DOI:
10.1039/C6RA02141G
(Paper)
RSC Adv., 2016,
6, 39444-39451
Influence of CH4–Ar ratios on the composition, microstructure and optical properties of Be2C films synthesized by DC reactive magnetron sputtering†
Received
24th January 2016
, Accepted 31st March 2016
First published on 4th April 2016
Abstract
Beryllium carbide (Be2C) films were first deposited on optical quartz substrates by DC reactive magnetron sputtering on a beryllium target with variable CH4–Ar ratios. The influence of CH4–Ar ratios on the composition, microstructure and optical properties were investigated by X-ray photoelectron spectroscopy, X-ray diffraction, high-resolution transmission electron microscope, atomic force microscopy, scanning electron microscope and UV-vis spectrum. The main component in the films prepared at lower CH4–Ar ratios (<5%) was Be2C, while hydrocarbon (CH) films were formed at higher CH4–Ar ratios (>15%). The films exhibited a nanocomposite structure consisting of Be2C nanocrystals (3 to 5 nm in size) embedded in amorphous hydrocarbon matrices. A smooth surface and columnar structure on the cross-sectional view were revealed. Besides, the depositing rates reached ∼125 nm h−1, which were significantly higher than that of RF reactive magnetron sputtering. High transparency (>50%) of the Be2C films in the visible region as well as an even higher transparency (>80%) in the near-infrared region were demonstrated. Finally, the dispersion of the optical constants of Be2C films is presented, and the optical bandgaps were evaluated to be ∼2 eV. The good properties of Be2C films prepared by DC reactive magnetron sputtering showed that this material could be a potential candidate for application to inertial confinement fusion targets.
1. Introduction
Recently, many concerns have been paid to the low Z (atomic number) materials due to their important usages in nuclear fusion research.1–3 Beryllium carbide (Be2C), as a typical low Z material, has been reported to be a promising wall material choice for the next generation experimental fusion reactor ITER because it can mitigate not only the pure beryllium but also the pure carbon sputtering yields.4,5 Another important application is for inertial confinement fusion (ICF) capsule materials. The most principle requirements for capsule materials are a low atomic number, smooth surface, uniform composition and structures and high density.6–8 Doped beryllium and hydrocarbon (CH) coatings are the most commonly selected target materials at the National Ignition Facility (NIF).6–11 Currently, the compound Be2C has caught researchers' interest because it combines the advantages of beryllium and hydrocarbons.12,13 Compared with many hydrocarbons, Be2C has a lower atomic number (Z = 4.7), and a higher density (ρ = 2.4 g cm−3) and mechanical strength (bulk modulus B0 = 217.05 GPa and shear modulus Cs = 165.95 GPa).2,3,14 Compared with beryllium, its antifluorite structure results in a smoother surface and grain refinement. It has yellowish-brown transparent crystals and a bandgap over 1.2–2 eV,3,14 which might indicate that the deuterium–tritium (D–T) ice layers in Be2C capsules could be characterized by optical methods and homogenized by infrared heating.15 However, comprehensive understanding of its optical properties are needed to evaluate this advantage.
