Next Article in Journal
Effect of Nano-TiO2 Addition on Some Properties of Pre-Alloyed CoCrMo Fabricated via Powder Technology
Previous Article in Journal
Experimental Study and Numerical Modeling of Inter-Pass Forging in Wire-Arc Additive Manufacturing of Inconel 718
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancing the Quality of Diamond Film Growth Through the Synergistic Addition of Nitrogen and Carbon Dioxide

1
College of Chemical Engineering, Zhejiang University of Technology, Hangzhou 310000, China
2
State Key Laboratory of Advanced Marine Materials, Zhejiang Key Laboratory of Extreme-environmental Material Surfaces and Interfaces, Ningbo Institute of Materials Technology and Engineering, Chinese Academy of Sciences, Ningbo 315201, China
*
Authors to whom correspondence should be addressed.
Materials 2026, 19(1), 183; https://doi.org/10.3390/ma19010183
Submission received: 21 November 2025 / Revised: 30 December 2025 / Accepted: 31 December 2025 / Published: 4 January 2026
(This article belongs to the Section Materials Physics)

Abstract

This study investigates the synergistic effects of co-doping with ultralow-concentration nitrogen and trace carbon dioxide on the growth of polycrystalline diamond films via microwave plasma chemical vapor deposition (MPCVD). The films were characterized using scanning electron microscopy, X-ray diffraction, Raman spectroscopy, and photoluminescence spectroscopy. Results indicate that trace nitrogen effectively promotes <111> oriented growth and enhances the deposition rate, whereas excessive nitrogen leads to the formation of defects such as pores and microcracks. The introduction of CO2 suppresses the formation of nitrogen-vacancy-related defects through a selective etching mechanism. Under co-doping conditions, diamond films with high growth rates, strong <111> texture, and superior thermal conductivity (up to 1863.94 W·m−1·K−1) were successfully synthesized, demonstrating significant potential for thermal management applications in high-power integrated circuits.

1. Introduction

Diamond films possess exceptional hardness, outstanding thermal conductivity, a broad optical transmission window, and excellent chemical stability. These remarkable properties make them promising candidates for a wide range of applications in fields such as mechanical processing, optical windows, semiconductors, and thermal management [1,2]. Chemical vapor deposition (CVD) has become the method of choice for fabricating high-quality diamond films [3]. However, conventional CVD processes, particularly those employing hydrogen and methane as the primary gas precursors, often face challenges in producing high-quality diamond films. The performance of polycrystalline diamond—an aggregate of micron- or nano-sized diamond grains sintered together—is dictated not only by the intrinsic hardness of the diamond crystallites but more critically by the quality of intergranular bonding, the grain size distribution, and the purity of the grain boundaries. Typically, PCD films suffer from microstructural defects such as highly disordered grain orientations, the presence of excessive impurities or non-diamond carbon (e.g., graphite, amorphous carbon) at the grain boundaries, and non-uniform grain sizes [3,4,5,6,7,8]. In MPCVD, the diamond growth kinetics are dictated by the underlying plasma chemistry, itself determined by a set of parameters including gas composition, process pressure, gas flow rate, and substrate temperature [9,10,11,12,13,14,15].
To optimize the growth quality of diamond films, the use of additive gases has been widely adopted as a well-established strategy. Introducing nitrogen gas (N2) is a common approach, which has been proven to effectively enhance the growth rate and modify the crystal morphology. However, the sole incorporation of nitrogen often introduces an excessive number of defects, such as nitrogen-vacancy (NV) centers, leading to unstable electrical properties and even a darkening of the film, which compromises its performance in optical applications [15]. Yang et al. [15] successfully synthesized highly textured polycrystalline diamond via ultra-low-level nitrogen doping, achieving rapid growth rates (~3–8 μm/h) and high infrared transmittance (~49.2–70.8%) across a broad spectrum of 3.0–15.0 μm. In contrast, Ashkinazi et al. [16] employed a microwave plasma CVD process with a high-concentration CH4-H2-N2 atmosphere, demonstrating that the morphology of the diamond film grown on a single-crystal substrate was strongly dependent on the substrate’s crystal orientation. However, a critical paradox emerges from these doping strategies. Analysis by scanning electron microscopy and Raman spectroscopy reveals that although nitrogen doping enhances the growth rate, an excessively high rate can induce detrimental microstructural defects, such as surface microcracks and intergranular voids. These defects, in turn, degrade the crystallinity and ultimately compromise the optical performance of the film. The incorporation of oxygen into the process provides a source of highly reactive oxygen atoms (O) and hydroxyl radicals (OH). These species preferentially etch sp2-bonded carbon by reacting with it to form volatile carbon monoxide (CO) and carbon dioxide (CO2), which are then evacuated from the system. In MPCVD, this is typically achieved by introducing gases such as carbon dioxide (CO2) [17]. This mechanism functions as an in situ etching process, which continuously and selectively suppresses the growth of graphitic phases and other defects during diamond deposition, thereby purifying the growth environment and ultimately yielding higher-quality diamond films [18,19]. The work by Gu et al. [20] showed that high-level oxygen doping, particularly at elevated pressures, significantly boosts the etching efficiency of graphite, thereby producing diamond films of higher purity and greater thickness. This enhanced growth was accompanied by an increase in deposition rate and resultant surface roughness. A key microstructural outcome was the substantial increase in the density of (111) facets on the film surface facilitated by the high-pressure conditions. In their study on boron-doped diamond films synthesized via microwave plasma chemical vapor deposition, Zhao et al. [21] observed that as the oxygen concentration (O/C ratio) increased from 0% to 5%, the film’s growth rate initially decreased and then increased. The introduction of oxygen was found to suppress boron incorporation, improve crystal quality, and significantly inhibit residual nitrogen. Hall effect measurements confirmed that at an O/C ratio of 3%, a high-mobility, high-quality diamond film was achieved with a growth rate of 9 μm/h. This work suggests that trace Carbon dioxide can compensate for the crystal strain induced by boron doping, thereby enhancing both crystalline and surface quality. Although a moderate Carbon dioxide addition can enhance film purity and crystallinity by selectively etching amorphous carbon and graphitic phases formed during growth, excessive etching introduces defects. Specifically, if the CO2 concentration becomes too high, even the diamond phase undergoes significant etching. This leads to a roughened growth surface, increased defect density, and grain boundary formation, ultimately degrading the film quality. While moderate Carbon dioxide doping enhances diamond film purity and crystallinity by selectively etching amorphous carbon and graphitic phases, excessive etching, particularly from high CO2 concentrations, becomes detrimental. It not only attacks the diamond phase itself—leading to a roughened surface, increased defects, and grain boundaries that degrade quality—but also disrupts the optimal CVD growth environment. A hydrogen-rich atmosphere is crucial, as atomic hydrogen stabilizes sp3 bonds and passivates surface dangling bonds. The introduction of excessive CO2 can upset this balance by consuming active hydrogen species (e.g., forming H2O or OH), thereby shifting the growth-etch equilibrium unfavorably and potentially promoting the deposition of non-diamond carbon films. Given the inherent limitations of single-gas additives in precision and efficacy, the synergistic effect of co-introducing gases presents a significant yet underexplored avenue.
This study proposes an innovative multi-component co-doping strategy that achieves precise modulation of diamond crystal facet competition by synchronously introducing trace amounts of carbon dioxide into a ppm-level nitrogen doping system during microwave plasma chemical vapor deposition (MPCVD). This approach effectively addresses longstanding technical challenges in conventional nitrogen-doped polycrystalline diamond fabrication, including disordered crystal orientation, lattice distortion, suboptimal crystallinity, and low thermal conductivity.
Research demonstrates that the introduction of carbon dioxide, through mechanisms such as selective etching of non-diamond phases and regulation of surface reaction kinetics pathways, significantly suppresses the proliferation of crystal defects induced by single-element nitrogen doping. This enables synergistic modulation rather than mere enhancement of the nitrogen doping effect. The nitrogen-doped polycrystalline diamond films fabricated under varying carbon dioxide flow parameters exhibit outstanding comprehensive performance: a highly <111>-oriented crystal structure is successfully achieved, accompanied by a growth rate of approximately 3–4 μm/h and a high thermal conductivity of 1863.94 W·m−1·K−1, while also offering advantages of short fabrication cycles and relatively low energy consumption.
This work systematically elucidates the dynamic balancing mechanism and crystal quality optimization principles of carbon dioxide within the co-doping system. The developed high-performance diamond films demonstrate significant potential for engineering applications in the thermal management of high-power integrated circuit chips operating under high-speed and extreme thermal conditions.

