A solution to electric-field screening in diamond quantum electrometers
Abstract
There are diverse interdisciplinary applications for nanoscale resolution electrometry of elementary charges under ambient conditions. These include characterization of 2D electronics, charge transfer in biological systems, and measurement of fundamental physical phenomena. The nitrogen-vacancy center in diamond is uniquely capable of such measurements, however electrometry thus far has been limited to charges within the same diamond lattice. It has been hypothesized that the failure to detect charges external to diamond is due to quenching and surface screening, but no proof, model, or design to overcome this has yet been proposed. In this work we affirm this hypothesis through a comprehensive theoretical model of screening and quenching within a diamond electrometer and propose a solution using controlled nitrogen doping and a fluorine-terminated surface. We conclude that successful implementation requires further work to engineer diamond surfaces with lower surface defect concentrations.
Introduction
Nanoscale charge imaging has been employed for diverse purposes including high-sensitivity biological and chemical sensorsCui et al. (2001); Patolsky et al. (2006), detectors within quantum devicesEizerman et al. (2004), and investigating fundamental physical phenomenaMartin et al. (2008); Yoo et al. (1997). Many techniques have been employed for precision electrometry with nanometer spatial resolutionYoo et al. (1997); Martin et al. (2008); Henning et al. (1995); Williams et al. (1989), elementary charge detectionYoo et al. (1997); Martin et al. (2008); Devoret and Schoelkopf (2000); Schönenberger and Alvarado (1990); Martin et al. (1988); Cleland and Roukes (1998); Bunch et al. (2007); Salfi et al. (2010); Lee et al. (2008), and the ability to operate at ambient temperatures and pressuresBunch et al. (2007); Lee et al. (2008). However, no device yet possess all three of these properties simultaneously. This capability would be extremely valuable for investigating biological systems, such as neuronsBarry et al. (2016); Hanlon et al. (2019), and as a critical characterization tool for the emerging field of two-dimensional electronicsSchwierz (2010); Radisavljevic et al. (2011); Mak and Shan (2016). For example, atomic-resolution imaging of silicene to aid development of a room-temperature transistorTao et al. (2015), probing novel charged quasiparticles in MOS2 filmsMak et al. (2012), and detection of polarization skyrmionsDas et al. (2019).
The nitrogen-vacancy (NV) centerDoherty et al. (2013) is currently the only system capable of nanoscale resolution electrometry of elementary charges under ambient conditions. This point defect in diamond consists of a substitutional nitrogen atom () situated adjacent to a carbon vacancy. Single NV centers have demonstrated room-temperature a.c. (d.c.) electric-field sensitivities reaching 202 (891 )Dolde et al. (2011, 2014). Ensembles of NV centers have achieved shot-noise limited a.c. sensitivities on the order of 1 Chen et al. (2017) and also been employed as in-situ electric field sensors within semiconductor heterojunctionsIwasaki et al. (2017).
The NV center’s proficiency for quantum sensing is due to a unique combination of capabilities. Firstly, the NV exhibits bright optical fluorescence allowing for identification of single defects that can be employed for measurements with nanoscale resolution. Secondly, the NV possesses a mechanism for optical spin initialization and readout which permits spin resonances of individual defects to be measured with high fidelityDoherty et al. (2013). Finally, the NV boasts the longest room-temperature coherence time for any solid-state defectBalasubramanian et al. (2009) allowing for high-resolution detection of spin resonances when combined with optical readout. In addition to electrometry these properties have been applied for precision nano-magnetometryTetienne et al. (2014); Thiel et al. (2016); Häberle et al. (2015); Arai et al. (2015), thermometryKucsko et al. (2013); Neumann et al. (2013); Toyli et al. (2013) and quantum computingWaldherr et al. (2014), as well as proposed for investigating fundamental physical phenomena such as magnetic phase changesCai et al. (2013) and coherent quantum transportOberg et al. (2019).
While the NV center can exist in several charge states, including neutral (NV0) and negative (NV-), only the latter possesses the aforementioned properties needed for electrometry. In particular the spin resonances of the NV- triplet ground state are susceptible to electric field induced changes in its electron spin-spin interactions. The resulting Stark shifts can be detected using optically detected magnetic resonance (ODMR) and used to determine the field magnitude. Furthermore, vector components of the electric field can be measured through rotation of a bias transverse magnetic fieldDolde et al. (2014).
Although single NV centers possess sufficiently high sensitivities, elementary charges external to diamond have not been detected with nanoscale resolution. We hypothesize that this is because of electric field screening and charge-state quenching of NV centers. For the former, recent experimental works have identified multiple screening sources inherent to diamond systems. These include Debye screening from bulk defectsBroadway et al. (2018), charge reorganization in primal sp2 surface defectsStacey et al. (2019) and polarization of adsorbed water vaporMertens et al. (2016). Moreover, p-type defects within bulk diamondLiao et al. (2008), surface acceptor defectsStacey et al. (2019) and surface terminations with negative electron affinitiesHauf et al. (2011) are known to quench the negative charge state, particularly for near-surface NV centersCui and Hu (2013); Ohno et al. (2012). However, the extent that these sources impact charge detection are unknown, and a comprehensive theoretical treatment is needed.
