Next Article in Journal
Preliminary Investigation on Degradation Behavior and Cytocompatibility of Ca-P-Sr Coated Pure Zinc
Previous Article in Journal
Wear Resistance of Aluminum Matrix Composites’ Coatings Added on AA6082 Aluminum Alloy by Laser Cladding
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancement of sp3 C Fraction in Diamond-like Carbon Coatings by Cryogenic Treatment

1
School of Materials Science and Engineering, South China University of Technology, Guangzhou 510640, China
2
School of Mechanical & Automotive Engineering, South China University of Technology, Guangzhou 510640, China
3
School of Medicine, South China University of Technology, Guangzhou 510006, China
*
Author to whom correspondence should be addressed.
Coatings 2022, 12(1), 42; https://doi.org/10.3390/coatings12010042
Submission received: 6 December 2021 / Revised: 22 December 2021 / Accepted: 23 December 2021 / Published: 30 December 2021
(This article belongs to the Section Surface Characterization, Deposition and Modification)

Abstract

:
Diamond-like carbon (DLC) coatings deposited onto high-speed-steel surfaces were subjected to deep cryogenic treatment (DCT) at temperatures of −120 to −196 °C to investigate the evolution of microstructure, bonding structure, and mechanical properties. The surface morphology and the bonding structure of the DLC coatings were studied using scanning electron microscopy, transmission electron microscopy, micro-Raman spectroscopy, and X-ray photoelectron spectroscopy. It is found that DCT affects the surface morphology, especially the size and the height of the aggregates. For those DLCs with more than 50% sp3 C fraction, the sp2 C → sp3 C transformation occurred in coatings treated at a temperature of −120 to −160 °C; and the maximum fraction of sp3 C was obtained after treatment at −140 °C. Almost keeping the wear resistance of DLCs, DCT can improve the adhesion strength, and surface hardness. The findings of this study indicate that DCT will be a potential post-treatment method to tune the microstructure and mechanical performance of DLC coatings.

Graphical Abstract

1. Introduction

Diamond-like carbon coatings (DLC) have attracted increasing attention because of their excellent performance [1,2]. Their many special characters have found extensive application, including high hardness, solid lubricant, wear resistance, low-temperature field emission, tunable bandgap, wide electrochemical window, a wide range of optical transmission, gas sensitivity and barrier, biocompatibility, and so on [3,4,5,6,7,8,9,10]. DLCs consist of a mixture of sp3 C and sp2 C bonded carbon atoms. At present physical and plasma-enhanced chemical vacuum vapor depositing are still the main methods to prepare DLCs in which the energic C+ impinges the top surface and implants into the subsurface to create a thermal spike site with high intrinsic stress and high local temperature, thus, stabilize the sp3 C hybrid [11]. Thus, the sp3 C fraction can be controlled by varying the depositing parameters, such as bias voltage, working gas, doping species, and even the multiple layer structure [12,13,14,15,16,17]. The sp2 C/sp3 C bonding ratio plays a crucial role in determining the performance of DLCs [18,19,20]. Usually, the sp3 C fraction is responsible for the mechanical performance of DLCs, while the sp2 C fraction determines the electrical and optical properties. Some post-treatments can also adjust the sp3 C fraction in carbon materials, for example, annealing [21,22,23] and high-pressure deforming [24,25,26]. However, the significant drawbacks of annealing are reduction in the fraction of sp3 C bonds and hardness degradation, and the high-pressure deforming methods are destructive.
Is there a suitable and non-destructive post-treatment to enhance the sp3 C fraction to strengthen DLCs? The direct sp2 C to sp3 C transformation must overcome a high energy barrier (~0.4 eV per atom). From the thermodynamical viewpoint, high pressures and/or temperatures benefit sp3 C stabilization. Theoretical simulations and experimental studies showed a high contact pressure induced sp2 C → sp3 C transformation in carbon-based materials [27,28,29,30,31]. When impinged by a C+ with a chosen kinetic energy, a low substrate temperature enhances the intrinsic stress of the local sites to stabilize the sp3 C phase [32]. In addition to thermodynamical methods, the sp2 C to sp3 C transformation can occur if the energy release from a system is large enough to overcome the above barrier. For example, it is kinetically driven at a low temperature [33]. To explore a non-destructive method to enhance the sp3 C fraction of DLC, we pay attention to the deep cryogenic treatment (DCT) in this study.
DCT has been extensively used in the machining industry to improve the mechanical performance of alloys by inducing phase transformation and microstructural changes [34,35,36,37]. DLCs have been reported to perform well under cryogenic service conditions, e.g., cryogenically treated carbon coatings exhibit significantly improved wear resistance [38,39,40,41,42,43]. However, there are few comprehensive studies on the effect of DCT on the bonding structure and microstructure of DLCs. We carried out a series of experiments at cryogenic conditions in this study. It was found that the surface morphology and the bonding structure of DLCs varied with the cryogenic temperature. The plausible reason behind them has been provided.