2. Materials and Methods

Free-standing diamond films were synthesized using a microwave plasma chemical vapor deposition (MPCVD) system, where a plasma was generated under strong electromagnetic fields to facilitate diamond deposition. The source gases were hydrogen (H2, 99.999%, Linde, Shanghai, China) and methane (CH4, 99.999%, Linde, Shanghai, China). Nitrogen concentration was controlled by introducing an H2-N2 mixture (1% N2/99.9% H2, Linde, Shanghai, China). A 2-inch diameter, 3 mm thick, p-type doped single-crystal silicon wafer with a (100) orientation served as the substrate. The CH4 to H2 flow ratio was fixed at 3:100, with the CH4 flow rate maintained at 12 sccm for a growth duration of 40 h. The nitrogen concentration during deposition was regulated by introducing 0–1.6 sccm of the H2-N2 mixture into the reaction chamber. The formula for calculating nitrogen concentration (relative to H2) is:
N2(ppm) = MN flow rate × 1% ÷ (H2 flow rate + MN flow rate × 99%) × 106
The substrate temperature was kept at 850 ± 10 °C, monitored by an infrared pyrometer (IR-AHS0, Chino, Japan), and the chamber pressure was set at 100 Torr. The relevant parameters are summarized in Table 1.
Substrate pretreatment is a critical step for the heteroepitaxial growth of diamond films, primarily aimed at surface cleaning, nucleation enhancement, and improved film adhesion. A three-step pretreatment process was employed on the silicon substrate: 1. Ultrasonic Cleaning: The substrate was ultrasonically cleaned in absolute ethanol to remove surface contaminants such as organic residues and dust particles, thereby preventing them from adversely affecting the epitaxial quality of the diamond film; 2. Nanodiamond Seeding and Secondary Ultrasonication: A colloidal suspension of W10 nanodiamond powder in ethanol was uniformly drop-coated onto the substrate surface. Subsequently, the sample was immersed in acetone and subjected to a second ultrasonic treatment. This step served to further purify the substrate surface and pre-deposit nanodiamond particles, which act as nucleation centers; 3. Mechanical Abrasion: The substrate surface was continuously rubbed in a unidirectional manner with a lint-free cloth. This mechanical action created uniform micro-scratches and embedded nanodiamond seeds, significantly increasing the density of nucleation sites. The treated substrate exhibits a clean surface with a uniform defect distribution, which effectively enhances the adhesion between the diamond film and the substrate. Insufficient adhesion may lead to film delamination or cracking during growth or subsequent processing, resulting in epitaxial failure. Therefore, substrate pretreatment serves as a critical guarantee for high-quality heteroepitaxial diamond growth. After baking in dry air, the silicon wafer was placed into a laboratory-developed disc-shaped resonant MPCVD (GMP60KIPT400, with a maximum power of 6 kw) chamber for further processing. The diamond film was then deposited according to the process parameters listed in Table 1. The heating and cooling rates for the sample were controlled at approximately 15 °C/min. Following the growth process, the sample was cooled down and subsequently removed from the chamber. To obtain a free-standing diamond film, the silicon substrate was selectively etched away using an HF/HNO3 mixture (with a volume ratio of HF:HNO3 = 3:1). Finally, the resulting diamond film was thoroughly rinsed with deionized water and dried.
The crystal orientation was analyzed using X-ray diffraction (XRD, Advance D8, Bruker, Billerica, MA, USA) with filtered Cu K-α radiation (λ = 1.5406 Å). The measurement was conducted at room temperature in continuous scanning mode, with a 2θ range of 30° to 130. The crystal orientation was analyzed using X-ray diffraction (XRD, Advance D8, Bruker, USA) with filtered Cu K-α radiation (λ = 1.5406 Å). The measurement was conducted at room temperature in continuous scanning mode, with a 2θ range of 30° to 130°, a step size of 0.02°, and a scanning speed of 2°/min. Phase identification was performed by referring to the Powder Diffraction File (PDF) database of the International Centre for Diffraction Data (ICDD), a step size of 0.02°, and a scanning speed of 2/min. Phase identification was performed by referring to the Powder Diffraction File (PDF) database of the International Centre for Diffraction Data (Specific card number: PDF#06-0675) [22]. Furthermore, the surface morphology and microstructure of the diamond films were examined by scanning electron microscopy (SEM, Regulus 8230, Hitachi, Tokyo, Japan).
To enable further investigation, the growth surface of the free-standing diamond film was mirror-polished using a polishing machine (Model: SPG40-II, Supower, Nanjing, China) with commercial diamond powder (median particle size ~2.5 μm) as the abrasive medium. The thermal properties of the diamond material were subsequently characterized using a laser flash analyzer (NETZSCH LFA 467, Yokohama, Japan).