Due to its important usages, the structural, electronic and optical properties of beryllium carbide have been adequately investigated by theoretical calculations;1–3,16 however, there are few literature reports on the experimental side. Beryllium carbide was synthesized and its structural properties were studied in ref. 14 and 17. Thereafter, Xie and Shih et al. deposited beryllium carbide coatings by plasma polymerization of diethyl-beryllium and by RF reactive magnetron sputtering for application to ICF.13,18,19 The H2 permeability, mechanical properties, chemical composition and electrical and thermal conductivities of beryllium carbide films were analysed and the results indicated that most of the requirements of the ignition target were satisfied. However, some shortcomings in these two methods have so far limited their application to ICF targets. First, the instability and long recovery time of the saturation vapour pressure of diethyl-beryllium20 makes it hard to control the composition and incurs a high cost. Moreover, doping is difficult to realize in deposition by plasma polymerization. Second, the maximum deposition rate is ∼56 nm h−1 by RF reactive magnetron sputtering,21 which is too low for depositing targets. To the best of our knowledge, the optical properties of beryllium carbide have not been researched experimentally to date, even though it is important for application to ICF targets. The scarcity of experimental optical studies is due to the nature of the Be2C samples: Be2C is a highly toxic component, which is also difficult to prepare and has a tendency to hydrolyse and oxidize.3
In the present work, Be2C films were first prepared by DC magnetron sputtering of beryllium into methane plasma. The compositions were controlled by the CH4–Ar ratio during the sputtering process. The carbon content in the films increased with the increase in the CH4–Ar ratio, with hydrocarbon films formed at a CH4–Ar ratio >15%. Based on this finding, the Be2C films were in situ covered by a layer of hydrocarbons (∼26 nm), which was demonstrated to be effective in preventing oxidation. In addition, the microstructures and morphologies were determined. Finally, the optical properties of Be2C films were measured by a double-beam spectrophotometer for the first time, indicating an optical bandgap of about 2 eV.
2. Experimental
2.1. Preparation of Be2C films
Beryllium carbide films were deposited on optical quartz substrates by DC reactive magnetron sputtering of a 76 mm-diameter beryllium target under various CH4–Ar ratios at room temperature. The flow rates of argon and methane were accurately controlled by mass flow controllers, with the flow rate of Ar fixed at 50 sccm while the flow rate of CH4 was varied (0.8, 1.0, 1.2, 1.5, 2.0, 2.5, 3.0, 5.0, 7.5 and 10 sccm, corresponding to CH4–Ar ratios of 1.6%, 2.0%, 2.4%, 3.0%, 4.0%, 5.0%, 6.0%, 10%, 15% and 20%, respectively). An angle of ∼45° existed between the plane of the target and substrate. The distance between the Be target and substrates was set at ∼90 mm. In order to reduce the possibility of target poisoning, argon was introduced near the target and methane was introduced near the substrates. The system was evacuated to a base pressure of 1.0 × 10−5 Pa. Before the deposition, the substrate shutter was closed and pre-sputtering was conducted on the Be target to remove any contaminants and to guarantee a fresh beryllium surface. During the growth, the working pressure was maintained at 0.5 Pa and the sputtering power was constant at 50 W. The substrate was rotated at a constant rate of 6 rpm to improve the uniformity. The depositing time was controlled at 4 hours. After the deposition, protective layers (∼26 nm hydrocarbon) were in situ covered on the surface of the Be2C films, which were prepared at a 1.6–5.0% CH4–Ar ratio, by simply increasing the CH4 flow rate up to 10 sccm and sustaining this for 8 minutes. A 26 nm hydrocarbon film was deposited on quartz substrate as a reference sample for the morphology and optical property measurements.
2.2. Characterization of the Be2C films
The elemental composition and the corresponding chemical states of different Be2C films were studied by X-ray (Mg-Kα 1253.6 eV) photoelectron spectroscopy (XPS). The microstructure was investigated by X-ray diffraction (XRD) with the grazing angle mode using Cu Kα (λ = 0.154056 nm) radiation, as well as high-resolution transmission electron microscope (HRTEM). For the HRTEM measurements, the film (deposited at a 2.4% CH4–Ar ratio) was scratched with a sharp razor blade in an ethanol environment. Then, it was dispersed in ethanol and placed on the TEM grid. The surface and cross-sectional morphologies were observed by atomic force microscopy (AFM) and scanning electron microscope (SEM), respectively. The root mean square (RMS) surface roughness of the films was evaluated over a 1 × 1 μm2 area. The deposition rates were calculated from the thickness (measured by SEM) and deposition time. The average densities were evaluated from the weight gain and thickness of the films with a definite area. The optical properties were analysed from the transmittance and reflectance spectrum measured by a double-beam spectrophotometer, which had two beams of light, one for measurement and the other for reference. When testing, the Be2C films were placed on the testing sample holder, and the fleshly deposited ∼26 nm hydrocarbon on quartz substrate (reference sample) was placed on the reference sample holder. Thus, the obtained spectrum only showed the response of the Be2C films.
