The quantum efficiency and mean transverse energy of electrons emitted from a cathode determine the quality of beams generated from photoinjectors. The nitrogen-incorporated ultrananocrystalline diamond, (N)UNCD, is a new class of robust semiconductor photocathodes, which has been considered in photoinjectors for high peak current extraction. In this work, we measure the spectral response in quantum efficiency, photoemission energy spectra, and mean transverse energy of the (N)UNCD photocathode using a photoemission electron microscope. The observed quantum efficiency was comparable to that of copper photocathodes. Photoemission spectra showed the evidence of scattering of electrons before emission. This relaxation of electrons due to scattering is also observed in the spectral response of the mean transverse energy. The mean transverse energy is limited to 70 meV at the threshold. We attribute this to the physical and chemical roughness of the (N)UNCD photocathode and, hence, smoother films will be required to further reduce the mean transverse energy obtained from the (N)UNCD photocathode.

High-density bunched electron beams required for accelerators in X-ray Free Electron Lasers (XFELs),1 Ultrafast Electron Diffraction and Microscopy experiments,2 Inverse Compton Scattering X-ray sources,3 etc. are typically produced using photoinjectors. Photoinjectors essentially consist of a photoemissive material (photocathode) placed in a large accelerating field, which is typically in the range of 5–100 MV/m depending on the design of the photoinjector. The performance of a photocathode is usually characterized by the emitted charge or, equivalently, the quantum efficiency (QE). The cathode must also have a long lifetime, fast response time, and low intrinsic emittance. QE is defined as the number of electrons emitted per incident photon. High-QE photocathodes are desirable for applications that require high average current or high bunch charges and to mitigate the effects of non-linear photoemission.

The response time of a photocathode is given in terms of the extracted electron bunch length when compared to the incoming laser pulse. Photocathodes require a fast response time to temporally shape the initial electron distribution and maintain the beam quality during acceleration in the photoinjector. A fast response time implies that the electrons are emitted quickly after the absorption of incident photon energy. However, electrons excited deep within the cathode surface can take tens of pico-seconds to escape into the vacuum leading to undesirable temporal structure. A fast response time (<1 ps) is desirable for various photoinjector applications.4 

The intrinsic emittance of a photocathode is expressed in terms of rms laser spot size and mean transverse energy (MTE) of the emitted electrons and is given by the following equation:

(1)

where εn,x is the normalized transverse emittance in the x–z plane, σl,x is the rms laser spot size in the x-direction, me is the mass of an electron, and c is the speed of light. The MTE of the electrons depends on the cathode material, its surface morphology, and the laser photon energy and fluence used for excitation.5 For the case of metal photocathodes, the MTE of the emitted electrons follows the Dowell–Schmerge model, which says that the MTE of the electrons scales as one-third of the excess energy (difference between the photon energy and work function) and at the threshold is limited by the thermal limit.6 For example, the smallest MTE of 25 meV was demonstrated for antimony films when operated near the threshold at room temperature.7 For the semiconductor photocathode such as GaAs activated to negative electron affinity using Cs and NF3, the observed value of MTE is 40 meV at room temperature.8 However, in most photoinjectors, the MTE extracted from photocathodes is in the range of 0.1–1 eV.5 

Another parameter that determines the MTE of the emitted electron beam is the physical roughness and work function variation (also termed as chemical roughness) of a photocathode. Nano-scale physical roughness on the cathode results in an increase in MTE and this can be attributed to two effects. First is the “slope effect” where the electrons are emitted from the local surface normal of the photocathode as opposed to its global surface normal. The second, known as the “field effect,” arises due to the generation of a transverse local electric field due to the roughness on the photocathode surface. In general, the field effect dominates as compared to the slope effect and ergo the slope effect is usually ignored in the analysis. A detailed analysis of MTE increase due to physical roughness has been previously reported.9–12 Assuming the sinusoidal variation in the surface roughness profile, the scaling of MTE to the first order is given by the following equation:

(2)

where E is the applied electric field, a is the amplitude, and λ is the period of the physical roughness profile. As we can see, the contribution to MTE due to physical roughness increases with increasing accelerating field gradient. The chemical roughness arises due to lateral potential variations, which is the result of the varying work function. It has been shown that the effect of chemical roughness becomes less prominent with increasing accelerating field gradient.13 The combined effect of physical and chemical roughness has been calculated by Gevorkyan et al. by decomposition of electric field components close to the cathode surface and numerically tracking the trajectory of the emitted electron in the applied electric field.14 

Another factor that determines the choice of a photocathode in a photoinjector is its lifetime/robustness. Many emerging applications such as Free Electron Lasers (FELs), Linear Accelerators (LINACs), and Relativistic Heavy Ion Collider (RHIC) experiments require high peak current along with high-quality electron beams.15 High-brightness photocathodes such as GaAs, which have low MTE and high QE, tend to have a longer lifetime when operated at low peak currents. However, it degrades very quickly when operated at high peak currents.16,17 For high peak current applications, one requires a photoemissive material with high conductivity. A wide-bandgap semiconductor such as diamond is a promising candidate for such applications because of its mechanical, thermal, and electrical stability along with an intrinsically fast response time (100 fs).18,19 Moreover, its vacuum robustness allows it to work under harsh conditions and, thus, makes it a promising candidate for photoinjector applications. A comparison between the emission properties of single-, micro-, nano-, and graphite-like nano-crystalline diamond films suggests that graphite-like nano-crystalline diamond has a better performance compared to the others. The QE of graphite-like nano-crystalline diamond was reported to be 104 at 200 nm.20 However, none of these films showed a good performance above 200 nm.