2. Materials and Methods

2.1. Substrate Treatment

Substrate coupons included EM 35 high speed steel disks with Φ 20.5 mm × 5 mm size and single-crystal Si wafers. The Si wafer substrate were without any treatment. EM35 alloy commercially from Eramet Group (Paris, France) is with composition of 0.93% C, 0.3% Mn, 0.35% Si, 4.1% Cr, 5.0% Mo, 6.2% W, 1.9% V, and balance F. The different groups and their processing steps are listed in Table 1. Quenching was performed by heating the samples in a VUQ-557H-10 furnace (Find-Shine Vacuum Tech. Co., Ltd., Shengyang, China) at 1170 °C for 15 min followed by cooling using nitrogen gas. Tempering was carried out thrice in a VPT-557H furnace (Find-Shine Vacuum Tech. Co., Ltd., Shengyang, China) at 560 °C for 1 h. The cooling and the heating rates during DCT were maintained at 5 °C /min using a cooling box (Cryometal-50, Aisike Instrum. Co., Ltd., Wuxi, China). Cryo-treat-I (−180 °C for 6 h) was used to stabilize the microstructure of the EM35 alloy (similar detail seen in [34]).

2.2. DLC Deposition and Post-Treatment

After those substrate EM 35 coupons treated above and Si wafers were mirror polished, orderly cleaned in ethanol and deionized water to remove surface contaminants, and dried using N2 stream, they are put into the vacuum chamber of Hauzer Flexicoat 850 PECVD system (IHI Corporation, Venlo, The Netherlands). The deposition processing steps are presented in Table 2. Prior to depositing the a-C:H top layer, a Cr+WC interlayer was prepared using WC and Cr targets (99.99 at.%) connected to DC power sources in the presence of a carrier gas (argon, 99.999% pure). Then, the a-C:H was immediately deposited in the vacuum chamber by the method of pulsed-DC PECVD with 40 kHz frequency and 740 V bias voltage, using acetylene (99.999 at.%) as the gas precursor. The thickness of the top a-C:H layer was 1.6–1.9 μm.
After the DLC deposition, the Cryo-treat-II step in Table 1 was carried out for some specimens at different temperatures, i.e., −120, −140, −160, −180 and −196 °C for 6 h. The Si wafer specimens with or without DLC were only used to measure the residual stress in the coatings.

2.3. Microstructure Characterization

Scanning electronic microscopy (SEM; Sigma 300, Zeiss Group, Oberkochen, Germany) and 300 kV transmission electron microscopy (TEM; Titan3 Themis G2, Thermo Fisher Scientific, Waltham, MA, USA) studies were conducted to analyze the microstructure of the DLC samples. The hybridization of carbon was determined by micro-Raman spectroscopy using a 532-nm laser (LabRAM Arami, Horiba Scientific, Palaiseau, France) and X-ray photoelectron spectroscopy (XPS) at an incident photon energy of 1486.6 eV (Thermo K-Alpha+, Thermo Fisher Scientific, Waltham, MA, USA). The XPS instrument was calibrated using Au4f7/2 core-level electron binding energy (84.00 eV) as the standard. Prior to each measurement, the specimen surface was etched for 60 s using Ar ions with 2 keV energy and a current density of 1 mA/mm2. The Shirley method was used for background subtraction and the data analysis was performed using the XPS Peak 4.1 software, using the Gaussian–Lorentzian (G–L) with a 20% L mode.