3. Results and discussion
3.1. Composition analysis
Fig. 1(a) shows the partial XPS survey spectra of the films deposited under different CH4–Ar ratios from 1.6% to 20% after 30 min Ar+ etching to remove any surface contaminants and protective layers. This confirmed the chemical purity of the films, as it only consisted of Be, C and O, and no peaks corresponding to other elements were detected. The binding energy of O1s peak was about 533.1 eV in all samples, which suggested that its chemical state remained unchanged. Nevertheless, appreciable shifts were observed in the C1s and Be1s peaks, indicating changes in the chemical states of carbon and beryllium. In order to identify the chemical states and relative amounts of each element in that state, high-resolution XPS spectra of Be, C and O were conducted, in which the overlapping bands were deconvoluted into separate peaks by Gaussian fitting using XPS peakfit software. The partial typical high-resolution XPS spectra of beryllium and carbon are illustrated in Fig. 1(b) and (c), respectively. The peaks centred at 111.23, 112.42 and 114.55 eV in the Be1s high-resolution spectra were assigned to metallic Be, Be2C and BeO, respectively.22 Also, the peaks at 282.46, 284.54, 286.23 and 288.17 eV in the C1s high-resolution spectra were attributed to Be2C, CH, C–O and C
O species, respectively.21,22 It was found that the presence of hydrocarbons increased with increasing CH4–Ar ratios. When the CH4–Ar ratios were below 5%, the major chemical state (>50 mol%) of the beryllium and carbon peaks was beryllium carbide. Beryllium was not incorporated in the depositions, and hydrocarbon films were formed when the CH4–Ar ratios were higher than 15% (shown in Fig. 1(c)). Besides, a small amount of metallic beryllium existed in films prepared at lower CH4–Ar ratios (2.0% and 1.6%).
 |
| Fig. 1 XPS spectra for films prepared at different CH4–Ar ratios after 30 min Ar+ etching: survey spectrum (a), deconvolution of Be1s peaks (b) and deconvolution of C1s peaks (c). | |
The atomic concentrations of elements and their relative amounts in each chemical state were calculated from the peak areas, with corresponding sensitivity factors, as plotted in Fig. 2. It was found that the beryllium content decreased while the carbon content increased gradually with the increasing CH4–Ar ratios in the range of 1.6–5% (Fig. 2(a)), leading to the reduction of Be–C ratios from 2.11 to 1.26 (Fig. 2(b)). Further increasing the CH4–Ar ratio resulted in a dramatic decrease of the beryllium content and increase of the carbon content, until pure hydrocarbon films were formed. The results, however, were not consistent with previous research,20 where the Be–C ratios reached 3.5 when the CH4–Ar ratio was 25% and was maintained at ∼1 even when the CH4–Ar ratio exceeded 50%. However, the variation trend of the beryllium and carbon content with the CH4–Ar ratios in our work was the same as in the previous research. However, taking into account the difference in the experiment configuration and parameters, the distinction is understandable. The higher sputtering power (200 W) used in the previous researchers experiments was more effective for beryllium sputtering, resulting in the higher Be–C ratio. The percentage of beryllium (Be*) and carbon (C*) bonded as Be2C as a function of the CH4–Ar ratios in the range of 1.6% to 5% are illustrated in Fig. 2(b). The major (>85%) chemical state of beryllium was in Be2C, meanwhile, the percentage of carbon bonded as Be2C decreased with the increase in CH4–Ar ratios. However, the atomic ratios of Be* and C* always remained close to 2, which fitted the stoichiometry of Be2C, suggesting the accuracy of the XPS analysis.
 |
| Fig. 2 (a) The atomic content of Be, C and O in the films with various CH4–Ar ratios. (b) The atomic ratios of Be–C, as well as the percentage of Be and C bonded as Be2C for the films deposited at different CH4–Ar ratios. | |
The XPS analysis also suggested that oxidation had been effectively controlled by the protective coverages since only a little oxygen existed in the films with CH4–Ar ratios < 5%, compared with the fact that the uncovered Be2C films were quickly oxidized and delaminated after exposure to ambient air (not shown). This was also confirmed by the films deposited at 6% and 10% (uncovered), where all beryllium was in its oxidation state and the oxygen content reached 34.3 at% and 32.2 at%, respectively.
