a [1̅1̅1] axis at 0 V (top panel) and under a reverse bias of 100 V (bottom panel). The gray lines are the experimental data. The black and red lines represent the fits for mI = 0 and mI = ±1, respectively. The blue lines are the envelopes of the fit curves.

distribution to be imaged with a spatial resolution as low as approximately 10 nm, thus enabling more precise investigations of the electric field concentration even in complex device structures, such as narrow channel transistors47,48 and super junctions.49 We note that the applications of electric field sensing are not limited to electronic devices. In an electrochemical device, an electric field is generated at the interface between the solution and semiconductor materials. In addition, the band bending inside the material surface generates an electric field. The electric fields are a key factor affecting the electrochemical reaction. Diamond is expected to be a good candidate for use in such applications50−52 because of its chemical stability, large potential window, and negative electron affinity. The direct detection of such electric fields will lead to a better understanding of the nanoscale mechanisms of electrochemical reactions. Finally, it is worth mentioning that fluorescent atomic structures can be fabricated in various materials such as GaN,53 AlN,54 and two-dimensional h-BN.55 A structure similar to the NV center is also predicted in c-BN by the first-principles calculation.56 Therefore, with the development of the fabrication method and the spin control, the internal electric field sensing in such wide-band-gap materials will be realized.

direction. An NV center can be oriented in one of four possible directions: [111], [11̅1̅], [1̅11̅], or [1̅1̅1], as shown in Figure 5a. In the (111) substrate used here, the three axes are rotated by 109.5° from the [111] direction normal to the substrate. Thus, the transverse electric field of such an NV center yields information on the vertical electric field or, more exactly, the electric field rotated by 19.5° from the z direction. We found an NV center with an inclined axis direction at (x,z) = (0.89 μm, 0.38 μm) in another diode. In the ODMR measurements, while the magnetic field direction was varied using the three-axis magnets, the split width of the dips reached a maximum when the magnetic field was parallel to [11̅ 1̅ ] (Figure 5b). Thus, the N−V axis was determined to lie along the [1̅1̅1] direction. The nuclear hyperfine structure of this NV center was measured at 0 V and under a reverse bias of 100 V with cancellation of the environmental magnetic field (Figure 5c). For this NV center, 2dgs⊥Π⊥0/h was found to be 1.70 MHz, indicating a much larger strain compared with the NV center in Figure 3b. The split width for mI = ±1 increased upon bias application. The corresponding transverse electric field, 1̅ 1̅ ] in Figure 5a, was estimated to be 130 ± 50 denoted by E[1 ⊥ kV/cm, considering the error caused by the effect of the strain. The simulated electric field was approximately 60 kV/cm, which is close to the lower value of the experimental estimation, ∼80 kV/cm (the case in which the strain field is in the same direction as the transverse electric field). It is important to suppress the strain around the NV center or to determine the strain direction12 to reduce the experimental error. In addition, the use of more sensitive pulse measurements would enable more precise detection of the electric field.11,12 We confirmed the sensing ability of the internal electric field in a diamond power device using a simple device structure and by comparison to the simulation results. In future studies, this sensing method could be applied to cases in which accurate simulations are difficult, such as for a device with an unexpected electric field concentration or a high leakage current. This method will be useful for understanding the mechanisms of such devices in greater detail and for promoting the development of diamond power devices. Furthermore, the utilization of multiple NV centers in a regular array formed by ion implantation techniques42,43 and super-resolution techniques44−46 will enable the position-dependent electric field

CONCLUSIONS In summary, we demonstrated the direct sensing of the internal electric field in vertical diamond p-i-n diodes using the single NV centers fabricated in the devices by ion implantation and subsequent annealing. The transverse electric field in SCR was estimated based on the variation in the splitting width between ODMR dips. At a reverse voltage of 150 V, the field magnitude reaches 350 kV/cm, in good agreement with the device simulation results. We showed that NV centers with differently oriented axes can be used to determine the electric fields in various directions. This technique will enable the imaging of the electric field distribution in diamond devices and will be extended to the sensing of more complex structures and various materials to accelerate the development of next-generation lowloss electronics. 1242

DOI: 10.1021/acsnano.6b04460 ACS Nano 2017, 11, 1238−1245

Article

ACS Nano

METHODS

(WPI) Initiative on Material Nanoarchitechtonics (MANA Foundry)” of Japan for the EB lithography.