2.4. Mechanical Performance Characterization

Friction tests were performed using a pin-on-disk instrument (SFT-2M, Zhongke Kaihua Co., Ltd., Lanzhou, China) by applying a normal load of 40 N in air at 25 °C and 40% relative humidity, and an alumina ball with a diameter of 4 mm was used as the pin. The wear volume was measured using a light-interference profiler (RTEC UP Dual Model, Rtec Instruments, San Jose, CA, USA).
The hardness was measured using a nanoindenter (Anton Paar TTX-NHT3, Anton Paar GmbH, Graz, Austria) by applying a load of 5 mN at a rate of 10 mN/min. For each specimen, the average hardness value over six tests was considered as the hardness of the coatings. The adhesion strength was estimated using a scratch instrument with a diamond stylus tip of radius 200 μm (MFT-4000, Zhongke Kaite, Lanzhou, China). A maximum load of 100 N was applied at a rate of 100 N/min for a scratch length of 5 mm. The adhesion was characterized by the critical load which induced the first detachment of the film. For each specimen, four tests were performed.
The residual stress in the coating deposited on the Si wafer was analyzed using a film stress instrument (FST1000, Supro Instruments, Shenzhen, China) and the Stoney’s Equation (1) [13,14,23]. In this formula, E is the Young’s modulus of the Si wafer, hs is the thickness of the silicon wafer, ν is the Poisson ratio of the Si wafer, hf is the thickness of the film, R0 is the radius of curvature of the Si wafer before film deposition, and R is the radius of curvature after film deposition. The residual stress was then taken as the average of the values for the three specimens with the same surface treatment.
σ x , y   =   Eh s 2 1 / R 0 1 / R / 6 h f 1 ν .

3. Results

3.1. Microstructure

All DLC-coated specimens were prepared in the same heat. The top surface morphology of specimen EM-0 without DCT is shown in Figure 1a, which presents smooth appearance with tens’ nanometer of cauliflower-like aggregates. The surface morphology changes significantly after DCT, and the aggregates with different sizes and heights are observed in Figure 1b–f. Many pinhole-like defects exist at the boundaries of the aggregates in Figure 1d. In this study, the size and the height of the aggregates vary inconsistently with the DCT temperature; smooth (Figure 1b,e) and rough (Figure 1c,d) surface morphology images are observed. DCT process leads to shrinkage and even elimination of various defects, smoothening the surface of DLC or graphite-like carbon (GLC) films [38,39]. Sometimes an unusual large submicrometer sp2 C-enriched aggregates could be produced when DLCs depositing onto a 20–100 K cryogenic substrate using pulsed laser [44], and it is unlike this study. If only considering the shrinkage effect induced by DCT, it is difficult to interpret the surface morphology change in this study. This result probably relates to phase transformation during DCT, and it will be discussed in the next sections.
The cross-sectional TEM images of specimens EM-0 and EM-140 are presented respectively in Figure 2a and (Figure 2b), indicating that the coatings are intact, and no microcracks are formed after DCT. There is no locally ordered region formation in the cryogenically treated coating in the high-resolution TEM (HRTEM) image (Figure 2c).

3.2. Chemical Bonding

The raw and fitted peaks of C 1s spectra of different specimens are shown in Figure 3a–f. The standard core level binding energies (BE) of sp2 C=C, sp3 C–C, and C–O bonds are 284.4 ± 0.2, 285.2 ± 0.2 and 286.5 ± 0.3 eV, respectively [6,12,19]. During fitting, the fitted BE deviated slightly from the above ideal value probably due to the collective effect of the chemical environment, defect density, together with residual stress [17]. It is shown that the area of sp3 C–C and sp2 C=C varies a little with DCT temperature.
Raman spectra varied with DCT temperature, as shown in Figure 4a, where the dot line linking those peak points deviated nonlinearly from the straight line. It meant microstructure change induced by DCT at different temperature. The so-called G peak and D peak, fitted based on two Gaussian curve shapes, lay at around 1580 and 1360 cm−1 respectively [17,45]. The G peak the D peak can be fitted [20,29,30]. The G peak is due to all sp2 C sites, and the D peak is only due to six-fold ring sp2 C sites [11]. Thus, the G peak position and the intensity ratio (ID/IG) depend inversely on the content of sp3 C, number of chain-like sp2 C sites; and relate to cluster size, residual stress, and disorder degree [20,44,46].
The variations of ID/IG and sp2/sp3 bond ratios with the DCT temperature are shown in Figure 4b. The sp2 C=C → sp3 C–C transformation takes place in the specimens treated at temperatures of −120 to −160 °C. This result was surprising, so we repeated these tests three times. Every time EM-140 specimen exhibits the highest fraction of sp3 C; which can reach approximately 4% more than EM-0 without DCT. The temperature dependence of ID/IG ratio differs from that of sp2 C/sp3 C bond ratio at temperatures below −180 °C, which may be partly attributed to the different depth resolutions of XPS and Raman techniques, as well as the different contributions from surface dangling bonds [23,29]. The C–O bond fraction differs in different DCT coatings in Figure 4, which were probably produced by the chemical absorption of O2 and H2O in the environment. The background signal of XPS spectra intensified when some kinds of pencil lead was wet etched [45], probably it was due to the similar chemical absorption.