One way to improve the performance of diamond at wavelengths >200 nm is to take advantage of the negative electron affinity (NEA) of the diamond. A way to do so would be by n-doping of the diamond films and surface treatment in a hydrogen environment.21 Recently, n-doping of micro, nano, and ultrananocrystalline diamond (UNCD) films using nitrogen has been performed by many groups.22–24 It has been shown that the addition of nitrogen to UNCD films increased its electrical conductivity up to 1854 S cm1 and the maximum current density to 8 mA cm2.25 This is advantageous for high peak current photoinjector applications. Such films with nitrogen incorporation are termed as nitrogen-incorporated ultrananocrystalline diamond (N)UNCD. (N)UNCD typically consists of sp2 as well as sp3 bonded carbon atoms. The sp2 bonded carbon phase, also known as the graphitic phase, resides in the grain boundary region of sp3 bonded carbon phase, also known as the diamond phase. It has been reported that the photoemission of electrons from such materials originates from the sub-nm scale grain boundaries, which are predominantly composed of sp2 bonded carbon atoms.26 In addition, the NEA can be introduced in (N)UNCD after synthesis via surface hydrogenation treatment leading to the surface C–H dipole formation. Such cathodes are known as hydrogen-terminated nitrogen-incorporated ultrananocrystalline diamond [(N)UNCD:H].

The QE of (N)UNCD:H as well as (N)UNCD has been reported by many groups. It was observed that the QE of (N)UNCD:H was two orders of magnitude higher than the QE of (N)UNCD, which was 105 at 254 nm.27 However, not many experimental measurements of the MTE of (N)UNCD have been reported. A recent experimental measurement of MTE of (N)UNCD:H by Chen et al. using solenoid scan technique at the electric field of 0.45 MV/m suggests that the MTE of (N)UNCD:H is constant at 266 meV with decreasing photon energy.28 Another experimental measurement of MTE of an (N)UNCD was performed in an RF gun environment at the electric field of 30 MV/m. The normally incident 262 nm laser had a Gaussian longitudinal distribution with a full width at half maximum pulse length of 300 fs. The measured MTE was reported to be 1000 meV.29 

In this work, we report proof-of-concept QE, photoemission spectra, and MTE measurements for the (N)UNCD photocathode grown on a molybdenum substrate. The measurements were carried out in the near UV range of 200–300 nm, standard for many photocathode applications. Further, the measurements were performed for (N)UNCD photocathode prior to bakeout and after the bakeout of the sample. The bakeout was performed at 120°C for 24 h. The QE of the baked sample was higher than the QE of the unbaked sample by an order of magnitude for the same range of photon energies. In the photoemission energy distribution spectra, the sum of work function and kinetic energy corresponding to the maxima in the emission spectra does not match the excitation energy. This indicates that the scattering of electrons influences the emission. The measured MTE approximately scales as one-sixth of the excess energy. It is proposed that this trend in MTE is due to the relaxation of electrons possibly via phonon scatterings in the conduction band before emission. The MTE of 70 meV close to the threshold is attributed to the physical and chemical roughness of the (N)UNCD photocathode.

The diamond deposition process for the synthesis of (N)UNCD films was performed by Advanced Diamond Technologies. Details of different steps for the preparation of (N)UNCD are reported elsewhere.30 Typically, it has 1 μm layer of conductive (N)UNCD films, followed by a layer of 15 μm microcrystalline diamond. A diamond film is brazed to a polished Mo substrate using the brazing material TiCuSil. Figure 1 shows the Raman spectrum of an (N)UNCD sample measured using a custom built Raman spectrometer in a 180° geometry. The sample was excited using a 150 mW Coherent Sapphire SF laser with a 532 nm laser wavelength. The laser power was controlled using a neutral density filters wheel and an initial laser power of 100 mW. The laser was focused onto the sample using a 50X super long working distance plane APO Mitutoyo objective with a numerical aperture of 0.42. The signal was discriminated from the laser excitation using an Ondax®SureBlock™ ultranarrow-band notch filter combined with two optigrate notch filters. The broad peaks at 1350 and 1550 cm1 correspond to the D and G bands of diamond and graphite, respectively.

FIG. 1.

Raman spectrum of the (N)UNCD sample showing a characteristic diamond (D) peak and graphite (G) peak.

FIG. 1.

Raman spectrum of the (N)UNCD sample showing a characteristic diamond (D) peak and graphite (G) peak.

Close modal

QE, photoemission spectra, and MTE measurements of the (N)UNCD sample were performed using a photoemission electron microscope (PEEM).31 The base pressure of the PEEM chamber was in the low 1010 Torr range during the measurements.