For the device fabrication, first, a 5 μm thick i-layer was deposited on a boron-doped p-type (111) diamond substrate with a doping concentration of 1 × 1017 cm−3 via microwave plasma chemical vapor deposition. The boron concentration in the i-layer was below the detection limit of secondary ion mass spectrometry (2 × 1014 cm−3). Then, a heavily phosphorus-doped (1 × 1019 cm−3) 350 nm thick n+-type diamond layer was synthesized using metal mask patterns. After the metal mask was removed, NV centers were fabricated in the device by 14N ion implantation with a dose of 1 × 109 cm−2 and an acceleration energy of 350 keV, giving rise to a projected depth of approximately 350 nm (N atoms are located at depths ranging up to ∼450 nm from the surface). Then, the sample was annealed at 750 °C for 30 min for NV formation. Thus, the NV centers were directly embedded in the device structure. Finally, the electrodes were formed on top of the n+-layers and at the bottom of the p-type substrate to apply the voltage. The diamond surface was oxygen-terminated for insulation. The electric field sensing was performed using a home-built confocal microscope system with a 532 nm excitation laser. The red fluorescence from the NV centers was measured using avalanche photodiodes (APDs). Two APDs were utilized for the HBT-type interferometry, confirming that an observed fluorescent spot was a single NV center. The sample was placed in the immersion oil throughout the measurements to increase the spatial resolution of the confocal fluorescence image and prevent the discharge of air between the electrodes under the application of a high voltage. The electric field measurements were performed by the ODMR measurements while applying voltages of 0−150 V to the cathode. During the ODMR measurements, we utilized three-axis electromagnets to compensate the magnetic field from the circumference environment. The voltage was applied using a semiconductor parameter analyzer. A Cu thin wire was used to radiate MW for magnetic resonance. The device simulation (Sentaurus TCAD, Synopsys) was performed for the electric field comparison. To approach the experimental measurement conditions, an insulator was placed around the diode to simulate the immersion oil. The precise composition of the immersion oil (Olympus Immersion Oil Type-F) is undisclosed. Therefore, we used a relative permittivity of 2−3 for the insulator based on the permittivity values of oil-based materials such as polybutene and paraffin oil.57

