Cross-sectional transmission electron microscopy has been used to characterize the morphological features of thin boron nitride films grown on single-crystal boron-doped diamond substrates (lattice mismatch of 1.36%) using electron cyclotron resonance-plasma enhanced chemical vapor deposition. The effect of gas precursor concentration, growth temperature, and substrate cleaning method on determining the BN phase (either cubic or turbostratic), the defect density, and the orientation relationship between c-BN domains and the diamond substrate were investigated. A nucleation step involving a hydrogen-limited gas mixture promoted etching of sp2-bonded BN phases and increased the fraction of cubic phase present in the films. A growth temperature of 820 °C resulted in larger BN grains and reduced defect densities particularly in regions away from the BN/diamond interface. Substrate cleaning with hydrogen plasma was found to be associated with twin-related BN growth rather than a simple epitaxial relationship. High-resolution electron micrographs showed complex contrast features caused by the presence of a high density of twin domains and stacking faults near the BN/diamond heterointerface for samples with predominant cubic phase. Electron-energy-loss spectroscopy was used to differentiate between regions of sp2 and sp3 bonding, and showed distinct transitions to the latter for the growth of cubic materials. Growth experiments at even higher temperatures and optimal substrate cleaning methods are needed for improved defect mitigation in BN films.

Cubic boron nitride (c-BN) is an ultra-wide bandgap (UWBG) semiconductor with the highest bandgap among the binary nitrides (∼6.4 eV), as well as high thermal conductivity (13 W cm−1 K−1), low dielectric constant (7.1), and high breakdown electric field (8 MV cm−1).1 Moreover, both n- and p-type doping have been achieved, thus making it highly attractive for high-power and high-temperature electronic applications, especially in harsh environments.1 Diamond, which has a high thermal conductivity (22 W/cm  K), is a promising substrate candidate for epitaxial growth of single-crystal c-BN due to the small lattice mismatch (1.36%), and its linear thermal expansion coefficient (3.1 × 10−6/K) is only slightly smaller than that of c-BN (4.7 × 10−6/K).2 Thus, c-BN/diamond heterostructures could potentially provide new and improved options for high-power electronic devices.3–5 However, the synthesis of wafer-scale c-BN single crystals or epitaxial heterostructures remains in an early stage of development.

Cubic BN has the zincblende crystal structure with a lattice constant a = 3.615 Å, and it belongs to the F 4 ¯ 3 m space group. Boron nitride is polymorphic and crystallizes in sp2-bonded hexagonal (h-BN), rhombohedral (r-BN), and turbostratic (t-BN) phases, as well as in sp3-bonded cubic (c-BN) and wurtzite (w-BN) phases.6 This polymorphism poses a major challenge for the growth of the pure cubic-phase material. Moreover, there is considerable ongoing discussion about which BN phase is thermodynamically the most stable.7–9 

Since the initial discovery of the c-BN phase,10 its stability has been closely tied to the method of growth. Aggressive growth techniques that induce compressive stress in the growing film, such as high-pressure high-temperature (HPHT),11–13 ion-assisted pulsed laser deposition (PLD),14 and physical vapor deposition (PVD),15–17 have proceeded on the basis that the initial nucleation of amorphous BN and the more stable t- or h-BN phases was inevitable. Thus, highly energetic ion bombardment was considered necessary to achieve the desired transition to the cubic phase.16,18,19 These results aligned well with the first-ever reported phase diagram of BN,20 which suggested that t-BN or h-BN were the two most stable BN phases under standard temperature and pressure (STP) conditions (273.15 K and 1 atm). Later reports indicated nucleation of the cubic phase under more relaxed growth condition,21,22 and these results were supported by a modified phase diagram based on theoretical thermodynamic calculations.23 More recently, advances in growth techniques have demonstrated significant progress in achieving high-quality c-BN films with the presence of minimal sp2-bonded BN phases. For example, fluorine-based chemical vapor deposition (CVD) has been utilized to grow c-BN films with a columnar structure and large crystal sizes (>1 μm) using silicon substrates.24 This method involved the formation of approximately 100 nm of amorphous or turbostratic BN at the interface, followed by the nucleation of a predominantly cubic phase, indicating a substantial reduction in the proportion of non-cubic phases compared to earlier growth methods. Moreover, molecular beam epitaxy (MBE) has achieved epitaxial growth of c-BN films with notable results. One recent study reported films containing over 99% c-BN, with only trace amounts of sp2-bonded BN detected at the interface and at the film surface,25 while another reported the successful growth of epitaxial c-BN films devoid of any sp2-bonded BN at the interface.26 However, simulations based on ab initio framework—fixed-node diffusion Monte Carlo (FNDMC) have suggested that h-BN is more stable than c-BN under STP conditions.27 

Despite reports of single crystal c-BN films being achieved, there remains a notable absence of any detailed structural analysis. Prior workers have used Fourier transform infrared spectroscopy (FTIR), x-ray diffraction (XRD), and transmission electron microscopy (TEM) for phase identification and texture analysis of nanostructured c-BN films.14,28–32 These studies have mostly focused on comparing film quality under different growth conditions, with little attention being paid to morphological features present in the films. This present paper describes a detailed TEM investigation of c-BN/diamond heterostructures grown by electron cyclotron resonance-plasma enhanced chemical vapor deposition (ECR PECVD) and focuses on the structural evolution of the BN films and evaluating the defect types and densities as a function of growth conditions. Chemical analysis using electron-energy-loss spectroscopy (EELS) has also been used to differentiate between regions of sp2 and sp3 bonding within the BN films.