XRD measurements were also performed on the films as part of our structural investigations. Fig. S1 (ESI†) shows the XRD patterns of the as-deposited Be2C films with various CH4–Ar ratios. There were no obvious diffraction peaks in the XRD patterns. Further measurements should thus be introduced to reveal the microstructure. In ref. 23 and 24, the microstructure of Be-related materials (Be1−xBx and BeN) were measured by TEM.
HRTEM observations were performed on films deposited at 2.4%. Fig. 3 represents the plane view HRTEM image and its corresponding fast Fourier transform (FFT) pattern (upper-right insert). The HRTEM image demonstrated the nanocomposite structure of the films, with Be2C nanocrystals (3–5 nm) distributed in amorphous matrices. The interplanar spacing was found to be 0.251 ± 0.005 nm, referring to the (111) plane of the Be2C phase (PDF-33-0191).25 The FFT pattern exhibited a clear diffraction ring, attributed to the (111) plane of Be2C. Although there was ∼19% hydrocarbon in the film (as determined by XPS analysis), no diffraction rings corresponding to carbon were detected, which might indicate that the amorphous matrices consisted of hydrocarbons. This nanocomposite structure was widely found in the films prepared by reactive sputtering at room temperature with a low sputtering power.26 Under these conditions, atoms ejected from the target surface with low kinetic energy could not gain sufficient extra energy from the substrates to achieve a competent diffusion rate when they were adsorbed on the surface of the substrates, resulting in the formation of a nanocomposite structure. The influence of the sputtering power and substrate temperature on the deposition was also investigated and will be discussed in other works.
 |
| Fig. 3 Plane view high-resolution TEM image and corresponding FFT pattern (inset upper-right) for the film obtained at a 2.4% CH4–Ar ratio. | |
Generally, the nanocrystalline phase is detectable in the XRD spectrum.27 However, there were no obvious peaks in the XRD patterns of these nanocomposite Be2C films. This might be attributed to two possible reasons: (a) the low atomic number of Be2C resulted in difficulty for XRD detection and28 (b) the films in this work were too thin and the grain size was too small for Be2C to produce obvious X-ray diffraction peaks.
3.2. Film morphologies and growth characteristics
The surface and cross-sectional morphologies of the films under various CH4–Ar ratios were studied using AFM and SEM, respectively. The partial typical surface and cross-sectional morphologies are displayed in Fig. 4. The AFM images of Be2C films at a 2.4% CH4–Ar ratio clearly show that many small spherical-shaped granules (20–70 nm in diameter and 0–20 nm in height) are homogenously distributed over the surfaces. With the increasing CH4–Ar ratio, small particles tended to merge into larger ones, resulting in the greater RMS roughness at higher CH4–Ar ratios. The surface morphologies and RMS roughness of the ∼26 nm hydrocarbon films (i.e. the reference sample) were also examined, and presented a non-characteristic surface with a very low surface roughness of ∼0.22 nm (shown in Fig. 4(a)). Thus, it was shown that the granular morphologies and roughness can be mainly attributed to the Be2C films instead of a hydrocarbon protective layer. The measured RMS roughness of the films deposited at 1.6%, 2.0%, 2.4%, 3.0%, 4.0% and 5.0% were 2.51, 2.19, 2.63, 2.82, 3.31 and 5.80 nm, respectively. This suggests that the surface roughness of the Be2C films increased with the increase of CH contents. Since the smooth surface is significant for its application in ICF targets, purer Be2C films are desirable. The typical surface morphologies (see the 3D views) for the ∼26 nm thick hydrocarbon film (i.e. the reference sample) and films prepared at 2.4% and 5.0% are illustrated in Fig. 4(a–c), respectively. Surface morphologies for the Be2C films at 1.6%, 2.0%, 3.0% and 4.0% CH4–Ar ratios are shown in the ESI, Fig. S2.† A typical 2D view of the morphology (deposition at the 2.4% ratio) is also given in Fig. 4(d). Columnar structures always existed when the CH4–Ar ratio was below 5%, as revealed by the cross-sectional views. As a contrast, a homogeneous morphology, without columns or other characteristic structures, was revealed in hydrocarbon films. Herein, we have only presented the typical cross-sectional morphologies for the samples prepared under 2.4% and 20% ratios, as illustrated in Fig. 4(e) and (f), respectively. The lamination in Fig. 4(f) resulted from a fractionated deposition due target poisoning under high CH4–Ar ratios.
