Next Article in Journal
Green Coating Polymers in Meat Preservation
Next Article in Special Issue
Potential Applications of Geopolymer Cement-Based Composite as Self-Cleaning Coating: A Review
Previous Article in Journal
Impacting Droplet Can Mitigate Dust from PDMS Micro-Post Array Surfaces
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Properties of Spark Plasma Sintered Compacts and Magnetron Sputtered Coatings Made from Cr, Mo, Re and Zr Alloyed Tungsten Diboride

1
Institute of Fundamental Technological Research Polish Academy of Sciences, 5B Pawinskiego St., 02-106 Warsaw, Poland
2
Faculty of Mechanical and Industrial Engineering, Warsaw University of Technology, Narbutta 85, 02-524 Warsaw, Poland
3
Łukasiewicz Research Network – Metal Forming Institute, 14 Jana Pawla II St., 61-139 Poznan, Poland
*
Author to whom correspondence should be addressed.
Coatings 2021, 11(11), 1378; https://doi.org/10.3390/coatings11111378
Submission received: 22 October 2021 / Revised: 3 November 2021 / Accepted: 5 November 2021 / Published: 10 November 2021
(This article belongs to the Special Issue Hard Transition Metal Compound Coatings with Increased Flexibility)

Abstract

:
To enhance the properties of tungsten diboride, we have synthesized and characterized solid solutions of this material with chromium, molybdenum, rhenium and zirconium. The obtained materials were subsequently deposited as coatings. Various concentrations of these transition metal elements, ranging from 0.0 to 24.0 at.%, on a metals basis, were made. Spark plasma sintering was used to synthesize these refractory compounds from the pure elements. Elemental and phase purity of both samples (sintered compacts and coatings) were examined using energy dispersive X-ray spectroscopy and X-ray diffraction. Microindentation was utilized to measure the Vickers hardness. X-ray diffraction results indicate that the solubility limit is below 8 at.% for Mo, Re and Zr and below 16 at.% for Cr. Above this limit both diborides (W,TM)B2 are created. Addition of transition metals caused decrease of density and increase of hardness and electrical conductivity of sintered compacts. Deposited coatings W1−xTMxBy (TM = Cr, Mo, Re, Zr; x = 0.2; y = 1.7–2) are homogenous, smooth and hard. The maximal hardness was measured for W-Cr-B films and under the load of 10 g was 50.4 ± 4.7 GPa. Deposited films possess relatively high fracture toughness and for WB2 coatings alloyed with zirconium it is K1c = 2.11 MPa m1/2.

1. Introduction

Nowadays the rise of a broad class of compounds comprising heavy transition metal (TM) and light-element atoms, like nitrides, carbides, and borides can be seen. Such compounds possess excellent mechanical properties such as high hardness and high wear resistance, refractory properties and also good thermal and electrical conductivity resulting from the concurrently high valence electron density and strong covalent bonding in these compounds [1,2]. These advanced ceramics can find applications in key technological fields, including cutting and drilling tools, wear resistant coatings, and engine components [3,4]. However, a major challenge remains in that the applications are to produce reliable tool components made of these materials in a relatively simple and time-consuming manner. This challenge is partly resolved by deposition of thin films [5,6,7,8] developing easily machinable materials [3] or new methods of sintering [9]. For better electro machining the good electrical conductivity is needed. Among such compounds, tungsten diborides (WB2) alloyed with other TM are especially promising for their very high hardness accompanied with increased toughness and good electrical conductivity [10,11]. The theoretical studies have showed that WB2 doped by TM can possess the very high hardness above 40 GPa [1,5].
One of best candidates of WB2 alloying is rhenium (Re). Diboride of this metal is superhard and can crystalize in similar to WB2 hexagonal AlB2-type structure [12]. The addition of rhenium into the WB4 and W2B5 phases have been reported by Mohammadi et al. [13] and by Feng et al. [14]. With the addition of 1 at.% Re, the Vickers microhardness increased to approximately 50 GPa at 0.49 N. Obtained tungsten tetraboride (WB4) with 1 at.% Re admixture is thermally stable up to approximately 400 °C in air. In the case of Re alloyed W2B5 theoretical studies showed that W1.5Re0.5B5 and W0.5Re1.5B5 are energetically and thermodynamically stable. The calculated hardness of W2B5 and W0.5Re1.5B5 was 16.11 and 17.91 GPa, respectively. On the other hand, ReB2 alloyed by tungsten was studied experimentally by Lech et al. [15]. The solid solutions of tungsten in ReB2 have been successfully synthesized by using an electric-arc furnace. The solubility limit for tungsten in ReB2 is nearly 48 at.%, which indicates a very high degree of solubility. The studies showed also that ReB2 structured compounds are superhard and may warrant further studies into additional solid solutions or ternary compounds taking this structure type. By using first-principles calculations, the structure and elastic properties of RexW1−xB2 alloys (x = 0.1–0.9) with the ReB2 like structure were studied by Yufei Tu et al. [16]. It was found that Re0.4W0.6B2 has both high bulk modulus (360 GPa) and high shear modulus (291 GPa). The authors show that there is a pseudogap at the Fermi level, which indicates that the Re0.4W0.6B2 is stable. The superior elastic property suggests the Re0.4W0.6B2 is desired as a potential hard material. Shaw [10] showed that addition of molybdenum (Mo) to WB2 significantly improves wear resistance. The formation of WB2 and MoB2 hard phases with AlB2-type structure is most likely the reason for the very low wear volume. In this case, the wear volume decreases from 1.00 × 107 (WB2) to 3.22 × 106 μm3 for WB2–MoB2 binary compound. Additionally, corrosion resistance of WB2 ceramic was improved. In the case of alloying with zirconium (Zr) [9], the spark plasma sintered compacts are characterized by a specific wear rate in the range of 1 × 10−6–3 × 10−5 mm3/Nm. The friction and wear test results reveal the formation of a boron-based film which seems to play the role of solid lubricant. This effect demonstrates the possibility to use sintered W-Zr-B materials to make machining tools and various components work in harsh conditions. With an increasing zirconium content, the electrical conductivity of the SPSed specimens increased both in W1−xZrxB4.5 and W1−xZrxB4.5 (0 < x ≤ 0.24). W0.76Zr0.24B2.5 sintered at a 24 min holding time is characterized by the highest electrical conductivity of 3.961 × 106 S/m, which is similar to the electrical conductivity of WC-Co cemented carbides [9]. In the case WB4 alloyed with chromium (Cr), Mohammadi et al. [13] showed that addition of 10 at.% Cr on a metals basis results in an increase in microhardness (at 0.49 N) from 43.3 ± 2.9 GPa for pure WB4 to 53.5 ± 1.9 GPa and the defect structure of this material may be responsible for the hardening trends observed for the solid solutions.
As it was shown by Garbiec et al. [9], it is possible to synthesize in one stage fast process the W1−xZrxB2 ceramic with properties competitive to the widely used WC-4Co sintered compacts. Additionally, until now there has been no information about the use of WBx alloyed by Re, Mo, Cr, Zr as a coatings deposited by magnetron sputtering (MS). However, addition of other transition metals such as tantalum (Ta) [11] or titanium (Ti) [6,17] to α-WB2 lattice can cause combining of high strength with high fracture toughness in deposited coatings. Deposited films are superhard (H > 40 GPa) and in the case of tantalum can exhibit fracture toughness K1c values of 3.7 to 3.0 MPa∙m1/2 for increasing Ta content (single-phased up to 26 at.% Ta) [11]. Psiuk et al. [18] studied W-Zr-B coatings deposited by a hybrid process combining pulsed laser deposition of ZrB2 and radio frequency magnetron sputtering of W2B5. Deposited films with ≈1% atomic content of Zr were superhard (H = 40 ± 4 GPa) and incompressible (Rs = 12 ± 1 GPa) but possessed a relatively low Young’s modulus (E = 330 ± 32 GPa) and a high elastic recovery (We = 0.9). The above promising results suggest the need for further research.
Currently, there are only two commercially relevant superhard materials: diamond and cubic boron nitride [3]. Although diamond can be used to cut through rock, it cannot be used to cut steel due to a chemical reaction that forms iron carbide. Instead, cubic boron nitride is used to cut and mill ferrous metals. In some ways, it is remarkable that humanity has achieved so much in terms of industrial machining, drilling, and polishing, relying on just two high-end compositions. In this work novel materials in the form of compacts and protective coatings are presented. As it is shown in the above literature and on the basis of our research it can be concluded that alloyed WB2 with Cr, Mo, Re or Zr can be competitive to widely used WC-Co. Proposed materials possess greater hardness and 20% lower density than WC. The alloying with transition metals additionally improves elastic and conductive properties of these materials. Because WC is quite brittle, binders such as cobalt are currently added to improve ductility, but such addition comes at the expense of hardness and additionally Co is toxic. The materials proposed by us may in the future replace tungsten carbide and may find many industrial applications, such as high-performance cutting tools and wear-protective coatings. In addition, W-TM-B coatings were investigated due to their excellent mechanical properties, i.e., very high hardness and relatively low Young’s modulus, which make them competitive to commercial protective TiN coatings [11].
In this article the synthesis of novel materials by spark plasma sintering and the possibility of subsequent using of compacts as a source for deposition of hard coatings with improved crack resistivity is presented. Till now there is no information about properties of WB2 SPSed compacts and deposited by magnetron sputtering method coatings doped by chromium, molybdenum, rhenium and zirconium.
The aim of this work is to investigate and compare the influence of Re, Mo, Cr and Zr dopants on the microstructure, mechanical properties and electrical conductivity of WB2 ceramics sintered by SPS method technique and explore the possibility of using them for deposition of functional coatings with magnetron sputtering method.