The BN films were grown on heavily B-doped single-crystal diamond substrates (Technical Institute of Superhard Novel Carbon Materials, TISNCM) using ECR PECVD. The substrates were loaded into the CVD vacuum system, and hydrogen plasma was used for cleaning the substrate surfaces at high temperatures and low pressures prior to initiation of growth (refer to Table I for details). Two different cleaning procedures, labeled here as C1 and C2, were followed. Deposition parameters for the samples are summarized in Table II. Deposition of BN was carried out using gas precursors H2, BF3, N2, and inert gases He and Ar, with a negative bias voltage of −60 V, and microwave power of 1.42 kW. Gas flow rates of N2, and He and Ar were kept constant at 12.5, 35, and 2.5 SCCM, respectively, during film growth. Substrate temperatures were varied between 735 and 820 °C and the chamber pressure was ∼1.1 × 10−4 Torr. More extensive details about the substrate cleaning procedure and the growth techniques are given elsewhere.33 

TABLE I.

Summary of the substrate cleaning method.

Cleaning methodH2 (SCCM)Pressure (Torr)Microwave power (W)Substrate temperature (°C)Duration (H2 plasma)
AnnealingH2 plasma
C1 20 9 × 10−5 365 760 760 1 h, 30 m 
C2 830 750 10 m 
Cleaning methodH2 (SCCM)Pressure (Torr)Microwave power (W)Substrate temperature (°C)Duration (H2 plasma)
AnnealingH2 plasma
C1 20 9 × 10−5 365 760 760 1 h, 30 m 
C2 830 750 10 m 
TABLE II.

Key growth parameters used for the synthesis of BN films.

Sample labelSubstrate orientationSubstrate cleaning methodH2/BF3 ratioGrowth temperature (Ts) (°C)
A1 (111) C1 0.75: H2 = 1.5 SCCM; BF3 = 2.0 SCCM (40 min) 1.0: H2 = BF3 = 2.0 SCCM (110 min) 735  
A2 (001) C1 1.0: H2 = BF3 = 2.0 SCCM (150 min) 735 
A3 (111) C1 0.75: H2 = 1.5 SCCM; BF3 = 2.0 SCCM (40 min) 1.0: H2 = BF3 = 2.0 SCCM (110 min) 820 
A4 (111) C2 0.75: H2 = 1.5 SCCM; BF3 = 2.0 SCCM (40 min) 820 
Sample labelSubstrate orientationSubstrate cleaning methodH2/BF3 ratioGrowth temperature (Ts) (°C)
A1 (111) C1 0.75: H2 = 1.5 SCCM; BF3 = 2.0 SCCM (40 min) 1.0: H2 = BF3 = 2.0 SCCM (110 min) 735  
A2 (001) C1 1.0: H2 = BF3 = 2.0 SCCM (150 min) 735 
A3 (111) C1 0.75: H2 = 1.5 SCCM; BF3 = 2.0 SCCM (40 min) 1.0: H2 = BF3 = 2.0 SCCM (110 min) 820 
A4 (111) C2 0.75: H2 = 1.5 SCCM; BF3 = 2.0 SCCM (40 min) 820 

Samples suitable for cross-sectional TEM observation were prepared by focused-ion beam (FIB) milling using a Thermo Fisher Helios 5UX dual-beam instrument, with initial thinning done at 30 keV and further thinning at 5 and 2 keV to minimize surface damage induced by ion milling. A Philips-FEI CM-200 FEG TEM operated at 200 kV, and an image-corrected FEI Titan 80–300 operated at 300 kV, were used for imaging, and a JEOL ARM200F operated at 200 kV was used for analysis by electron-energy-loss spectroscopy (EELS).

Cross-sectional TEM images of samples A1 and A2, which compare BN film nucleation at the BN/diamond interface, followed by growth in the intermediate region, and near the top surface, are shown in Fig. 1. The image of sample A1 grown with H2 = 1.5, BF3 = 2 SCCM in the initial growth stage, and H2 = 2, BF3 = 2 SCCM in the later stage, while maintaining H/F ratios of 0.75 and 1, respectively, demonstrates the presence of overlapped c-BN twin domains at the interface, and well-oriented c-BN grains in the intermediate and near-top surface regions of the film. The presence of the cubic BN phase is also confirmed by the {111} and {002} diffraction spots visible in the fast Fourier transforms (FFTs) of the respective images. In contrast, sample A2, grown with H2 = 4 SCCM, BF3 = 4 SCCM throughout the experiment, while maintaining a H/F ratio = 1, shows nucleation of t-BN at the heterointerface. The disordered t-BN phase then continued to grow all the way to the top surface of the BN film.