QE measurements were performed using a 500 kHz repetition rate femtosecond pulsed laser with a pulse length of 150 fs from a pulsed Optical Parametric Amplifier (Light Conversion Orpheus pumped by Light Conversion Pharos). The LASER was made incident onto the sample at 65° angle of incidence with respect to the normal of the sample surface and was focused by a lens down to the spot size of 100×250μm2. One millimeter real space field of view (significantly larger than the LASER spot size) was selected to image all the emitted photo-electrons using a double micro channel plate (MCP) detector at the end of the PEEM column. The photoelectron current was determined by recording counts/second on the MCP detector and using a predetermined calibration factor to convert it into photocurrent at each wavelength.

The photoemission spectra were obtained by using the imaging energy filter (IEF) capability of the PEEM. The IEF unit is placed after the PEEM column and before the imaging unit of the microscope. It consists of two pre-retardation lens and two retardation grids that form the retarding field analyzer (RFA). The retarding field acts as the high pass energy filter that allows for energy filtered imaging. Background subtracted energy filtered images were obtained for the (N)UNCD sample using RFA of the PEEM with sub 65 meV energy resolution. The contrast aperture (CA) of diameter 1750 μm placed in the backfocal plane of the objective lens in the PEEM column was selected to ensure transmission of all the emitted electrons from the sample surface. After acquiring the energy filtered images, a numerical differentiation of high pass spectrum of the RFA was performed and the normalized counts were plotted against the kinetic energy of emitted electrons for each photon energy for both baked and unbaked samples.

MTE measurements were performed by operating PEEM in k-space mode with a sub 8-mÅ1k-space resolution. A 35±5μm region of the sample was selected by the adjustable iris aperture that was placed in the first intermediate image plane of the PEEM column while imaging in the real space mode. After limiting the area under investigation in the real space mode by adjustable iris aperture, its counterpart in the k-space mode was imaged by changing the projective settings of the PEEM. Figure 2 shows k-space images taken at 240 nm with FOV of 1.9 and 3.9 Å1 and the Gaussian fit applied to the kx and ky distributions obtained from 1.9 Å1 FOV. The MTE was then calculated using the second moment obtained from the Gaussian fit applied to the measured k-space distribution of the emitted electrons using the following equation:

(3)

Here, is the reduced Planck’s constant, m is the mass of an emitted electron, and σ=σkx2+σky2, where σkx and σky are the second moments obtained from the Gaussian fit applied to kx and ky distributions of the emitted electrons as shown in Figs. 2(c) and 2(d).

FIG. 2.

k-space images taken at 240 nm with FOV of (a) 1.9 and (b) 3.9 Å1. (c) Gaussian fit applied to the kx distribution obtained from 1.9 Å1 FOV and (d) Gaussian fit applied to the ky distribution obtained from 1.9 Å1 FOV.

FIG. 2.

k-space images taken at 240 nm with FOV of (a) 1.9 and (b) 3.9 Å1. (c) Gaussian fit applied to the kx distribution obtained from 1.9 Å1 FOV and (d) Gaussian fit applied to the ky distribution obtained from 1.9 Å1 FOV.

Close modal

For all the measurements, the fluence of the laser pulses incident on the sample was kept small enough to ensure linear electron counts and avoid effects of space charge and non-linear photoemission. The QE was measured at the extraction field of 0.005 MV/m and the photoemission spectra and MTE were measured at the extraction field of 0.5 MV/m. For the case of semiconductor, according to Stratton,32 the Schottky work function lowering is given by the following equation:

(4)

where α is the fine structure constant, e is the electron charge, and Ks is the dielectric constant. Assuming Ks=5.7, typical for diamond,33 and the maximum extraction field of 0.5 MV/m on the (N)UNCD photocathode surface in the PEEM, Eq. (4) results in a Schottky work function lowering effect of only 22.47 meV that is comparable to the thermal energy at the room temperature. QE, photoemission spectra, and MTE measurements were performed in the near UV range of 200–300 nm standard for many photocathode applications.

According to the Dowell–Schmerge model,6 for photoemission near threshold, the quantum efficiency is given by the following equation:

(5)

where ω is the photon energy and Φ is the work function of a photocathode. Hence, on plotting the square root of QE vs photon energy, the x-intercept gives the work function of a photocathode sample. Figure 3 shows experimentally determined square root of QE values for the (N)UNCD sample. There are four main features in Fig. 3. First, the QE of the baked sample is higher than the QE of the unbaked sample for the same range of photon energies. This enhancement in QE can be attributed to the desorption of contaminant layers from the (N)UNCD surface during the bake-out. Second, the QE of the unbaked (N)UNCD sample is of the order of 108 and that of the baked sample is of the order of 107 at 4.9 eV photon energy. These values are comparable to the previously reported values.27 Third, the photoemission threshold for both unbaked and baked samples is observed at the photon energy of 4.4±0.1 eV. This is comparable to the previously reported values.28,29 Last, there is a significant rise in QE beyond 5.4 eV photon energy. Since the bandgap of diamond is 5.4 eV,34 this increase in QE can be attributed to a higher number of electrons being excited in the conduction band when the photon energy is greater than the bandgap of diamond.