3.3. Mechanical Properties

The macro-residual stress in the DLCs after DCT is similar to that in the untreated DLC (Figure 5a). The adhesion of all the specimens except EM-196 is improved, as shown in Figure 5b. Specimen EM-140 exhibits the highest hardness in Figure 5c. The variation of hardness with DCT temperature agree with the above change of ID/IG and sp2/sp3 bond ratios. Because specimens with or without DCT kept almost the same residual stress, and the nano-indention depth was less than 1/8 thickness of the DLC coatings, thus, the hardness improvement of EM-140 surely was responsible for sp3 C fraction enhancement [2,12,20]. One hand, the mechanical and micro-structural change of EM 35 alloy substrate induced by DCT [34,35,37] provided strong support for the coating; another hand, the sp2/sp3 C ratio of the coating affected the interface strength between DLC and the interlayer.
The friction curves and wear volumes of the specimens are shown in Figure 6a,b, respectively. The steady-state friction coefficient of EM-0 is 0.08; the other specimens exhibit higher or lower values, c.f., 0.1 for EM-196, and 0.08 for EM-140. Because the sp3 C fraction the hardness differs a little for these coatings, friction coefficient seems positively related to the surface roughness under high load 40 N. From Figure 1 the surface roughness can be ranked in the order: EM-180 < EM-140 < EM-120 < EM-0 < EM-160 < EM-196. The asperity on the surface affects the effective contact Hertz stress [23,29]. The factors that influence the friction performance of DLCs include surface morphology, adhesion, and hardness [1,2,7,46]. Contrasting with EM-0, cryogenic treatment almost holds the wear resistance of DLCs. Probably rough surface and higher sp2 C in the EM-196 coating, the tribofilm could be formed easily, which presented the minimum wear volume. Detailed analyses of these factors will be carried out in our future work.

4. Discussion

The surface morphology is determined by the bombardment energy and the hydrogen content of the DLC [17,47]. The change in the surface morphology after DCT is not only due to the cooling contraction, but also due to phase transformation. Based on the hybrid structure analysis in Section 3.2, it shows that the sp2 C → sp3 C transformation can be fulfilled under a combination of suitable cryogenic temperature and original DLC microstructure. Owing to the presence of large amounts of unbonded H in the coating [14,17], it is expected that H2 outdiffuses during cooling. This may have resulted in the formation of pinhole-like defects in EM-160.
High pressure deformation, high pressure torsion, and high contact pressure of friction can result in phase transformation of carbon materials to enhance the sp3 C fraction [26,27,28,30]. According to the carbon P–T phase [32], after overcoming the energy barrier (−0.4 eV per carbon atom [33]), the pressure required to stabilize the diamond phase decreases with decrease in temperature. Recently the bulk ta-C could be synthesized at room temperature by pressing the glassy carbon (sp2 C) if the pressure was more than 8.5 GPa [25]. In this study the macro residual compressive stress in DLCs was approximately 1.8 GPa (Figure 5a), which is not enough to promote sp2 C → sp3 C transformation.
In order to plausibly interpret the phase transformation induced by DCT, we propose that the following conditions need to be satisfied to realize sp2 C → sp3 C. First, very high pressure should be created at local sites in the coatings. Second, the release of elastic energy must be greater than the energy barrier. In the present study, the fraction of sp3 C bond in the as-deposited DLCs is more than 50% (Figure 4b). This results in the formation of a rigid network of sp3 C bonds. If sp2 C=C clusters embedded in the rigid network slide under strain, the confinement from the rigid network generates high local pressure. During DCT, the mismatch between the rigid network and the clusters provides the needed local strain if the rigid network is stiff enough. In summary, the perquisites to realize the sp2 C → sp3 C transformation are: (1) enough stiffness of the rigid network of sp3 C bonds; and (2) suitable strain range in the clusters. The above-mentioned requirements are fulfilled when the DLCs are treated at temperatures of −120 to −160 °C. The very high local pressure at temperatures lower than −160 °C collapses the rigid network; the local pressure is too low when treated at temperatures higher than −120 °C. In addition, the sp3 CHn fraction in a DLC will affect the rigidity of the sp3 C network, and influence the sp2 C to sp3 C transformation. Verification of the above hypothesis is undergoing by molecular dynamic simulation and systematical experiments.