FIG. 1.

High magnification TEM images of samples A1, A2, and A3 at the BN top surface, intermediate, and BN/diamond interface regions, respectively. Corresponding FFTs from the images of A1 are shown beside (a), (d), and (g). All scale bars are 3 nm in length.

FIG. 1.

High magnification TEM images of samples A1, A2, and A3 at the BN top surface, intermediate, and BN/diamond interface regions, respectively. Corresponding FFTs from the images of A1 are shown beside (a), (d), and (g). All scale bars are 3 nm in length.

Close modal

Figure 1 also compares images taken from near-interface, intermediate, and near-surface regions of samples A1 and A3, which were grown at temperatures of 735 and 820 °C, respectively, together with the corresponding fast Fourier transforms (FFTs) of sample A1. The BN layers near the substrate surfaces of both samples show a complex appearance. Moreover, the corresponding FFTs show a pattern of spots that could at first sight be interpreted as corresponding to hexagonal-close-packing (HCP). However, further examination and analysis reveals that this appearance is yet another example of a well-known imaging artefact originating from the overlap of twins and stacking faults (SFs) in projection along the beam direction in nanograined materials, as recently explained in detail by Cayron.34 A more detailed explanation of the origin of this complex contrast in the case of c-BN during HRTEM imaging will be given elsewhere.35 Intermediate and near-surface regions in the film showed no apparent sign of any t- or h-BN phases and the predominance of the cubic phase in these regions was again confirmed by the {111} and {002} spots visible in the FFTs. Areas further away from the BN/diamond interface in sample A3 had FFTs (not shown here) with sharper {111} and {002} spots and no additional spots or streaking effects, indicating improved film quality in the intermediate and near-surface regions.

Figure 2 shows cross-sectional TEM images of sample A1 and sample A3. Both samples exhibit extensive regions of BN with darker contrast just above the heterointerface, corresponding to highly defective layers that contain twins and stacking faults. These regions occupy roughly comparable proportions of the two films. Thus, it appears that the increased growth temperature for sample A3 had minimal effect on reducing defect densities in regions near the BN/diamond interface. Furthermore, the film growth rate was unaffected by the change in temperature.

FIG. 2.

Cross-sectional TEM images: (a) sample A1 and (b) sample A3.

FIG. 2.

Cross-sectional TEM images: (a) sample A1 and (b) sample A3.

Close modal

Samples A3 and A4 were prepared using almost identical conditions, as found here to be the most suitable for growth of the less defective cubic phase, but the substrates were subjected to different hydrogen plasma treatment. The substrate temperatures during surface cleaning were almost the same for sample C1 and sample C2 (760 and 750 °C). However, the duration of the H2-plasma treatments was 90 and 10 min, respectively. The initial c-BN layers (⪅30 nm) were compared to assess any differences in morphology. Regions below and above the BN/diamond heterointerface in samples A3 and A4 can be compared in Figs. 3(a) and 3(b), respectively. The regions above the heterointerface in sample A3 displays a higher proportion of film with relatively darker contrast compared to sample A4 which are associated with high density of twins and stacking faults.

FIG. 3.

Cross-sectional TEM images: (a) sample A3 and (b) sample A4.

FIG. 3.

Cross-sectional TEM images: (a) sample A3 and (b) sample A4.

Close modal

Figure 4 shows a comprehensive EELS spectrum analysis from the diamond-BN interface region of sample A3. Figure 4(a) is a spectrum image that combines EELS data for energies over the range from 180 to 460 eV. It is overlaid with the sp2-B (blue) and sp3-B (green) signal. The white bars in the image, labeled A, B, and C, are positioned on c-BN, interface, and diamond regions, respectively, and shadow the pixels that were used for generating the spectra shown in Fig. 4(b). The spectra confirm the presence of boron and nitrogen atoms to the left of the interface (growth direction), and carbon to the right of the interface, and also highlight the different bonding states of these elements. The prominent π* and σ* peaks of boron at around 191 and 197 eV, carbon at around 286 and 294 eV, and nitrogen at around 401 and 405 eV, which are characteristic of sp2- and sp3-bonding, respectively, are then used to provide sp2–sp3 maps of the region. The signal intensities of these maps are summed perpendicular to the growth direction and then normalized to produce the plots shown in Fig. 4(c). The sp2 and sp3 curves of boron and nitrogen both reveal transitions in the type of bonding on moving from nucleation layer to the epilayer. Similarly, the carbon bonding changes from the sp3-bonded state of diamond cubic material to an sp2-bonded state close to the BN/diamond heterointerface. Notably, bluish regions observed near the interface in Fig. 4(a) correspond to the higher proportion of lower intensity sp2 EELS signal, which reveals some local variability in the width of the transition region.

FIG. 4.