REFERENCES (1) Willander, M.; Friesel, M.; Wahab, Q.; Straumal, B. Silicon Carbide and Diamond for High Temperature Device Applications. J. Mater. Sci.: Mater. Electron. 2006, 17, 1−25. (2) Henning, A. K.; Hochwitz, T.; Slinkman, J.; Never, J.; Hoffmann, S.; Kaszuba, P.; Daghlian, C. Two-Dimensional Surface Dopant Profiling in Silicon Using Scanning Kelvin Probe Microscopy. J. Appl. Phys. 1995, 77, 1888−1896. (3) Slinkman, J. A.; Williams, C. C.; Abraham, D. W.; Wickramasinghe, H. K. Lateral Dopant Profiling in MOS Structures on a 100 nm Scale Using Scanning Capacitance Microscopy. Proc. IEDM 1990, 73−76. (4) Wrachtrup, J.; Jelezko, F. Processing Quantum Information in Diamond. J. Phys.: Condens. Matter 2006, 18, S807−S824. (5) Koehl, W. F.; Buckley, B. B.; Heremans, F. J.; Calusine, G.; Awschalom, D. D. Room Temperature Coherent Control of Defect Spin Qubits in Silicon Carbide. Nature 2011, 479, 84−87. (6) Kraus, H.; Soltamov, V. A.; Fuchs, F.; Simin, D.; Sperlich, A.; Baranov, P. G.; Astakhov, G. V.; Dyakonov, V. Magnetic Field and Temperature Sensing With Atomic-Scale Spin Defects in Silicon Carbide. Sci. Rep. 2014, 4, 5303. (7) Falk, A. L.; Klimov, P. V.; Buckley, B. B.; Ivady, V.; Abrikosov, I. A.; Calusine, G.; Koehl, W. F.; Gali, A.; Awschalom, D. D. Electrically and Mechanically Tunable Electron Spins in Silicon Carbide Color Centers. Phys. Rev. Lett. 2014, 112, 187601. (8) Maze, J. R.; Stanwix, P. L.; Hodges, J. S.; Hong, S.; Taylor, J. M.; Cappellaro, P.; Jiang, L.; Dutt, M. V. G.; Togan, E.; Zibrov, A. S.; Yacoby, A.; Walsworth, R. L.; Lukin, M. D. Nanoscale Magnetic Sensing With an Individual Electronic Spin in Diamond. Nature 2008, 455, 644−647. (9) Balasubramanian, G.; Chan, I. Y.; Kolesov, R.; Al-Hmoud, M.; Tisler, J.; Shin, C.; Kim, C.; Wojcik, A.; Hemmer, P. R.; Krueger, A.; Hanke, T.; Leitenstorfer, A.; Bratschitsch, R.; Jelezko, F.; Wrachtrup, J. Nanoscale Imaging Magnetometry With Diamond Spins Under Ambient Conditions. Nature 2008, 455, 648−651. (10) Taylor, J. M.; Cappellaro, P.; Childress, L.; Jiang, L.; Budker, D.; Hemmer, P. R.; Yacoby, A.; Walsworth, R.; Lukin, M. D. HighSensitivity Diamond Magnetometer With Nanoscale Resolution. Nat. Phys. 2008, 4, 810−816. (11) Van Oort, E.; Glasbeek, M. Electric-Field-Induced Modulation of Spin Echoes of N-V Centers in Diamond. Chem. Phys. Lett. 1990, 168, 529−532. (12) Dolde, F.; Fedder, H.; Doherty, M. W.; Noebauer, T.; Rempp, F.; Balasubramanian, G.; Wolf, T.; Reinhard, F.; Hollenberg, L. C. L.; Jelezko, F.; Wrachtrup, J. Electric-Field Sensing Using Single Diamond Spins. Nat. Phys. 2011, 7, 459−463. (13) Doherty, M. W.; Michl, J.; Dolde, F.; Jakobi, I.; Neumann, P.; Manson, N. M.; Wrachtrup, J. Measuring the Defect Structure Orientation of a Single Centre in Diamond. New J. Phys. 2014, 16, 063067. (14) Doherty, M. W.; Struzhkin, V. V.; Simpson, D. A.; McGuinness, L. P.; Meng, Y.; Stacey, A.; Karle, T. J.; Hemley, R. J.; Manson, N. B.; Hollenberg, L. C. L.; Prawer, S. Electronic Properties and Metrology Applications of the Diamond NV− Center under Pressure. Phys. Rev. Lett. 2014, 112, 047601. (15) Acosta, V. M.; Bauch, E.; Ledbetter, M. P.; Waxman, A.; Bouchard, L. − S.; Budker, D. Temperature Dependence of the Nitrogen-Vacancy Magnetic Resonance in Diamond. Phys. Rev. Lett. 2010, 104, 070801. (16) Toyli, D. M.; Christle, D. J.; Alkauskas, A.; Buckley, B. B.; Van de Walle, C. G.; Awschalom, D. D. Measurement and Control of Single Nitrogen-Vacancy Center Spins above 600 K. Phys. Rev. X 2012, 2, 031001. (17) Neumann, P.; Jakobi, I.; Dolde, F.; Burk, C.; Reuter, R.; Waldherr, G.; Honert, J.; Wolf, T.; Brunner, A.; Shim, J. H.; Suter, D.; Sumiya, H.; Isoya, J.; Wrachtrup, J. High-Precision Nanoscale

ASSOCIATED CONTENT S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsnano.6b04460. Estimation of the NV depth, cancellation of the circumference magnetic field, effect of the leakage current on ODMR, simulation of the axial electric field in the device, and nuclear hyperfine structure under electric field (PDF)

AUTHOR INFORMATION Corresponding Author

*E-mail: [email protected]. ORCID

Takayuki Iwasaki: 0000-0001-6319-7718 Notes

The authors declare no competing financial interest.