The first aim of this paper is to develop a physical model of screening due to the external environment, internal diamond, and diamond surface. This is performed in Section I where we identify that screening due to charge rearrangement amongst sp2 surface defects is the greatest impediment to electrometry. We propose a solution by saturating the charge traps through a sacrificial -doped layer of . The effectiveness of this idea is explored in Section II in which an analytical toy model is developed for a simplified electrometer. Finally, in Section III this toy model is adapted into a more sophisticated device compatible with NV quantum sensing. The electrostatic properties of the device are modeled computationally, and the physical parameters optimized for charge detection. We conclude that this design successfully mitigates screening for concentrations of sp2 surface defects below approximately , two orders of magnitude lower than that currently demonstrated on fluorine-passivated diamondStacey et al. (2019).
I Screening and quenching within diamond
Screening and quenching effects within diamond can be decomposed into three coupled systems; the external atmosphere, the surface and the internal diamond. These three environments and their associated screening/quenching sources are depicted in Figure 1. In the following subsections we model each system individually and assess the associated impacts for diamond electrometry.

I.1 External atmosphere
We first address screening due to the external atmosphere at ambient temperature and pressure. As air possesses a relative permittivity of approximately unity it causes negligible electrical screening. In contrast, the physisorption of water vapor on the diamond surface may be a detrimental source of screening given water’s high relative permittivity. This subsection reviews the current understanding of wetting on diamond surfaces and applies these results to form a cohesive theory of screening due to water adsorption.
Considering the most common diamond surface terminations, oxygen is known to be strongly hydrophilic and therefore not suitable for precision electrometryWidmann et al. (2014). While hydrogen terminated surfaces exhibit polar hydrophobicityMayrhofer et al. (2016); Mertens et al. (2016), they are not viable for electrometry with near-surface NV centers. This is because they possess a negative electron affinity which introduces subsurface holes that quench the NV- charge stateHauf et al. (2011). However, this is not the case for fluorine-terminated diamond, which is chemically inert and room-temperature stableCui and Hu (2013) with a positive electron affinityRietwyk et al. (2013). Importantly, the fluorine surfaces exhibits strong hydrophobicityWidmann et al. (2014); Kissa (2001), reflected by its high wetting angle and a small physisorption energy of 0.07 eV for F-C(111) as determined using ab-initio calculationsMayrhofer et al. (2016). Ideally these adsorption energies could be used in conjunction with a suitable isotherm equation to determine the water coverage. Unfortunately, this is not possible as current studies neglect Gibbs contributions to adsorption and so cannot accurately predict coverages under ambient conditions.
Regardless, it can be expected that surface coverage is much lower than a monolayer. Simple analysis with the Brunauer-Emmett-Teller isotherm indicates negligible coverage as the condensation energy of water far exceeds that of physisorption (neglecting entropic and enthalpic contributions)Brunauer et al. (1938). Comparisons can also be made to a recent spectroscopic study of water coverage on hydrophobic H-Si(111), which shares a similar wetting angle to F-C(111)Mertens et al. (2016); Widmann et al. (2014); Mayrhofer et al. (2016); Lange et al. (2009); Silvestrelli et al. (2006). This observed that at ambient temperatures and pressures, surface coverage is an increasing function of humidity that never exceeds a monolayerChen et al. (2018). Microscopically, wetting behaviour on hydrophobic surfaces is far more complicated. Graphene templating in conjunction with atomic force microscopy has revealed complex water structures on H-Si(111), with nano-droplets up to 20 nm wide and 0.5 nm in height accumulating on surface defects and step edgesCao et al. (2011).
We now investigate whether adsorbed water produces a major or minor screening effect. Given the ambiguity of the nature and extent of wetting on hydrophobic diamond surfaces, we will model the dielectric permittivity as a function of adsorbed water adlayers. This is denoted by , the fractional amount of water monolayers adsorbed on the surface between 0 and 1. While surface-adsorbed vapor likely possesses a high rate of diffusion at room temperature, here we only consider the time-averaged response of the induced permittivity. As the electrical properties of the F-C(111) surface have been well-characterized theoreticallyMayrhofer et al. (2016), it shall be considered as the model surface for electrometry. However, the results presented here are likely applicable to other cuts of fluorine-terminated diamond surfaces as well. The polar C-F surface bonds induce an electric field which orientates the dipoles of the physisorbed water. This generates a net polarization in the direction of the C-F dipoles which can be calculated at thermal equilibrium. The polarization density is given byAtkins and de Paula (2014)
(1) |
where D is the dipole moment of water, , is the electric field generated by the diamond surface and
(2) |
for the number density of water molecules and the Langevin function given by
A value for may be estimated by treating each F as a single adsorption site. This is justified as ab-initio calculations indicate that the energy of a physisorbed molecule is minimized when aligned laterally with a terminating F atomMayrhofer et al. (2016). In the direction perpendicular to the surface (denoted ) we may bound the linear density of water molecules by that of liquid water, yielding .
We now consider the electrostatic response of the dipolar water layer to some perturbing field generated by the charged source, . Denoting the electric field generated by the C-F surface dipoles as , the total electric field can be written as
The polarization of the water layer due to can be characterized through a first order Taylor expansion of equation (1) about as
(3) |
where is the linear electric susceptibility tensor given by
Consequently, may be written as the sum of an isotropic () and anisotropic () component as
(4) |
where
Therefore the change in polarization of the adlayer induced by is given by
(5) |
Taking V/Å derived from ab-initio calculationsMayrhofer et al. (2016), we find that at 300 K and therefore the anisotropic term may be neglected. The isotropic susceptibility is presented as a linear function of surface coverage in Figure 2 where we obtain for . The susceptibility of the water adlayer is therefore relatively large but diminished in respect to liquid water. This is because the ability of the adlayer molecules to re-orientate in response to an electric field is constrained by the surface. In the following section these susceptibilities will be use to model the screening fraction due to both the water adlayer and bulk diamond.