2. Material and Methods

2.1. Spark Plasma Sintering

Tungsten (purity: 99.9%, APS: 25 µm), zirconium (purity: 99.8%, APS: 250–350 µm), molybdenum (purity: 99.9%, APS: <10 µm), rhenium (purity: 99.8%, APS: 20 µm), chromium (purity: 99%, APS: <5 µm) and amorphous boron (purity: 95%, APS: 1 µm) powders were mixed for 30 min using a Turbula T2F shaker-mixer (WAB, Muttenz, Switzerland) in the compositions presented in Table 1. The obtained powder mixtures were SPSed in vacuum using an HP D 25/3 (FCT Systeme, Frankenblick, Germany) furnace. The SPS process parameters are listed in Table 2. Samples with a diameter of 25.4 mm and thickness of ≈3.5 mm were produced.
Optimal sintering parameters were selected on the basis of [9], where the influence of the holding time (2.5–30 min), sintering temperature and compacting pressure on the densification behavior, microstructure evolution and development of the properties of W-Zr-B compounds were studied. An example of sintering curves (chromium) is presented in Figure 1. It is clearly seen that the densification process is complete and after 24 min the displacement of punch is nearly constant.

2.2. Radio Frequency Magnetron Sputtering (RF-MS)

W-TM-B (where TM = Cr, Mo, Re, Zr) films were deposited by RF magnetron sputtering on Si (100) plates. The substrates were cleaned with subsequent rinses in acetone, alcohol and deionized water. Each of the substrates was mounted on a rotating substrate holder and heated to 520 °C. Heating was performed for better layer-to-substrate adhesion [19].
The deposition was performed in a vacuum chamber pumped (PREVAC, Rogów, Poland) to residual pressure of 10−3 Pa and then filled with argon to a working pressure of 0.9 Pa. The argon flow was 19 mL/min and it was adjusted with a mass flow controller. For deposition the W0.76TM0.24B2.5 targets were chosen because of the best properties of coatings obtained from target with titanium dopant [6]. The water-cooled W-TM-B targets were mounted in the 1″ magnetron source (Kurt J. Lesker TORUS Magnetron Sputtering Cathode, Jefferson Hills, PA, USA) and positioned at a distance of 40 mm from the substrate. The RF sputtering power was 60 W. The bias potential on the substrate was floating. Pre-sputtering of the target was carried out for 5 min prior to each deposition. Then, the substrate holder was positioned next to the magnetron. The deposition time was 120 min. The deposition conditions were chosen due to the best mechanical and tribological properties of films presented in [6].