 |
| Fig. 4 Typical surface morphologies (AFM 3D images) of: (a) ∼26 nm hydrocarbon film on the quartz substrate as a reference sample, Be2C films deposited at a (b) 2.4%, (c) 5.0% CH4–Ar ratio and (d) AFM 2D image for Be2C films deposited at a 2.4% CH4–Ar ratio. Cross-sectional morphologies (SEM images) for films deposited at a (e) 2.4% and (f) 20% CH4–Ar ratio. | |
The coating thickness was evaluated from the cross-sectional views and the deposition rates were calculated from the thickness in a defined deposition time. The variation of the deposition rates with CH4–Ar ratios could imply changes in the sputtering modes, which thus should also be monitored by the target voltages. Fig. 5 shows the deposition rates and target voltages as a function of CH4–Ar ratios. The introduction of a small volume of methane (1.6%) resulted in a remarkable deterioration of the deposition rates. The deposition rates were relatively stable (∼125 nm h−1) over the ratio range of 1.6% to 5.0% and then increased to ∼182 nm h−1 in the ratio range of 15–20%. The deposition rates in our work (DC reactive magnetron sputtering) were much higher that reported in ref. 21 (RF reactive magnetron sputtering), and could be even further improved by increasing the sputtering power (not shown in this work). The higher deposition rate was beneficial to depositing ICF capsules. The oxidation in films deposited at 6% and 10% ratios led to an increasing thickness, resulting in the abnormally high deposition rates. Combined with XPS analysis, we suspected that there were three sputtering modes during the depositions as the CH4–Ar ratio varied from 0–20%: (a) a metallic sputtering mode, in which the depositions mainly consisted of metallic beryllium, (b) a compound sputtering mode, where a thin layer of carbide was formed on the target surface and the depositions mainly consisted of beryllium carbide, (c) a non-metallic sputtering mode, in which hydrocarbon was coated on the target surface and the depositions comprised hydrocarbons. The formation of hydrocarbons made it more complicated than other reactive sputtering techniques (involving oxides or nitrides). The variations of target voltages followed a similar trend to that of the deposition rates. Target voltages in the metallic mode were higher than that in the compound mode but lower than that in the non-metallic mode, as shown in Fig. 5. The results had good repeatability. Therefore, we suspected that the target voltage can be used as a proxy for judging the sputtering status.
 |
| Fig. 5 Deposition rates and target voltages versus CH4–Ar ratios. | |
Based on the thickness and weight gain of the films in a defined area (15 × 15 mm2), the average density was calculated. The average densities were found to decrease from 1.94 to 1.68 g cm−3 with the increasing CH4–Ar ratios from 1.6% to 5.0%, which were smaller than for Be2C films prepared by PECVD13 but higher than that of hydrocarbon films.9 The comparative low density might be caused by the incorporation of hydrocarbons, but could be improved by optimizing the experiment parameters, such as the sputtering power and annealing temperature (not studied in this work).