I.2 Internal diamond
Debye screening due to internal defects presents a major impediment to diamond electrometry. This form of screening occurs when charges are free to rearrange within a spatial continuum of donors and acceptorsRidley (2013). This causes external fields to decay exponentially within the lattice as they are counteracted by the induced response of the charges. Within diamond, uncontrolled N defects are a common and potent source of Debye screening with a characteristic decay length of 15 nm at low doping concentrationsBroadway et al. (2018). In addition to screening, p-type defects such as boron are detrimental to electrometry as they introduce holes which quench the NV- centerLiao et al. (2008). Consequently, only pure diamond is compatible with precision electrometry in which isotropic polarizability is the sole source of internal field decay.
Ignoring presently the impact of surface defects, the magnitude of screening due to the adsorbed water layer and pure diamond can be determined analytically. As depicted in the inset of Figure 3, a basic electrometer may be modeled by three stacked planar dielectrics consisting of a thin water layer sandwiched between diamond and air. If an elementary charge is placed above the surface, the electrostatic problem can be solved using the method of imagesBarrera et al. (1978); Pont and Serra (2015). This technique solves Poisson’s equation by introducing an infinite series of fictitious ‘image’ charges which reproduce the boundary conditions of the dielectric stack.
In Figure 3 we present the electric field screening ratio of a point charge located at varying heights above the surface. This represents the magnitude of the screened field relative to the unscreened field sampled at a depth of 10 nm into the lattice, corresponding to a possible location of a near-surface NV center. A lower estimate for the thickness of the water layer is taken to be 3 Å, approximately the Van der Waals radius of a single molecule. We obtain screening fractions of , independent of surface coverage and the height of the charge above the surface. This is comparable to the screening ratio purely due to the dielectric response of diamond and indicates that adsorption of water vapor is not detrimental to nanoscale electrometry. While a 70% reduction in field strength appears considerable, this is unavoidable, and minor compared to screening induced by surface charge rearrangement as discussed below.

I.3 Surface
Uncontrolled, diamond surface defects present a detrimental source of screening and charge quenching for NV-based electrometers. Recent work has identified a family of primal sp2 defects universal to all diamond surface terminationsStacey et al. (2019). These introduce acceptor states into diamond approximately 2.2 eV above the valence band which readily quench the NV- charge state at 2.9 eVAslam et al. (2013). Furthermore, partial occupation of these traps results in intense field screening through re-arrangement of surface charge. For fluorine terminated diamond, an sp2 surface trap density of 4% has been observed following passivation using SF6 plasmaStacey et al. (2019). This corresponds to a trap concentration of roughly , comparable to the surface density of free electrons in copper. While the mechanisms for conductivity differ in these materials (i.e., localized defects vs. conduction electrons), such high trap concentrations in diamond effectively render the surface conducting under partial occupation. Surface screening therefore presents the greatest impediment to diamond electrometry and must be addressed in any realistic device.
Uncontrolled, diamond surface defects present a detrimental source of screening and charge quenching for NV-based electrometers. Recent work has identified a family of primal sp2 defects universal to all diamond surface terminationsStacey et al. (2019). These introduce acceptor states into diamond approximately 2.2 eV above the valence band which readily quench the NV- charge state at 2.9 eVAslam et al. (2013). Furthermore, partial occupation of these traps results in intense field screening through permitting unimpeded rearrangement of surface charge. There is an analogy to a conductor where free carriers may also rearrange to screen an external field. The screening strength is determined by the accessible density of charges that can rearrange. For fluorine terminated diamond, an sp2 surface trap density of 4% has been observed following passivation using SF6 plasmaStacey et al. (2019), corresponding to a trap concentration of roughly . To draw the analogy with a conductor, this is a similar density of free charges as copper. Hence such a diamond surface with a high density of surface traps can be considered as a conductor that screens external electric fields. Surface defects therefore present the greatest impediment to diamond electrometry and must be addressed in any realistic device.
II Toy model electrometer
A solution to both surface-induced screening and charge instability is to saturate the surface traps using sacrificial donors within the diamond. We propose fabricating a -doped layer of which introduces a donor level 3.8 eV above the valence band. In the limit that the concentration of exceeds that of the sp2 defects, the Fermi-level will be pinned to the donor level and prevent quenching of the NV-. This is possible as present doping techniques allow for precision control of the -layer height and defect concentrations up to 1000 ppmOhno et al. (2012); Chandran et al. (2016). In this section we demonstrate the effectiveness of this idea through an analytical toy model that explores the complex interactions of screening and quenching within a highly coupled system.