FIG. 3.

Experimentally determined QE from the (N)UNCD samples: for the unbaked sample and baked sample which was baked at 120°C for 24 h. The dotted line is a linear fit indicating 4.4±0.1 eV threshold from the unbaked (brown) and baked (black) samples.

FIG. 3.

Experimentally determined QE from the (N)UNCD samples: for the unbaked sample and baked sample which was baked at 120°C for 24 h. The dotted line is a linear fit indicating 4.4±0.1 eV threshold from the unbaked (brown) and baked (black) samples.

Close modal

Figure 4 shows experimentally determined photoemission spectra for the (N)UNCD sample before and after the bakeout of the sample. For a given photon energy, the energy spectra appear identical for both baked and unbaked samples. For both baked and unbaked samples, the photoemission spectra broaden in kinetic energy with increasing photon energy. The top x-axis in Fig. 4 shows the sum of kinetic energy (Ek) of the emitted electrons and the work function (Φ) of the (N)UNCD photocathode. Considering the work function of the (N)UNCD to be 4.4 eV, as obtained from the QE spectral response in Fig. 3, the sum of the maximum kinetic energy of the emitted electrons and the work function approximately equal to the photon energy. Also, from the top x-axis in Fig. 4, we can see that the sum of the work function and the kinetic energy corresponding to the maximum of the energy distribution curves does not coincide with that of the photon energy. This indicates that the electrons are scattered possibly by phonons before being emitted from the (N)UNCD sample. A similar electron scattering mechanism is observed for nitrogen-doped diamond films35 and other semiconductor photocathodes.36 In the case of an unbaked sample, at 4.13 eV photon energy, the emission spectrum is broadened due to non-linear photoemission. This non-linear photoemission is also reflected in the MTE plot shown in Fig. 5, which results in an increase in MTE at 4.13 eV photon energy.37 

FIG. 4.

Photoemission spectra at different incident photon energies for the (N)UNCD samples: for (a) unbaked sample and (b) baked sample which was baked at 120°C for 24 h. The bottom x-axis represents the kinetic energy of the emitted electrons and the top x-axis represents the sum of the kinetic energy of the emitted electrons and the work function (Φ) of the photocathode sample.

FIG. 4.

Photoemission spectra at different incident photon energies for the (N)UNCD samples: for (a) unbaked sample and (b) baked sample which was baked at 120°C for 24 h. The bottom x-axis represents the kinetic energy of the emitted electrons and the top x-axis represents the sum of the kinetic energy of the emitted electrons and the work function (Φ) of the photocathode sample.

Close modal
FIG. 5.

Experimentally determined MTE from the (N)UNCD samples: for unbaked sample and baked sample which was baked at 120°C for 24 h.

FIG. 5.

Experimentally determined MTE from the (N)UNCD samples: for unbaked sample and baked sample which was baked at 120°C for 24 h.

Close modal

1. MTE spectral response

Figure 5 shows experimentally determined MTE values for the (N)UNCD samples. There are three main features in Fig. 5. First, the MTE of the (N)UNCD decreases with a decrease in excess energy and shows a flat trend close to the threshold, which is observed at a photon energy 4.4 eV. This is in contrast to the constant MTE of 266 meV for (N)UNCD:H reported by Chen et al.28 for a similar spectral range. Also, the MTE at 260 nm (4.76 eV photon energy) was observed to be 100 meV. This is an order of magnitude smaller when compared to the MTE observed by Ref. 29 at 262 nm and 30 MV/m electric field gradient on the surface of the cathode. Second, the scaling of MTE with excess energy is not in accordance with the Dowell–Schmerge model.6 This behavior can be attributed to the relaxation of electrons possibly with phonon scattering in the conduction band before emission. A similar behavior is observed for other semiconductor photocathodes.36 Last, the MTE at the threshold does not reach 25 meV (thermal limit) but flattens at 70 meV. This is attributed to the physical and chemical roughness of (N)UNCD photocathodes and is discussed in detail in Sec. III C2.

2. Dependence of MTE on physical and chemical roughness

A detailed discussion of the effect of physical and chemical roughness of the photocathode on the MTE of the emitted electrons is done in Sec. I. The effect of physical roughness on the MTE due to the applied electric field is given by Eq. (2). In the case of UNCDs, the amplitude of physical roughness varies between 5 and 10 nm38–40 and the period of the roughness corresponds to the grain size which is generally reported to be in the range of 5–15 nm.41–43 The contribution of MTEfield as a function of surface roughness is plotted for different periods typical of UNCD photocathodes in Fig. 6. The MTEfield was plotted for two values of the electric field. First, 30 MV/m, which was the field gradient at the surface of the (N)UNCD reported in Ref. 29. Second, 0.5 MV/m (the inset of Fig. 6), which was the field gradient at the surface of the (N)UNCD films during our measurements of MTE. The MTEfield obtained at 30 MV/m is larger than the thermal limit (25 meV) and is comparable to the previously reported value.29 In addition, the contribution of MTEfield at 0.5 MV/m shown in the inset of Fig. 6 is 25 meV for a=10 nm and λ=10 nm.