3.3. Optical properties
The deposited Be2C films varied from deep reddish-brown (1.6%) to light orange (5.0%) in colour. In order to evaluate the optical properties of the Be2C films, the optical transmittance and reflectance spectra were measured over the wavelength range of 200–2000 nm. Fig. 6 shows the optical transmittance spectrum of the Be2C films deposited with various CH4–Ar ratios (1.6–5.0%). An apparent interference feature could be observed in every spectrum, which was universal in transparent films and implied a smooth and homogeneous surface. The optical transmittance increased with the increasing CH4–Ar ratios. For Be2C films without metallic beryllium, the average transmittance was greater than 55% in the visible region (500 to 760 nm) and over 80% in the near-infrared region. This suggested that it was feasible to characterize the deuterium–tritium (D–T) ice layers by optical methods and to homogenize them by infrared heating the Be2C capsules.
 |
| Fig. 6 Optical transmittance spectra of the Be2C films prepared with different CH4–Ar ratios. | |
The dispersion of the absorption coefficient was calculated using transmittance and reflectance data combined with the thickness data. The transmittance and reflectance spectra were fitted by envelope methods before being adopted to eliminate interferences; one typical fitting curve is shown in the inset of Fig. 6. The absorption coefficient was computed by:29
|
 | (1) |
where
d is the thickness measured by SEM,
R(
λ) is the reflectance and
T(
λ) is the transmittance.
The dispersion behaviour of the optical constants, as is well known, plays an important role in the research of optical materials and can be readily calculated from the following equations:29,30
|
 | (2) |
|
 | (3) |
|
εr(λ) = n2(λ) − k2(λ)
| (4) |
where
α(
λ) is the absorption coefficient calculated from
eqn (1),
k(
λ) is the extinction coefficient,
n(
λ) is the refractive index and
εr(
λ) is the real parts and
εi(
λ) is the imaginary parts of the complex dielectric constant. The dispersion curves of the extinction coefficient (
k), the refractive index (
n) and the variation of the real (
εr) and imaginary (
εi) parts of the dielectric constant are illustrated in
Fig. 7. As can be seen, the extinction coefficient (
k) and refractive index (
n) tend to increase when the CH
4–Ar ratios decrease. The variations in the refractive index could be attributed to the difference in densities. According to the Lorentz–Lorenz formula,
31 
(where
α(
λ) is the mean polarizability,
N is the number density of molecules), the refractive index (
n) increases with the increase in densities. In addition, for the purer Be
2C films (deposited at 2.4%) the refractive index (
n) and the real parts of the dielectric constant (
εr) in the infrared region are 2.52 and 6.34, respectively, which are extraordinarily close to the theoretical calculation values of 2.50 for
n, and 6.25 for
εr.
2
 |
| Fig. 7 (a) Dispersion curves of the extinction coefficient (k), refractive index (n) and (b) the variation of the real (εr), imaginary (εi) parts of the dielectric constant with wavelength for Be2C films deposited at 2.0–5.0% CH4–Ar ratios. | |
The optical bandgaps were determined from the absorption coefficient via the Tauc relation26 by extrapolating the linear region of the (αhν)1/2–hν plot to the energy axis, where the exponent 1/2 is utilized for the indirect bandgap. Fig. 8 shows the Tauc analysis, which gives optical bandgaps of 2.43, 2.21, 2.10, 2.03, 1.7 and 0.37 eV for films deposited with CH4–Ar ratios of 5.0%, 4.0%, 3.0%, 2.4%, 2.0% and 1.6%, respectively. Theoretical calculations on Be2C using the GGA and LDA functionals revealed an indirect bandgap of about 1.2 eV, located between Γ and Χ.1,2,16 However, the electron energy loss spectra (EELS) data determined the need for an experimental correction to the gap of ∼2 eV.3 This value was close to the obtained optical bandgaps. A decrease in the optical bandgaps with a decrease in the CH4–Ar ratio was found, which might be related to the compositions. The larger optical bandgaps for films with CH4–Ar ratios of 5.0% and 4.0% might result from the presence of beryllia (bandgap ∼ 10.6 eV (ref. 32)). The Be2C film deposited at 1.6% has the smallest bandgap (0.37 eV), which is probably due to the existence of metallic beryllium in the films, as revealed by XPS analysis. The comparison with the other experimental literature has proven to be difficult as the optical properties of beryllium carbide films have been barely investigated.