Consider the schematic for a simplified electrometer presented in Figure 4 (a). The -doped layer is positioned at a depth below the diamond surface () while the NV spin-probe is placed between them at a depth . At electrostatic equilibrium the occupation of the sp2 defects leads to the accumulation of an isotropic and homogeneous charge density on the surface, . This charge density subsequently generates a surface potential which is related as per
(6) |
where is the device capacitance, is the electron charge, is the density of surface traps, is the Fermi-Dirac distribution, eV is the energy of and eV is the energy of an sp2 surface defectStacey et al. (2019). The two layers – surface and -doping – effectively form a parallel plate capacitor and hence where is the dielectric permittivity of diamond. This induces a linear potential between the capacitor plates such that the potential at the NV center is given by
(7) |

The magnitude of the potential induced at the surface has major ramifications for NV- charge stability. Figure 4 (b) depicts the energies of , NV and sp2 defects within the simplified electrometer. Upon charging, the sp2 defect and NV energies are raised by an amount and respectively. To avoid transition to NV0, the condition
(8) |
must be maintained. Inserting equation (7) into the inequality (8) places a constraint on the maximum surface potential that prevents quenching, given by
(9) |
The surface potential also has major implications for screening. Note that is ultimately limited above by eV. When the defect energy is pinned to that of the donors and screening effects dominate; equation (6) indicates that and hence the surface is effectively conducting. Clearly, electrometry requires that
(10) |
for under such conditions (the linear regime) and the surface charge cannot reorganize in response to an external electric field. Fortunately, the inequality (10) can be determined precisely and the robustness of the linear regime to charge screening can be demonstrated quantitatively.
Consider the screening field induced by a perturbing potential at the diamond surface, . Within the linear regime we have that
(11) |
where we have denoted and expanded to first order in . Suppose that the perturbing potential is due to a point charge positioned a height above the electrometer surface and aligned with the NV center. Then equation 11 indicates that the induced charge density is given by
(12) |
as a function of radial distance from the point charge . This produces a screening field at the NV center given by
(13) |
Considering that the field generated by the point charge at a depth in the absence of screening is
(14) |
the magnitude of the screening ratio is given by
(15) |
Equation 15 demonstrates a further bound on the maximum surface potential compatible with electrometry. Taking nm and we obtain a screening ratio of . Screening effects will therefore dominate if the surface potential is not sufficiently deep within the linear regime. In this specific example, a ratio only occurs for eV at room temperature.
To summarize, our toy model has identified the fundamental limitations to precision electrometry. The potential of the surface must be controlled such that all charge traps are saturated while the NV energy is maintained below the Fermi level. In general, one should aim to reduce to avoid the detrimental effects of screening and quenching. Equation 6 indicates that the surface potential is a function of only two variables; the capacitance and the density of charge traps. For the simplified electrometer design, within the linear regime and therefore should be minimized. However, is limited below by physical constraints such as the thickness of the -doped layer and its proximity to the surface and NV center. Hence the surface potential is largely dictated by the density of surface traps. Whereas the capacitance can be controlled, surface defects are an undesirable byproduct of diamond surface passivationStacey et al. (2019).
The performance of a diamond electrometer is strongly dependent on the surface trap density. This relationship is demonstrated in Figure 5 where nm, nm and K have been chosen as a realistic example of device parameters. Three different electrometer operating regions can be distinguished. Region (i) represents values of which correspond to surface voltages within the linear regime. Here electrometry is viable as the charge traps are fully saturated and the NV maintains its negative charge state. Region (ii) also represents values of for which the NV- center remains charge stable. However, the sp2 defects are only partially occupied and so electrometry is impossible due to surface screening. Similarly, in region (iii) is so large that the NV center has been quenched.

Consequently, the presence of a -doped layer can simultaneously maintain NV- stability and prevent surface screening within the linear regime. However, the capabilities of the device are fundamentally limited by the density of surface traps. As capacitance is a geometric property, equation (6) indicates that this result is universal to any device which employs donors to saturate surface defects. The parallel plate capacitor has one of the greatest capacitances over small length scales and hence Figure 5 demonstrates that electrometry is only compatible with surface densities on the order of . This is two orders of magnitude lower than defect densities currently observed on fluorine terminated surfaces passivated using SF6 plasma. However, fluorine is a relatively new surface termination and many alternative passivation techniques exist and continue to be developedCui and Hu (2013); Rietwyk et al. (2013); Salvadori et al. (2010). The sp2 defect yields on these surfaces are yet to be characterized and may well be low enough to permit precision electrometry.
III Electrometer design
The simple electrometer presented in the previous section has several deficiencies which make it impractical for quantum sensing. Fortunately, these can be overcome with a simple modification to the electrometer design. In this section we discuss the shortcomings of the toy-model electrometer and their solutions, culminating in the presentation of an effective and physically realizable device.
Initialization and readout of the NV center requires optical control using a 532 nm green laserDoherty et al. (2013); Dolde et al. (2011, 2014). As depicted in Figure 6, performing electrometry with the toy device requires the optical path to pass through the -doped layer. This introduces several complications. Firstly, layers typically contain some density of erroneous NV centers. These are capable of producing background counts during read-out which decrease measurement contrast and lead to lower sensitivity. Secondly, the 532 nm laser ionizes defects to form which modulates the charge density within the -doped layerIakoubovskii and Adriaenssens (2000). During the optical steady state this results in a local positive potential which reduces charge stability of the center and further diminishes signal contrast. Moreover, following initialization of the NV- spin-state the laser is deactivated and the induced charge relaxes back to the equilibrium it adopted before the laser was turned on. If this relaxation time is slow ( ns) compared to the NV sensing period ( s)Dolde et al. (2011, 2014), this relaxation will influence the sensing measurement. Ionization of also introduces free charge carriers into the lattice, and hence there also exists a low probability of Auger electrons scattering against the NV- and causing quenching. Finally, the -doped layer acts as a spin-bath which can cause decoherence of the NV spin at close proximitiesBalasubramanian et al. (2009).