2.3. Characterization

The density of the SPSed samples was measured using the Archimedes method according to the ISO 18754: 2013 standard using an EX 2225DM (Ohaus, Parsippany, NJ, USA) laboratory scale with a resolution of 0.00001 g. The phase composition of the samples was examined by X-ray diffraction (XRD) with a D8 Discover (Bruker, Billerica, MA, USA) using CuKα radiation (λ = 1.5418 Å). Microscopic observations of the microstructure and elemental microanalysis to study the elemental distribution of tungsten, zirconium and boron in the investigated samples were performed using a JSM6010PLUS/LV (JEOL, Tokyo, Japan) scanning electron microscope (SEM) equipped with an energy-dispersive X-ray spectroscope (EDS). Accelerating voltage 10 kV was used. For sintered compacts, Vickers hardness measurements were performed using Wilson VH1102 microhardness tester (Buehler, Lake Bluff, IL, USA). Load of 200 gf was used to measure hardness of targets and loads of 10, 25, 50 g were used to measure hardness of deposited coatings; 10 indentations on each sample were performed. For lower loads, a laser confocal microscope VK-X100 (Keyence, Osaka, Japan) was used to measure the indents. Vickers hardness was recalculated to account the influence of substrate by using Equation (1) [20].
H c = H s + [ 2 C t D C 2 ( t D ) 2 ] ( H f H s )
where: Hc is the measured Vickers hardness of deposited composite bilayer, Hs is the hardness of silicon substrate, C = 2sin2(f/2), f is the angle of the tip sides with the surface (f = 22° for Vickers intender), t is the film thickness, D is the indentation depth (basing on indent diameter and ideal geometry of indenter) and Hf is the hardness of the film. Palmqvist method [21] was used to analyze changes in fracture toughness (K1c). For each coating, based on a group of 10 indentations, the value of crack l, from the corner of the indent to end of crack, was determined. Equation (2) was used to evaluate the fracture toughness of deposited coatings.
K I C = 0.0028 × H V × P T
where: K1c is fracture toughness (MPa m1/2), HV is Vickers hardness in N/mm2 (or MPa), P is the load (N) and T is total crack length (mm) measured from each corner of the indent (4 × l). The electrical conductivity was measured according to the ASTM E1004 standard using a Sigma Check (Ether NDE, St Albans, UK) eddy current conductivity meter.

3. Results and Discussion

3.1. Targets

3.1.1. Density

The densification behavior of the powder mixtures during the SPS process (see Figure 1) was analyzed based on the upper punch displacement registered with resolution of 0.01 mm during the compacting (I), heating (II), holding (III) and cooling (IV) stage. Regardless of the dopant, the profile of the sintering curves is similar. In the stage (I), when the powder mixtures are compacted, this punch moved by ≈1.6–1.7 mm in positive (+) direction. At the end of the compacting stage, when the compacting pressure of 50 MPa was reached, the powder mixtures were heated from the ambient temperature to 1800 °C with heating rate of 400 °C/min (stage (II)). At the beginning of this stage, due to the thermal expansion of the graphite tool setup, punch displacement in negative (−) direction is clearly seen. As the temperature increases, this effect is compensated by the further densification of the powder mixtures. When sintering temperature of 1800 °C is reached, further densification is also clearly seen in the upper punch displacement in positive direction. In Figure 1 it is clearly seen that sintering for 2.5 min is not enough to complete the densification process and sintered compacts with open porosity were obtained. The minimum holding time required to obtained near full dense sintered compacts is 8 min and a couple of seconds before this time the plateau of the sintering curves is clearly seen. This means that sintering is completed under certain conditions. Nevertheless, a further increase in the holding time (up to 15.5 min) resulted in further upper punch movement. This is related to the reactive sintering and creation of diborides from pure elements resulted in an increase in the density of the sintered body and finished between 14 and 15 min without any further densification (plateau of curves). Increasing the holding time to 30 min does not cause significant changes in densification of the samples. The dependence of density on composition change is presented in Figure 2.
It is clearly seen from Figure 2 that as the amount of dopant increased, the density decreased and this is related to the lower atomic mass of dopant and also phase transformation occurring during the holding time from pure elements to (W,TM)B2 phase, whose density is lower than the WB2 phase. In turn, the density of the W0.92Cr0.08B2.5 increases in comparison to WB2.5. This may be connected with creation of solid solution of WB2 with Cr where chromium takes not occupied by tungsten atoms positions in hexagonal structure of WB2. For higher concentrations, when the solubility limit is reached, the CrB2 phase begins to precipitate. Chromium diboride is much softer than the WB2 phase and has a much lower density (5.21 and 12.76 g/cm3 [22] respectively).