FIG. 6.

Variation of MTEfield with a. The calculations are done for λ=5, 10, and 15 nm and electric field gradients of 30 and 0.5 MV/m (inset figure).

FIG. 6.

Variation of MTEfield with a. The calculations are done for λ=5, 10, and 15 nm and electric field gradients of 30 and 0.5 MV/m (inset figure).

Close modal

As discussed in Sec. I, another source of increase in MTE is the chemical roughness of the photocathode. From Fig. 1, there are two phases present in the (N)UNCD sample. First, the diamond phase and second, the graphite phase. These two phases can have work function difference of several 100s meV. If we consider 1 nm graphite regions between 10 nm diamond grains, the variation of several 100s meV in surface potential can have an additional contribution to the MTE at 0.5 MV/m electric field.13,14

In general, for practical photocathodes with physical and chemical roughness, the MTE of the emitted electrons is given by

(6)

where MTEkT is the contribution due to thermal limit which is 25 meV at room temperature. Assuming a=10 nm and λ=10 nm, which is typical for UNCDs, the contribution of MTEfield at 0.5 MV/m is 25 meV. Lastly, MTEWF is the contribution due to the chemical roughness arising from several 100s meV variation in surface potential due to the presence of graphite phase and diamond phase in the (N)UNCD sample. Hence, the total MTE which is the sum of MTEkT, MTEfield, and MTEWF is 70 meV at the threshold for the (N)UNCD photocathodes which is in agreement with what is shown in Fig. 5.

We measured the QE, photoemission spectra, and MTE of a (N)UNCD photocathode in the UV range of 200–300 nm for unbaked and baked samples. The QE was measured to be 107 for a baked sample and 108 for an unbaked sample at 250 nm. From the photoemission spectra, we observed that the sum of work function and kinetic energy corresponding to the maxima in the emission spectrum does not match the excitation energy. This indicates that the scattering of electrons influences the emission. The MTE decreased with the decrease in excess energy and showed a flat trend near the threshold. This behavior is in contrast to the constant MTE of 266 meV reported for an (N)UNCD:H for a similar spectral range.28 The MTE at 260 nm (4.76 eV photon energy) was observed to be 100 meV. This is an order of magnitude smaller than the MTE reported by Ref. 29. We attribute this difference to the physical roughness of the (N)UNCD photocathode. The scaling of MTE with excess energy is not in accordance with the Dowell–Schmerge model.6 This observed trend in MTE is attributed to the relaxation of electrons possibly with phonon scattering in the conduction band before emission.36 This scattering of electrons before emission is also evident in the photoemission spectra shown in Fig. 4. Furthermore, the MTE at the threshold does not go to 25 meV (thermal limit) but flattens at 70 meV. This is attributed to the physical roughness due to the fabrication process and chemical roughness, which is a result of the work function variation due to the presence of graphite phase and diamond phase in the (N)UNCD sample.13,14 The work function of the (N)UNCD obtained from the spectral response of QE and MTE was 4.4±0.1 eV. This is in agreement with the previously reported values.28,29

In summary, we have reported the spectral responses of QE, photoemission spectra and MTE of the (N)UNCD photocathode. With the QE comparable to that of metal photocathodes and MTE of 70 meV at the threshold, (N)UNCD photocathodes are suitable for various high peak current photoinjector applications. However, the use of these photocathodes at high electric field gradients can significantly increase the MTE of the emitted electrons owing to their surface roughness. Consequently, it can lead to a decrease in brightness of the emitted electron beams. Even at low electric fields, the contribution to MTE due to physical and chemical roughness is significant enough to increase MTE beyond the thermal limit at the threshold. One way to reduce MTE and consequently increase the brightness of (N)UNCD photocathodes is by optimizing the growth techniques to produce a smoother film. One way to do this would involve growing (N)UNCD films via tungsten interlayers.44 As a result, future work would involve growing (N)UNCD films with reduced surface roughness and additional fundamental understanding of their emission mechanism.

This work is supported by the National Science Foundation Center for Bright Beams under Award No. PHY-1549132, the (U.S.) Department of Energy under Grant No. DE-SC0021092, and the Los Alamos National Laboratory (LANL) Laboratory Directed Research and Development (LDRD) Program. This work was performed, in part, at the Center for Integrated Nanotechnologies, an Office of Science User Facility operated for the U.S. Department of Energy (DOE) Office of Science. Los Alamos National Laboratory, an affirmative action equal opportunity employer, is managed by Triad National Security, LLC for the U.S. Department of Energy’s National Nuclear Security Administration, under Contract No. 89233218CNA000001.

The authors have no conflicts to disclose.