ACKNOWLEDGMENTS We thank F. Jelezko and M. Shimizu for discussions. This work was supported in part by TEPCO Memorial Foundation. We thank the “World Premier International Research Center 1243

DOI: 10.1021/acsnano.6b04460 ACS Nano 2017, 11, 1238−1245

Article

ACS Nano Temperature Sensing Using Single Defects in Diamond. Nano Lett. 2013, 13, 2738−2742. (18) Kucsko, G.; Maurer, P. C.; Yao, N. Y.; Kubo, M.; Noh, H. J.; Lo, P. K.; Park, H.; Lukin, M. D. Nanometre-Scale Thermometry in a Living Cell. Nature 2013, 500, 54−58. (19) Doherty, M. W.; Acosta, V. M.; Jarmola, A.; Barson, M. S. J.; Manson, N. B.; Budker, D.; Hollenberg, L. C. L. Temperature Shifts of the Resonances of the NV− Center in Diamond. Phys. Rev. B: Condens. Matter Mater. Phys. 2014, 90, 041201. (20) Clevenson, H.; Trusheim, M. E.; Teale, C.; Schroeder, T.; Braje, D.; Englund, D. Broadband Magnetometry and Temperature Sensing With a Light-Trapping Diamond Waveguide. Nat. Phys. 2015, 11, 393−397. (21) Staudacher, T.; Shi, F.; Pezzagna, S.; Meijer, J.; Du, J.; Meriles, C. A.; Reinhard, F.; Wrachtrup, J. Nuclear Magnetic Resonance Spectroscopy on a (5-Nanometer)3 Sample Volume. Science 2013, 339, 561−563. (22) Mamin, H. J.; Kim, M.; Sherwood, M. H.; Rettner, C. T.; Ohno, K.; Awschalom, D. D.; Rugar, D. Nanoscale Nuclear Magnetic Resonance with a Nitrogen-Vacancy Spin Sensor. Science 2013, 339, 557−560. (23) Mueller, C.; Kong, X.; Cai, J. − M.; Melentijevic, K.; Stacey, A.; Markham, M.; Twitchen, D.; Isoya, J.; Pezzagna, S.; Meijer, J.; Du, J. F.; Plenio, M. B.; Naydenov, B.; McGuinness, L. P.; Jelezko, F. Nuclear Magnetic Resonance Spectroscopy With Single spin Sensitivity. Nat. Commun. 2014, 5, 4703. (24) Haeberle, T.; Schmid-Lorch, D.; Reinhard, F.; Wrachtrup, J. Nanoscale Nuclear Magnetic Imaging With Chemical Contrast. Nat. Nanotechnol. 2015, 10, 125−127. (25) Rugar, D.; Mamin, H. J.; Sherwood, M. H.; Kim, M.; Rettner, C. T.; Ohno, K.; Awschalom, D. D. Proton Magnetic Resonance Imaging Using a Nitrogen−Vacancy Spin Sensor. Nat. Nanotechnol. 2014, 10, 120−124. (26) DeVience, S. J.; Pham, L. M.; Lovchinsky, I.; Sushkov, A. O.; Bar-Gill, N.; Belthangady, C.; Casola, F.; Corbett, M.; Zhang, H.; Lukin, M.; Park, H.; Yacoby, A.; Walsworth, R. L. Nanoscale NMR Spectroscopy and Imaging of Multiple Nuclear Species. Nat. Nanotechnol. 2015, 10, 129−134. (27) Dolde, F.; Doherty, M. W.; Michl, J.; Jakobi, I.; Naydenov, B.; Pezzagna, S.; Meijer, J.; Neumann, P.; Jelezko, F.; Manson, N. B.; Wrachtrup, J. Nanoscale Detection of a Single Fundamental Charge in Ambient Conditions Using the NV− Center in Diamond. Phys. Rev. Lett. 2014, 112, 097603. (28) Grotz, B.; Hauf, M. V.; Dankerl, M.; Naydenov, B.; Pezzagna, S.; Meijer, J.; Jelezko, F.; Wrachtrup, J.; Stutzmann, M.; Reinhard, F.; Garrido, J. A. Charge State Manipulation of Qubits in Diamond. Nat. Commun. 2012, 3, 729. (29) Schreyvogel, C.; Wolfer, M.; Kato, H.; Schreck, M.; Nebel, C. Tuned NV Emission by In-Plane Al-Schottky Junctions on Hydrogen Terminated Diamond. Sci. Rep. 2014, 4, 3634. (30) Kato, H.; Wolfer, M.; Schreyvogel, C.; Kunzer, M.; MuellerSebert, W.; Obloh, H.; Yamasaki, S.; Nebel, C. Tunable Light Emission from Nitrogen-Vacancy Centers in Single Crystal Diamond PIN Diodes. Appl. Phys. Lett. 2013, 102, 151101. (31) Shimizu, M.; Makino, T.; Iwasaki, T.; Hasegawa, J.; Tahara, K.; Naruki, W.; Kato, H.; Yamasaki, S.; Hatano, M. Charge State Modulation of Nitrogen Vacancy Centers in Diamond by Applying a Forward Voltage Across a p−i−n Junction. Diamond Relat. Mater. 2016, 63, 192−196. (32) Mizuochi, N.; Makino, T.; Kato, H.; Takeuchi, D.; Ogura, M.; Okushi, H.; Nothaft, M.; Neumann, P.; Gali, A.; Jelezko, F.; Wrachtrup, J.; Yamasaki, S. Electrically Driven Single-Photon Source at Room Temperature in Diamond. Nat. Photonics 2012, 6, 299−303. (33) Klimov, P. V.; Falk, A. L.; Buckley, B. B.; Awschalom, D. D. Electrically Driven Spin Resonance in Silicon Carbide Color Centers. Phys. Rev. Lett. 2014, 112, 087601. (34) Lohrmann, L.; Iwamoto, N.; Bodrog, Z.; Castelletto, S.; Ohshima, T.; Karle, T. J.; Gali, A.; Prawer, S.; McCallum, J. C.;