3. Result and Discussion

3.1. Diamond Film Growth Rate and Its Quality

To investigate the film growth rates under different process conditions, the thicknesses of the free-standing diamond films were measured before and after deposition. The growth rate was calculated based on the thickness difference. As summarized in Table 2, the introduction of nitrogen into the process gas significantly enhanced the growth rate of the diamond film. Specifically, the growth rate increased by a factor of 3.1, from 1.13 μm/h to 3.48 μm/h [23,24].
In the MPCVD growth of diamond thin films, surface hydrogen desorption is the rate-determining step. A pristine diamond surface is passivated by hydrogen atoms. Growth species (e.g., CH3 radicals) must locate an active site (a dangling bond) on the surface that is not covered by hydrogen to facilitate bonding. In the absence of doping, the density of active sites (dangling bonds) is very low because hydrogen atoms are strongly bound (strong C–H bond) and difficult to desorb, which severely limits the deposition rate. When nitrogen gas (N2) is introduced and activated into reactive nitrogen species in the plasma, nitrogen atoms become incorporated into the surface layer of the growing diamond film. As revealed in the referenced study, the formation of C–N groups alters the local charge and bonding environment [25]. These C–N sites can serve as energetically more favorable “reaction platforms” compared to pristine C–H sites. Specifically, the C–H bonds near C–N sites are weakened, effectively lowering the activation energy barrier for hydrogen desorption—the rate-determining step. Additionally, C–N groups act as dominant active sites, which can “shorten the migration path of CH2 and other radicals.” In the context of MPCVD, this implies that carbon-containing growth species adsorbed on the surface can more rapidly locate stable bonding positions (such as another active site or a step edge), thereby reducing the chance of re-passivation by hydrogen before bonding occurs and improving the efficiency of carbon incorporation into the diamond lattice. These combined effects collectively contribute to a marked increase in the film’s deposition rate [26,27,28].
Similarly, as shown in Figure 1,the introduction of carbon dioxide (CO2) also led to an increase in the film growth rate at the same nitrogen concentration. This enhancement is attributed to the provision of a more efficient carbon decomposition pathway by Carbon dioxide atoms (O·) and hydroxyl radicals (OH·) derived from CO2. The representative reactions, CH4 + O·→ CH3· + OH· and CH4 + OH· → CH3· + H2O, typically proceed at higher rates than those involving H·. Consequently, under identical power and methane concentration, the gas-phase concentration of reactive methyl radicals (CH3) is significantly elevated. This abundance of key growth precursors directly accelerates the diamond deposition rate [20].
As shown in Figure 2a, the Raman spectra of samples prepared under different parameters all exhibit a strong characteristic peak at 1332 cm−1. The absence of detectable D or G bands around 1350 cm−1 and 1580 cm−1 [29], which are indicative of amorphous carbon or graphitic impurities, confirms the high phase purity of the diamond. Furthermore, with the introduction of nitrogen, the spectra show no significant impurity peaks other than the emerging feature at 1425 cm−1, attributed to the NV0 center [15]. This further attests to the high purity and good crystallinity of the synthesized product. Close examination of Figure 2a reveals that the NV0 peak is not distinctly discernible at low nitrogen concentrations. However, its intensity becomes progressively more pronounced with increasing nitrogen flow. This trend is directly linked to the formation mechanism of nitrogen-vacancy (NV) centers: the incorporation of nitrogen leads to the formation of substitutional nitrogen atoms, which can subsequently bind with adjacent lattice vacancies to create NV centers. A higher nitrogen concentration in the process gas thereby increases the population of these NV centers, resulting in the enhanced visibility of their characteristic Raman signature [26]. The influence of CO2 introduction is revealed in the Raman spectra of Figure 2b, where a marked suppression of the NV0 peak at 1425 cm−1 is observed across all nitrogen levels. This is explained by the etching action of OH radicals, produced from CO2, on the nitrogen-vacancy (NV) centers. Figure 2d highlights that this etching effect is strongest in films grown with high nitrogen influx. Meanwhile, the crystalline quality, as probed by the FWHM of the diamond Raman peak (Figure 2c), shows a non-monotonic dependence on nitrogen concentration. The fitted FWHM values of 3.111, 2.713, 2.744, 3.393, 2.527, 7.382, and 4.174 first decline and then rise. This progression suggests that low-level nitrogen doping enhances crystallinity, whereas further increasing the nitrogen concentration ultimately compromises the crystalline perfection of the diamond film. Furthermore, at nitrogen concentrations above 24.93 ppm, the fitted full width at half maximum (FWHM) values for samples grown with CO2 introduction are consistently lower than those without CO2. This result indicates that the etching effect significantly reduces crystal defects, which are primarily caused by lattice expansion induced by nitrogen incorporation. The improvement in both bulk crystal and surface quality is particularly evident at high nitrogen concentrations.
The residual stress in the diamond film can be determined based on the peak shift in the Raman spectrum, specifically by measuring the difference between the observed diamond peak position and the standard stress-free value of 1332 cm−1 (2) [30]
σ = −1.08 cm−1/Gpa (ν1 − ν0)
where −1.08 cm−1/GPa is the Raman stress factor for diamond, ν1 denotes the measured Raman shift under stress, and ν0 = 1332 cm−1 represents the Raman shift of stress-free diamond.