 |
| Fig. 8 Plots of (αhν)1/2 versus photon energy hν and extrapolating for optical bandgaps (Eg) for Be2C films under various CH4–Ar ratios. | |
4. Conclusions
In summary, Be2C films were firstly prepared by DC reactive magnetron sputtering on optical quartz substrates with various CH4–Ar ratios and characterized by suitable analytical techniques. The XPS results revealed that the hydrocarbon content in films increased with increasing CH4–Ar ratios; meanwhile, beryllium carbide was the main component (>50 mol%) in films prepared with lower CH4–Ar ratios (<5%), while hydrocarbon films were obtained when the CH4–Ar ratios exceeded 15%. Based on this finding, protective coatings (∼26 nm CH) were in situ deposited on Be2C films, which was proven to be effective in preventing oxidation. HRTEM images clearly showed that nanocrystal Be2C particles were embedded in the amorphous hydrocarbon matrices. The films exhibited a smooth surface and columnar structure on the cross-sectional view as presented by AFM and SEM. In our experiments, the deposition rates (∼125 nm h−1) were higher than that of RF sputtering, which was favourable for depositing inertial confinement fusion (ICF) capsules. The optical measurements revealed the high transparency of Be2C films in visible light to the near-infrared region (500–2000 nm), suggesting the feasibility of characterizing the deuterium–tritium (D–T) ice layers by optical methods and homogenizing by infrared heating in Be2C capsules. In addition, the dispersion of the optical constant was investigated. The obtained optical bandgap (Eg), refractive index (n) and real part of the dielectric constant (εr) for the purer Be2C films (deposited at a CH4–Ar ratio of 2.4%) were 1.98 eV, 2.52 and 6.34, respectively, and are in accordance with the theoretical calculations. The results showed that this material could be a potential candidate for application to ICF capsules.
Acknowledgements
This study was financially supported by the National Key Science Foundation of China (Grant NO. 2014YQ090709). The authors would like to acknowledge Yong Zeng and Dr Qi Yang for their help in testing. The authors also are grateful to Xiulan Tan and Dr Zhengwei Xiong for their beneficial discussions.
References
- C. H. Lee, W. R. L. Lambrecht and B. Segall, Phys. Rev. B: Condens. Matter Mater. Phys., 1995, 51(16), 10392–10398 CrossRef CAS.
- S. Laref and A. Laref, Comput. Mater. Sci., 2008, 44(2), 664–669 CrossRef CAS.
- C.-T. Tzeng, K.-D. Tsuei and W.-S. Lo, Phys. Rev. B: Condens. Matter Mater. Phys., 1998, 58(11), 6837–6843 CrossRef CAS.
- M. Mehine, C. Björkas, K. Vörtler, K. Nordlund and M. I. Airilaet, J. Nucl. Mater., 2011, 414(1), 1–7 CrossRef CAS.
- R. Mateus, P. A. Carvalho, N. Franco, L. C. Alves, M. Fonseca, C. Porosnicu and E. Alves, J. Nucl. Mater., 2013, 442(1), S320–S324 CrossRef CAS.
- A. N. Simakov, D. C. Wilson, A. Y. Sunghwan, J. L. Kline, D. S. Clark and S. H. Batha, Phys. Plasmas, 2014, 21(2), 022701 CrossRef.
- J. E. T. Team, J. Nucl. Mater., 1990, 176, 3–13 CrossRef.
- A. J. MacKinnon, N. B. Meezan, J. S. Ross, S. L. Pape and L. B. Hopkins, et al., Phys. Plasmas, 2014, 21, 056318 CrossRef.
- S. W. Haan, J. D. Salmonson, D. S. Clark, D. D. Ho, B. A. Hammel and D. A. Callahan, et al., Fusion Sci. Technol., 2011, 59(1), 1–7 CAS.