Johnson, B. C. Single-Photon Emitting Diode in Silicon Carbide. Nat. Commun. 2015, 6, 7783. (35) Bourgeois, E.; Jarmola, A.; Siyushev, P.; Gulka, M.; Hruby, J.; Jelezko, F.; Budker, D.; Nesladek, M. Photoelectric Detection of Electron Spin Resonance of Nitrogen-Vacancy Centres in Diamond. Nat. Commun. 2015, 6, 8577. (36) Brenneis, A.; Gaudreau, L.; Seifert, M.; Karl, H.; Brandt, M. S.; Huebl, H.; Garrido, J. A.; Koppens, F. H. L.; Holleitner, A. W Ultrafast Electronic Readout of Diamond Nitrogen-Vacancy Centres Coupled to Graphene. Nat. Nanotechnol. 2014, 10, 135−139. (37) Hrubesch, F. M.; Braunbeck, G.; Stutzmann, M.; Reinhard, F.; Brandt, S. Efficient Electrical Spin Readout of NV− Centers in Diamond. arXiv:1608.02459 2016. (38) Hauf, M. V.; Grotz, B.; Naydenov, B.; Dankerl, M.; Pezzagna, S.; Meijer, J.; Jelezko, F.; Wrachtrup, J.; Stutzmann, M.; Reinhard, F.; Garrido, J. A. Chemical Control of the Charge State of NitrogenVacancy Centers in Diamond. Phys. Rev. B: Condens. Matter Mater. Phys. 2011, 83, 081304. (39) Petrakova, V.; Nesladek, M.; Taylor, A.; Fendrych, F.; Cigler, P.; Ledvina, M.; Vacik, J.; Stursa, J.; Kucka, J. Luminescence Properties of Engineered Nitrogen Vacancy Centers in a Close Surface Proximity. Phys. Status Solidi A 2011, 208, 2051−2056. (40) Leifgen, M.; Schroeder, T.; Gaedeke, F.; Riemann, R.; Metillon, V.; Neu, E.; Hepp, C.; Arend, C.; Becher, C.; Lauritsen, K.; Benson, O. Evaluation of Nitrogen- and Silicon-Vacancy Defect Centres as Single Photon Sources in Quantum Key Distribution. New J. Phys. 2014, 16, 023021. (41) Doherty, M. W.; Dolde, F.; Fedder, H.; Jelezko, F.; Wrachtrup, J.; Manson, N. B.; Hollenberg, L. C. L. Theory of the Ground-State Spin of the NV− Center in Diamond. Phys. Rev. B: Condens. Matter Mater. Phys. 2012, 85, 205203. (42) Ofori-Okai, B. K.; Pezzagna, S.; Chang, K.; Loretz, M.; Schirhagl, R.; Tao, Y.; Moores, B. A.; Groot-Berning, K.; Meijer, J.; Degen, C. L. Spin Properties of Very Shallow Nitrogen Vacancy Defects in Diamond. Phys. Rev. B: Condens. Matter Mater. Phys. 2012, 86, 081406. (43) Feng, F.; Wang, J.; Zhang, W.; Zhang, J.; Lou, L.; Zhu, W.; Wang, G. Efficient Generation of Nanoscale Arrays of NitrogenVacancy Centers With Long Coherence Time in Diamond. Appl. Phys. A: Mater. Sci. Process. 2016, 122, 944. (44) Arroyo-Camejo, S.; Adam, M.-P.; Besbes, M.; Hugonin, J.-P.; Jacques, V.; Greffet, J.-J.; Roch, J.-F.; Hell, S. W.; Treussart, F. Stimulated Emission Depletion Microscopy Resolves Individual Nitrogen Vacancy Centers in Diamond Nanocrystals. ACS Nano 2013, 7, 10912−10919. (45) Pfender, M.; Aslam, N.; Waldherr, G.; Neumann, P.; Wrachtrup, J. Single-Spin Stochastic Optical Reconstruction Microscopy. Proc. Natl. Acad. Sci. U. S. A. 2014, 111, 14669−14674. (46) Arai, K.; Belthangady, C.; Zhang, H.; Bar-Gill, N.; DeVience, S. J.; Cappellaro, P.; Yacoby, A.; Walsworth, R. L. Fourier Magnetic Imaging With Nanoscale Resolution and Compressed Sensing SpeedUp Using Electronic Spins in Diamond. Nat. Nanotechnol. 2015, 10, 859−864. (47) Iwasaki, T.; Hoshino, Y.; Tsuzuki, K.; Kato, H.; Makino, T.; Ogura, M.; Takeuchi, D.; Matsumoto, T.; Okushi, H.; Yamasaki, S.; Hatano, M. Diamond Junction Field-Effect Transistors with Selectively Grown n+-Side Gates. Appl. Phys. Express 2012, 5, 091301. (48) Suwa, T.; Iwasaki, T.; Sato, K.; Kato, H.; Makino, T.; Ogura, M.; Takeuchi, D.; Yamasaki, S.; Hatano, M. Normally-Off Diamond Junction Field-Effect Transistors with Submicron Channel. IEEE Electron Device Lett. 2016, 37, 209−211. (49) Fujihira, T. Theory of Semiconductor Superjunction Devices. Jpn. J. Appl. Phys. 1997, 36, 6254−6262. (50) Einaga, Y.; Foord, J. S.; Swain, G. M. Diamond ElectrodesDiversity and Maturity. MRS Bull. 2014, 39, 525−532. (51) Yang, N.; Uetsuka, H.; Osawa, E.; Nebel, C. E. Vertically Aligned Nanowires from Boron-Doped Diamond. Nano Lett. 2008, 8, 3572−3576. 1244