Table 3 lists the obtained Raman peak positions and the corresponding residual stresses for the synthesized samples. For the characteristic diamond peak at 1332 cm−1, samples S0 to S3 all exhibit a positive shift from the standard position. This positive shift indicates the presence of compressive stress within the diamond lattice [31]. The residual stresses calculated accordingly are 0.69 GPa, 1.57 GPa, 1.27 GPa, and 1.35 GPa, respectively. It is evident that the residual stress in nitrogen-doped diamond crystals (S1–S3) is significantly higher than that in the undoped sample (S0). When a trace amount of Carbon dioxide was introduced during film growth, the corresponding doped diamond crystals (S4–S6) exhibited residual stresses of 1.35 GPa, 1.31 GPa, and 0.98 GPa, respectively, all of which demonstrate a certain degree of reduction compared to their Carbon dioxide-free counterparts with similar doping. The residual stress primarily originates from two sources: first, thermal stress resulting from the difference in the coefficient of thermal expansion (CTE) between the substrate and the diamond film; and second, intrinsic stress induced by the lattice mismatch between the dopant atoms and carbon atoms [32]. Nitrogen doping introduces substitutional nitrogen atoms that form C–N bonds during growth. Since the C–N bond differs in length from the original C–C bond, this substitution induces local lattice expansion, consequently leading to an increase in intrinsic stress and localized lattice distortion [33]. The observed stress reduction upon the introduction of trace Carbon dioxide is attributed to the etching effect of OH radicals on the crystal surface. This etching process helps to partially relieve lattice strain and suppress excessive lattice expansion, a mechanism consistent with the Raman FWHM results presented in Figure 2c. However, the residual stress in Carbon dioxide-co-doped samples remains higher than that in the undoped sample, indicating that the lattice expansion effect induced by nitrogen doping still dominates the overall stress state.
Figure 3a presents the photoluminescence (PL) spectra of diamond films prepared with varying nitrogen concentrations. A sharp characteristic diamond Raman peak is evident at 572 nm in the PL spectra of all films, regardless of the nitrogen concentration [15]. When the nitrogen concentration was 0 ppm, sample S0 exhibited the flattest PL baseline, with no detectable emission peaks corresponding to the NV0 center at 575 nm or the NV center at 637 nm. In contrast, as the nitrogen concentration increased to 12.48 ppm, a significant enhancement in the intensity of the characteristic diamond peak was observed, indicating improved film crystallinity. A concomitant rise in the spectral baseline was also noted, suggesting a slight increase in film impurities [33]. The luminescence peak for the NV0 center at 575 nm remains faint, whereas the NV peak at 637 nm has intensified markedly. When the nitrogen concentration reaches 24.93 ppm, the spectral baseline exhibits a slight increase. Concurrently, the relative intensity of the characteristic diamond Raman peak shows a minor enhancement, accompanied by a notable strengthening of the NV-related luminescence peaks. As the nitrogen concentration is further elevated to 39.84 ppm, the relative intensity of the diamond Raman peak experiences a slight decrease. Meanwhile, both the spectral baseline and the intensity of NV-related luminescence peaks surge significantly, demonstrating a substantial increase in nitrogen-related defects within the film. Furthermore, the integrated area of the respective luminescence peaks in the PL spectra is directly proportional to the concentration of NV0 and NV defects in the diamond film [33]. Figure 3b displays the photoluminescence (PL) spectra of diamond films grown with CO2 addition under varying nitrogen concentrations. The diamond Raman peak at 572 nm exhibits a similar trend of initial increase followed by a decrease, as observed in films without CO2. Similarly, the intensities of the NV0 peak at 575 nm and the NV peak at 637 nm increase with higher nitrogen concentrations. However, the rate of this increase is markedly lower compared to the samples without CO2, and this suppressive effect is more pronounced at higher nitrogen levels. These results demonstrate that the etching action by OH radicals, introduced via CO2, effectively reduces the incorporation of nitrogen impurities in the films, which is consistent with the Raman spectroscopy data in Figure 2. Furthermore, the integrated areas of the NV0 (575 nm) and NV (637 nm) luminescence peaks serve as indicators for the respective concentrations of these nitrogen-vacancy defects in the diamond film [34]. As demonstrated in Figure 3c,d, the concentration of nitrogen-vacancy (NV) defects exhibits a positive correlation with increasing nitrogen concentration. Furthermore, under conditions without CO2 introduction, the integrated areas of the luminescence peaks for both NV and NV0 centers in the prepared diamond films are consistently larger than the corresponding values from samples grown with CO2.
Figure 4a,b presents the X-ray diffraction (XRD) patterns of diamond growth surfaces under different processing conditions, along with the relative intensities and corresponding variations in diffraction peaks from various crystal planes. The results show that the primary diffraction peaks of the as-grown diamond films within the 2θ range are located at 43.9°, 75.3°, 91.5°, and 119.7°, corresponding to the four main crystal planes: <111>, <220>, <311>, and <400>, respectively. It can be clearly observed that no distinct <400> diffraction peak is present in any of the samples. For the undoped sample S0 prepared at approximately 850 °C, the crystal orientation consists mainly of <111> and <220>, accompanied by a minor <311> orientation, while the overall crystal morphology appears relatively disordered. This indicates that the process parameters (temperature and nitrogen concentration) may be unfavorable for the nucleation and growth of the <100> crystal plane in diamond. Furthermore, with increasing MN flow rate (i.e., nitrogen concentration), the intensities of the <220> and <311> diffraction peaks decrease significantly, whereas the <111> diffraction peak intensity shows a marked enhancement. This suggests that at a growth temperature around 850 °C, the introduction of trace nitrogen can suppress the growth of other crystal planes, thereby promoting the preferential growth of the <111> orientation. In addition, as shown in Figure 4b, after the introduction of carbon dioxide, the reduction in the intensities of the <220> and <311> diffraction peaks becomes more pronounced. Moreover, under the same nitrogen concentration, the peak intensities of the films with CO2 addition are consistently lower than those without.