A. Kachwala: Conceptualization (Supporting); Data Curation (Lead); Formal Analysis (Lead); Investigation (Equal); Methodology (Equal); Project Administration (Supporting); Resources (Equal); Software (Equal); Validation (Lead); Visualization (Lead); Original Draft (Lead); Reviewing and Editing (Supporting) O. Chubenko: Conceptualization (Supporting); Data Curation (Supporting); Formal Analysis (Supporting); Investigation (Equal); Methodology (Equal); Project Administration (Supporting); Resources (Equal); Software (Equal); Validation (Supporting); Visualization (Supporting); Reviewing and Editing (Supporting) D. Kim: Conceptualization (Supporting); Data Curation (Supporting); Formal Analysis (Supporting); Investigation (Equal); Methodology (Equal); Project Administration (Supporting); Resources (Equal); Software (Equal); Validation (Supporting); Visualization (Supporting); Reviewing and Editing (Supporting) E.I. Simakov: Conceptualization (Supporting); Data Curation (Supporting); Formal Analysis (Supporting); Funding Acquisition (Supporting); Investigation (Equal); Methodology (Equal); Project Administration (Supporting); Resources (Equal); Software (Equal); Supervision (Supporting); Validation (Supporting); Visualization (Supporting); Reviewing and Editing (Supporting) S. Karkare: Conceptualization (Lead); Data Curation (Supporting); Formal Analysis (Supporting); Funding Acquisition (Lead); Investigation (Equal); Methodology (Equal); Project Administration (Lead); Resources (Equal); Software (Equal); Supervision (Lead); Validation (Supporting); Visualization (Supporting); Reviewing and Editing (Lead)