One possible solution to these issues is to spatially separate the layer and the NV center. The optical focus can then be maintained on the NV while the -doping is subject to a negligible laser intensity. We estimate this would require a separation distance of approximately m. However, this severely limits the sensitivity of the spin-probe to external charges as equation (9) necessitates that m and therefore m to maintain charge stability. A more practical solution is to introduce of a hole within the -doped layer. Consider the schematic presented in Figure 7 in which a disk of pure diamond has been fabricated around the NV center. This hole permits optical access to the spin probe while simultaneously minimizing the number of ionized defects. Furthermore, the hole reduces the probability of optically addressing multiple NV centers and so increases the yield of electrometer fabrication.

The capabilities of this realistic electrometer design for elementary charge detection were investigated using COMSOL Multiphysics software. The potential at the NV center and surface as well as the occupation of charge traps were simulated as a function of the device parameters; the NV depth , layer depth and hole radius . This was achieved by solving Poisson’s equation self-consistently for a surface charge density given by equation 6 and assuming a grounded -doped layer. The device parameters were then optimized to identify the greatest possible trap density compatible with precision electrometry. As discussed in Section II, these criterion are charge stability of the NV ( eV) and a surface potential which leads to less than 1% surface screening as per equation (15).
Figure 8 presents the largest viable surface trap density as a function of the -doping depth and hole radius. Trap densities greater than those presented lead to field screening in excess of 1%. For the parameters sampled here ( nm and nm) we find that electrometry is feasible for sp2 surface densities within the regime of and that no charge quenching occurs for NV centers level with the -doped layer. Devices with smaller radii and -doping depth possess a greater capacitance and are therefore compatible with larger defect densities. Note that the optical spot size is diffraction limited by nm, and hence some degree of ionization will occur for hole radii less than 100 nm. The impact of the induced positive charge density for NV quantum sensing is unknown and left as an avenue for future work. If the effects are significantly detrimental, Figure 8 indicates that the hole size can simply be increased beyond the optical spot size without drastically reducing the achievable sp2 defect concentrations.

IV Conclusion
Sub-nanometer resolution electrometry of elementary charges under ambient conditions would allow for investigation of diverse electrical phenomenon ranging from biological systems to fundamental physics. The NV center is the only known system capable of such a feat, however measurements are currently limited to charges internal to diamond. In this paper we have applied theoretical modeling to conclusively demonstrate that external charge detection is not yet feasible due to field screening. While screening due to the atmosphere and internal defects can be mitigated using fluorine-passivated and ultra-pure diamond, electrometry is ultimately frustrated by charge rearrangement amongst surface defects.
We have proposed a solution to surface screening through introduction of a sacrificial -doped layer. Fabrication of a deficit hole surrounding the NV center allows for optical access to the spin probe while minimizing the read-out of erroneous NV centers and ionization of free charges. This electrometry device is technologically feasible and computational simulations have demonstrated that it can successfully mitigate screening effects for surface trap densities up to . Although this is two orders of magnitude below currently observed sp2 defect densities on fluorine-terminated diamond, the outcome of this work is a clear pathway towards nanoscale imaging of external charges at ambient conditions. Electrometry cannot be achieved until surface passivation technologies realize lower defect concentrations on fluorine-terminated diamond.
Acknowledgments
We acknowledge funding from the Australian Research Council (DP170102735). We thank Patrick Maletinsky for providing insightful feedback on this manuscript.
References
- Cui et al. (2001) Y. Cui, Q. Wei, H. Park, and C. M. Lieber, Nanowire nanosensors for highly sensitive and selective detection of biological and chemical species, Science 293, 1289 (2001).
- Patolsky et al. (2006) F. Patolsky, G. Zheng, and C. M. Lieber, Fabrication of silicon nanowire devices for ultrasensitive, label-free, real-time detection of biological and chemical species, Nature Protocols 1, 1711 (2006).
- Eizerman et al. (2004) J. M. Eizerman, R. Hanson, L. H. Van Beveren, B. Witkamp, L. M. Vandersypen, and L. P. Kouwenhoven, Single-shot read-out of an individual electron spin in a quantum dot, Nature 430, 431 (2004).
- Martin et al. (2008) J. Martin, N. Akerman, G. Ulbricht, T. Lohmann, J. H. Smet, K. Von Klitzing, and A. Yacoby, Observation of electron-hole puddles in graphene using a scanning single-electron transistor, Nature Physics 4, 144 (2008).
- Yoo et al. (1997) M. J. Yoo, T. A. Fulton, H. F. Hess, R. L. Willett, L. N. Dunkleberger, R. J. Chichester, L. N. Pfeiffer, and K. W. West, Scanning single-electron transistor microscopy: Imaging individual charges, Science 276, 579 (1997).
- Henning et al. (1995) A. K. Henning, T. Hochwitz, J. Slinkman, J. Never, S. Hoffmann, P. Kaszuba, and C. Daghlian, Two-dimensional surface dopant profiling in silicon using scanning Kelvin probe microscopy, Journal of Applied Physics 77, 1888 (1995).
- Williams et al. (1989) C. C. Williams, J. Slinkman, W. P. Hough, and H. K. Wickramasinghe, Lateral dopant profiling with 200 nm resolution by scanning capacitance microscopy, Applied Physics Letters 55, 1662 (1989).
- Devoret and Schoelkopf (2000) M. H. Devoret and R. J. Schoelkopf, Amplifying quantum signals with the single-electron transistor, Nature 406, 1039 (2000).