3.1.2. Microstructure

Figure 3 shows SEM micrographs and chemical analysis of (W,TM)B2.5 sintered compacts surface alloyed with 8 at.% and 24 at.% of TM, where TM: Cr (Figure 3a), Mo (Figure 3b), Re (Figure 3c) and Zr (Figure 3d). In the left figures results of 8 at.% of addition were shown, at right—24 at.% respectively. In the middle position of Figure 3 an EDS analysis of tungsten and TM content in 24 at.% samples is presented. The rectangles with elements symbols point to the zones with maximal content of element. As it can be seen, the micrographs were done with different magnification. It is connected with different microstructure of compacts. In the case of chromium (Figure 3a), the samples are the most homogenous. For 8 at.% Cr the places with a predominance of chromium are not observed. The solubility limit is not reached and the chromium atoms are located in the crystal lattice of WB2 which is supported by XRD spectra of this sample (Figure 4a). The increase of chromium content causes the CrB2 to appear and can be observed in micrographs in the form of grains with irregular shapes. Similar results are observed in the case of molybdenum (Figure 3b) where grains of MoB2 are observed. Addition of rhenium caused ReB2 to be observed also for lower content of this element. The zones with rhenium diboride (ReB2) are greater than for Cr and Mo. However, it is hard to indicate zones with a predominance of Re using SEM with backscattered electron (BSE) mode and EDS because the atomic mases of tungsten and rhenium are similar, 183.85 and 186.20 u respectively, and the contrast between both phases is very low. Zirconium is two times lighter and it is easy to see that it possesses the greatest grains among studied elements. It can be explained by the size of the used powders. The dimension of zirconium powders is about ten times greater than the size of tungsten powders. In this case the zirconium diboride is created mainly on the grain boundary (Figure 3d).
The diborides formation can be proved by using the XRD technique. Figure 4a shows the XRD patterns of Wx−1CrxB2 compound. These spectra show that the solubility of Cr in hexagonal P63/mmc WB2 (the upper spectra) is greater than 16 at.% and above CrB2 appears as a second phase. An atomic radius of chromium is lower than the other used TM metals, and in agreement with this fact, the solubility of chromium in WB2 appears to be the highest. The XRD spectra of Mo, Re and Zr added compounds are shown in Figure 4b–d respectively. The solubility of these elements in WB2 is below 8 at.%. At and above 8 at.% TM addition, trigonal (R3m) MoB2, hexagonal P63/mmc ReB2 and hexagonal P6/mmm ZrB2 starts to show up as an additional phase. The lattice parameters of the WB2 solid solutions are a = 2.985 and c = 13.900 Å, respectively. It can be noted that the addition of small amounts of Cr, Mo, Re and Zr does not change the lattice parameters of WB2. Only a small shift of the strongest peak (0 0 0 4) to higher angles is observed (0.08°, Figure 3) and may result from the stresses caused by the filling vacancies by atoms with different radii.
The microstructure of obtained compacts strongly depends on crystal structure of synthetized borides and also diameter of used powders. The structure of tungsten diboride allows it to accommodate metal atoms in vacancies in addition to ordinary substitution of metal (tungsten) positions, which enables alloys with other transition metals (Zr, Mo and Cr) to have enhanced mechanical properties (hardness), oxidation resistance and electrical conductivity [5]. Therefore, in addition to accommodating the metals in the voids inside the structure (as well as substitute in tungsten positions in the case of W1−xMoxB2.5) WB2 can also form a traditional solid solution, and therefore, the Hume-Rothery rules (atomic radii difference, similarity of crystal structure, oxidation states and electronegativity) are fully applicable to these cases [23]. Such situation can be observed for Mo, Cr and Zr where the most stable hexagonal structure P6/mmm is created during reactive sintering. For ReB2, the phase P63/mmc is formed at all concentrations of rhenium. In this case, rhenium forms a very favorable, stable phase and does not readily alloy with WB2. In result the spectra of this phase have higher intensity than other dopants (Figure 4c). In the case of zirconium, the diameter of used powders is crucial. The diameter of used zirconium powder is above ten times higher than tungsten. During sintering only surface of zirconium particles was reacted with boron and zirconium diboride was created (Figure 3d). Similar observation was done by Garbiec et al. [9]. As it will be shown in the next paragraphs, it highly influences hardness and electrical conductivity of compacts.

3.1.3. Hardness

In Figure 5 the results of Vickers hardness are presented. The increase of hardness can be observed with an increase in the Cr, Mo, Re and Zr content. At low contents (below 8 at.%), by sitting at the positions of missing tungsten atoms, these elements could induce either tungsten vacancy at the WB2 structure. It may cause the defect structure of material and may be responsible for the hardening trends observed for the solid solutions. For higher concentrations, XRD results (Figure 3) indicate the possibility of grain boundary strengthening mechanisms outside the solubility limits of three elements (Mo, Re and Zr). The greatest increase of hardness is observed for Re. It can be explained by the greatest hardness of ReB2 (HV = 30.40–48.0 GPa [12,22]) among all diborides. With the increase of content of rhenium, the amount of rhenium diboride grows and can cause dispersion hardening. Due to the fact that the differences in the microhardness of ZrB2 [24] or MoB2 [10] and WB2 [22] are insignificant, at lower contents, i.e., 0 and 16 at.% the hardness practically does not change and is about 23 ± 1 GPa. In the case of zirconium, according to the phase diagram of the W2B5–ZrB2 pseudobinary system [25], a further increase in the amount of dopant results in achieving the composition with the lowest melting point (2180 ± 30 °C for the eutectic composition of 80 mol% W2B5 + 20 mol% ZrB2). This phenomenon promotes greater consolidation of the material and thus also an increase in hardness. Moreover, on the basis of the analysis of the microstructure of the samples with the higher zirconium content (Figure 3d), both the unbound boron and ZrB2 grains are smaller and more evenly distributed, which promotes dispersion strengthening. A similar relationship was observed by Hirota et al. [26] when ZrB2 was doped with small amounts of tungsten. For monolithic Zr1−xWxB2 (0 < x ≤ 0.12), W-doped ZrB2 solid solutions with a lower hardness were formed and the hardness grew and stabilized for composites consisting of ZrB2 solid solutions and WB2 when x > 0.12 due to the formation of solid solution, homogeneously dispersed needle-like WB2 particles [26]. An inverse relationship for hardness can be observed for the higher chromium content, where the hardness decreases with an increase in the chromium content. In this case chromium diboride is the softest material among investigated borides and increase of amount of CrB2 causes the decrease in hardness.

3.1.4. Electrical Conductivity

The electrical conductivity of the WxTMx−1B2.5, (0 < x ≤ 0.24 TM = Cr, Mo, Re, Zr) samples is shown in Figure 6. It is clearly seen that the electrical conductivity of the samples with zirconium content is more or less two times higher than the samples without admixture and increases at ambient temperature with an increasing zirconium content up to 3.961 × 106 S/m. In the case of molybdenum and rhenium the electrical conductivity is almost constant with increasing content of dopant. This is a result of similar conductivity of MoB2 and ReB2 to WB2 and better electrical conductivity of zirconium diboride, respectively. Exemplary values of electrical conductivities for clean diborides are presented in Table 3.
Furthermore, the increased content of chromium decreased the electrical conductivity. For these materials, the addition of chromium reduces the electrical conductivity of the sintered compacts from 1.393 × 106 to 0.954 × 106 S/m. It should be noted that the electrical conductivity of WB2.5 alloyed with Mo or Re (≈2.5 × 106 S/m) is greater than the electrical conductivity of 304 stainless steel (1.45 × 106 S/m) and in the case of 24 at.% zirconium is similar to the electrical conductivity of WC-Co cemented carbides (4.76 × 106 S/m) [31].