- B. C. Luo, K. Li, X. L. Tan, J. Q. Zhang, J. S. Luo, W. D. Wu and Y. J. Tang, J. Alloys Compd., 2014, 607, 150–156 CrossRef CAS.
- B. A. Hammel and The National Ignition Campaign Team, Plasma Phys. Controlled Fusion, 2006, 48(12B), B497–B506 CrossRef CAS.
- R. Brusasco, S. Letts, P. Miller, M. Saculla and R. Cook, J. Vac. Sci. Technol., A, 1996, 14(3), 1019–1024 CAS.
- W. S. Shih, R. B. Stephens and W. J. James, Fusion Sci. Technol., 2000, 37(1), 24–31 CAS.
- M. W. Mallett, E. A. Durbin, M. C. Udy, D. A. Vaughan and E. J. Center, J. Electrochem. Soc., 1954, 101(6), 298–305 CrossRef CAS.
- B. C. Luo, K. Li, Y. D. He, G. Niu, J. Q. Zhang, J. S. Luo, W. D. Wu and Y. J. Tang, High Power Laser Part. Beams, 2013, 25(12), 2359–2365 Search PubMed.
- F. Kalarasse and B. Bennecer, J. Phys. Chem. Solids, 2008, 69(7), 1775–1781 CrossRef CAS.
- J. H. Coobs and W. J. Koshuba, J. Electrochem. Soc., 1952, 99(3), 115–120 CrossRef CAS.
- Y. X. Xie, R. B. Stephens, N. C. Morosoff and W. J. James, J. Fusion Energy, 1998, 17(3), 259–260 CrossRef CAS.
- W. S. Shih, N. E. Barr, W. J. James, N. C. Morosoff and R. B. Stephens, Fusion Sci. Technol., 1997, 31(4), 442–448 CAS.
- R. Brusasco, S. Letts, P. Miller, M. Saculla and R. Cook, J. Vac. Sci. Technol., A, 1996, 14(3), 1019–1024 CAS.
- Y. X. Xie, Plasma deposition of beryllium carbide via magnetron sputtering [D], Univ. of Missouri, Rolla, 1998 Search PubMed.
- Y. X. Xie, N. C. Morosoff and W.
J. James, J. Nucl. Mater., 2001, 289, 48–51 CrossRef CAS.
- A. F. Jankowski, M. A. Wall and T. G. Nieh, J. Non-Cryst. Solids, 2003, 317(1), 129–136 CrossRef CAS.
- A. F. Jankowski, M. A. Wall, A. W. Van Buuren, T. G. Nieh and J. Wadsworth, Acta Mater., 2002, 50, 4791–4800 CrossRef CAS.
- Powder Diffraction File (PCPDFWIN v.2.02), JCPDS-International Centre for Diffraction Data, 33–0191, 1999.
- J. Nazon, J. Sarradin, V. Flaud, J. C. Tedenac and N. Frety, J. Alloys Compd., 2008, 464(1), 526–531 CrossRef CAS.
- S. Mehraj, M. S. Ansari and Alimuddin, Thin Solid Films, 2015, 589, 57–65 CrossRef CAS.
- Y. D. He, J. S. Luo, J. Li, L. B. Meng, B. C. Luo, J. Q. Zhang, Y. Zeng and W. D. Wu, Fusion Eng. Des., 2016, 103, 118–124 CrossRef CAS.
- Y. Xin, H. Zhou, X. Ni, Y. Pan, X. Zhang, J. Zheng, S. Bao and P. Jin, RSC Adv., 2015, 5(71), 57757–57763 RSC.
- J. M. Khoshman, A. Khan and M. E. Kordesch, J. Appl. Phys., 2007, 101, 103532 CrossRef.
- K. E. Oughstun and N. A. Cartwright, Opt. Express, 2003, 11(13), 1541–1546 CrossRef PubMed.
- A. L. Ivanovskii, I. R. Shein, Y. N. Makurin, V. S. Kiiko and M. A. Gorbunova, Inorg. Mater., 2009, 45(3), 223–234 CrossRef CAS.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra02141g |
|
This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.