Cross-sectional EELS spectrum analysis of Sample A3. (a) Image formed using high-loss (180–450 eV) EELS signal with the sp2 (blue) and sp3 (green) B signals overlaid. White bars labeled A, B, and C are positioned on c-BN, interface, and diamond regions, respectively, indicate where EELS signal have been summed to provide the corresponding EELS spectra at different positions relative to the interface, as shown in (b). (c) Signal intensities corresponding to π* and σ* peaks of boron, carbon, and nitrogen were normalized and plotted as a function of position (x). (d) Fitting of sp2- and sp3-B curves from (c) is shown.

FIG. 4.

Cross-sectional EELS spectrum analysis of Sample A3. (a) Image formed using high-loss (180–450 eV) EELS signal with the sp2 (blue) and sp3 (green) B signals overlaid. White bars labeled A, B, and C are positioned on c-BN, interface, and diamond regions, respectively, indicate where EELS signal have been summed to provide the corresponding EELS spectra at different positions relative to the interface, as shown in (b). (c) Signal intensities corresponding to π* and σ* peaks of boron, carbon, and nitrogen were normalized and plotted as a function of position (x). (d) Fitting of sp2- and sp3-B curves from (c) is shown.

Close modal

In order to estimate the effective width of the nucleation layer, the boron curves were selected for curve fitting since lower energy ranges in an EELS spectrum yield higher signal counts with lower background noise. Figure 4(d) shows the fitted gaussian and sigmoidal curves based on normalized signal intensities of the sp2- and sp3-bonded boron across the interface. The corresponding adjusted R-squared values are 0.991 and 0.999, respectively. Using the fitted sigmoidal curve based on the sp3-bonded boron signal intensities, the lateral separation from 10% to 90% of the highest intensity level was estimated to be 2.5 ± 0.3 nm. Since the effect of probe size and beam spreading have not been taken into account, this figure represents an upper estimate for the width of the nucleation layer where there are transitions in bonding characteristics for each of boron, nitrogen, and carbon.

In this study, it was not expected that different substrate orientations would contribute to a change in phase of the BN films because of the very small lattice mismatch with the substrate. Thus, the structure of the BN films grown on (111) surface (sample A1) and (001) surface (sample A2) can be compared based solely on the effect of different gas precursor concentrations and ratio.

Gas precursor concentration has been shown previously to be critical in establishing the phase of bulk BN films.36,37 Hydrogen and fluorine gas precursors are generally considered as key reactants for diamond growth since they help to maintain the sp3 bonding configuration of the carbon surface during growth.38 Similarly, they have been used here to help attain the cubic phase of BN. Fluorine etches the sp2-bonded B present in t- or h-BN six times faster than it etches sp3-bonded B in c-BN.39 It also stabilizes the B-layer of the c-BN (111) plane for further growth.40 The respective roles of the H2 and BF3 precursors in the BN growth was investigated here by changing their respective concentrations in the Ar–N2–BF3–H2 gas mixture. These results are consistent with the fact that varying the concentrations of the gas-phase hydrogen and BF3 could establish two different growth regimes, as described by the following set of chemical equations:41 
(1)
(2)

An increase in the H2 concentration increases the concentration of NHx species, which is essential for BN film growth, while it decreases the concentration of free F radicals, which are crucial for the activation of H-terminated N surface atoms and for etching away of any non-cubic BN constituent.42,43 The experimental results here confirm that a nucleation step with optimized H/F ratio and absolute concentrations of gas species during the growth step are both crucial for achieving the desired plasma chemistry that is needed to grow BN films with a dominant cubic phase.

The effect of etching due to different H2/BF3 ratios in the gas mixture was also apparent from the different film thicknesses of samples A1 and A2. Sample A1 had a greater concentration of F radicals in the gas mixture with H2/BF3 = 0.75 compared to sample A2 (H2/BF3 = 1), during the initial growth stage, and it had lower film thickness overall (150 nm vs 250 nm).

It has been established that a hydrogen-limited environment is more conducive for growth of the BN cubic phase.9,41 Thus, the substrate temperature was varied here to determine any other morphological changes while keeping the same gas precursor concentrations. The primary defects observed in the resulting c-BN films were twins and stacking faults, albeit in different proportions.

Figure 2 presents a comparative analysis of the c-BN films in samples A1 and A3. Notably, the darker contrast observed in regions just above the BN/diamond interface is indicative of a high density of twin-related c-BN domains and stacking faults. Despite the higher growth temperature for sample A3, a comparative assessment of the size of these darker, defect-rich regions showed no significant difference. This observation suggests that the types and densities of defects occurring near the BN/diamond heterointerface are minimally affected by changes in growth temperature. In contrast, as observed in multiple TEM images of the two samples, regions farther from the heterointerface in A3 exhibited larger domain sizes compared to A1. This enlargement is most likely due to the enhanced thermal energy available to B and N adatoms at higher temperatures, which increases their in-plane mobility on the growth surface, leading to more ordered and hence less defective cubic phase growth. This difference was further shown by the sharper diffraction spots in the corresponding FFTs of these images, taken from identically sized regions while moving away from the substrate. To confirm this qualitative distinction, additional observations across extensive cross sections of both films are required.