3.2. Coatings

Targets with an admixture of 24 at.%. TM were selected to show the differences in the properties of the deposited layers. After 2 h of deposition all the coatings were smooth with a roughness below 22 ± 5 nm. No delamination was observed. An exemplary SEM micrograph for rhenium is presented in Figure 7a. Thickness of coatings differ from each other (Table 2). The greatest thickness is found in films with rhenium addition (3.18 μm) and the thickest coatings are for chromium (1.67 μm). As can be easily seen, the thickness of the layers decreases with molecular weight of the dopant. Rhenium has the greatest mass (186 u), comparable to tungsten (184 u), and chromium the lowest (52 u). It can influence dynamics of plasma during magnetron sputtering. The SEM-EDS analysis indicates that chemical elements of WxTM1−xBy coatings were homogeneously distributed (see Figure 7b–e).
The quantitative analysis of chemical composition showed that stoichiometry of target is not preserved (Table 2). In all cases the TM and boron content dropped in the comparison to target composition. The decrease in TM content, as in the case of boron, is predicted to be due to scattering of the TM atoms on the tungsten atoms in the plasma plume during deposition. In addition, resputtering of atoms from surface deposited coatings is also possible. A decrease in light elements was already observed in References [18,32]. The boron content B/(Zr + W) is in the range of 1.54–2.06 (see Table 2) which means boron decrease of 38.4–17.7 % respectively compared to the target composition. In the case of transition metals losses are much smaller, which can be explained by their greater mass in relation to boron. In all coatings the presence of oxygen was detected. The lower content of oxygen was in films with chromium admixture (2.16 at.%) and the higher was for rhenium (5.24 at.%) respectively (Table 4).
In Figure 8, XRD spectra of phase composition in deposited coatings is presented. Only spectra were recorded in the range of 24°–31°. XRD spectra of phase composition corresponds to the phase WBy [18]. The comparison of XRD pattern of targets (Figure 4) and deposited coatings indicates the change of crystal structure of WB2 from softer ω-WB2 (P63/mmc) to harder α-WB2 (P6/mmm). Regarding the synthesis via physical vapor deposition, it is a well-known fact that metastable structures can be captured via this synthesis route [5]. Admixture of transition metals does not cause of appearance of new phases of TM borides. XRD spectra of W-TM-By do not differ significantly from WBy. However, careful analysis of the plot shows that depending on the used element, the peak positions are shifted. The black arrows in Figure 8 denote the direction of shift of (0 0 0 1) peak position. Only alloying with zirconium is significant and changes the direction of sift. Comparing the radii of the used elements with radii of tungsten (1.394 Å [33]), it can be observed that zirconium has a much larger radius than the other elements (Zr—1.597 Å; Cr—1.357 Å; Mo—1.386 Å; Re—1.373 Å [33]). The transition metal atoms take the places of tungsten in the crystal lattice, thus creating a solid solution with WB2. Since zirconium possesses the radius larger than tungsten only, this makes the unit cell larger and causes the shift to lower angles. Detailed deconvolution of XRD spectra leads to the determination of cell parameters. The WB2 phase has lattice parameters a = 2.96 Å and c = 3.089 Å and addition of zirconium increases them to a = 3.028 Å and c = 3.192 Å respectively.
All deposited coatings are very hard and the results of hardness measurement under load 10, 25 and 50 g are presented in Table 5. Film alloyed with Cr possesses the greatest hardness; after recalculation that takes into account coatings’ thickness (Equation (1)) and conversion to SI unit, it is 50.4 ± 4.7 GPa (at 98 mN). In the case of Mo, Zr and Re the hardness is lower and is 45.4 ± 6.7, 43.9 ± 3.3 and 32.6 ± 6.4 GPa respectively. These values are lower than hardness of unalloyed WB2 presented in literature (H = 46 ± 2 GPa [18]). However, it should be noted that hardness results presented by Psiuk et al. [18] were measured by nanoindentation method under load of 3 mN. The lowest hardness, when rhenium is added, can be explained by the degree of crystallization conducted from XRD spectra. In this case the intensity of (0 0 0 1) peak (Figure 8) is also the lowest, which suggests that coatings are more amorphous. Additionally, decrease in boron content and creation of vacancies can influence this mechanical property.
Deposited coatings are characterized by relatively high fracture toughness (K1c). Because of difficulties with measurements of indents’ dimension under load 10 g for calculation K1c, the load 50 g was selected. Exemplary micrographs of indents for zirconium and molybdenum are shown in Figure 9. As can be seen in Figure 9b, the molybdenum alloyed layers cracked during indentation. This proves low adhesion of this coating to the substrate. In the other cases the dimensions of indents are comparable but the length of the cracks are different. The shortest cracks were observed for coatings with chromium and the longest for rhenium. Based on the common Palmqvist method [21] the values of fracture toughness for coatings alloyed with zirconium, chromium and rhenium are 2.11, 1.40 and 1.77 MPa m1/2 respectively. Obtained results of K1c are similar to values of fracture toughness commonly used as protective coatings titanium nitrides [11]. It should be noted that in Palmqvist method K1c is related to hardness (Equation (2)) and is influenced by relatively soft silicon substrate. Additionally, fracture toughness of deposited coatings might be influenced by brittle silicon substrate (<1 MPa m1/2).

4. Conclusions

To improve the properties of tungsten diboride, solid solutions of this material with chromium, molybdenum, rhenium and zirconium in the form of spark plasma sintered compacts and magnetron sputtered coatings were synthesized and characterized. Various concentrations of transition metal elements, ranging from 0.0 to 24.0 at.%, on a metals basis, were made. Spark plasma sintering was used to synthesize these refractory compounds from the pure elements. Elemental and phase purity of the both samples (sintered compacts and coatings) were examined using energy-dispersive X-ray spectroscopy and X-ray diffraction, and microindentation was utilized to measure the Vickers hardness under applied load of 2 N for compacts and 0.1 N for coatings respectively. The main observations are:
  • X-ray diffraction results of SPS-ed compacts indicate that the solubility limit is below 8 at.% for molybdenum, rhenium and zirconium and below 16 at.% for chromium. Above this limit, both diborides (W,TM)B2 are created.
  • The addition of transition metal caused decrease of density and increase of hardness and electrical conductivity of sintered compacts.
  • Deposited coatings W1−xTMxBy (TM = Cr, Mo, Re, Zr; x = ~0.2; y = 1.7–2.0) are homogenous, smooth and hard. The maximal hardness was measured for W-Cr-B2 coatings and under the load 10 g was 50.4 ± 4.7 GPa. Deposited films possess relatively high fracture toughness and for WB2 coatings alloyed with zirconium it was K1c = 2.11 MPa m1/2. The connection of very high hardness with high fracture toughness places these coatings in a group of novel materials with improved mechanical properties.
Due to the fact that the proposed materials possess reduced density and at the same time electrical and mechanical properties comparable or better than tungsten carbide, the presented research suggests the possibility of their implementation in everyday life. The simple and fast manufacturing process of the proposed materials should also be emphasized.