5. Conclusions

The effects of DCT at temperatures of −120 to −196 °C on the microstructure, bonding structure, and mechanical properties of DLCs were investigated. DCT affects the surface morphology, especially the size and the height of the aggregates. The hardness of DLC improves after treatment at −140 °C due to sp2 C → sp3 C transformation. DCT at suitable conditions is beneficial for enhancing the fraction of sp3 bonded carbon and the surface hardness of DLC. Thus, DCT will be a potential method for tuning the microstructure and mechanical properties of DLCs.

Author Contributions

Conceptualization, J.P.; methodology, J.P.; validation, J.L., Y.P., Y.X. and J.H.; formal analysis, J.L. and Y.P.; investigation, J.L., J.H. and L.L.; data curation, J.L. and Y.X.; writing—original draft preparation, J.P.; writing—review and editing, J.P.; supervision, J.P.; project administration, J.P.; funding acquisition, J.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Guangzhou Science and Technology project (Grant 201902010018 and 201807010091).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article.

Acknowledgments

This study was also supported by the financial aid from Conprofe Green Tools Co.: ltd., Guangzhou, China.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Vetter, J. 60 years of DLC coatings: Historical highlights and technical review of cathodic arc processes to synthesize various DLC types, and their evolution for industrial applications. Surf. Coat. Technol. 2014, 257, 213–240. [Google Scholar] [CrossRef]
  2. Bewilogua, K.; Hofmann, D. History of diamond-like carbon films—From first experiments to worldwide applications. Surf. Coat. Technol. 2014, 242, 214–225. [Google Scholar] [CrossRef]
  3. Birney, R.; Cumming, A.; Campsie, P.; Gibson, D.; Hammond, G.; Hough, J.; Martin, I.; Reid, S.; Rowan, S.; Song, S.; et al. Coatings and surface treatments for enhanced performance suspensions for future gravitational wave detectors. Class. Quantum Grav. 2017, 34, 235012. [Google Scholar] [CrossRef] [Green Version]
  4. Saha, B.; Toh, W.; Liu, E.; Tor, S.; Hardt, D.; Lee, J. A review on the importance of surface coating of micro/nano-mold in micro/nano-molding processes. J. Micromech. Microeng. 2016, 26, 013002. [Google Scholar] [CrossRef] [Green Version]
  5. Markwitz, A.; Leveneur, J.; Gupta, P.; Suschke, K.; Futter, J.; Rondeau, M. Transition metal ion implantation into diamond-like carbon coatings: Development of a base material for gas sensing applications. J. Nanomater. 2015, 2015, 706417. [Google Scholar] [CrossRef] [Green Version]
  6. Son, M.; Zhang, T.; Jo, Y.; Kim, K. Enhanced electrochemical properties of the DLC films with an arc interlayer, nitrogen doping and annealing. Surf. Coat. Technol. 2017, 329, 77–85. [Google Scholar] [CrossRef]
  7. Fan, X.; Xue, Q.; Wang, L. Carbon-based solid-liquid lubricating coatings for space applications—A review. Friction 2015, 3, 191–207. [Google Scholar] [CrossRef] [Green Version]
  8. Yu, Q.; Chen, X.; Zhang, C.; Luo, J. Influence factors on mechanisms of superlubricity in DLC films: A Review. Front. Mech. Eng. 2020, 6, 65. [Google Scholar] [CrossRef]
  9. Nakaya, M.; Uedono, A.; Hotta, A. Recent progress in gas barrier thin film coatings on PET bottles in food and beverage applications. Coatings 2015, 5, 987–1001. [Google Scholar] [CrossRef] [Green Version]
  10. Voevodin, A.; Zabinski, J. Naocomposite and nanostructured tribological materials for space applications. Compos. Sci. Technol. 2005, 65, 741–748. [Google Scholar] [CrossRef]
  11. Robertson, J. Diamond amorphous carbon. Mater. Sci. Eng. 2002, 37R, 129–281. [Google Scholar] [CrossRef] [Green Version]