- Schönenberger and Alvarado (1990) C. Schönenberger and S. F. Alvarado, Observation of single charge carriers by force microscopy, Physical Review Letters 65, 3162 (1990).
- Martin et al. (1988) Y. Martin, D. W. Abraham, and H. K. Wickramasinghe, High-resolution capacitance measurement and potentiometry by force microscopy, Applied Physics Letters 52, 1103 (1988).
- Cleland and Roukes (1998) A. N. Cleland and M. L. Roukes, A nanometre-scale mechanical electrometer, Nature 392, 160 (1998).
- Bunch et al. (2007) J. S. Bunch, A. M. Van Der Zande, S. S. Verbridge, I. W. Frank, D. M. Tanenbaum, J. M. Parpia, H. G. Craighead, and P. L. McEuen, Electromechanical resonators from graphene sheets, Science 315, 490 (2007).
- Salfi et al. (2010) J. Salfi, I. G. Savelyev, M. Blumin, S. V. Nair, and H. E. Ruda, Direct observation of single-charge-detection capability of nanowire field-effect transistors, Nature Nanotechnology 5, 737 (2010).
- Lee et al. (2008) J. Lee, Y. Zhu, and A. Seshia, Room temperature electrometry with SUB-10 electron charge resolution, Journal of Micromechanics and Microengineering 18, 10.1088/0960-1317/18/2/025033 (2008).
- Barry et al. (2016) J. F. Barry, M. J. Turner, J. M. Schloss, D. R. Glenn, Y. Song, M. D. Lukin, H. Park, and R. L. Walsworth, Optical magnetic detection of single-neuron action potentials using quantum defects in diamond, Proceedings of the National Academy of Sciences of the United States of America 113, 14133 (2016).
- Hanlon et al. (2019) L. Hanlon, V. Gautam, J. D. A. Wood, P. Reddy, M. S. J. Barson, M. Niihori, A. R. J. Silalahi, B. Corry, J. Wrachtrup, M. J. Sellars, V. R. Daria, P. Maletinsky, G. J. Stuart, and M. W. Doherty, Diamond nano-pillar arrays for quantum microscopy of neuronal signals (2019), arXiv:1901.08743 [quant-ph] .
- Schwierz (2010) F. Schwierz, Graphene transistors, Nature Nanotechnology 5, 487 (2010).
- Radisavljevic et al. (2011) B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti, and A. Kis, Single-layer MoS2 transistors, Nature Nanotechnology 6, 147 (2011).
- Mak and Shan (2016) K. F. Mak and J. Shan, Photonics and optoelectronics of 2D semiconductor transition metal dichalcogenides, Nature Photonics 10, 216 (2016).
- Tao et al. (2015) L. Tao, E. Cinquanta, D. Chiappe, C. Grazianetti, M. Fanciulli, M. Dubey, A. Molle, and D. Akinwande, Silicene field-effect transistors operating at room temperature, Nature Nanotechnology 10, 227 (2015).
- Mak et al. (2012) K. F. Mak, K. He, C. Lee, G. H. Lee, J. Hone, T. F. Heinz, and J. Shan, Tightly bound trions in monolayer MoS2, Nature Materials 12, 207 (2012).
- Das et al. (2019) S. Das, Y. L. Tang, Z. Hong, M. A. Gonçalves, M. R. McCarter, C. Klewe, K. X. Nguyen, F. Gómez-Ortiz, P. Shafer, E. Arenholz, V. A. Stoica, S. L. Hsu, B. Wang, C. Ophus, J. F. Liu, C. T. Nelson, S. Saremi, B. Prasad, A. B. Mei, D. G. Schlom, J. Íñiguez, P. García-Fernández, D. A. Muller, L. Q. Chen, J. Junquera, L. W. Martin, and R. Ramesh, Observation of room-temperature polar skyrmions, Nature 568, 368 (2019).
- Doherty et al. (2013) M. W. Doherty, N. B. Manson, P. Delaney, F. Jelezko, J. Wrachtrup, and L. C. Hollenberg, The nitrogen-vacancy colour centre in diamond, Physics Reports 528, 1 (2013), arXiv:1302.3288 .
- Dolde et al. (2011) F. Dolde, H. Fedder, M. W. Doherty, T. Nöbauer, F. Rempp, G. Balasubramanian, T. Wolf, F. Reinhard, L. C. Hollenberg, F. Jelezko, and J. Wrachtrup, Electric-field sensing using single diamond spins, Nature Physics 7, 459 (2011).
- Dolde et al. (2014) F. Dolde, M. W. Doherty, J. Michl, I. Jakobi, B. Naydenov, S. Pezzagna, J. Meijer, P. Neumann, F. Jelezko, N. B. Manson, and J. Wrachtrup, Nanoscale detection of a single fundamental charge in ambient conditions using the NV - Center in diamond, Physical Review Letters 112, 10.1103/PhysRevLett.112.097603 (2014).
- Chen et al. (2017) E. H. Chen, H. A. Clevenson, K. A. Johnson, L. M. Pham, D. R. Englund, P. R. Hemmer, and D. A. Braje, High-sensitivity spin-based electrometry with an ensemble of nitrogen-vacancy centers in diamond, Physical Review A 95, 10.1103/PhysRevA.95.053417 (2017).
- Iwasaki et al. (2017) T. Iwasaki, W. Naruki, K. Tahara, T. Makino, H. Kato, M. Ogura, D. Takeuchi, S. Yamasaki, and M. Hatano, Direct Nanoscale Sensing of the Internal Electric Field in Operating Semiconductor Devices Using Single Electron Spins, ACS Nano 11, 1238 (2017).