Author Contributions

Conceptualization, T.M. and R.P.; data analysis, T.M. and R.P.; investigation, R.P., J.R. and M.W.; writing—original draft, T.M.; writing—review and editing, T.M., R.P., J.R. and D.G.; project administration, T.M. and D.G.; funding acquisition, T.M. and D.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work was co-financed by the National Science Centre (NCN, Poland) under project No: UMO-2017/25/B/ST8/01789 and supported by the National Centre for Research and Development (NCBR, Poland) under project No. TECHMATSTRATEGIII/0017/2019.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gu, X.; Liu, C.; Guo, H.; Zhang, K.; Chen, C. Sorting transition-metal diborides: New descriptor for mechanical properties. Acta Mater. 2021, 207, 116685. [Google Scholar] [CrossRef]
  2. Ivanovskii, A.L. The search for novel superhard and incompressible materials on the basis of higher borides of s, p, d metals. J. Superhard Mater. 2011, 33, 73–87. [Google Scholar] [CrossRef]
  3. Yeung, M.T.; Mohammadi, R.; Kaner, R.B. Ultraincompressible, superhard materials. Annu. Rev. Mater. Res. 2016, 46, 465–485. [Google Scholar] [CrossRef]
  4. Wuchina, E.; Opila, E.; Opeka, M.; Fahrenholtz, B.; Talmy, I. UHTCs: Ultra-high temperature ceramic materials for extreme environment applications. Electrochem. Soc. Interface 2007, 16, 30–36. [Google Scholar] [CrossRef]
  5. Moraes, V.; Riedl, H.; Fuger, C.; Polcik, P.; Bolvardi, H.; Holec, D.; Mayrhofer, P.H. Ab initio inspired design of ternary boride thin films. Sci. Rep. 2018, 8, 9288. [Google Scholar] [CrossRef] [PubMed]
  6. Moscicki, T.; Psiuk, R.; Słomińska, H.; Levintant-Zayonts, N.; Garbiec, D.; Pisarek, M.; Bazarnik, P.; Nosewicz, S.; Chrzanowska-Giżyńska, J. Influence of overstoichiometric boron and titanium addition on the properties of RF magnetron sputtered tungsten borides. Surf. Coat. Technol. 2020, 390, 125689. [Google Scholar] [CrossRef]
  7. Wicher, B.; Chodun, R.; Trzciński, M.; Lachowski, A.; Kubiś, M.; Nowakowska-Langier, K.; Zdunek, K. Design of pulsed neon injection in the synthesis of W-B-C films using magnetron sputtering from a surface-sintered single powder cathode. Thin Solid Film. 2020, 716, 138426. [Google Scholar] [CrossRef]
  8. Radziejewska, J.; Psiuk, R.; Mościcki, T. Characterization and wear response of magnetron sputtered W–B and W–Ti–B coatings on WC–Co tools. Coatings 2020, 10, 1231. [Google Scholar] [CrossRef]
  9. Garbiec, D.; Wiśniewska, M.; Psiuk, R.; Denis, P.; Levintant-Zayonts, N.; Leshchynsky, V.; Rubach, R.; Mościcki, T. Zirconium alloyed tungsten borides synthesized by spark plasma sintering. Arch. Civ. Mech. Eng. 2021, 21, 37. [Google Scholar] [CrossRef]
  10. Shaw, A.H. Physical Properties of Various Conductive Diborides and Their Binaries. Graduate Theses and Dissertations, Iowa State University, USA, 2015; p. 14496. Available online: https://lib.dr.iastate.edu/etd/14496 (accessed on 4 November 2021).
  11. Fuger, C.; Moraes, V.; Hahn, R.; Bolvardi, H.; Polcik, P.; Riedl, H.; Mayrhofer, P.H. Influence of Tantalum on phase stability and mechanical properties of WB2. MRS Commun. 2019, 9, 375–380. [Google Scholar] [CrossRef]
  12. Maździarz, M.; Mościcki, T. Structural, mechanical, optical, thermodynamical and phonon properties of stable ReB2 polymorphs from density functional calculations. J. Alloys Compd. 2016, 657, 878–888. [Google Scholar] [CrossRef]
  13. Mohammadi, R.; Lech, A.T.; Xie, M.; Weaver, B.E.; Yeung, M.; Tolbert, S.H.; Kaner, R.B. Tungsten tetraboride, an inexpensive superhard material. Proc. Natl. Acad. Sci. USA 2011, 108, 10958–10962. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Feng, S.; Li, X.; Su, L.; Li, H.; Yang, H.; Cheng, X. Ab initio study on structural, electronic properties, and hardness of re-doped W2B5. Solid State Commun. 2016, 245, 60–64. [Google Scholar] [CrossRef]
  15. Lech, A.T.; Turner, C.L.; Lei, J.; Mohammadi, R.; Tolbert, S.H.; Kaner, R.B. Superhard rhenium/tungsten diboride solid solutions. J. Am. Chem. Soc. 2016, 138, 14398–14408. [Google Scholar] [CrossRef] [PubMed]
  16. Tu, Y.; Wang, Y. First-principles study of the elastic properties of OsxW1−xB2 and RexW1−xB2 alloys. Solid State Commun. 2011, 151, 238–241. [Google Scholar] [CrossRef]
  17. Smolik, J.; Kacprzyńska-Gołacka, J.; Sowa, S.; Piasek, A. The analysis of resistance to brittle cracking of tungsten doped tib2 coatings obtained by magnetron sputtering. Coatings 2020, 10, 807. [Google Scholar] [CrossRef]
  18. Psiuk, R.; Milczarek, M.; Jenczyk, P.; Denis, P.; Jarząbek, D.M.; Bazarnik, P.; Pisarek, M.; Mościcki, T. Improved mechanical properties of W-Zr-B coatings deposited by hybrid RF magnetron—PLD method. Appl. Surf. Sci. 2021, 570, 151239. [Google Scholar] [CrossRef]
  19. Chrzanowska, J.; Kurpaska, Ł.; Giżyński, M.; Hoffman, J.; Szymański, Z.; Mościcki, T. Fabrication and characterization of superhard tungsten boride layers deposited by radio frequency magnetron sputtering. Ceram. Int. 2016, 42, 12221–12230. [Google Scholar] [CrossRef]