It was observed that growth at higher temperatures resulted in larger c-BN domains and a reduction in defect density. However, the initial layers of c-BN were predominantly characterized by high densities of twin domains and stacking faults. EELS experiments were thus conducted to provide an understanding of the nucleation and early stages of c-BN layer growth, and its dependence on substrate surface quality.

The EELS analysis shown in Fig. 4 established that the initial layers of the film predominantly contained sp2-bonded BN, and transitioned to almost entirely sp3-bonded BN later in the growth process. Similarly, there is a shift of carbon bonding from exclusively sp3 in diamond to a mixture of sp2 and sp3 in the nucleation layer. These observations suggest the possibility that the diamond surface undergoes some disordering either as a result of the cleaning process or during the early stages of BN growth. Other potential causes for the transition could be the influence of very low levels of residual hydrogen and fluorine contaminants or irradiation with H+ ions.

A mechanism has been proposed previously explaining how growth temperature can affect the development of multi-domain heteroepitaxial c-BN (111) structures grown on diamond (111) substrates.26 This same mechanism can be adapted here to explain the formation and relative density of twin-related c-BN domains depending on the substrate cleaning method. The c-BN (111) plane is polar, while the bond energies for C–B and C–N are 444 and 305 kJ/mol, respectively,44,45 suggesting that C–B bonds are more likely to form at the c-BN/diamond interface.

Figure 5(a) shows an HRTEM image of c-BN multi-domain structures near the c-BN/diamond interface in sample A4. Epitaxial c-BN domains are enclosed with yellow dashed lines and a rotated c-BN domain is demarcated with a teal dashed line. The red line segment(s) highlights the structural relationship and the presence of a lattice defect in the region. For N-polar c-BN (111), as sketched in Fig. 5(b), the stacking sequence would depend on the location of incoming nitrogen atoms. For sp3-bonded BN, the zincblende structure is more stable than the metastable wurtzite structure.46,47 Considering probabilistic events, nitrogen atoms can attach at specific sites (a′) for the wurtzite structure during growth, resulting in stacking faults (SFs) that would, in turn, lead to the formation of rotational twins. This process can be considered as equivalent to inserting a single layer of wurtzite structure with the 2H stacking sequence into the 3C stacking sequence, as sketched in Fig. 5(c). The SF at the heterointerface inverts the 3C stacking sequence from “…cbacba” to “abcabc….” For a specific growth temperature, surface amorphization due to H2-plasma cleaning is postulated to increase the potential barrier for the migration of N atoms to c sites (zincblende structure) from the a′ sites (wurtzite structure). This phenomenon could account for the observed higher proportion of twin-related c-BN domains (darker contrast regions above the heterointerface), especially in sample A3, which was subjected to prolonged cleaning, potentially leading to greater surface roughness.

FIG. 5.

(a) XTEM image showing the cross section of sample A4 with epitaxially grown c-BN domains (yellow) and rotated twin c-BN domains (teal). Red line segment(s) represents structural relationship or presence of a defect; (b) ball-and-stick atomic model illustrating growth in the [111] direction with surface sites that would be suitable for growth of zincblende (c) or wurtzite (a') structure: (c) Model for SF formation at the heterointerface resulting in growth rotational twin.

FIG. 5.

(a) XTEM image showing the cross section of sample A4 with epitaxially grown c-BN domains (yellow) and rotated twin c-BN domains (teal). Red line segment(s) represents structural relationship or presence of a defect; (b) ball-and-stick atomic model illustrating growth in the [111] direction with surface sites that would be suitable for growth of zincblende (c) or wurtzite (a') structure: (c) Model for SF formation at the heterointerface resulting in growth rotational twin.

Close modal

The morphology of BN films grown on diamond substrates using ECR PECVD has been investigated using cross-sectional TEM, focusing on phase formation, defect characteristics, and structural evolution during the initial stages of growth. Within the studied range of growth temperatures (735–820 °C) and using a specific hydrogen plasma substrate cleaning method, it was established that the concentrations of gas precursors played a crucial role in determining whether the primary phase of the BN film was cubic or turbostratic. H2-deficient growth (nucleation step) leads to essentially phase-pure c-BN. Twins and stacking faults were predominantly found in the bulk of c-BN films grown with a nucleation step. Higher growth temperatures resulted overall in a reduction in defect densities and larger c-BN domains, in regions of the film away from the interface. Surface cleaning proved to be critical for the nucleation of heteroepitaxial c-BN domains. Aggressive cleaning tended to roughen the substrate surface and contributed to a higher proportion of rotational twins. A plausible mechanism correlating surface roughness with rotation twin formation was presented. In order to further improve film quality and achieve single-domain epitaxial c-BN formation, it is recommended that the BN films should be grown at higher temperatures. The surface preparation should also be improved in order to minimize the potential barriers for N adatoms to migrate to stable sites more suitable for the growth of the zincblende structure.

This work was supported as part of ULTRA, an Energy Frontier Research Center funded by the U.S. Department of Energy, Office of Science, Basic Energy Sciences under Award No. DE-SC0021230. The authors acknowledge the use of facilities within the John M. Cowley Center for High Resolution Electron Microscopy at Arizona State University. We also thank Professor Martha R. McCartney for assistance with some microscopy.