DOI: 10.1021/acsnano.6b04460 ACS Nano 2017, 11, 1238−1245

Article

ACS Nano (52) Zhu, D.; Zhang, L.; Ruther, R. E.; Hamers, R. J. PhotoIlluminated Diamond as a Solid-State Source of Solvated Electrons in Water for Nitrogen Reduction. Nat. Mater. 2013, 12, 836−841. (53) Demchenko, D. O.; Diallo, I. C.; Reshchikov, M. A. Yellow Luminescence of Gallium Nitride Generated by Carbon Defect Complexes. Phys. Rev. Lett. 2013, 110, 087404. (54) Ishikawa, R.; Lupini, A. R.; Oba, F.; Findlay, S. D.; Shibata, N.; Taniguchi, T.; Watanabe, K.; Hayashi, H.; Sakai, T.; Tanaka, I.; Ikuhara, Y.; Pennycook, S. J. Atomic Structure of Luminescent Centers in High-Efficiency Ce-doped w-AlN Single Crystal. Sci. Rep. 2014, 4, 3778. (55) Tran, T. T.; Bray, K.; Ford, M. J.; Toth, M.; Aharonovich, I. Quantum Emission from Hexagonal Boron Nitride Monolayers. Nat. Nanotechnol. 2015, 11, 37−41. (56) Abtew, T. A.; Gao, W.; Sun, Y. Y.; Zhang, S. B.; Zhang, P. Theory of Oxygen-Boron Vacancy Defect in Cubic Boron Nitride. Phys. Rev. Lett. 2014, 113, 136401. (57) Electrical Insulating Liquids, Bartnikas, R., Ed.; ASTM International, 1994.

1245

DOI: 10.1021/acsnano.6b04460 ACS Nano 2017, 11, 1238−1245