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

1.
P.
Emma
,
R.
Akre
,
J.
Arthur
,
R.
Bionta
,
C.
Bostedt
,
J.
Bozek
,
A.
Brachmann
,
P.
Bucksbaum
,
R.
Coffee
,
F.-J.
Decker
, and
Y.
Ding
, “
First lasing and operation of an ångstrom-wavelength free-electron laser
,”
Nat. Photonics
4
,
641
647
(
2010
).
2.
H. W.
Kim
,
N. A.
Vinokurov
,
I. H.
Baek
,
K. Y.
Oang
,
M. H.
Kim
,
Y. C.
Kim
,
K.-H.
Jang
,
K.
Lee
,
S. H.
Park
,
S.
Park
, and
J.
Shin
, “
Towards jitter-free ultrafast electron diffraction technology
,”
Nat. Photonics
14
,
245
249
(
2020
).
3.
W.
Graves
,
J.
Bessuille
,
P.
Brown
,
S.
Carbajo
,
V.
Dolgashev
,
K.-H.
Hong
,
E.
Ihloff
,
B.
Khaykovich
,
H.
Lin
,
K.
Murari
, and
E. A.
Nanni
, “
Compact x-ray source based on burst-mode inverse compton scattering at 100 kHz
,”
Phys. Rev. Spec. Top.-Accel. Beams
17
,
120701
(
2014
).
4.
E.
Lessner
,
X.
Wang
, and
P.
Musumeci
,
Future of Electron Sources
(
US Department of Energy
,
Washington, DC
,
2017
).
5.
D.
Dowell
,
I.
Bazarov
,
B.
Dunham
,
K.
Harkay
,
C.
Hernandez-Garcia
,
R.
Legg
,
H.
Padmore
,
T.
Rao
,
J.
Smedley
, and
W.
Wan
, “
Cathode R&D for future light sources
,”
Nucl. Instrum. Methods Phys. Res., Sect. A
622
,
685
697
(
2010
).
6.
D. H.
Dowell
and
J. F.
Schmerge
, “
Quantum efficiency and thermal emittance of metal photocathodes
,”
Phys. Rev. Spec. Top.-Accel. Beams
12
,
074201
(
2009
).
7.
J.
Feng
,
J.
Nasiatka
,
W.
Wan
,
S.
Karkare
,
J.
Smedley
, and
H. A.
Padmore
, “
Thermal limit to the intrinsic emittance from metal photocathodes
,”
Appl. Phys. Lett.
107
,
134101
(
2015
).
8.
I. V.
Bazarov
,
B. M.
Dunham
,
Y.
Li
,
X.
Liu
,
D. G.
Ouzounov
,
C. K.
Sinclair
,
F.
Hannon
, and
T.
Miyajima
, “
Thermal emittance and response time measurements of negative electron affinity photocathodes
,”
J. Appl. Phys.
103
,
054901
(
2008
).
9.
Z.
Zhang
and
C.
Tang
, “
Analytical study on emittance growth caused by roughness of a metallic photocathode
,”
Phys. Rev. Spec. Top.-Accel. Beams
18
,
053401
(
2015
).
10.
H.
Qian
,
C.
Li
,
Y.
Du
,
L.
Yan
,
J.
Hua
,
W.
Huang
, and
C.
Tang
, “
Experimental investigation of thermal emittance components of copper photocathode
,”
Phys. Rev. Spec. Top.-Accel. Beams
15
,
040102
(
2012
).
11.
J.
Feng
,
S.
Karkare
,
J.
Nasiatka
,
S.
Schubert
,
J.
Smedley
, and
H.
Padmore
, “
Near atomically smooth alkali antimonide photocathode thin films
,”
J. Appl. Phys.
121
,
044904
(
2017
).
12.
J.
Smedley
,
M.
Gaowei
,
J.
Sinsheimer
,
K.
Attenkofer
,
J.
Walsh
,
J.
Kuhn
,
E.
Muller
,
Z.
Ding
,
H.
Frisch
,
H.
Bhandari
, and
H.
Lingertat
, “Sputter growth of alkali antimonide photocathodes: An in operando materials analysis,” in Proceeding of the 2015 International Particle Accelerator Conference (IPAC, 2015).
13.
S.
Karkare
and
I.
Bazarov
, “
Effects of surface nonuniformities on the mean transverse energy from photocathodes
,”
Phys. Rev. Appl.
4
,
024015
(
2015
).
14.
G.
Gevorkyan
,
S.
Karkare
,
S.
Emamian
,
I.
Bazarov
, and
H.
Padmore
, “
Effects of physical and chemical surface roughness on the brightness of electron beams from photocathodes
,”
Phys. Rev. Accel. Beams
21
,
093401
(
2018
).
15.
A.
Drees
, “
Heavy ions in RHIC—How to get to higher luminosities?
,”
Acta Phys. Hung. A
25
,
475
486
(
2006
).
16.
J.
Grames
,
R.
Suleiman
,
P.
Adderley
,
J.
Clark
,
J.
Hansknecht
,
D.
Machie
,
M.
Poelker
, and
M.
Stutzman
, “
Charge and fluence lifetime measurements of a dc high voltage GaAs photogun at high average current
,”
Phys. Rev. Spec. Top.-Accel. Beams
14
,
043501
(
2011
).
17.
R.
Suleiman
,
P.
Adderley
,
J.
Grames
,
J.
Hansknecht
,
M.
Poelker
, and
M.
Stutzman
, “
High current polarized electron source
,”
AIP Conf. Proc.
1970
,
050007
(
2018
).
18.
L.
Liu
,
X.
Ouyang
,
J.
Zhang
,
X.
Zhang
, and
Y.
Zhong
, “
Polycrystalline CVD diamond detector: Fast response and high sensitivity with large area
,”
AIP Adv.
4
,
017114
(
2014
).
19.
M. E.
Schneider
, “Advancing field emission technology for high power injectors operating in GHz and beyond,” Ph.D. thesis (Michigan State University, 2022).
20.
M.
Nitti
,
M.
Colasuonno
,
E.
Nappi
,
A.
Valentini
,
E.
Fanizza
,
F.
Bénédic
,
G.
Cicala
,
E.
Milani
, and
G.
Prestopino
, “
Performance analysis of poly-, nano- and single-crystalline diamond-based photocathodes
,”
Nucl. Instrum. Methods Phys. Res., Sect. A
595
,
131
135
(
2008
).
21.
J.
van der Weide
,
Z.
Zhang
,
P.
Baumann
,
M.
Wensell
,
J.
Bernholc
, and
R.
Nemanich
, “
Negative-electron-affinity effects on the diamond (100) surface
,”
Phys. Rev. B
50
,
5803
(
1994
).
22.
M.-A.
Pinault
,
J.
Barjon
,
T.
Kociniewski
,
F.
Jomard
, and
J.
Chevallier
, “
The n-type doping of diamond: Present status and pending questions
,”
Phys. B: Condens. Matter
401
,
51
56
(
2007
).
23.
U.
Mengui
,
R.
Campos
,
K.
Alves
,
E.
Antunes
,
M.
Hamanaka
,
E.
Corat
, and
M.
Baldan
, “
Combined effect of nitrogen doping and nanosteps on microcrystalline diamond films for improvement of field emission
,”
Appl. Surf. Sci.
334
,
222
226
(
2015
).
24.
L.
Shen
and
Z.
Chen
, “
An investigation of grain size and nitrogen-doping effects on the mechanical properties of ultrananocrystalline diamond films
,”
Int. J. Solids Struct.
44
,
3379
3392
(
2007
).
25.
W.
Yuan
,
L.
Fang
,
Z.
Feng
,
Z.
Chen
,
J.
Wen
,
Y.
Xiong
, and
B.
Wang
, “
Highly conductive nitrogen-doped ultrananocrystalline diamond films with enhanced field emission properties: Triethylamine as a new nitrogen source
,”
J. Mater. Chem. C
4
,
4778
4785
(
2016
).
26.
R. L.
Harniman
,
O. J.
Fox
,
W.
Janssen
,
S.
Drijkoningen
,
K.
Haenen
, and
P. W.
May
, “
Direct observation of electron emission from grain boundaries in CVD diamond by peakforce-controlled tunnelling atomic force microscopy
,”
Carbon
94
,
386
395
(
2015
).
27.
K. J.
Pérez Quintero
,
S.
Antipov
,
A. V.
Sumant
,
C.
Jing
, and
S. V.
Baryshev
, “
High quantum efficiency ultrananocrystalline diamond photocathode for photoinjector applications
,”
Appl. Phys. Lett.
105
,
123103
(
2014
).
28.
G.
Chen
,
G.
Adhikari
,
L.
Spentzouris
,
K. K.
Kovi
,
S.
Antipov
,
C.
Jing
,
W.
Andreas Schroeder
, and
S. V.
Baryshev
, “
Mean transverse energy of ultrananocrystalline diamond photocathode
,”
Appl. Phys. Lett.
114
,
093103
(
2019
).
29.
G.
Chen
,
L.
Spentzouris
,
C.
Jing
,
M.
Conde
,
G.
Ha
,
W.
Liu
,
J.
Power
,
E.
Wisniewski
,
A. V.
Sumant
,
S.
Antipov
, and
E.
Gomez
, “
Demonstration of nitrogen-incorporated ultrananocrystalline diamond photocathodes in a RF gun environment
,”
Appl. Phys. Lett.
117
,
171903
(
2020
).
30.
D.
Kim
,
H. L.
Andrews
,
B. K.
Choi
, and
E. I.
Simakov
, “Fabrication of micron-scale diamond field emitter arrays for dielectric laser accelerators,” in 2018 IEEE Advanced Accelerator Concepts Workshop (AAC) (IEEE, 2018), pp. 1–3.
31.
See https://www.focus-gmbh.com for “FOCUS-IS-IEF-PEEM.”
32.
R.
Stratton
, “
Theory of field emission from semiconductors
,”
Phys. Rev.
125
,
67
(
1962
).
33.
See http://www.chm.bris.ac.uk/motm/diamond/diamprop.htm for “Principal Properties of Diamond.”
34.
C. J.
Wort
and
R. S.
Balmer
, “
Diamond as an electronic material
,”
Mater. Today
11
,
22
28
(
2008
).
35.
T.
Sun
,
F. A.
Koeck
,
C.
Zhu
, and
R. J.
Nemanich
, “
Combined visible light photo-emission and low temperature thermionic emission from nitrogen doped diamond films
,”
Appl. Phys. Lett.
99
,
202101
(
2011
).
36.
S.
Karkare
,
D.
Dimitrov
,
W.
Schaff
,
L.
Cultrera
,
A.
Bartnik
,
X.
Liu
,
E.
Sawyer
,
T.
Esposito
, and
I.
Bazarov
, “
Monte Carlo charge transport and photoemission from negative electron affinity GaAs photocathodes
,”
J. Appl. Phys.
113
,
104904
(
2013
).
37.
C.
Knill
,
S.
Karkare
, and
H.
Padmore
, “Near-threshold nonlinear photoemission from Cu (100),” in 12th International Particle Accelerator Conference (IPAC’21), Campinas, SP, Brazil, 24–28 May 2021 (JACOW Publishing, Geneva, 2021), pp. 2822–2825.
38.
L.
Booth
,
S. A.
Catledge
,
D.
Nolen
,
R. G.
Thompson
, and
Y. K.
Vohra
, “
Synthesis and characterization of multilayered diamond coatings for biomedical implants
,”
Materials
4
,
857
868
(
2011
).
39.
A.
Gujrati
,
A.
Sanner
,
S. R.
Khanal
,
N.
Moldovan
,
H.
Zeng
,
L.
Pastewka
, and
T. D.
Jacobs
, “
Comprehensive topography characterization of polycrystalline diamond coatings
,”
Surf. Topogr.: Metrol. Prop.
9
,
014003
(
2021
).
40.
O.
Auciello
,
S.
Renou
,
K.
Kang
,
D.
Tasat
, and
D.
Olmedo
, “
A biocompatible ultrananocrystalline diamond (UNCD) coating for a new generation of dental implants
,”
Nanomaterials
12
,
782
(
2022
).
41.
S.
Tóth
,
J.
Budai
,
M.
Veres
, and
M.
Koós
, “Optical properties of nano-and ultrananocrystalline diamond thin layers in the UV and visible spectral range,” in IOP Conference Series: Materials Science and Engineering (IOP Publishing, 2018), Vol. 426, p. 012049.
42.
R.
Rani
,
K.
Panda
,
N.
Kumar
,
K. A.
Titovich
,
K. V.
Ivanovich
,
S. A.
Vyacheslavovich
, and
I.-N.
Lin
, “
Tribological properties of ultrananocrystalline diamond films: Mechanochemical transformation of sliding interfaces
,”
Sci. Rep.
8
,
1
16
(
2018
).
43.
O.
Chubenko
,
S. S.
Baturin
, and
S. V.
Baryshev
, “
Theoretical evaluation of electronic density-of-states and transport effects on field emission from n-type ultrananocrystalline diamond films
,”
J. Appl. Phys.
125
,
205303
(
2019
).
44.
N. N.
Naguib
,
J. W.
Elam
,
J.
Birrell
,
J.
Wang
,
D. S.
Grierson
,
B.
Kabius
,
J. M.
Hiller
,
A. V.
Sumant
,
R. W.
Carpick
,
O.
Auciello
, and
J. A.
Carlisle
, “
Enhanced nucleation, smoothness and conformality of ultrananocrystalline diamond (UNCD) ultrathin films via tungsten interlayers
,”
Chem. Phys. Lett.
430
,
345
350
(
2006
).