- Balasubramanian et al. (2009) G. Balasubramanian, P. Neumann, D. Twitchen, M. Markham, R. Kolesov, N. Mizuochi, J. Isoya, J. Achard, J. Beck, J. Tissler, V. Jacques, P. R. Hemmer, F. Jelezko, and J. Wrachtrup, Ultralong spin coherence time in isotopically engineered diamond, Nature Materials 8, 383 (2009).
- Tetienne et al. (2014) J. P. Tetienne, T. Hingant, J. V. Kim, L. Herrera Diez, J. P. Adam, K. Garcia, J. F. Roch, S. Rohart, A. Thiaville, D. Ravelosona, and V. Jacques, Nanoscale imaging and control of domain-wall hopping with a nitrogen-vacancy center microscope, Science 344, 1366 (2014).
- Thiel et al. (2016) L. Thiel, D. Rohner, M. Ganzhorn, P. Appel, E. Neu, B. Müller, R. Kleiner, D. Koelle, and P. Maletinsky, Quantitative nanoscale vortex imaging using a cryogenic quantum magnetometer, Nature Nanotechnology 11, 677 (2016), arXiv:1511.02873 .
- Häberle et al. (2015) T. Häberle, D. Schmid-Lorch, F. Reinhard, and J. Wrachtrup, Nanoscale nuclear magnetic imaging with chemical contrast, Nature Nanotechnology 10, 125 (2015).
- Arai et al. (2015) K. Arai, C. Belthangady, H. Zhang, N. Bar-Gill, S. J. DeVience, P. Cappellaro, A. Yacoby, and R. L. Walsworth, Fourier magnetic imaging with nanoscale resolution and compressed sensing speed-up using electronic spins in diamond, Nature Nanotechnology 10, 859 (2015).
- Kucsko et al. (2013) G. Kucsko, P. C. Maurer, N. Y. Yao, M. Kubo, H. J. Noh, P. K. Lo, H. Park, and M. D. Lukin, Nanometre-scale thermometry in a living cell, Nature 500, 54 (2013), arXiv:1304.1068 .
- Neumann et al. (2013) P. Neumann, I. Jakobi, F. Dolde, C. Burk, R. Reuter, G. Waldherr, J. Honert, T. Wolf, A. Brunner, J. H. Shim, D. Suter, H. Sumiya, J. Isoya, and J. Wrachtrup, High-precision nanoscale temperature sensing using single defects in diamond, Nano Letters 13, 2738 (2013), arXiv:1304.0688 .
- Toyli et al. (2013) D. M. Toyli, C. F. De Las Casas, D. J. Christle, V. V. Dobrovitski, and D. D. Awschalom, Fluorescence thermometry enhanced by the quantum coherence of single spins in diamond, Proceedings of the National Academy of Sciences of the United States of America 110, 8417 (2013).
- Waldherr et al. (2014) G. Waldherr, Y. Wang, S. Zaiser, M. Jamali, T. Schulte-Herbrüggen, H. Abe, T. Ohshima, J. Isoya, J. F. Du, P. Neumann, and J. Wrachtrup, Quantum error correction in a solid-state hybrid spin register, Nature 506, 204 (2014), arXiv:1309.6424 .
- Cai et al. (2013) J. Cai, A. Retzker, F. Jelezko, and M. B. Plenio, A large-scale quantum simulator on a diamond surface at room temperature, Nature Physics 9, 168 (2013), arXiv:1208.2874 .
- Oberg et al. (2019) L. M. Oberg, E. Huang, P. M. Reddy, A. Alkauskas, A. D. Greentree, J. H. Cole, N. B. Manson, C. A. Meriles, and M. W. Doherty, Spin coherent quantum transport of electrons between defects in diamond, Nanophotonics 8, 1975 (2019).
- Broadway et al. (2018) D. A. Broadway, N. Dontschuk, A. Tsai, S. E. Lillie, C. T.-K. Lew, J. C. McCallum, B. C. Johnson, M. W. Doherty, A. Stacey, L. C. L. Hollenberg, and J.-P. Tetienne, Spatial mapping of band bending in semiconductor devices using in situ quantum sensors, Nature Electronics 1, 502 (2018).
- Stacey et al. (2019) A. Stacey, N. Dontschuk, J. P. Chou, D. A. Broadway, A. K. Schenk, M. J. Sear, J. P. Tetienne, A. Hoffman, S. Prawer, C. I. Pakes, A. Tadich, N. P. de Leon, A. Gali, and L. C. Hollenberg, Evidence for Primal sp 2 Defects at the Diamond Surface: Candidates for Electron Trapping and Noise Sources, Advanced Materials Interfaces 6, 10.1002/admi.201801449 (2019).
- Mertens et al. (2016) M. Mertens, M. Mohr, K. Brühne, H. J. Fecht, M. Łojkowski, W. Świȩszkowski, and W. Łojkowski, Patterned hydrophobic and hydrophilic surfaces of ultra-smooth nanocrystalline diamond layers, Applied Surface Science 390, 526 (2016).
- Liao et al. (2008) M. Liao, Y. Koide, J. Alvarez, M. Imura, and J. P. Kleider, Persistent positive and transient absolute negative photoconductivity observed in diamond photodetectors, Physical Review B - Condensed Matter and Materials Physics 78, 10.1103/PhysRevB.78.045112 (2008).