The authors have no conflicts to disclose.

Saurabh Vishwakarma: Conceptualization (lead); Data curation (lead); Formal analysis (lead); Investigation (lead); Writing – original draft (lead). Avani Patel: Investigation (equal); Writing – review & editing (supporting). Manuel R. Gutierrez: Investigation (supporting); Writing – review & editing (supporting). Robert J. Nemanich: Investigation (equal); Supervision (equal); Writing – review & editing (equal). David J. Smith: Conceptualization (equal); Investigation (equal); Supervision (equal); Writing – original draft (equal); Writing – review & editing (equal).

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

1.
J. Y.
Tsao
,
S.
Chowdhury
,
M. A.
Hollis
,
D.
Jena
,
N. M.
Johnson
,
K. A.
Jones
,
R. J.
Kaplar
,
S.
Rajan
,
C. G.
Van de Walle
,
E.
Bellotti
,
C. L.
Chua
,
R.
Collazo
,
M. E.
Coltrin
,
J. A.
Cooper
,
K. R.
Evans
,
S.
Graham
,
T. A.
Grotjohn
,
E. R.
Heller
,
M.
Higashiwaki
,
M. S.
Islam
,
P. W.
Juodawlkis
,
M. A.
Khan
,
A. D.
Koehler
,
J. H.
Leach
,
U. K.
Mishra
,
R. J.
Nemanich
,
R. C. N.
Pilawa-Podgurski
,
J. B.
Shealy
,
Z.
Sitar
,
M. J.
Tadjer
,
A. F.
Witulski
,
M.
Wraback
, and
J. A.
Simmons
, “
Ultrawide-bandgap semiconductors: Research opportunities and challenges
,”
Adv. Electron. Mater.
4
(
1
),
1600501
(
2018
).
2.
X.
Zhang
and
J.
Meng
, “
Chapter 4 - Recent progress of boron nitrides
,” in
Materials Today: Ultra-Wide Bandgap Semiconductor Materials
, edited by M. Liao, B. Shen, and Z. Wang (Elsevier,
2019
), pp. 347–419.
3.
J. T.
Mullen
,
J. A.
Boulton
,
M.
Pan
, and
K. W.
Kim
, “
Electronic properties of c-BN/diamond heterostructures for high-frequency high-power applications
,”
Diam. Relat. Mater.
143
,
110920
(
2024
).
4.
M. H.
Wong
,
O.
Bierwagen
,
R. J.
Kaplar
, and
H.
Umezawa
, “
Ultrawide-bandgap semiconductors: An overview
,”
J. Mater. Res.
36
(
23
),
4601
4615
(
2021
).
5.
S.
Jia
,
S.
Fu
,
Y.
Liu
,
N.
Gao
,
H.
Li
, and
M.
Liao
, “
Interfacial charge transfer and electronic structure of diamond/c-BN heterointerface
,”
Comput. Mater. Sci.
218
,
111947
(
2023
).
6.
M. J.
Purnomo
,
Y.
Febrita
,
O.
Dinaryanto
,
W.
Gierlotka
, and
I. S.
Yu
, “
The stability prediction and epitaxial growth of boron nitride nanodots on different substrates
,”
Molecules
29
(
6
),
1313
(
2024
).
7.
V. L.
Solozhenko
,
I. A.
Petrusha
, and
A. A.
Svirid
, “
Thermal phase stability of rhombohedral boron nitride
,”
High Pressure Res.
15
(
2
),
95
103
(
1996
).
8.
W. J.
Yu
,
W. M.
Lau
,
S. P.
Chan
,
Z. F.
Liu
, and
Q. Q.
Zheng
, “
Ab initio study of phase transformations in boron nitride
,”
Phys. Rev. B
67
(
1
),
1
9
(
2003
).
9.
C.
Cazorla
and
T.
Gould
, “
Polymorphism of bulk boron nitride
,”
Sci. Adv.
5
(
1
),
1
5
(
2019
).
10.
R. H.
Wentorf
, “
Cubic form of boron nitride
,”
J. Chem. Phys.
26
(
4
),
956
(
1957
).
11.
O.
Mishima
,
S.
Yamaoka
, and
O.
Fukunaga
, “
Crystal growth of cubic boron nitride by temperature difference method at ∼55
kbar and ∼1800 °C
,”
J. Appl. Phys.
61
(
8
),
2822
2825
(
1987
).
12.
T.
Taniguchi
,
T.
Teraji
,
S.
Koizumi
,
K.
Watanabe
, and
S.
Yamaoka
, “
Appearance of n-type semiconducting properties of cBN single crystals grown at high pressure
,”
Jpn. J. Appl. Phys., Part 2 Lett.
41
(
2A
),
L109
L111
(
2002
).
13.
L. C.
Nistor
,
S. V.
Nistor
,
G.
Dinca
,
P.
Georgeoni
,
J.
Van Landuyt
,
C.
Manfredotti
, and
E.
Vittone
, “
Microstructure and spectroscopy studies on cubic boron nitride synthesized under high-pressure conditions
,”
J. Phys.: Condens. Matter
14
(
44 Spec. Iss.