  20. Wang, H.L.; Chiang, M.J.; Hon, M.H. Determination of thin film hardness for a film/substrate system. Ceram. Int. 2001, 27, 385–389. [Google Scholar] [CrossRef]
  21. Palmqvist, S. Occurrence of crack formation during Vickers indentation as a measure of the toughness of hard materials. Arch. Eisenhuettenwes 1962, 33, 629–633. [Google Scholar]
  22. Akopov, G.; Yeung, M.; Kaner, R.B. Rediscovering the crystal chemistry of borides. Adv. Mater. 2017, 29, 1604506. [Google Scholar] [CrossRef] [PubMed]
  23. Kittel, C. Introduction to Solid State Physics, 8th ed.; Johnson, S., Ed.; John Wiley & Sons: Hoboken, NJ, USA, 2005. [Google Scholar]
  24. Maździarz, M.; Mościcki, T. New zirconium diboride polymorphs—First-principles calculations. Materials 2020, 13, 3022. [Google Scholar] [CrossRef] [PubMed]
  25. Ordan’Yan, S.S.; Boldin, A.A.; Suvorov, S.S.; Smirnov, V.V. Phase diagram of the W2B5-ZrB2 system. Inorg. Mater. 2005, 41, 232–234. [Google Scholar] [CrossRef]
  26. Ken, H.; Endo, T.; Masaki, K.; Nakane, S.; Nishimura, T.; Morisada, Y.; Mizuuchi, K. Simultaneous synthesis and consolidation of W-added ZrB2 by pulsed electric current pressure sintering and their mechanical properties. Mater. Sci. Forum 2007, 561–565, 527–530. [Google Scholar] [CrossRef]
  27. Wang, C.; Song, L.; Xie, Y. Mechanical and electrical characteristics of WB2 synthesized at high pressure and high temperature. Materials 2020, 13, 1212. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Bauer, A.; Regnat, A.; Blum, C.G.F.; Gottlieb-Schönmeyer, S.; Pedersen, B.; Meven, M.; Wurmehl, S.; Kunes, J.; Pfleiderer, C. Low-temperature properties of single-crystalCrB2. Phys. Rev. B 2014, 90, 064414. [Google Scholar] [CrossRef] [Green Version]
  29. Levine, J.B. Synthesis and Characterization of Ultra-Incompressible Superhard Borides; Proquest, Umi Dissertation Publishing: Ann Arbor, MI, USA, 2008. [Google Scholar]
  30. Rahman, M.; Wang, C.C.; Chen, W.; Akbar, S.A.; Mroz, C. Electrical resistivity of titanium diboride and zirconium diboride. J. Am. Ceram. Soc. 1995, 78, 1380–1382. [Google Scholar] [CrossRef]
  31. Guimarães, B.; Fernandes, C.; Figueiredo, D.; Cerqueira, M.F.; Carvalho, O.; Silva, F.; Miranda, G. A novel approach to reduce in-service temperature in WC-Co cutting tools. Ceram. Int. 2020, 46, 3002–3008. [Google Scholar] [CrossRef]
  32. Chrzanowska-Giżyńska, J.; Denis, P.; Giżyński, M.; Kurpaska, Ł.; Mihailescu, I.; Ristoscu, C.; Szymański, Z.; Mościcki, T. Thin WBx and WyTi1−yBx films deposited by combined magnetron sputtering and pulsed laser deposition technique. Appl. Surf. Sci. 2019, 478, 505–513. [Google Scholar] [CrossRef]
  33. Pauling, L. Atomic Radii and Interatomic Distances in Metals. J. Am. Chem. Soc. 1947, 69, 542–553. [Google Scholar] [CrossRef]
Figure 1. Sintering curves of W1−xCrxB2.5 (0 < x ≤ 0.24) samples spark plasma sintered for 24 min. Dotted lines are temporal distribution of sintering temperature.
Figure 1. Sintering curves of W1−xCrxB2.5 (0 < x ≤ 0.24) samples spark plasma sintered for 24 min. Dotted lines are temporal distribution of sintering temperature.
Coatings 11 01378 g001
Figure 2. Changes in the density of the SPSed samples in function of amount and type of admixture TM (where TM = Cr, Mo, Re, Zr).
Figure 2. Changes in the density of the SPSed samples in function of amount and type of admixture TM (where TM = Cr, Mo, Re, Zr).
Coatings 11 01378 g002
Figure 3. SEM micrographs and chemical analysis of (W,TM)B2.5 sintered compact surface alloyed with 8 at.% and 24 at.% TM, where TM: (a) Cr, (b) Mo, (c) Re, (d) Zr. Left figures show results of 8 at.% of addition, right 24 at.% respectively and middle is an EDS analysis of W and TM content in 24 at.% samples. Only one element is presented in each picture. Black colour means 0% at. and pink 100% at. respectively.
Figure 3. SEM micrographs and chemical analysis of (W,TM)B2.5 sintered compact surface alloyed with 8 at.% and 24 at.% TM, where TM: (a) Cr, (b) Mo, (c) Re, (d) Zr. Left figures show results of 8 at.% of addition, right 24 at.% respectively and middle is an EDS analysis of W and TM content in 24 at.% samples. Only one element is presented in each picture. Black colour means 0% at. and pink 100% at. respectively.
Coatings 11 01378 g003
Figure 4. XRD spectra of phase composition of samples W1−xTMxB2.5 with molar ratio of x = TM/(TM + W) where x = 0, 8, 16 and 24 at.% and TM: (a) Cr, (b) Mo, (c) Re, (d) Zr.
Figure 4. XRD spectra of phase composition of samples W1−xTMxB2.5 with molar ratio of x = TM/(TM + W) where x = 0, 8, 16 and 24 at.% and TM: (a) Cr, (b) Mo, (c) Re, (d) Zr.
Coatings 11 01378 g004
Figure 5. Hardness of WxTMx−1B2.5, (0 < x ≤ 0.24 TM = Cr, Mo, Re, Zr); specimens spark plasma sintered at holding times 24 min.
Figure 5. Hardness of WxTMx−1B2.5, (0 < x ≤ 0.24 TM = Cr, Mo, Re, Zr); specimens spark plasma sintered at holding times 24 min.