3.2. Microstructural Analysis

The crystal structure of polycrystalline diamond films plays a decisive role in determining their thermal conductivity, mechanical strength, surface roughness, and optical properties. In such polycrystalline systems, the transport behavior of both photons and phonons is strongly influenced by microstructural features within the material, such as pores, secondary phase inclusions, impurity elements, and the presence of grain boundaries. To investigate the influence of different process parameters on the microstructure of free-standing diamond films, this section examines the surface morphology of nitrogen-doped diamond films before and after the addition of trace amounts of carbon dioxide. The lattice structure of the as-deposited diamond films was further studied using high-resolution electron microscopy. Figure 5 shows scanning electron microscopy (SEM) images of the growth surfaces of unpolished samples S0 to S6. As seen in Figure 5a, in the absence of nitrogen, the grains on the growth surface of S0 are composed of dense and uniform pyramid-like and prismatic particles with sizes ranging approximately from 23.5 μm to 25.3 μm. From Figure 5b–d, it can be observed that at a growth temperature of approximately 850 °C, with increasing nitrogen flow rate, the grains on the growth surfaces of S1–S3 become significantly larger compared to the undoped sample. Most of the exposed grain facets are pyramid-like particles with a predominant (111) orientation. This transition toward larger grain size can be attributed primarily to the step-bunching phenomenon induced by accelerated growth rates due to the introduction of nitrogen [35]. Meanwhile, as the nitrogen concentration increased progressively from 12.48 ppm to 39.84 ppm, the grain size of the diamond particles also exhibited a continuous enlargement, growing from approximately 23.5–45.6 μm for sample S1 to about 46.5–65.7 μm for S3. However, as shown in Figure 5d, excessively large grains can lead to an insufficiently dense microstructure, resulting in the formation of residual micropores or intergranular pores along the grain boundaries. According to reports [36], for polycrystalline optical ceramics, such microscopic or intergranular pores can lead to degradation of optical performance. Furthermore, as shown in Figure 5e,f, the grain sizes of diamond particles on the growth surfaces of S4–S6 are 20.6–42.7 μm, 39.7–43.0 μm, and 40.9–46.7 μm, respectively. Compared to samples S1–S3 at the same nitrogen concentrations, these values indicate a certain degree of reduction in both grain size and intergranular porosity. This suggests that the etching effect induced by the introduction of carbon dioxide effectively suppresses the degradation of film quality caused by excessive nitrogen doping, which is consistent with the decrease in Raman FWHM observed in Figure 2.

3.3. Thermodynamic Properties of Diamond Materials

Assuming transient heat transfer within the sample and neglecting heat loss during testing, a laser pulse was emitted to irradiate the bottom surface of the substrate. The subsequent absorption of heat led to a temperature increase in the sample, and the corresponding temperature rise on the top surface was measured. The thermal diffusivity of samples S0–S6, as listed in Table 4, was determined based on the time required for the temperature rise to reach half of its maximum value. The thermal conductivity of diamond was calculated using the formula [37,38]:
K = α × ρ × Cp
In this formula, α represents the thermal diffusivity under different growth conditions, while ρ and Cp denote the density of diamond (3.52 g·cm−3) and the specific heat capacity (0.519 J·g−1·K−1).
Figure 6 shows the thermal conductivity of samples prepared under different processes as a function of nitrogen concentration. It can be clearly observed that the thermal conductivity of the films decreases with increasing nitrogen concentration. This phenomenon can be attributed to the following mechanism: The high thermal conductivity of diamond primarily originates from the efficient propagation of lattice vibrations (i.e., phonons). Any defects that disrupt the perfection of the crystal lattice can scatter phonons, thereby reducing thermal conductivity. The introduction of nitrogen exacerbates phonon scattering mainly through the formation of complex defects and the induction of internal stress. Nitrogen atoms in diamond tend to combine with other defects, such as vacancies, forming more complex defect centers (e.g., nitrogen-vacancy color centers). These defects simultaneously introduce internal stress into the lattice, further disturbing its periodicity and intensifying phonon scattering [39]. We aim to elucidate the synergistic effect introduced by trace CO2 incorporation under fixed nitrogen doping levels. The specific data analysis is as follows: compared to the nitrogen-only doped sample S1, the thermal conductivity of the co-doped sample S4 increased by 125.47 W·m−1·K−1; compared to sample S2, sample S5 increased by 293.17 W·m−1·K−1; and compared to sample S3, sample S6 increased by 146.79 W·m−1·K−1. These results indicate that the introduction of trace CO2 co-doping on top of nitrogen doping can significantly enhance the thermal conductivity of the films, with an improvement ranging approximately between 125 and 300 W·m−1·K−1. Particularly under low nitrogen concentration (12.48 ppm), the degradation in thermal conductivity induced by nitrogen doping was almost completely suppressed. This phenomenon can be attributed to the fact that introduced CO2 effectively etched the film during growth, thereby suppressing the formation of complex defect centers such as NV centers caused by nitrogen doping. The elimination of such defects mitigates excessive lattice expansion and associated internal stress, which is consistent with the observations from Raman and photoluminescence spectroscopy discussed previously.

4. Conclusions

  • Advantage of introducing trace amounts of carbon dioxide: This study successfully demonstrates that the co-doping strategy utilizing N2 and CO2 during MPCVD growth is a highly effective method for fabricating high-quality, optical-grade polycrystalline diamond films. This approach ingeniously leverages the dual functions of N2 in promoting growth rate and CO2 in purifying the crystal lattice.
  • Performance Optimization: Employing this method enabled the attainment of diamond films featuring high thermal conductivity (1863.94 W·m−1·K−1) and a highly <111>-textured morphology, while maintaining a high growth rate (~3–4 µm/h) and effectively mitigating defects introduced by nitrogen impurities.
  • Defect Control Mechanism: The introduction of CO2, through its etching action, significantly reduces the concentration of complex defects such as NV colour centres and alleviates lattice strain. This mechanism is identified as the fundamental reason for the observed enhancements in both thermal conductivity and optical properties.
  • Application Prospect: This work provides an effective solution to the long-standing challenge of reconciling crystal quality with performance in nitrogen-doped diamond. The fabricated ultra-low nitrogen-doped diamond films exhibit broad application prospects in fields such as high-efficiency thermal management.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma19010183/s1.