  12. Almeida, L.; Souza, A.; Costa, L.; Rangel, E.; Manfrinato, M.; Rossino, L. Effect of nitrogen in the properties of diamond-like carbon (DLC) coating on Ti6Al4V substrate. Mater. Res. Express 2020, 7, 065601. [Google Scholar] [CrossRef]
  13. Zhang, D.; Li, S.; Zuo, X.; Guo, P.; Ke, P.; Wang, A. Structural and mechanism study on enhanced thermal stability of hydrogenated diamond-like carbon films doped with Si/O. Diam. Relat. Mater. 2020, 108, 107923. [Google Scholar] [CrossRef]
  14. Peng, J.; Yang, M.; Zeng, J.; Su, D.; Liao, J.; Yick, M. Influence of nitrogen doping on the thermal stability of hydrogenated amorphous diamond coating. Thin Solid Films 2020, 709, 138188. [Google Scholar] [CrossRef]
  15. Wang, J.; Pu, J.; Zhang, G.; Wang, L. Architecture of superthick diamond-like carbon films with excellent high temperature wear resistance. Tribol. Int. 2015, 81, 129–138. [Google Scholar] [CrossRef]
  16. Braak, R.; May, U.; Onuseit, L.; Repphun, G.; Guenther, M.; Schmid, C.; Durst, K. Accelerated thermal degradation of DLC-coatings via growth defects. Surf. Coat. Technol. 2018, 349, 272–278. [Google Scholar] [CrossRef]
  17. Peng, J.; Yang, M.; Bi, J.; Sheng, R.; Li, L. Hydrogen existence state of a hydrogenated amorphous carbon coating and its thermal stability. Diam. Relat. Mater. 2019, 99, 107535. [Google Scholar] [CrossRef]
  18. Yamamoto, S.; Kawana, A.; Ichimura, H.; Masuda, C. Relationship between tribological properties and sp3/sp2 structure of nitrogenated diamond-like carbon deposited by plasma CVD. Surf. Coat. Technol. 2012, 210, 1–9. [Google Scholar] [CrossRef]
  19. Muguruma, T.; Iijima, M.; Kawaguchi, M.; Mizoguchi, I. Effects of sp2/sp3 ratio and hydrogen content on in vitro bending and frictional performance of DLC-coated orthodontic stainless steels. Coatings 2018, 8, 199. [Google Scholar] [CrossRef] [Green Version]
  20. Li, A.; Li, X.; Wang, Y.; Lu, Z.; Wang, Y.; Zhang, G.; Wu, Z. Investigation of mechanical and tribological properties of super-thick DLC films with different modulation ratios prepared by PECVD. Mater. Res. Express 2019, 6, 086433. [Google Scholar] [CrossRef]
  21. Takabayashi, S.; Okamoto, K.; Nakatani, T. Influence of post-annealing on a diamondlike carbon film analyzed by Raman spectroscopy. Surf. Interface Anal. 2018, 50, 441–447. [Google Scholar] [CrossRef]
  22. Grenadyorov, A.; Solovyev, A.; Oskomov, K.; Oskirko, V.; Semenov, V. Thermal stability of anti-reflective and protective a-C:H:SiOx coating for infrared optics. Appl. Surf. Sci. 2020, 510, 145433. [Google Scholar] [CrossRef]
  23. Peng, J.; Liao, J.; Yang, M.; Liao, J.; Bi, J. Effect of annealing treatment on the tribological performance of hydrogenated amorphous diamond coatings doped with nitrogen. Ceram. Intern. 2021, 47, 13423–13431. [Google Scholar] [CrossRef]
  24. Durandurdu, M. Formation of a very high-density amorphous phase of carbon and its crystallization into a simple cubic structure at high pressure. Comput. Mater. Sci. 2021, 200, 110822. [Google Scholar] [CrossRef]
  25. Tan, L.; Sheng, H.; Lou, H.; Cheng, B.; Xuan, Y.; Prakapenka, V.; Greenberg, E.; Zeng, Q.; Peng, F.; Zeng, Z. High-pressure tetrahedral amorphous carbon synthesized by compressing glassy carbon at room temperature. J. Phys. Chem. C 2020, 124, 5489–5494. [Google Scholar] [CrossRef]
  26. Edalati, K.; Horita, Z. Phase transformations during high-pressure torsion (HPT) in titanium, cobalt and graphite. IOP Conf. Ser.: Mater. Sci. Eng. 2014, 63, 012099. [Google Scholar] [CrossRef] [Green Version]
  27. Li, X.; Wang, A.; Lee, K. Insights on low-friction mechanism of amorphous carbon films from reactive molecular dynamics study. Tribol. Int. 2019, 131, 567–578. [Google Scholar] [CrossRef]
  28. Bai, Y.; Yu, Z.; Liu, R.; Li, N.; Yan, S.; Yang, K.; Liu, B.; Wei, D.; Wang, L. Pressure-induced crystallization and phase transformation of para-xylene. Sci. Rep. 2017, 7, 5321. [Google Scholar] [CrossRef] [Green Version]
  29. Manimunda, P.; Al-Azizi, A.; Kim, S.; Chromik, R. Shear-induced structural changes and origin of ultralow friction of hydrogenated diamond-like carbon (DLC) in dry environment. ACS Appl. Mater. 2017, 9, 16704–16714. [Google Scholar] [CrossRef] [PubMed]
  30. Fukumasu, N.; Bernardes, C.; Ramirez, M.; Trava-Airoldi, V.; Souza, R.; Machado, I. Local transformation of amorphous hydrogenated carbon coating induced by high contact pressure. Tribol. Int. 2018, 124, 200–208. [Google Scholar] [CrossRef]
  31. Chen, L.; Liu, K.; Wei, X.; Lu, Z.; Ren, N.; Zhang, G.; Xue, Q. Enhancement in the tribological properties of Cr/DLC multilayers in methane: Structural transformation induced by sliding. SN Appl. Sci. 2019, 1, 1483. [Google Scholar] [CrossRef] [Green Version]
  32. Kalinichenko, A.; Perepelkin, S.; Strel’nitskij, V. Dependence of intrinsic stress and structure of ta-C film on ion energy and substrate temperature in model of the non-local thermoelastic peak. Diam. Relat. Mater. 2010, 19, 996–998. [Google Scholar] [CrossRef]
  33. Esmeryan, K.; Castano, C.; Bressler, A.; Abolghasemibizaki, M.; Fergusson, C.; Roberts, A.; Mohammad, R. Kinetically driven graphite-like to diamond-like carbon transformation in low temperature laminar diffusion flames. Diam. Relat. Mater. 2017, 75, 58–68. [Google Scholar] [CrossRef]
  34. Liao, J.; Li, L.; Peng, J.; Gao, J. Effects of deep cryogenic treatment on the microstructure and friction performance of M35 high-speed steel. J. Phys. Conf. Ser. 2020, 1676, 012098. [Google Scholar] [CrossRef]
  35. Jovicevic-Klug, P.; Jenko, M.; Jovicevic-Klug, M.; Batic, B.; Kovac, J.; Podgornik, B. Effect of deep cryogenic treatment on surface chemistry and microstructure of selected high-speed steels. Appl. Surf. Sci. 2021, 548, 149257. [Google Scholar] [CrossRef]
  36. Dhokey, N.; Hake, A.; Kadu, S.; Bhoskar, I.; Dey, G. Influence of cryoprocessing on mechanism of carbide development in cobalt-bearing high-speed steel (M35). Metall. Mater. Trans. A 2014, 45, 1508–1516. [Google Scholar] [CrossRef]
  37. Gavriljuk, V.; Theisen, W.; Sirosh, V.; Polshin, E.; Kortmann, A.; Mogilny, G.; Petrov, Y.; Tarusin, Y. Low-temperature martensitic transformation in tool steels in relation to their deep cryogenic treatment. Acta Mater. 2013, 61, 1705–1715. [Google Scholar] [CrossRef]
  38. Kong, Y.; Chen, Q.; Wang, L. Microstructures and tribological properties of GLC coated 100Cr6 bearing steels. Mater. Res. Express 2017, 4, 116404. [Google Scholar] [CrossRef]
  39. Zhao, Y.; Nie, C.; Jin, Y.; Zhu, W.; Liu, X.; Nie, Y. Effect of cryogenic treatment on friction and wear property of DLC film/304 stainless steel. Adv. Mater. Res. 2014, 900, 767–772. [Google Scholar] [CrossRef]
  40. Wang, J.; Jia, Q.; Yuan, X.; Wang, S. Experimental study on friction and wear behaviour of amorphous carbon coatings for mechanical seals in cryogenic environment. Appl. Surf. Sci. 2012, 258, 9531–9535. [Google Scholar] [CrossRef]
  41. Aggleton, M.; Burton, J.; Taborek, P. Cryogenic vacuum tribology of diamond and diamond-like carbon films. J. Appl. Phys. 2009, 106, 013504. [Google Scholar] [CrossRef] [Green Version]
  42. Gradt, T.; Borner, H.; Schneider, T. Low temperature tribometers and the behaviour of ADLC coatings in cryogenic environment. Tribol. Intern. 2001, 34, 225–230. [Google Scholar] [CrossRef]
  43. Ostrovskaya, Y.; Strelnitskij, V.; Kuleba, V.; Gamulya, G. Friction and wear behaviour of hard and superhard coatings at cryogenic temperatures. Tribol. Intern. 2001, 34, 255–263. [Google Scholar] [CrossRef]
  44. Hu, A.; Alkhesho, I.; Duley, W.; Zhou, H. Cryogenic graphitization of submicrometer grains embedded in nanostructured tetrahedral amorphous carbon films. J. Appl. Phys. 2006, 100, 084319. [Google Scholar] [CrossRef]
  45. Knápek, A.; Sobola, D.; Burda, D.; Danhel, A.; Mousa, M.; Kolarík, V. Polymer graphite pencil lead as a cheap alternative for classic conductive SPM probes. Nanomaterials 2019, 9, 1756. [Google Scholar] [CrossRef] [Green Version]
  46. Tambe, N.; Bhushan, B. Nanoscale friction-induced phase transformation of diamond-like carbon. Scr. Mater. 2005, 52, 751–755. [Google Scholar] [CrossRef]
  47. Catena, A.; McJunkin, T.; Agnello, S.; Gelardi, F.; Wehner, S.; Fischer, C. Surface morphology and grain analysis of successively industrially grown amorphous hydrogenated carbon films (a-C:H) on silicon. Appl. Surf. Sci. 2015, 347, 657–667. [Google Scholar] [CrossRef]
Figure 1. SEM images of top-surface morphologies of different specimens (af).
Figure 1. SEM images of top-surface morphologies of different specimens (af).
Coatings 12 00042 g001aCoatings 12 00042 g001b
Figure 2. TEM images of the cross-sections (a,b) and plan-view HRTEM image (c).
Figure 2. TEM images of the cross-sections (a,b) and plan-view HRTEM image (c).
Coatings 12 00042 g002
Figure 3. (af): C 1s XPS spectra of different specimens.
Figure 3. (af): C 1s XPS spectra of different specimens.
Coatings 12 00042 g003
Figure 4. Raman spectra (a), ID/IG and sp2/sp3 bond ratios (b) of specimens with the cryogenic temperature.
Figure 4. Raman spectra (a), ID/IG and sp2/sp3 bond ratios (b) of specimens with the cryogenic temperature.
Coatings 12 00042 g004
Figure 5. Residual stress (a), adhesion strength (b), and hardness (c) of specimens subjected to different treatments.
Figure 5. Residual stress (a), adhesion strength (b), and hardness (c) of specimens subjected to different treatments.
Coatings 12 00042 g005
Figure 6. Friction coefficient (a) and wear volume (b) of specimens subjected to different treatments.
Figure 6. Friction coefficient (a) and wear volume (b) of specimens subjected to different treatments.
Coatings 12 00042 g006
Table 1. Specimen groups and their processing steps.
Table 1. Specimen groups and their processing steps.
GroupPre-TreatDLC CoatPost-Treat
SubstrateQuenchingTemperingCryo-Treat-ICryo-Treat-II
EM-0EM35YESYESYESYESNO
EM-TEM35YESYESYESYESYES (at T)
Si-0Si waferNONONOYESNO
Si-TSi waferNONONOYESYES (at T)
Table 2. Summary of a-C:H film deposition parameters.
Table 2. Summary of a-C:H film deposition parameters.
ItemItem Parameter
Substrate cleaningBase pressure 5 × 10−3 Pa; voltage bias 200 V; Ar gas flow rate 250 sccm; duration 90 min
Cr + WC interlayerTarget power 5 kW; heating 150 °C; duration 50 min
a-C:H layerBias voltage 740 V; coil current 2 A; C2H2 gas flow rate 250 sccm; heating 300 °C; duration 80 min; work pressure 0.8–1 Pa
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Peng, J.; Liao, J.; Peng, Y.; Xiao, Y.; Huang, J.; Li, L. Enhancement of sp3 C Fraction in Diamond-like Carbon Coatings by Cryogenic Treatment. Coatings 2022, 12, 42. https://doi.org/10.3390/coatings12010042

AMA Style

Peng J, Liao J, Peng Y, Xiao Y, Huang J, Li L. Enhancement of sp3 C Fraction in Diamond-like Carbon Coatings by Cryogenic Treatment. Coatings. 2022; 12(1):42. https://doi.org/10.3390/coatings12010042

Chicago/Turabian Style

Peng, Jihua, Jingwen Liao, Yinglong Peng, Yang Xiao, Jinhai Huang, and Liejun Li. 2022. "Enhancement of sp3 C Fraction in Diamond-like Carbon Coatings by Cryogenic Treatment" Coatings 12, no. 1: 42. https://doi.org/10.3390/coatings12010042

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