Coatings 11 01378 g005
Figure 6. Electrical conductivity of WxTMx−1B2.5, (0 < x ≤ 0.24 TM = Cr, Mo, Re, Zr); samples spark plasma sintered at holding time 24 min.
Figure 6. Electrical conductivity of WxTMx−1B2.5, (0 < x ≤ 0.24 TM = Cr, Mo, Re, Zr); samples spark plasma sintered at holding time 24 min.
Coatings 11 01378 g006
Figure 7. (a) SEM micrograph, EDS maps of elements: (b) Re, (c) W, (d) B, (e) O in the coating deposited from the W0.76Re0.24B2.5 target.
Figure 7. (a) SEM micrograph, EDS maps of elements: (b) Re, (c) W, (d) B, (e) O in the coating deposited from the W0.76Re0.24B2.5 target.
Coatings 11 01378 g007
Figure 8. XRD spectra of phase composition of WBy [18] and W-TM-By coatings deposited from targets alloyed with 24 at.% TM (TM = Cr, Mo, Re, Zr). Black arrows denote direction of shift of (0 0 0 1) peak position.
Figure 8. XRD spectra of phase composition of WBy [18] and W-TM-By coatings deposited from targets alloyed with 24 at.% TM (TM = Cr, Mo, Re, Zr). Black arrows denote direction of shift of (0 0 0 1) peak position.
Coatings 11 01378 g008
Figure 9. Confocal microscope micrographs of indented (HV0.025) surface of coatings. (a) W-Cr-B, (b) W-Mo-B, (c) W-Re-B and (d) W-Zr-B.
Figure 9. Confocal microscope micrographs of indented (HV0.025) surface of coatings. (a) W-Cr-B, (b) W-Mo-B, (c) W-Re-B and (d) W-Zr-B.
Coatings 11 01378 g009
Table 1. Weight content of tungsten, transition metals (chromium, molybdenum, rhenium, zirconium) and boron in W-TM-B powder mixtures.
Table 1. Weight content of tungsten, transition metals (chromium, molybdenum, rhenium, zirconium) and boron in W-TM-B powder mixtures.
Material CompositionTungsten, (g)TM, (g)Boron, (g)
WB2.512.202-1.795
W0.92Cr0.08B2.511.8170.2911.890
W0.84Cr0.16B2.511.3890.6141.995
W0.76Cr0.24B2.510.9110.9752.112
W0.92Mo0.08B2.511.6130.5271.857
W0.84Mo0.16B2.510.9821.0921.924
W0.76Mo0.24B2.510.3051.6981.995
W0.92Re0.08B2.511.2160.9881.794
W0.84Re0.16B2.510.2321.9741.792
W0.76Re0.24B2.59.2492.9581.791
W0.92Zr0.08B2.511.6350.5021.861
W0.84Zr0.16B2.511.0251.0421.931
W0.76Zr0.24B2.510.3661.6242.007
Table 2. Spark plasma sintering process parameters of W/B and W/Zr/B powder mixtures.
Table 2. Spark plasma sintering process parameters of W/B and W/Zr/B powder mixtures.
Material CompositionSintering Temperature, °CHeating Rate, °C/minHolding Time, minCompacting Pressure, MPa
W1−xTMxB2.518004002450
Table 3. Electrical conductivity of TM (TM = W, Cr, Mo, Re, Zr) diborides and tungsten carbide.
Table 3. Electrical conductivity of TM (TM = W, Cr, Mo, Re, Zr) diborides and tungsten carbide.
BorideWB2CrB2MoB2ReB2ZrB2WC-Co
Electrical conductivity (S/m)1.1–4.5 × 106 [27]0.95–1.42 × 106 [28]7.8 × 106 [10]2.45 × 106 [29]10.9 × 106 [30]4.76 × 106 [31]
Table 4. Chemical composition and thickness of deposited coatings W1−xTMxBy alloyed with TM (TM = Cr, Mo, Re, Zr).
Table 4. Chemical composition and thickness of deposited coatings W1−xTMxBy alloyed with TM (TM = Cr, Mo, Re, Zr).
TMX = TM/(TM + W)y = B/(TM + W)at.% O (%)Thickness (μm)
Cr0.2131.6982.162.31
Mo0.2172.0573.702.34
Re0.1851.5405.243.18
Zr0.2021.9193.422.25
Table 5. Measured and recalculated hardness (Equation (1)) of deposited coatings.
Table 5. Measured and recalculated hardness (Equation (1)) of deposited coatings.
TMHV0.01HV0.01 (Recalculated)HV0.025HV0.025 (Recalculated)HV0.05HV0.05 (Recalculated)
Cr3610 ± 3905140 ± 4802760 ± 1905570 ± 3902140 ± 1205490 ± 380
Mo3250 ± 5004630 ± 6802330 ± 1204610 ± 2702170 ± 2105540 ± 660
Zr3050 ± 2404480 ± 3302690 ± 1305540 ± 2602010 ± 1105150 ± 370
Re2710 ± 5403330 ± 6501940 ± 2103070 ± 4001750 ± 603360 ± 180
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mościcki, T.; Psiuk, R.; Radziejewska, J.; Wiśniewska, M.; Garbiec, D. Properties of Spark Plasma Sintered Compacts and Magnetron Sputtered Coatings Made from Cr, Mo, Re and Zr Alloyed Tungsten Diboride. Coatings 2021, 11, 1378. https://doi.org/10.3390/coatings11111378

AMA Style

Mościcki T, Psiuk R, Radziejewska J, Wiśniewska M, Garbiec D. Properties of Spark Plasma Sintered Compacts and Magnetron Sputtered Coatings Made from Cr, Mo, Re and Zr Alloyed Tungsten Diboride. Coatings. 2021; 11(11):1378. https://doi.org/10.3390/coatings11111378

Chicago/Turabian Style

Mościcki, Tomasz, Rafał Psiuk, Joanna Radziejewska, Maria Wiśniewska, and Dariusz Garbiec. 2021. "Properties of Spark Plasma Sintered Compacts and Magnetron Sputtered Coatings Made from Cr, Mo, Re and Zr Alloyed Tungsten Diboride" Coatings 11, no. 11: 1378. https://doi.org/10.3390/coatings11111378

APA Style

Mościcki, T., Psiuk, R., Radziejewska, J., Wiśniewska, M., & Garbiec, D. (2021). Properties of Spark Plasma Sintered Compacts and Magnetron Sputtered Coatings Made from Cr, Mo, Re and Zr Alloyed Tungsten Diboride. Coatings, 11(11), 1378. https://doi.org/10.3390/coatings11111378

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