Author Contributions

Research Concept and Design: Z.S., J.Y., N.J. and L.Z. (Lingxia Zheng); Diamond Growth Experiment: Z.S. and D.K.; Structural Characterization and Performance Testing: X.C., J.Y., L.Z. (Lei Zhao), Y.L. and R.Z.; Data Analysis and Charting: Z.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Key Research and Development Program of China (2024YFB3816600).

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Materials. Further inquiries can be directed to the corresponding authors.

Acknowledgments

The authors have reviewed and edited the output and take full responsibility for the content of this publication.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Wu, K.; Chang, G.; Ye, J.; Zhang, G. Significantly Enhanced Interfacial Thermal Conductance across GaN/Diamond Interfaces Utilizing AlxGa1xN as a Phonon Bridge. ACS Appl. Mater. Interfaces 2024, 16, 58880–58890. [Google Scholar] [CrossRef]
  2. Guo, S.; Yang, L.; Dai, B.; Geng, F.; Ralchenko, V.; Han, J.; Zhu, J. Past Achievements and Future Challenges in the Development of Infrared Antireflective and Protective Coatings. Phys. Status Solidi (a) 2020, 217, 2000149. [Google Scholar] [CrossRef]
  3. Kotomin, E.A.; Kuzovkov, V.N.; Lushchik, A.; Popov, A.I.; Shablonin, E.; Scherer, T.; Vasil’chenko, E. The Annealing Kinetics of Defects in CVD Diamond Irradiated by Xe Ions. Crystals 2024, 14, 546. [Google Scholar] [CrossRef]
  4. Kaboli, S.; Burnley, P.C. Direct observations of crystal defects in polycrystalline diamond. Mater. Charact. 2018, 142, 154–161. [Google Scholar] [CrossRef]
  5. Mussi, A.; Eyidi, D.; Shiryaev, A.; Rabier, J. TEM observations of dislocations in plastically deformed diamond. Phys. Status Solidi (a) 2012, 210, 191–194. [Google Scholar] [CrossRef]
  6. Mizuochi, N.; Ogura, M.; Watanabe, H.; Isoya, J.; Okushi, H.; Yamasaki, S. EPR study of hydrogen-related defects in boron-doped p-type CVD homoepitaxial diamond films. Diam. Relat. Mater. 2004, 13, 2096–2099. [Google Scholar] [CrossRef]
  7. Mizuochi, N.; Watanabe, H.; Isoya, J.; Okushi, H.; Yamasaki, S. Hydrogen-related defects in single crystalline CVD homoepitaxial diamond film studied by EPR. Diam. Relat. Mater. 2004, 13, 765–768. [Google Scholar] [CrossRef]
  8. Alvarez, J.; Kleider, J.P.; Bergonzo, P.; Tromson, D.; Snidero, E.; Mer, C. Study of deep defects in polycrystalline CVD diamond from thermally stimulated current and below-gap photocurrent experiments. Diam. Relat. Mater. 2003, 12, 546–549. [Google Scholar] [CrossRef]
  9. Silva, F.; Hassouni, K.; Bonnin, X.; Gicquel, A. Microwave engineering of plasma-assisted CVD reactors for diamond deposition. J. Phys. Condens. Matter 2009, 21, 364202. [Google Scholar] [CrossRef]
  10. Liang, Y.; Liu, K.; Liu, B.; Li, Y.; Fan, S.; Dai, B.; Zhang, Y.; Zhu, J. Vapor phase nucleation and sedimentation of dispersed nanodiamonds by MPCVD. Powder Technol. 2024, 436, 119507. [Google Scholar] [CrossRef]
  11. Yamada, H.; Chayahara, A.; Mokuno, Y.; Soda, Y.; Horino, Y.; Fujimori, N. Modeling and numerical analyses of microwave plasmas for optimizations of a reactor design and its operating conditions. Diam. Relat. Mater. 2005, 14, 1776–1779. [Google Scholar] [CrossRef]
  12. Teng, Y.; Yang, K.; Tang, K.; Zhao, W.; Gu, L.; Zhao, G.; Zhu, S.; Ye, J.; Gu, S. The regulation mechanism of oxygen additives on diamond growth and residual boron by microwave plasma chemical vapor deposition. Diam. Relat. Mater. 2025, 160, 113065. [Google Scholar] [CrossRef]
  13. Wang, Q.-L.; Lu, X.-Y.; Li, L.-A.; Cheng, S.-H.; Li, H.-D. Growth and Characteristics of Freestanding Hemispherical Diamond Films by Microwave Plasma Chemical Vapor Deposition. Chin. Phys. Lett. 2010, 27, 047802. [Google Scholar] [CrossRef]
  14. Yang, Z.; An, K.; Liu, Y.; Guo, Z.; Shao, S.; Liu, J.; Wei, J.; Chen, L.; Wu, L.; Li, C. Edge effect during microwave plasma chemical vapor deposition diamond-film: Multiphysics simulation and experimental verification. Int. J. Miner. Metall. Mater. 2024, 31, 2287–2299. [Google Scholar] [CrossRef]
  15. Yang, G.; Sun, P.; Zhu, T.; Wang, Y.; Li, S.; Liu, C.; Yang, G.; Yang, K.; Yang, X.; Lian, W.; et al. Fabrication, microstructure and optical properties of <110> textured CVD polycrystalline diamond infrared materials. Diam. Relat. Mater. 2024, 141, 110600. [Google Scholar] [CrossRef]
  16. Ashkinazi, E.; Khmelnitskii, R.; Sedov, V.; Khomich, A.; Khomich, A.; Ralchenko, V. Morphology of Diamond Layers Grown on Different Facets of Single Crystal Diamond Substrates by a Microwave Plasma CVD in CH4-H2-N2 Gas Mixtures. Crystals 2017, 7, 166. [Google Scholar] [CrossRef]
  17. Izak, T.; Babchenko, O.; Varga, M.; Potocky, S.; Kromka, A. Low temperature diamond growth by linear antenna plasma CVD over large area. Phys. Status Solidi (b) 2012, 249, 2600–2603. [Google Scholar] [CrossRef]
  18. Liu, D.-Y.; Tang, K.; Zhu, S.-M.; Rong, Z.; Zheng, Y.-D.; Shu-Lin, G. Effects of oxygen/nitrogen co-incorporation on regulation of growth and properties of boron-doped diamond films. Chin. Phys. B 2023, 32, 118102. [Google Scholar] [CrossRef]