),
10983
10988
(
2002
).
14.
D. L.
Medlin
,
T. A.
Friedmann
,
P. B.
Mirkarimi
,
G. F.
Cardinale
, and
K. F.
McCarty
, “
Crystallographic texture in cubic boron nitride thin films
,”
J. Appl. Phys.
79
(
7
),
3567
3571
(
1996
).
15.
M. A.
Djouadi
,
O.
Banakh
,
A.
Soltani
,
R.
Sanjines
, and
F.
Levy
, “
Physical properties of cubic boron nitride films synthesized by PVD using a bi-step process
,”
Thin Solid Films
398
(
399
),
205
209
(
2001
).
16.
H.
Hofsäss
,
S.
Eyhusen
, and
C.
Ronning
, “
On the mechanisms of cubic boron nitride film growth
,”
Diamond Relat. Mater.
13
(
4–8
),
1103
1110
(
2004
).
17.
W. J.
Zhang
,
Y. M.
Chong
,
I.
Bello
, and
S. T.
Lee
, “
Nucleation, growth and characterization of cubic boron nitride (cBN) films
,”
J. Phys. D: Appl. Phys.
40
(
20
),
6159
6174
(
2007
).
18.
M.
Satou
, “
Formation of cubic boron nitride films by boron evaporation and nitrogen ion beam bombardment
,”
J. Jpn Soc. Precis. Eng.
53
(
10
),
1540
1543
(
1987
).
19.
D. J.
Kester
and
R.
Messier
, “
Phase control of cubic boron nitride thin films
,”
J. Appl. Phys.
72
(
2
),
504
513
(
1992
).
20.
F. R.
Corrigan
and
F. P.
Bundy
, “
Direct transitions among the allotropic forms of boron nitride at high pressures and temperatures
,”
J. Chem. Phys.
63
(
9
),
3812
3820
(
1975
).
21.
J.
Wang
and
Y. K.
Yap
, “
Growth of adhesive cubic phase boron nitride films without argon ion bombardment
,”
Diam. Relat. Mater.
15
(
2–3
),
444
447
(
2006
).
22.
H. S.
Yang
,
C.
Iwamoto
, and
T.
Yoshida
, “
Nanostructures of the turbostratic BN transition layer in cubic BN thin films deposited by low-pressure inductively coupled plasma-enhanced chemical vapor deposition
,”
J. Appl. Phys.
91
(
10 I
),
6695
6699
(
2002
).
23.
V. L.
Solozhenko
,
V. Z.
Turkevich
, and
W. B.
Holzapfel
, “
Refined phase diagram of boron nitride
,”
J. Phys. Chem. B
103
(
15
),
2903
2905
(
1999
).
24.
W. J.
Zhang
,
I.
Bello
,
Y.
Lifshitz
, and
S. T.
Lee
, “
Recent advances in cubic boron nitride deposition
,”
MRS Bull.
28
(
3
),
184
188
(
2003
).
25.
D. F.
Storm
,
S. I.
Maximenko
,
A. C.
Lang
,
N.
Nepal
,
T. I.
Feygelson
,
B. B.
Pate
,
C. A.
Affouda
, and
D. J.
Meyer
, “
Mg-facilitated growth of cubic boron nitride by ion beam-assisted molecular beam epitaxy
,”
Phys. Status Solidi RRL
16
(
7
),
2200036
(
2022
).
26.
K.
Hirama
,
Y.
Taniyasu
,
H.
Yamamoto
, and
K.
Kumakura
, “
Structural analysis of cubic boron nitride (111) films heteroepitaxially grown on diamond (111) substrates
,”
J. Appl. Phys.
125
(
11
),
115303
(
2019
).
27.
Y.
Nikaido
,
T.
Ichibha
,
K.
Hongo
,
F. A.
Reboredo
,
K. C. H.
Kumar
,
P.
Mahadevan
,
R.
Maezono
, and
K.
Nakano
, “
Diffusion Monte Carlo study on relative stabilities of boron nitride polymorphs
,”
J. Phys. Chem. C
126
(
13
),
6000
6007
(
2022
).
28.
D. J.
Kester
,
K. S.
Ailey
,
D. J.
Lichtenwalner
, and
R. F.
Davis
, “
Growth and characterization of cubic boron nitride thin films
,”
J. Vac. Sci. Technol. A
12
(
6
),
3074
3081
(
1994
).
29.
E. S.
Lee
,
J. K.
Park
,
W. S.
Lee
,
T. Y.
Seong
, and
Y. J.
Baik
, “
Effect of deposition temperature on cubic boron nitride thin film deposited by unbalanced magnetron sputtering method with a nanocrystalline diamond buffer layer
,”
Met. Mater. Int.
19
(
6
),
1323
1326
(
2013
).
30.
G. L.
Doll
,
J. A.
Sell
,
C. A.
Taylor
, and
R.
Clarke
, “
Growth and characterization of epitaxial cubic boron nitride films on silicon
,”
Phys. Rev. B
43
(
8
),
6816
6819
(
1991
).
31.
A. K.
Ballal
,
L.
Salamanca-Riba
,
C. A.
Taylor
, and
G. L.
Doll
, “