- Hauf et al. (2011) M. V. Hauf, B. Grotz, B. Naydenov, M. Dankerl, S. Pezzagna, J. Meijer, F. Jelezko, J. Wrachtrup, M. Stutzmann, F. Reinhard, and J. A. Garrido, Chemical control of the charge state of nitrogen-vacancy centers in diamond, Physical Review B - Condensed Matter and Materials Physics 83, 10.1103/PhysRevB.83.081304 (2011).
- Cui and Hu (2013) S. Cui and E. L. Hu, Increased negatively charged nitrogen-vacancy centers in fluorinated diamond, Applied Physics Letters 103, 10.1063/1.4817651 (2013).
- Ohno et al. (2012) K. Ohno, F. Joseph Heremans, L. C. Bassett, B. A. Myers, D. M. Toyli, A. C. Bleszynski Jayich, C. J. Palmstrøm, and D. D. Awschalom, Engineering shallow spins in diamond with nitrogen delta-doping, Applied Physics Letters 101, 10.1063/1.4748280 (2012).
- Mayrhofer et al. (2016) L. Mayrhofer, G. Moras, N. Mulakaluri, S. Rajagopalan, P. A. Stevens, and M. Moseler, Fluorine-Terminated Diamond Surfaces as Dense Dipole Lattices: The Electrostatic Origin of Polar Hydrophobicity, Journal of the American Chemical Society 138, 4018 (2016).
- Kissa (2001) E. Kissa, Fluorinated Surfactants and Repellents, 2nd ed., Vol. 97 (CRC Press, 2001).
- Rietwyk et al. (2013) K. J. Rietwyk, S. L. Wong, L. Cao, K. M. Odonnell, L. Ley, A. T. Wee, and C. I. Pakes, Work function and electron affinity of the fluorine-terminated (100) diamond surface, Applied Physics Letters 102, 10.1063/1.4793999 (2013).
- Ridley (2013) B. K. Ridley, Quantum Processes in Semiconductors, Journal of Chemical Information and Modeling 53, 1689 (2013), arXiv:arXiv:1011.1669v3 .
- Widmann et al. (2014) C. J. Widmann, C. Giese, M. Wolfer, S. Kono, and C. E. Nebel, F-and Cl-terminations of (100) oriented single crystalline diamond, Physica Status Solidi (A) Applications and Materials Science 211, 2328 (2014).
- Brunauer et al. (1938) S. Brunauer, P. H. Emmett, and E. Teller, Adsorption of Gases in Multimolecular Layers, Journal of the American Chemical Society 60, 309 (1938).
- Lange et al. (2009) B. Lange, R. Posner, K. Pohl, C. Thierfelder, G. Grundmeier, S. Blankenburg, and W. G. Schmidt, Water adsorption on hydrogenated Si(1 1 1) surfaces, Surface Science 603, 60 (2009).
- Silvestrelli et al. (2006) P. L. Silvestrelli, F. Toigo, and F. Ancilotto, Interfacial water on Cl- and H-terminated Si(111) surfaces from first-principles calculations, Journal of Physical Chemistry B 110, 12022 (2006).
- Chen et al. (2018) L. Chen, X. He, H. Liu, L. Qian, and S. H. Kim, Water Adsorption on Hydrophilic and Hydrophobic Surfaces of Silicon, Journal of Physical Chemistry C 122, 11385 (2018).
- Cao et al. (2011) P. Cao, K. Xu, J. O. Varghese, and J. R. Heath, The microscopic structure of adsorbed water on hydrophobic surfaces under ambient conditions, Nano Letters 11, 5581 (2011).
- Atkins and de Paula (2014) P. Atkins and J. de Paula, Atkins’ Physical chemistry, 10th ed. (Oxford University Press, 2014).
- Barrera et al. (1978) R. G. Barrera, O. Guzmán, and B. Balaguer, Point charge in a three-dielectric medium with planar interfaces, American Journal of Physics 46, 1172 (1978).
- Pont and Serra (2015) F. M. Pont and P. Serra, Comment on “Point charge in a three-dielectric medium with planar interfaces” [Am. J. Phys. 46 , 1172–1179 (1978)] , American Journal of Physics 83, 475 (2015).
- Aslam et al. (2013) N. Aslam, G. Waldherr, P. Neumann, F. Jelezko, and J. Wrachtrup, Photo-induced ionization dynamics of the nitrogen vacancy defect in diamond investigated by single-shot charge state detection, New Journal of Physics 15, 10.1088/1367-2630/15/1/013064 (2013), arXiv:1209.0268 .
- Chandran et al. (2016) M. Chandran, S. Michaelson, C. Saguy, and A. Hoffman, Fabrication of a nanometer thick nitrogen delta doped layer at the sub-surface region of (100) diamond, Applied Physics Letters 109, 10.1063/1.4971312 (2016).
- Salvadori et al. (2010) M. C. Salvadori, W. W. Araújo, F. S. Teixeira, M. Cattani, A. Pasquarelli, E. M. Oks, and I. G. Brown, Termination of diamond surfaces with hydrogen, oxygen and fluorine using a small, simple plasma gun, Diamond and Related Materials 19, 324 (2010).
- Iakoubovskii and Adriaenssens (2000) K. Iakoubovskii and G. J. Adriaenssens, Optical transitions at the substitutional nitrogen centre in diamond, Journal of Physics Condensed Matter 12, 10.1088/0953-8984/12/6/102 (2000).