  19. Giussani, A.; Janssens, S.D.; Vázquez-Cortés, D.; Fried, E. Evolution of nanodiamond seeds during the chemical vapor deposition of diamond on silicon substrates in oxygen-rich plasmas. Appl. Surf. Sci. 2022, 581, 152103. [Google Scholar] [CrossRef]
  20. Gu, Y.; Zhang, E.; Wang, J.; Shen, H.; Lei, X. Effect of oxygen doping on cutting performance of CVD diamond coated milling cutters in zirconia ceramics machining. Int. J. Refract. Met. Hard Mater. 2024, 124, 106853. [Google Scholar] [CrossRef]
  21. Zhao, G.; Tang, K.; Teng, Y.; Zhao, W.; Yang, K.; Zhu, S.; Gu, S. The regulation effect of trace amount of oxygen on the properties of p-type boron-doped diamond. J. Mater. Res. 2024, 39, 1313–1323. [Google Scholar] [CrossRef]
  22. Podgursky, V.; Bogatov, A.; Sedov, V.; Sildos, I.; Mere, A.; Viljus, M.; Buijnsters, J.G.; Ralchenko, V. Growth dynamics of nanocrystalline diamond films produced by microwave plasma enhanced chemical vapor deposition in methane/hydrogen/air mixture: Scaling analysis of surface morphology. Diam. Relat. Mater. 2015, 58, 172–179. [Google Scholar] [CrossRef]
  23. Li, L.; An, K.; Xu, G.; Wu, H.; Yang, Z.; Xi, Z.; Liu, P.; Zhang, Y.; Zhang, Y.; Zhang, X.; et al. Twins and dark features in MPCVD diamond films. Appl. Surf. Sci. 2025, 701, 163222. [Google Scholar] [CrossRef]
  24. Yiming, Z.; Larsson, F.; Larsson, K. Effect of CVD diamond growth by doping with nitrogen. Theor. Chem. Acc. 2013, 133, 1432. [Google Scholar] [CrossRef]
  25. dos Santos, R.B.; Rivelino, R.; Mota, F.d.B.; Gueorguiev, G.K. Effects of N doping on the electronic properties of a small carbon atomic chain with distinctsp2terminations: A first-principles study. Phys. Rev. B 2011, 84, 075417. [Google Scholar] [CrossRef]
  26. Ashfold, M.N.R.; Mankelevich, Y.A. Two-dimensional modeling of diamond growth by microwave plasma activated chemical vapor deposition: Effects of pressure, absorbed power and the beneficial role of nitrogen on diamond growth. Diam. Relat. Mater. 2023, 137, 110097. [Google Scholar] [CrossRef]
  27. Yang, J.X.; Zhang, H.D.; Li, C.M.; Chen, G.C.; Lu, F.X.; Tang, W.Z.; Tong, Y.M. Effects of nitrogen addition on morphology and mechanical property of DC arc plasma jet CVD diamond films. Diam. Relat. Mater. 2004, 13, 139–144. [Google Scholar] [CrossRef]
  28. Yan, X.; Wei, J.; An, K.; Liu, J.; Chen, L.; Zheng, Y.; Zhang, X.; Li, C. High temperature surface graphitization of CVD diamond films and analysis of the kinetics mechanism. Diam. Relat. Mater. 2021, 120, 108647. [Google Scholar] [CrossRef]
  29. Ager, J.W.; Drory, M.D. Quantitative measurement of residual biaxial stress by Raman spectroscopy in diamond grown on a Ti alloy by chemical vapor deposition. Phys. Rev. B 1993, 48, 2601–2607. [Google Scholar] [CrossRef]
  30. Gries, T.; Vandenbulcke, L.; Simon, P.; Canizares, A. Anisotropic biaxial stresses in diamond films by polarized Raman spectroscopy of cubic polycrystals. J. Appl. Phys. 2008, 104, 023524. [Google Scholar] [CrossRef]
  31. Haddad, M.; Kurtulus, O.; Mertens, M.; Brühne, K.; Glüche, P.; Fecht, H. Optimization of residual stresses inside diamond thin films grown by hot filament chemical vapor deposition (HFCVD). Diam. Relat. Mater. 2023, 131, 109564. [Google Scholar] [CrossRef]
  32. Smith, W.V.; Sorokin, P.P.; Gelles, I.L.; Lasher, G.J. Electron-Spin Resonance of Nitrogen Donors in Diamond. Phys. Rev. 1959, 115, 1546–1552. [Google Scholar] [CrossRef]
  33. Tarun, A.; Lee, S.J.; Yap, C.M.; Finkelstein, K.D.; Misra, D.S. Impact of impurities and crystal defects on the performance of CVD diamond detectors. Diam. Relat. Mater. 2016, 63, 169–174. [Google Scholar] [CrossRef]
  34. Tan, X.; Wang, J.; Li, X.; Wei, X.; Yang, Q.; He, Z.; Qi, H.; Zhang, B.; Meng, D.; Liang, Z.; et al. Preparation of step-flow grown diamond films and directionally oriented NV center ensembles. Diam. Relat. Mater. 2025, 156, 112387. [Google Scholar] [CrossRef]
  35. Tallaire, A.; Collins, A.T.; Charles, D.; Achard, J.; Sussmann, R.; Gicquel, A.; Newton, M.E.; Edmonds, A.M.; Cruddace, R.J. Characterisation of high-quality thick single-crystal diamond grown by CVD with a low nitrogen addition. Diam. Relat. Mater. 2006, 15, 1700–1707. [Google Scholar] [CrossRef]
  36. Krell, A.; Klimke, J.; Hutzler, T. Transparent compact ceramics: Inherent physical issues. Opt. Mater. 2009, 31, 1144–1150. [Google Scholar] [CrossRef]
  37. Chen, Y.; Zhang, Q.; Li, B.; Chen, Z.; Liu, S.; Ma, X.; Tofil, S.; Yao, J. Effect of Laser on the Interface and Thermal Conductivity of Metallized Diamond/Cu Composite Coatings Deposited by Supersonic Laser Deposition. Materials 2024, 17, 5174. [Google Scholar] [CrossRef]
  38. Zhang, M.; Tan, Z.; Huang, Y.; Huang, Z.; Liu, X.; Zhang, K. Preparation and interfacial microstructure of high thermal conductivity diamond/SiC composites. Ceram. Int. 2024, 50, 23754–23762. [Google Scholar] [CrossRef]