Structural characterization of preferentially oriented cubic BN films grown on Si (001) substrates
,”
Thin Solid Films
224
(
1
),
46
51
(
1993
).
32.
Y. H.
Choi
,
J. Y.
Huh
,
J. K.
Park
,
W. S.
Lee
, and
Y. J.
Baik
, “
Microstructure of ⟨111⟩-textured cubic boron nitride film deposited under oxygen-containing atmosphere
,”
Adv. Eng. Mater.
25
(
21
),
1
6
(
2023
).
33.
A.
Patel
,
S.
Vishwakarma
,
D. J.
Smith
, and
R. J.
Nemanich
, Surface chemistry of c-BN epitaxial growth on diamond substrates using fluorine-assisted ECR PECVD,” unpubl. results (n.d.).
34.
C.
Cayron
, “
Diffraction artefacts from twins and stacking faults, and the mirage of hexagonal, polytypes or other superstructures
,”
Scr. Mater.
194
,
113629
(
2021
).
35.
S.
Vishwakarma
,
A.
Patel
,
R. J.
Nemanich
, and
D. J.
Smith
, Understanding complex contrast features in HRTEM imaging of cubic boron nitride,” unpubl. results (n.d.).
36.
E. A.
Il’inchik
and
I. S.
Merenkov
, “
X-ray photoelectron study of the effect of the composition of the initial gas phase on changes in the electronic structure of hexagonal boron nitride films obtained by PECVD from borazine
,”
J. Struct. Chem.
57
(
4
),
670
678
(
2016
).
37.
S. V.
Nguyen
,
T.
Nguyen
,
H.
Treichel
, and
O.
Spindler
, “
Plasma-assisted chemical vapor deposition and characterization of boron nitride films
,”
J. Electrochem. Soc.
141
(
6
),
1633
1638
(
1994
).
38.
W. J.
Zhang
and
S.
Matsumoto
, “
The roles of hydrogen and fluorine in the deposition of cubic boron nitride films in the Ar-N2-BF3-H2 system
,”
Chem. Phys. Lett.
330
(
3–4
),
243
248
(
2000
).
39.
W.
Kalss
,
R.
Haubner
, and
B.
Lux
, “
Preparation of BN films in the B–N–F system
,”
Diam. Relat. Mater.
7
(
2–5
),
369
375
(
1998
).
40.
Z.
Sun
,
N.
Gao
, and
H.
Li
, “
Structural and electronic properties of c-BN (111) surface with hydrogen/fluorine functionalization and nitrogen-based small-molecule adsorption
,”
J. Phys.: Condens. Matter
32
(
26
),
265002
(
2020
).
41.
J. M.
Brown
,
S.
Vishwakarma
,
D. J.
Smith
, and
R. J.
Nemanich
, “
Nucleation of cubic boron nitride on boron-doped diamond via plasma enhanced chemical vapor deposition
,”
J. Appl. Phys.
133
(
21
),
215303
(
2023
).
42.
W. J.
Zhang
,
C. Y.
Chan
,
X. M.
Meng
,
M. K.
Fung
,
I.
Bello
,
Y.
Lifshitz
,
S. T.
Lee
, and
X.
Jiang
, “
The mechanism of chemical vapor deposition of cubic boron nitride films from fluorine-containing species
,”
Angew. Chem., Int. Ed.
44
(
30
),
4749
4753
(
2005
).
43.
W.
Zhang
,
I.
Bello
,
Y.
Lifshitz
,
K. M.
Chan
,
X.
Meng
,
Y.
Wu
,
C. Y.
Chan
, and
S. T.
Lee
, “
Epitaxy on diamond by chemical vapor deposition: A route to high-quality cubic boron nitride for electronic applications
,”
Adv. Mater.
16
(
16
),
1405
1408
(
2004
).
44.
P. W.
Atkins
and
J.
de Paula
,
Physical Chemistry
(
Oxford University Press
,
2009
), Vol. 9, p.
976
.
45.
B. D.
Darwent
, “
Bond dissociation energies in simple molecules. National standard reference data series, national bureau of standards, 31
,”
Natl. Stand. Ref. Data Syst.
31
,
23
(
1970
).
46.
H.
Neumann
, “
J. H. Edgar (ed.). properties of group III nitrides. (EMIS Datareviews Series No. 11). INSPEC, The Institution of Electrical Engineers, London 1994. 302 Seiten, 121 Abbildungen, 77 Tabellen. ISBN 0-85296-818-3
,”
Cryst. Res. Technol.
30
(
7
),
910
(
1995
).
47.
S. L.
Qiu
and
P. M.
Marcus
, “
Structure and stability under pressure of cubic and hexagonal diamond crystals of C, BN and Si from first principles
,”
J. Phys.: Condens. Matter
23
(
21
),
215501
(
2011
).