  39. Ali, N.K.; Omar, M.S.; Yousf, S.O. On the nitrogen concentration and crystal size dependence of lattice thermal conductivity in diamond thin and nanofilms. Diam. Relat. Mater. 2025, 155, 112340. [Google Scholar] [CrossRef]
Figure 1. Growth rates of diamond films under different nitrogen gas flow rates before and after adding Carbon dioxide.
Figure 1. Growth rates of diamond films under different nitrogen gas flow rates before and after adding Carbon dioxide.
Materials 19 00183 g001
Figure 2. Raman spectra of diamond films with different nitrogen doping concentrations, (a) no carbon dioxide introduced and (b) adding carbon dioxide and their (c) their Raman FWHM. (d) Comparison of Raman spectra with high-concentration nitrogen with or without added Carbon dioxide.
Figure 2. Raman spectra of diamond films with different nitrogen doping concentrations, (a) no carbon dioxide introduced and (b) adding carbon dioxide and their (c) their Raman FWHM. (d) Comparison of Raman spectra with high-concentration nitrogen with or without added Carbon dioxide.
Materials 19 00183 g002
Figure 3. Photoluminescence spectra of diamond films with different nitrogen doping concentrations, (a) no carbon dioxide and (b) adding carbon dioxide, and the peak area of (c) NV0 and (d) NV−.
Figure 3. Photoluminescence spectra of diamond films with different nitrogen doping concentrations, (a) no carbon dioxide and (b) adding carbon dioxide, and the peak area of (c) NV0 and (d) NV−.
Materials 19 00183 g003
Figure 4. X-ray diffraction patterns (XRD) of diamond films with different nitrogen doping concentrations (a) with trace carbon dioxide introduced and (b) without carbon dioxide introduced, (a) with trace carbon dioxide introduced and (b) without carbon dioxide introduced.
Figure 4. X-ray diffraction patterns (XRD) of diamond films with different nitrogen doping concentrations (a) with trace carbon dioxide introduced and (b) without carbon dioxide introduced, (a) with trace carbon dioxide introduced and (b) without carbon dioxide introduced.
Materials 19 00183 g004
Figure 5. SEM images of the growth surfaces of samples (a) S0, (b) S1, (c) S2, (d) S3, (e) S4, (f) S5 and (g) S6 under different nitrogen concentrations.
Figure 5. SEM images of the growth surfaces of samples (a) S0, (b) S1, (c) S2, (d) S3, (e) S4, (f) S5 and (g) S6 under different nitrogen concentrations.
Materials 19 00183 g005
Figure 6. Thermal conductivity of diamond films under different nitrogen gas flow rates before and after adding oxygen.
Figure 6. Thermal conductivity of diamond films under different nitrogen gas flow rates before and after adding oxygen.
Materials 19 00183 g006
Table 1. Experiment parameters of deposition processes.
Table 1. Experiment parameters of deposition processes.
SamplesH2/sccmCH4/sccmMN/sccmCO2/sccmPower/kwPressure/TorrTemperature/°C
S040012003.8100850 ± 10
S14000.50
S240010
S34001.60
S44000.51
S540011
S64001.61
MN:H2-N2 mixture gas with a nitrogen content of 1%.
Table 2. Characteristics parameters of thickness and growth rates of the diamond samples.
Table 2. Characteristics parameters of thickness and growth rates of the diamond samples.
SchemesThickness/μmGrowth Rate μm/h
S0451.125
S1842.100
S21112.775
S31393.475
S4952.375
S51253.125
S61503.750
Table 3. Raman shifts and residual stress in the diamond samples.
Table 3. Raman shifts and residual stress in the diamond samples.
SamplesRaman Shift/cm−1Residual Stress/GPa
S01332.64−0.69
S11333.45−1.57
S21333.18−1.27
S31333.25−1.35
S41333.21−1.31
S51333.08−1.17
S61332.91−0.98
Table 4. Thermal conductivity of diamond samples.
Table 4. Thermal conductivity of diamond samples.
SamplesThermal Diffusivity/mm2·SThermal Conductivity/W·m−1·K−1
S01038.341896.93
S1951.611738.47
S2752.391374.52
S3714.791305.84
S41020.291863.94
S5912.861667.69
S6795.141452.63
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sheng, Z.; Cui, X.; Zhao, L.; Lv, Y.; Zhang, R.; Kon, D.; Jiang, N.; Yi, J.; Zheng, L. Enhancing the Quality of Diamond Film Growth Through the Synergistic Addition of Nitrogen and Carbon Dioxide. Materials 2026, 19, 183. https://doi.org/10.3390/ma19010183

AMA Style

Sheng Z, Cui X, Zhao L, Lv Y, Zhang R, Kon D, Jiang N, Yi J, Zheng L. Enhancing the Quality of Diamond Film Growth Through the Synergistic Addition of Nitrogen and Carbon Dioxide. Materials. 2026; 19(1):183. https://doi.org/10.3390/ma19010183

Chicago/Turabian Style

Sheng, Zhanpeng, Xuejian Cui, Lei Zhao, Yihan Lv, Rongchen Zhang, Defang Kon, Nan Jiang, Jian Yi, and Lingxia Zheng. 2026. "Enhancing the Quality of Diamond Film Growth Through the Synergistic Addition of Nitrogen and Carbon Dioxide" Materials 19, no. 1: 183. https://doi.org/10.3390/ma19010183

APA Style

Sheng, Z., Cui, X., Zhao, L., Lv, Y., Zhang, R., Kon, D., Jiang, N., Yi, J., & Zheng, L. (2026). Enhancing the Quality of Diamond Film Growth Through the Synergistic Addition of Nitrogen and Carbon Dioxide. Materials, 19(1), 183. https://doi.org/10.3390/ma19010183

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop