Next Article in Journal
Physiological Functions of Carbon Dots and Their Applications in Agriculture: A Review
Next Article in Special Issue
Pillar Growth by Focused Electron Beam-Induced Deposition Using a Bimetallic Precursor as Model System: High-Energy Fragmentation vs. Low-Energy Decomposition
Previous Article in Journal
Constructing FeS and ZnS Heterojunction on N,S-Codoped Carbon as Robust Electrocatalyst toward Oxygen Reduction Reaction
Previous Article in Special Issue
Self-Organized Memristive Ensembles of Nanoparticles Below the Percolation Threshold: Switching Dynamics and Phase Field Description
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sintering, Mechanical and Optical Properties of TiB2 Composites with and without High-Energy Milling

1
Institute of Science, Technology and Sustainability for Ceramic, National Research Council of Italy, Via Granarolo 64, I-48018 Faenza, Italy
2
Department of Chemistry, Life Sciences and Environmental Sustainability, University of Parma, Parco Area delle Scienze 11/a, I-43124 Parma, Italy
3
National Institute of Optics, National Research Council of Italy, Largo E. Fermi, 6, I-50125 Firenze, Italy
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(19), 2683; https://doi.org/10.3390/nano13192683
Submission received: 8 September 2023 / Revised: 26 September 2023 / Accepted: 28 September 2023 / Published: 30 September 2023

Abstract

:
TiB2 is a promising material for several fields including impact-resistant armor, wear-resistant coatings, cutting tools and crucibles given its physical, mechanical and chemical properties, especially due to the combination of high hardness and exceptional wear resistance. It is however very difficult to sinter below 2000 °C, even under mechanical pressure; moreover, the low fracture toughness limits the applicability of the ceramic material. By using sintering additives, it is possible to improve the sintering process and increase the mechanical properties since the additives react with oxidized layers and form secondary phases. In this study, different preparation methods and various combinations of additives (B4C, Si3N4 and MoSi2) via hot pressing sintering have been explored. Through the synergy between optimized process and tailored composition, an almost fully dense material was obtained at 1700 °C with hardness of 24.4 ± 0.2 GPa and fracture toughness of 5.4 ± 0.2 MPa m1/2. However, the highest hardness (24.5 ± 0.2 GPa) and density values were obtained for only the high-energy-milled sample with WC-Co media, featuring a core–shell grain structure. Finally, optical properties for selected samples were measured, identifying the high-energy-milled TiB2 as the sample with the highest spectral selectivity α/ε and solar absorptance.

1. Introduction

Among non-oxide ceramics, TiB2 [1,2,3,4,5,6,7,8,9,10] is particularly studied for several reasons. Firstly, it is a hard and stiff ceramic. Moreover, it possesses a very high melting point (3225 °C) [11] and exhibits a level of electrical conductivity similar to that of metals [12]. This feature can be effectively utilized in electro-discharge machining, which is a more cost-effective technique compared to conventional diamond tool machining for shaping ceramic components.
Despite the numerous advantages of using TiB2, achieving a fully dense material is highly challenging. This is primarily due to the low self-diffusion coefficient and relatively high vapor pressure of its constituents, [13] resulting from the presence of both ionic and covalent bonds [14]. Furthermore, the presence of an oxide layer (TiO2 and B2O3) on the surface of TiB2 powder promotes excessive grain growth through vapor or surface transport mechanisms during the sintering process, leading to the formation of porosity that becomes trapped within the grains [15,16].
For all the previously addressed issues, among the possible densification routes, pressureless sintering (PLS) of TiB2 seems to be an unattractive choice, since it requires relatively high temperatures (2000–2200 °C) and many hours of dwell time to achieve a good densification. This can lead to exaggerated grain growth and consequently to lower mechanical properties due to the anisotropy of the hexagonal grain structure that results in deleterious internal stresses [6,15]. Grain growth can also cause microcracking when average grain size exceeds a critical value, which experimental observations suggest to be ~15 µm [14,17,18,19,20].
Densification by hot pressing (HP) can lower the sintering temperature (1800–2000 °C), increase the sintering rate and achieve a uniform microstructure of sintered materials [13,21,22,23,24]. Spark plasma sintering (SPS) is an alternative strategy where the combination of applied external pressure and current with a very high intensity leads to rapid heating, lower working temperature and shorter dwell time. The advantages of the HP and SPS techniques are a higher density and finer final microstructure. However, both processes are limited by a simple sample geometry [10,25,26,27,28,29,30].
One more way to improve the sintering of TiB2 is through addition of sintering aids, that can be either metallic or non-metallic. Metallic additives like Ti [31], Co [32], Cr [33], Ni [34], Fe [35], Al [36] and Ta [37] may lead to high relative density through liquid phase sintering (LPS); indeed, these transition metals can react with TiB2 and B2O3 to form various metal borides with a low melting point and a good wetting behavior. Although the liquid phase increases mass transport and decreases the sintering temperature (1500–1700 °C), accelerated grain growth is observed [38,39,40]. Moreover, metallic additives have been found to reduce the corrosion resistance, wear properties and high temperature properties of TiB2 [14,26,40,41].
Therefore, non-metallic additives such as carbides (B4C [42], SiC [43], TiC [44], TaC [45], WC [46]), nitrides (Si3N4 [9], AlN [10,47], TiN [29,43]), disilicides (TiSi2 [48], MoSi2 [7,49,50,51]) and borides (ZrB2 [52]) are preferred for applications at elevated temperatures.
These additives react with the oxide surface layer of TiB2 particles and form secondary phases, improving sinterability without promoting grain growth and possibly leading to lower sintering temperatures (1700–2000 °C) and better mechanical properties [26]. On the other hand, the production of secondary phases may result in a significant reduction in purity of the TiB2, thus, leading to a decrease in electric conductivity and thermal conductivity [25,47].
The purpose of this work is to optimize the densification of TiB2 by hot pressing and introducing small amounts of B4C, Si3N4 or MoSi2 as a sintering aid. In addition, the effect of high-energy milling with WC milling beads is also considered, in combination or not with these sintering aids. Densification mechanisms and microstructural characterization are carefully considered. In addition, hardness and toughness measurements were carried out. The optical properties of TiB2 ceramics were also investigated in order to explore their possible application as solar absorbers. In fact, a single previous study on the optical properties of TiB2 bulks can be found [53], which reported a higher solar absorptance with respect to other monophasic borides (ZrB2 and TaB2) and a comparable spectral selectivity with respect to MoSi2-added dense ZrB2 and TaB2, definitely suggesting that TiB2 is worthy of further studies for its use in solar thermal technologies. The cited work dealt with a monophasic and fully dense TiB2 pellet. However, it is known that the material composition with the possible presence of additives and byproducts, as well as the porosity, have a critical influence on the overall optical properties of UHTC ceramics [54,55,56,57].

2. Materials and Methods

Commercial powders of TiB2 (H.C. Starck, Grade F, D90 4.0–7.0 μm, D50 2.5–3.5 μm, <4.5 μm, impurities (wt. %): 0.4 C, 2.5 O, 0.5 N, 0.1 Fe), were used to produce powder mixture. The following commercial powders were used as additives:
-
B4C (H.C. Starck Grade HS-A, D90 2.0–4.0 μm, D50 0.6–1.2 μm, B:C ratio 3.7, impurities (wt. %): 0.7 N, 1.7 O, 0.05 Fe, 0.15 Si, 0.05 Al);
-
Si3N4 (α-Si3N4, H.C. Starck Grade M 11, α > 90%, D50 0.6 μm, D90 1.3 μm, D10 0.3 μm, impurities (wt. %): 0.5 O, 0.5 C, 0.08 Al, 0.01 Ca, 0.008 Fe);
-
MoSi2 (Sigma Aldrich, purity > 99% average particle size < 2 μm, metallic impurities (<2000 PPM): 400 ppm Al, 12 ppm Ba, 16ppm Cr, 400 Fe, 12 ppm K, 39 ppm W).
Preliminarily, the as-received TiB2 powder was milled via high-energy milling (HEM) in a WC jar using 0.5 mm WC-6Co spheres in air for 10, 20 and 30 min to analyze the change in the particle size distribution. Each ten minutes, small powder batches were analyzed using the sedimentation method (SEDIGRAPH III plus, Micromeritics) to determine the particle size distribution of the baseline TiB2 (with no sintering additives) with scanning electron microscopy (FE-SEM, Carl Zeiss Sigma NTS GmbH, Oberkochen, Germany) equipped with an energy dispersive microanalysis (EDS, Model INCA energy 300; Oxford Instruments, Abingdon, UK) system to qualitatively track the main contaminants and XRD (Bruker D8 Advance, Karlsruhe, Germany).
Then, powder mixtures were prepared according to the following compositions:
-
TiB2 + 5 vol.% B4C, labeled as TC;
-
TiB2 + 5 vol.% Si3N4, labeled as TS;
-
TiB2 + 5 vol.% MoSi2, labeled as TM.
For each composition, one batch was soft homogenized by ball milling in ethyl alcohol (EtOH), and the second one was additionally milled via high-energy milling for 30 min (labeled as TC-h, TS-h, TM-h). Two additional batches were the as-received TiB2 and the high-energy-milled TiB2 (labeled as T and T-h).
The eight powder mixtures were then sintered in a graphite mould using an induction-heated hot-pressing machine at 1700–1950 °C, with applied pressure of 30–35 MPa and holding time of 10 min in a vacuum (0.1–0.2 mbar).
X-ray diffraction analysis was carried out on dense samples to determine the crystalline phases. After sintering, the bulk densities were measured using Archimedes’ method. The theoretical densities of the samples were calculated using the rule of mixtures taking into account the incorporated WC and Co phases.
Fracture and polished sections were analyzed with SEM-EDS. Image analysis (Image-Pro Plus® version 7, Media Cybernetics, Silver Springs, MD, USA) carried out on the polished sections was used to determine the mean grain size of the sintered samples.
As for the mechanical properties, the hardness was tested with Vickers indentation, with a load of 1 Kg (Innovatest Falcon 505, Rupac, The Netherlands). Young’s modulus was determined through the resonance frequency method.
Spectra were collected on unpolished sample cross-sections cut with diamond blades, ground, and carefully cleaned. Optical reflectance spectra at room temperature in the 0.25–2.5 µm wavelength region were acquired using a double-beam spectrophotometer (Lambda900, Perkin Elmer, Waltham, Massachusetts) equipped with a 150 mm diameter integration sphere for the measurement of the hemispherical reflectance. The spectra in the wavelength region 2.5–15.5 µm were acquired using a Fourier Transform spectrophotometer (FT-IR Excalibur, Bio-Rad, Hercules, California) equipped with a gold-coated integrating sphere and a liquid-nitrogen-cooled detector. In all cases, the reflectance spectra were acquired for a quasi-normal incidence angle.

3. Results and Discussion

3.1. Effect of Milling with WC-Co Beads

The amount of incorporated WC was determined by weighing the powders before and after the milling procedure. The plot of Figure 1 shows a significant decrease in mean grain size and an increase in WC-Co contamination (~1.8 vol.%) after 30 min of high-energy milling. The SEM images in Figure 2 confirm the progressive decrease in the size of TiB2 particles and the introduction of submicrometric particles of WC-Co depicted as white particles. The reduction rate of the mean particle diameter decreased progressively during HEM without reaching a plateau, while the mass from the grinding media seems to have increased linearly within the first 30 min.
Furthermore, HEM resulted in a progressive increase in the number of particles with a diameter below 0.5 µm. The sharp peak observed in the distribution of Figure 3 was attributed to the lower limit of resolution, which prevented the distinction of particles with a diameter of 0.25 microns or less. This final peak was caused by the submicrometric particles of WC-Co introduced during HEM, and it increased as the milling process progressed.
The coarse fraction of the mixture was reduced, as evidenced by the change in the D90 value of the starting TiB2 from the nominal 5.8 µm (the coarsest) to 3.6 µm in the 30 min planetary milled mixture. It was found that TiB2 has a significant amount of oxygen impurities. According to previous analyses on B4C-TiB2 composites, SEM/TEM images of the powder mixtures after milling revealed that TiB2 particles were surrounded with a continuous amorphous layer of about 5.5 ± 1 nm, containing Ti, W and O elements [58]. SEM analyses confirmed that high-energy milling resulted in smaller particle sizes and contamination with WC/Co-Cr due to wear of WC-Co balls. This contamination could be attributed to a mechanical alloying phenomenon [59].
Additionally, alloy nanoparticles with different chemical compositions, such as W-rich (W-Ti-Cr-Co) and Co-rich (Co-Cr-Ti-W), were observed to be attached to TiB2 due to the mechanical alloying process during high-energy milling. These nanoparticles were formed as a result of contamination from other metallic species, such as W, Co and Cr [58].

3.2. Sintering Behavior and Microstructural Features

The details of the sintering cycles and quantitative data on microstructural features are summarized in Table 1. The hot-pressing schedule was determined during the heating process based on the recorded shrinkage measured by the displacement of the rams. Due to the reactions between the matrix, sintering aids and oxide phase, it was not possible to calculate a theoretical density for the different compositions, except for the T sample, which consisted solely of as-received TiB2. The relative density of samples was calculated as (1-P) where P represents the percent porosity determined through image analysis on secondary electron micrographs of the polished sections.
The densification behavior was clearly affected by both the addition of sintering agents and the high-energy milling of the powders, as shown in Figure 4a,b which present the relative density plotted against time. From this type of graph, it can be observed at which temperature the shrinkage starts and the maximum temperature needed for complete densification.
It can be observed that for the as-received TiB2 powder, densification was very slow and only began after 10 min at 1950 °C. In contrast, for the powder mixtures containing the sintering additives, densification started at lower temperatures: 1630 °C for TC (addition of B4C), 1770 °C for TS (addition of Si3N4) and 1710 °C for TM-h (addition of MoSi2); see curve in Fig 4a. The addition of boron carbide increased densification rate, but the final density remained below 80%. Si3N4 and MoSi2 notably improved sintering, as indicated by a steep increase in the densification curve and the presence of SiO2 pockets after sintering for both samples. The steep change in the slope of the curve could be an indication of liquid phase formation and the increase in the densification rate due to particle rearrangement.
In the case of high-energy-milled powders, densification initiated at 1570 °C for T-h (reference, no sintering aids), 1530 °C for TC-h (addition of B4C), 1560 °C for TS-h (addition of Si3N4) and 1250 °C for TM-h (addition of MoSi2). For only the high-energy-milled TiB2 ceramic, the increase in the shrinkage rate was more gradual. In contrast, a steep increase similar to what was observed for the TS and TM samples was also noticed for the three systems with B4C, Si3N4 or MoSi2.
X-ray diffraction patterns (Figure 5) recorded TiB2 (TiB2: JCPDF card 35-0741) in all samples and WB for the high-energy-milled ones, along with minor secondary phases depending on the sintering aid and the preparation method employed. After sintering, no B4C or Si3N4 peaks were detected, while MoSi2 was still present in the softly homogenized sample. Moreover, TiN was observed in the Si3N4 samples (TS and TS-h) originating from the reaction between Si3N4 and TiO2.
W2CoB2 was only found in the high-energy-milled sample of TiB2 without additives (T-h), TM-h (MoSi2) exhibited phases of MoB, TiC and TiSi.
An overview of the microstructural features of both the ball-milled and high-energy-milled samples is provided in Table 1. It can be observed that all samples treated with high-energy milling displayed a higher degree of densification, and the highest densification grades were obtained for TiB2 samples densified with MoSi2 as a sintering aid (TM and TM-h), which reached final relative densities equal to 95.1% for the soft homogenized sample and 94.7% for the HEM sample.
A detailed description of the microstructure is given below.
T and T-h. The EDS spectrum of the three different phases confirmed the absence of the other metals in the TiB2 cores and the presence of W and Co in TiB2 solid solution rims. Moreover, the presence of Cr and C can be observed in the mixed boride from WC-Co grinding media: chromium was indeed present in lower quantities than Co, and carbon is attributable to residual unreacted carbide.
TC and TC-h. TiB2 without sintering aids (T) did not sinter; in fact, as can be seen in Figure 6, no necks formed between the grains. Adding 5 vol.% B4C (Figure 7) improved sintering and densification rate, but after 15 min at 1900 °C, although the necks had already formed, the pores were still interconnected; therefore, the material was in an intermediate stage where pores were closing, and the relative density was less than 80%. With the additional aid of high-energy milling, a microstructure with closed porosities and a relative density of around 89% was reached.
Although XRD spectra did not record the presence of residual crystalline B4C, from the polished surface SEM images of TC sample it is possible to notice a dark phase which the EDS spectrum confirmed to be boron carbide (Figure 7). In the HEM sample (TC-h) B4C is not present, and wider rims than other HEM samples can be noted. Moreover, EDS revealed that the white contrasting phases in TC-h micrographs are WC with a lower Co content and not WB.
The core rim microstructure had two types of rims: an inner one of a lighter color, therefore heavier atomically, and an outer one of a darker gray color. The EDS spectra of these rims showed no significant differences in the composition.
TS and TS-h. The SEM micrograph of the TS (5 vol.% Si3N4) polished surface shows a bulk with 14% residual porosity; however, the SEM image of the fracture appears to be totally dense and without any porosity (Figure 8). This is due to the large presence of amorphous compounds like BN and SiO2 products from the reaction between Si3N4 and oxide phases (B2O3 and TiO2).
Porosities in the TS sample’s polished section reveal the presence of secondary phases when observed with the InLens mode. In detail, phases of BN were recognized, with the characteristic lamellar structure, SiO2, a darker phase, and TiN, which has a higher molecular weight than TiB2, appearing as a light gray phase in the SE2 mode. These phases were confirmed by EDS spectra (Figure 8). The SEM micrograph of the HEM sample with 5 vol.% Si3N4 (TS-h) showed the typical core–shell structure of TiB2 with WB inclusions in addition to the phases already seen in the soft homogenized sample. Despite the presence of WC contamination in the powders which reacted with the oxide phases to form WB, the formation of SiO2 and BN phases can still be noted. In addition, an amorphous phase containing B, O, N and Si was identified among the grains.
TM and TM-h: As already observed for TS, the TM sample (with 5 vol.% MoSi2 soft homogenized) also does not have empty pores, but silica pockets produced by the reaction between MoSi2 and B2O3, a solid solution containing Mo instead of W with TiB2, were formed, leading to a similar core–shell structure to that of the HEM sample. No MoB was observed, but XRD and EDS confirmed the presence of unreacted MoSi2, which appears as a white phase.
In the HEM sample (TM-h), WC and MoSi2 phases combine, resulting in a core–shell structure where the cores of TiB2 are surrounded by a rim composed of a solid solution of mixed boride of Ti, W, Mo (and Co in the minor part) with inclusions of mixed borides of W and Mo (and Co). Again, complete densification of the matrix is interrupted by the formation of silica pockets (Figure 9). The rims in this sample are very narrow, and light gray colored TiSi phases can be observed.

Densification Mechanisms

The densification behavior was highly affected by the sintering aid and the high-energy milling procedure. Starting from the as-received powder, the effect of milling with WC media was twofold: on one hand, high-energy milling reduced the mean particle size, activating the particle surface; on the other, contamination from WC led to elimination of surface oxides through surface reactions on the TiB2 particles (TiO2 and B2O3) [46,60,61].
The most probable reactions are [61,62,63]:
TiO2 (s) + 3WC (s) → TiC (s) + 3W (s) + 2CO (g)
3TiB2 (s) + TiO2 (s) + 6WC (s) → 4TiC (s) + 6WB (s) + 6CO (g)
3WC (s) + B2O3 (s) → 2WB (s) + W + 3CO (g)
WC particles favored the elimination of surface oxides, while release of W resulted in the formation of core–shell structures in TiB2. Although the formation of TiC seems favored, it was not detected with EDS and XRD methods. The large presence of WB inclusions and the absence of TiC therefore suggest a greater relevance of Reaction (3) between WC and B2O3.
The addition of sintering aids made the picture even more complex with other reactions taking place. In addition to Reactions (1)–(3), B4C reduced the oxygen-carrying chemical species present on the surfaces of the TiB2 powder particles through the following reaction:
7TiO2 (s) + 5B4C (s) → 7TiB2 (s) + 5CO (g) + 3B2O3 (l)
WC inclusions can also react with B4C via the reaction:
2WC (s) + B4C (s) → 2WB2 (s) + 3C (s)
The main effect of Si3N4 was to remove surface oxides from TiB2 powder particles and form liquid phases according to [64]:
Si3N4 (s) + 2B2O3 (l) → 4BN (s) + 3SiO2 (l)
2Si3N4 (s) + 6TiO2 (s) → 6TiN (s) + 6SiO2 (l) + N2 (g)
2TiB2 (s) + 3N2 (g) → 2TiN (s) + 4BN (s)
Subsequently, the liquid formed at about 1600 °C was very likely to be due to the reaction of Si3N4 remaining from Reactions (6) and (7) with SiO2 and TiB2. During cooling, BN, TiN and SiO2 pockets precipitated from the liquid phase (Figure 8), but some of them were retained along the grain boundaries in form of amorphous Si-O-B-N-Ti films.
Similar to other borides, for MoSi2 addition, densification is presumed to be assisted by a transient liquid phase, due to the reaction between MoSi2 and surface oxide impurities on the diboride particles [65,66]. Several reaction paths were indicated, as suggested by the variety of crystalline and amorphous phases identified in the dense microstructures. The presence of crystalline MoB (only for the HEM sample) and pockets of amorphous SiO2 suggests the following reaction occurs:
2MoSi2 (s) + B2O3 (l) → 2MoB (s) + 2.5Si (l) + 1.5SiO2 (l)
Reaction (9) implies the removal of B2O3 from boride particle surface and formation of Si- or Si-O-based liquids. According to the Mo-Si-B phase diagram (Figure S1) [67], liquids could form at temperatures as low as 1350 °C. The formation of liquid promotes mass transfer mechanisms via partial dissolution of the boride matrix. The observed formation of epitaxial core–shell structures suggests a solution re-precipitation mechanism even if diffusion of Mo into the diboride lattice could also occur. During cooling, the transient liquid phase solidifies, resulting in the formation of discrete crystalline phases.
The mixing of TiB2, WB and MoB resulted in a core–shell microstructure (Figure 9) with TiB2 core, a mixed boride rim with the three metals (Ti, W, Mo)BX and (W, Mo)BX particles.

3.3. Mechanical Properties

Hardness and fracture toughness measurements were conducted for all the samples except T and TC, as their density was below 80% (refer to Table 1). The hardness of the samples was negatively influenced by porosity and the presence of softer phases, while it increased with a decrease in the mean grain size. In this batch of materials, the highest hardness value, 24.5 GPa, was observed for T-h, indicating that high-energy milling (HEM) has a comparable effect on hardness as the addition of additives. Moreover, both samples with the addition of MoSi2 (TM and TM-h) exhibited similar hardness values due to their higher relative density and relatively small mean grain size. The lowest hardness value, 20 GPa, was recorded for TC-h, which had the highest porosity among the high-energy-milled samples.
Also, fracture toughness increased with the decrease in grain size and increase in density. However, it was sensitive to the presence of defects, secondary phases and their dispersion in the material. The highest fracture toughness value, 5.4 MPa m1/2, was measured for TS and TM samples, which were soft homogenized with Si3N4 and MoSi2, respectively. These samples exhibited large pockets of silica embedded within the material. T-h also showed a similar level of toughness, while the HEM samples with aids displayed lower values, approximately 4 MPa m1/2 or below, despite their high density.
The highest hardness and fracture toughness values obtained in this study are comparable to the limits reported in the literature for TiB2 samples with very high density, see Table 2, achieved through various sintering techniques such as hot pressing, spark plasma sintering and microwave sintering.

3.4. Optical Properties

The optical properties were measured for selected samples, e.g., T, T-h, TS and TC.
From optical spectra (Figure 10), it can be seen that all samples keep the original trend of the pure and dense TiB2 matrix [53], with the increase in reflectance values from UV–visible to the mid-infrared (MIR), although the step-like curve of the latter appears to have translated into more smoothly increasing curves. The reason can be primarily ascribed to the lower density of the samples investigated in the present work. In fact, the distorting effect of porosity on the reflectance curves of materials has been reported in the literature [70,71,72]. This is also confirmed if T, T-h and TS spectra are compared.
The densest (T-h) sample shows the smoothest curve, while the sample T is the most porous and complex one within this group, and the TS sample shows intermediate porosity and behavior. On the other hand, the presence of (W, Co, Cr)BX phases and a (Ti,W)BX solid solution detected by EDS in the T-h sample does not seem to significantly affect the T-h spectrum, as qualitatively similar spectral features (the shoulder around 2.6–3 μm wavelength) are also detected in the spectrum of TS, which was not subjected to high-energy milling.
From the experimental hemispherical reflectance spectra, it is possible to estimate the solar absorptance α and the thermal emittance ε(T) at the temperature T, as widely accepted in the literature [73], according to the following relationships:
α = λ m i n λ m a x ( 1 ρ λ ) · S λ d λ λ m i n λ m a x S λ d λ
ε = λ 1 λ 2 ( 1 ρ λ ) · B λ , T d λ λ 1 λ 2 B λ , T d λ
where ρ (λ) is the experimental spectral hemispherical reflectance, S(λ) the AM 1.5 solar spectrum [74], B(λ,T) is the blackbody spectral radiance at the temperature T of interest and the integration bounds are λmin = 0.3 µm and λmax = 3.0 µm for Equation (10) and λ1 = 0.3 µm and λ2 = 16.0 µm for Equation (11). It is worth noting that, being obtained form room-temperature spectra, the values obtained from Equations (10) and (11) can be seen as estimated values, useful for a comparative evaluation of samples when they do not undergo chemical or physical changes at the temperatures of interest [75]. The comparison is particularly robust if the samples are quite similar to each other, e.g., they belong to the same family or they have compositions with minor differences, like in our case. As for the discrepancy we could expect between the calculated and experimental values of at high temperatures, from the literature it is possible to infer that ε is generally underestimated (see for instance [55,76]).
The temperature-dependent emittance curves are shown in Figure 11a; the solar absorptance spectra are indicated in the legend, while the spectral selectivity α/ε is shown in Figure 11b.
If the solar absorptance is considered, the sample showing the highest values is T-h. It has an α value of around 0.79, which is an excellent value compared to that of the reference material in most advanced solar absorbers, i.e., SiC [57]. For all samples investigated here, α is significantly higher than that of the dense TiB2 previously reported [53].
For all samples, the estimated thermal emittance increases with temperature due to the interplay between the changes in the spectral distribution of the blackbody radiation as the temperature increases and the shape of the spectral reflectance of samples. In particular, as the temperature increases, the peak wavelength of the blackbody radiation is moved towards shorter wavelengths, where the samples’ reflectance is lower (Figure 11), arising in a higher value of the ratio in Equation (11). The increase in ε(T) with temperature is thus strictly related to the optical properties of the material under investigation. In the majority of cases, the emittance increases with temperature like in this case but not always (see for instance the materials investigated in [77]).
Spectral selectivity α/ε is one of the figures of merit that can be used to evaluate the characteristics of a solar absorber [78]. It gives information about the capability of the material to retain the thermal energy collected through the sunlight absorption before thermal radiation intervenes to decrease the collected energy. As such, a higher α/ε value is desired, providing that, for maximizing the absorber efficiency ηabs, it is not achieved through an insufficient α. From Figure 11b, the samples with the highest α/ε are T-h and TS, among which T-h can be considered as the most promising one due to its significantly higher solar absorptance. On the other hand, T-h favorably compares to TC as well due to the better α value, despite the similar emittance.

4. Conclusions

TiB2 ceramics were prepared through hot-pressing sintering after ball milling or high-energy milling using WC-Co milling media for four different compositions: TiB2 as received and with 5 vol.% of B4C, Si3N4 or MoSi2 as a sintering aid. The addition of sintering aids, except for B4C which only had a slight improvement in sintering, accelerated the densification process but also resulted in the formation of secondary phases. Specifically, a liquid phase was formed at around 1600 °C when using Si3N4 or MoSi2 as aids, which improved the densification rate but left silica pockets within the bulk material after sintering.
High-energy milling proved to be highly effective in reducing the size of the starting powders, leading to improved densification. The presence of carbon from WC doping exerted a cleaning action, contributing to elimination of surficial oxide phases. W remnants were incorporated in the TiB2 lattice, resulting in the formation of core–shell structures, the development of which was influenced by temperature and holding time.
Even without the addition of sintering aids, high-energy milling of TiB2 alone resulted in a density of 93.4% and the highest recorded Vickers hardness of 24.5. The best material was the one with added 5 vol.% MoSi2 and high-energy-milled, which achieved almost complete densification (although limited by the presence of silica pockets) at 1700 °C. This material also exhibited a Vickers hardness of 24.4 GPa, indicating its superior hardness properties.
Optical properties were measured for T, T-h, TS and TC. T-h showed both the highest α value around 0.79, and the highest spectral selectivity α/ε, indicating additive-free high-energy-milled TiB2 as the most promising sample.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano13192683/s1, Figure S1: Melting diagram of Mo–Si–B system [67].

Author Contributions

Conceptualization, D.S.; Methodology, S.T., D.S. and E.S.; Investigation, S.T., S.F. and E.S.; Writing—Original Draft Preparation, S.T.; Writing—Review and Editing, S.T., D.S., S.F. and E.S.; Supervision, D.S. and E.S. All authors have read and agreed to the published version of the manuscript.

Funding

A part of this research was carried out within the project with title “Network 4 Energy-Sustainable Transition—NEST”, funded under the National Recovery and Resilience Plan (NRRP), Mission 4 Component 2 Investment 1.3—Call for tender No. 1561 of 11 October 2022 of Ministero dell’Università e della Ricerca (MUR), funded by the European Union—NextGenerationEU, Project code PE0000021, Concession Decree No. 1561 of 11 October 2022 adopted by Ministero dell’Università e della Ricerca (MUR), CUP B53C22004060006, according to attachment E of Decree No. 1561/2022.

Data Availability Statement

The raw/processed data required to reproduce these findings cannot be shared at this time due to technical or time limitations.

Acknowledgments

Industrie Bitossi S.p.A and CNR are acknowledged for supporting the PhD of S. Taraborelli. ES would like to thank the CNR-INO technicians M. Pucci and M. D’Uva for technical assistance.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mandorf, V.; Hartwig, J.; Seldin, E.J. High Temperature Properties of Titanium Diboride. In High Temperature Materials II; Ault, G.M., Barclay, W.F., Munger, H.P., Eds.; Gordon and Breach: New York, NY, USA, 1963; pp. 455–467. [Google Scholar]
  2. Wiley, D.E.; Manning, W.R.; Hunter, O. Elastic Properties of Polycrystalline TiB2, ZrB2 and HfB2 from Room Temperature to 1300 °K. J. Less Common Met. 1969, 18, 149–157. [Google Scholar] [CrossRef]
  3. Baik, S.; Becher, P.F. Effect of Oxygen Contamination on Densification of TiB2. J Am. Ceram. Soc. 1987, 70, 527–530. [Google Scholar] [CrossRef]
  4. Matsushita, J.; Suzuki, T.; Sano, A. High Temperature Strength of TiB2 Ceramics. J. Ceram. Soc. Jpn. 1993, 101, 1074–1077. [Google Scholar] [CrossRef]
  5. Jüngling, T.; Sigl, L.S.; Oberacker, R.; Thümmler, F.; Schwetz, K.A. New Hardmetals Based on TiB2. Int. J. Refract. Met. Hard Mater. 1993, 12, 71–88. [Google Scholar] [CrossRef]
  6. Munro, R.G. Material Properties of Titanium Diboride. J. Res. Natl. Inst. Stand. Technol. 2000, 105, 709–720. [Google Scholar] [CrossRef]
  7. Murthy, T.S.R.C.; Basu, B.; Balasubramaniam, R.; Suri, A.K.; Subramanian, C.; Fotedar, R.K. Processing and Properties of TiB2 with MoSi2 Sinter-Additive: A First Report. J. Am. Ceram. Soc. 2006, 89, 131–138. [Google Scholar] [CrossRef]
  8. Murthy, T.S.R.C.; Sonber, J.K.; Subramanian, C.; Fotedar, R.K.; Kumar, S.; Gonal, M.R.; Suri, A.K. A New TiB2+CrSi2 Composite—Densification, Characterization and Oxidation Studies. Int. J. Refract. Met. Hard Mater. 2010, 28, 529–540. [Google Scholar] [CrossRef]
  9. Hamidzadeh Mahaseni, Z.; Dashti Germi, M.; Ahmadi, Z.; Shahedi Asl, M. Microstructural Investigation of Spark Plasma Sintered TiB2 Ceramics with Si3N4 Addition. Ceram. Int. 2018, 44, 13367–13372. [Google Scholar] [CrossRef]
  10. Ahmadi, Z.; Mahaseni, Z.H.; Germi, M.D.; Asl, M.S. Microstructure of Spark Plasma Sintered TiB2 and TiB2–AlN Ceramics. Adv. Ceram. Prog. 2019, 5, 36–40. [Google Scholar]
  11. Subramanian, C.; Murthy, T.S.R.C.; Suri, A.K. Synthesis and Consolidation of Titanium Diboride. Int. J. Refract. Met. Hard Mater. 2007, 25, 345–350. [Google Scholar] [CrossRef]
  12. Golla, B.R.; Bhandari, T.; Mukhopadhyay, A.; Basu, B. Titanium Diboride. In Ultra-High Temperature Ceramics; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2014; pp. 316–360. ISBN 978-1-118-70085-3. [Google Scholar]
  13. Basu, B.; Raju, G.B.; Suri, A.K. Processing and Properties of Monolithic TiB2 Based Materials. Int. Mater. Rev. 2006, 51, 352–374. [Google Scholar] [CrossRef]
  14. Raju, G.B.; Basu, B. Development of High Temperature TiB2-Based Ceramics. Key Eng. Mater. 2009, 395, 89–124. [Google Scholar] [CrossRef]
  15. Sabahi Namini, A.; Seyed Gogani, S.N.; Shahedi Asl, M.; Farhadi, K.; Ghassemi Kakroudi, M.; Mohammadzadeh, A. Microstructural Development and Mechanical Properties of Hot Pressed SiC Reinforced TiB2 Based Composite. Int. J. Refract. Met. Hard Mater. 2015, 51, 169–179. [Google Scholar] [CrossRef]
  16. Masi, A.; Alvani, C.; Armenio, A.A.; Bellusci, M.; Celentano, G.; Zignani, C.F.; Fabbri, F.; Barbera, A.L.; Padella, F.; Pentimalli, M.; et al. Effect of Oxygen Contamination on Densification of Fe(Se,Te). J. Phys. Conf. Ser. 2020, 1559, 012051. [Google Scholar] [CrossRef]
  17. Tennery, V.J.; Finch, C.B.; Yust, C.S.; Clark, G.W. Structure-Property Correlations for TiB2-Based Ceramics Densified Using Active Liquid Metals. In Science of Hard Materials; Viswanadham, R.K., Rowcliffe, D.J., Gurland, J., Eds.; Springer: Boston, MA, USA, 1983; pp. 891–909. ISBN 978-1-4684-4319-6. [Google Scholar]
  18. Ferber, M.K.; Becher, P.F.; Finch, C.B. Effect of Microstructure on the Properties of TiB2 Ceramics. J. Am. Ceram. Soc. 1983, 66, C-2–C-3. [Google Scholar] [CrossRef]
  19. Baumgartner, H.; Steiger, R.A. Sintering and Properties of Titanium Diboride Made from Powder Synthesized in a Plasma-Arc Heater. J. Am. Ceram. Soc. 1984, 67, 207–212. [Google Scholar] [CrossRef]
  20. Pan, M.-J.; Hoffman, P.A.; Green, D.J.; Hellmann, J.R. Elastic Properties and Microcracking Behavior of Particulate Titanium Diboride–Silicon Carbide Composites. J. Am. Ceram. Soc. 1997, 80, 692–698. [Google Scholar] [CrossRef]
  21. Königshofer, R.; Fürnsinn, S.; Steinkellner, P.; Lengauer, W.; Haas, R.; Rabitsch, K.; Scheerer, M. Solid-State Properties of Hot-Pressed TiB2 Ceramics. Int. J. Refract. Met. Hard Mater. 2005, 23, 350–357. [Google Scholar] [CrossRef]
  22. Murthy, T.S.R.C.; Basu, B.; Srivastava, A.; Balasubramaniam, R.; Suri, A.K. Tribological Properties of TiB2 and TiB2–MoSi2 Ceramic Composites. J. Eur. Ceram. Soc. 2006, 26, 1293–1300. [Google Scholar] [CrossRef]
  23. Gu, M.L.; Xu, H.J.; Zhang, J.H.; Wei, Z. Influence of Hot Pressing Sintering Temperature and Time on Microstructure and Mechanical Properties of TiB2/Al2O3 Tool Materials. Adv. Mater. Res. 2012, 500, 629–633. [Google Scholar] [CrossRef]
  24. Turan, A.; Cinar Sahin, F.; Goller, G.; Yucel, O. Spark Plasma Sintering of Self-Propagating High-Temperature Synthesized TiB2 Ceramics. Nano Stud. 2017, 52, 27–34. [Google Scholar]
  25. Zhang, Z.H.; Shen, X.B.; Wang, F.C.; Lee, S.K.; Wang, L. Densification Behavior and Mechanical Properties of the Spark Plasma Sintered Monolithic TiB2 Ceramics. Mater. Sci. Eng. A 2010, 527, 5947–5951. [Google Scholar] [CrossRef]
  26. Balcı, Ö.; Burkhardt, U.; Schmidt, M.; Hennicke, J.; Barış Yağcı, M.; Somer, M. Densification, Microstructure and Properties of TiB2 Ceramics Fabricated by Spark Plasma Sintering. Mater. Charact. 2018, 145, 435–443. [Google Scholar] [CrossRef]
  27. Delbari, S.A.; Sabahi Namini, A.; Shahedi Asl, M. Hybrid Ti Matrix Composites with TiB2 and TiC Compounds. Mater. Today Commun. 2019, 20, 100576. [Google Scholar] [CrossRef]
  28. Sabahi Namini, A.; Motallebzadeh, A.; Nayebi, B.; Shahedi Asl, M.; Azadbeh, M. Microstructure–Mechanical Properties Correlation in Spark Plasma Sintered Ti–4.8 Wt.% TiB2 Composites. Mater. Chem. Phys. 2019, 223, 789–796. [Google Scholar] [CrossRef]
  29. Shayesteh, F.; Delbari, S.A.; Ahmadi, Z.; Shokouhimehr, M.; Shahedi Asl, M. Influence of TiN Dopant on Microstructure of TiB2 Ceramic Sintered by Spark Plasma. Ceram. Int. 2019, 45, 5306–5311. [Google Scholar] [CrossRef]
  30. Nguyen, T.P.; Hamidzadeh Mahaseni, Z.; Dashti Germi, M.; Delbari, S.A.; Le, Q.V.; Ahmadi, Z.; Shokouhimehr, M.; Shahedi Asl, M.; Sabahi Namini, A. Densification Behavior and Microstructure Development in TiB2 Ceramics Doped with H-BN. Ceram. Int. 2020, 46, 18970–18975. [Google Scholar] [CrossRef]
  31. Vajdi, M.; Sadegh Moghanlou, F.; Ahmadi, Z.; Motallebzadeh, A.; Shahedi Asl, M. Thermal Diffusivity and Microstructure of Spark Plasma Sintered TiB2SiCTi Composite. Ceram. Int. 2019, 45, 8333–8344. [Google Scholar] [CrossRef]
  32. Ağaoğulları, D.; Gökçe, H.; Duman, İ.; Öveçoğlu, M.L. Influences of Metallic Co and Mechanical Alloying on the Microstructural and Mechanical Properties of TiB2 Ceramics Prepared via Pressureless Sintering. J. Eur. Ceram. Soc. 2012, 32, 1949–1956. [Google Scholar] [CrossRef]
  33. Matsushita, J.; Sano, A. Sinterability of TiB2 Ceramics Containing Cr and C as the Sintering Aids. J. Ceram. Soc. Jpn. 1992, 100, 593–595. [Google Scholar] [CrossRef]
  34. Cymerman, K.; Oleszak, D.; Rosinski, M.; Michalski, A. Structure and Mechanical Properties of TiB2/TiC–Ni Composites Fabricated by Pulse Plasma Sintering Method. Adv. Powder Technol. 2018, 29, 1795–1803. [Google Scholar] [CrossRef]
  35. Missiaen, J.M.; Chaix, J.M. The Homogeneity of Phase Repartition in TiB2–Fe Composites Using Variance and Covariance Analysis. J. Microsc. 1994, 175, 195–204. [Google Scholar] [CrossRef]
  36. Tian, S.F.; Jiang, L.T.; Guo, Q.; Wu, G.H. Effect of Surface Roughness on Tribological Properties of TiB2/Al Composites. Mater. Des. 2014, 53, 129–136. [Google Scholar] [CrossRef]
  37. Bača, Ľ.; Lenčéš, Z.; Jogl, C.; Neubauer, E.; Vitkovič, M.; Merstallinger, A.; Šajgalík, P. Microstructure Evolution and Tribological Properties of TiB2/Ni–Ta Cermets. J. Eur. Ceram. Soc. 2012, 32, 1941–1948. [Google Scholar] [CrossRef]
  38. Kang, E.S.; Jang, C.W.; Lee, C.H.; Kim, C.H.; Kim, D.K. Effect of Iron and Boron Carbide on the Densification and Mechanical Properties of Titanium Diboride Ceramics. J. Am. Ceram. Soc. 1989, 72, 1868–1872. [Google Scholar] [CrossRef]
  39. Hunt, M.P.; Logan, K.V. Pressureless Sintering of Titanium Diboride Powders. Dev. Strateg. Mater. Ceram. Eng. Sci. Proc. 2008, 29, 211–221. [Google Scholar]
  40. Fu, Z.; Koc, R. Sintering and Mechanical Properties of TiB2-TiC-Ni Using Submicron Borides and Carbides. Mater. Sci. Eng. A 2016, 676, 278–288. [Google Scholar] [CrossRef]
  41. Karthiselva, N.S.; Murty, B.S.; Bakshi, S.R. Low Temperature Synthesis of Dense TiB2 Compacts by Reaction Spark Plasma Sintering. Int. J. Refract. Met. Hard Mater. 2015, 48, 201–210. [Google Scholar] [CrossRef]
  42. Tuffé, S.; Dubois, J.; Fantozzi, G.; Barbier, G. Densification, Microstructure and Mechanical Properties of TiB2-B4C Based Composites. Int. J. Refract. Met. Hard Mater. 1996, 14, 305–310. [Google Scholar] [CrossRef]
  43. Shahedi Asl, M.; Delbari, S.A.; Shayesteh, F.; Ahmadi, Z.; Motallebzadeh, A. Reactive Spark Plasma Sintering of TiB2–SiC–TiN Novel Composite. Int. J. Refract. Met. Hard Mater. 2019, 81, 119–126. [Google Scholar] [CrossRef]
  44. Zhao, H.; Cheng, Y.-B. Formation of TiB2–TiC Composites by Reactive Sintering. Ceram. Int. 1999, 25, 353–358. [Google Scholar] [CrossRef]
  45. Demirskyi, D.; Nishimura, T.; Sakka, Y.; Vasylkiv, O. High-Strength TiB2–TaC Ceramic Composites Prepared Using Reactive Spark Plasma Consolidation. Ceram. Int. 2016, 42, 1298–1306. [Google Scholar] [CrossRef]
  46. Telle, R. Analysis of Pressureless Sintering of Titanium Diboride Ceramics with Nickel, Cobalt, and Tungsten Carbide Additives. J. Eur. Ceram. Soc. 2019, 39, 2266–2276. [Google Scholar] [CrossRef]
  47. Kim, H.-J.; Choi, H.-J.; Lee, J.-G. Mechanochemical Synthesis and Pressureless Sintering of TiB2–AlN Composites. J. Am. Ceram. Soc. 2002, 85, 1022–1024. [Google Scholar] [CrossRef]
  48. Jain, D.; Reddy, K.M.; Mukhopadhyay, A.; Basu, B. Achieving Uniform Microstructure and Superior Mechanical Properties in Ultrafine Grained TiB2–TiSi2 Composites Using Innovative Multi Stage Spark Plasma Sintering. Mater. Sci. Eng. A 2010, 528, 200–207. [Google Scholar] [CrossRef]
  49. Murthy, T.S.R.C.; Balasubramaniam, R.; Basu, B.; Suri, A.K.; Mungole, M.N. Oxidation of Monolithic TiB2 and TiB2–20wt.% MoSi2 Composite at 850 °C. J. Eur. Ceram. Soc. 2006, 26, 187–192. [Google Scholar] [CrossRef]
  50. Mukhopadhyay, A.; Raju, G.B.; Basu, B. Understanding Influence of MoSi2 Addition (5 Weight Percent) on Tribological Properties of TiB2. Met. Mater. Trans. A 2008, 39, 2998–3013. [Google Scholar] [CrossRef]
  51. Raju, G.B.; Basu, B.; Suri, A.K. Oxidation Kinetics and Mechanisms of Hot-Pressed TiB2–MoSi2 Composites. J. Am. Ceram. Soc. 2008, 91, 3320–3327. [Google Scholar] [CrossRef]
  52. Fahrenholtz, W.G.; Hilmas, G.E.; Talmy, I.G.; Zaykoski, J.A. Refractory Diborides of Zirconium and Hafnium. J. Am. Ceram. Soc. 2007, 90, 1347–1364. [Google Scholar] [CrossRef]
  53. Sani, E.; Meucci, M.; Mercatelli, L.; Balbo, A.; Musa, C.; Licheri, R.; Orrù, R.; Cao, G. Titanium Diboride Ceramics for Solar Thermal Absorbers. Sol. Energy Mater. Sol. Cells 2017, 169, 313–319. [Google Scholar] [CrossRef]
  54. Sani, E.; Mercatelli, L.; Sans, J.-L.; Silvestroni, L.; Sciti, D. Porous and Dense Hafnium and Zirconium Ultra-High Temperature Ceramics for Solar Receivers. Opt. Mater. 2013, 36, 163–168. [Google Scholar] [CrossRef]
  55. Sani, E.; Mercatelli, L.; Meucci, M.; Silvestroni, L.; Balbo, A.; Sciti, D. Process and Composition Dependence of Optical Properties of Zirconium, Hafnium and Tantalum Borides for Solar Receiver Applications. Sol. Energy Mater. Sol. Cells 2016, 155, 368–377. [Google Scholar] [CrossRef]
  56. Sani, E.; Landi, E.; Sciti, D.; Medri, V. Optical Properties of ZrB2 Porous Architectures. Sol. Energy Mater. Sol. Cells 2016, 144, 608–615. [Google Scholar] [CrossRef]
  57. Silvestroni, L.; Sciti, D.; Zoli, L.; Balbo, A.; Zanotto, F.; Orrù, R.; Licheri, R.; Musa, C.; Mercatelli, L.; Sani, E. An Overview of Ultra-Refractory Ceramics for Thermodynamic Solar Energy Generation at High Temperature. Renew. Energy 2019, 133, 1257–1267. [Google Scholar] [CrossRef]
  58. Sciti, D.; Failla, S.; Turan, S.; Savaci, U.; Galizia, P. Properties and Ballistic Tests of Strong B4C-TiB2 Composites Densified by Gas Pressure Sintering. J. Eur. Ceram. Soc. 2023, 43, 1334–1342. [Google Scholar] [CrossRef]
  59. Suryanarayana, C.; Klassen, T.; Ivanov, E. Synthesis of Nanocomposites and Amorphous Alloys by Mechanical Alloying. J. Mater. Sci. 2011, 46, 6301–6315. [Google Scholar] [CrossRef]
  60. Failla, S.; Melandri, C.; Zoli, L.; Zucca, G.; Sciti, D. Hard and Easy Sinterable B4C-TiB2-Based Composites Doped with WC. J. Eur. Ceram. Soc. 2018, 38, 3089–3095. [Google Scholar] [CrossRef]
  61. Silvestroni, L.; Failla, S.; Gilli, N.; Melandri, C.; Savacı, U.; Turan, S.; Sciti, D. Disclosing Small Scale Length Properties in Core-Shell Structured B4C-TiB2 Composites. Mater. Des. 2021, 197, 109204. [Google Scholar] [CrossRef]
  62. Zou, J.; Sun, S.-K.; Zhang, G.-J.; Kan, Y.-M.; Wang, P.-L.; Ohji, T. Chemical Reactions, Anisotropic Grain Growth and Sintering Mechanisms of Self-Reinforced ZrB2–SiC Doped with WC. J. Am. Ceram. Soc. 2011, 94, 1575–1583. [Google Scholar] [CrossRef]
  63. Fahrenholtz, W.G.; Wuchina, E.J.; Lee, W.E.; Zhou, Y. Ultra-High Temperature Ceramics: Materials for Extreme Environment Applications; John Wiley & Sons: Hoboken, NJ, USA, 2014; ISBN 978-1-118-92441-9. [Google Scholar]
  64. Park, J.-H.; Koh, Y.-H.; Kim, H.-E.; Hwang, C.S.; Kang, E.S. Densification and Mechanical Properties of Titanium Diboride with Silicon Nitride as a Sintering Aid. J. Am. Ceram. Soc. 1999, 82, 3037–3042. [Google Scholar] [CrossRef]
  65. Sciti, D.; Silvestroni, L.; Celotti, G.; Melandri, C.; Guicciardi, S. Sintering and Mechanical Properties of ZrB2–TaSi2 and HfB2–TaSi2 Ceramic Composites. J. Am. Ceram. Soc. 2008, 91, 3285–3291. [Google Scholar] [CrossRef]
  66. Silvestroni, L.; Kleebe, H.-J.; Lauterbach, S.; Müller, M.; Sciti, D. Transmission Electron Microscopy on Zr- and Hf-Borides with MoSi2 Addition: Densification Mechanisms. J. Mater. Res. 2010, 25, 828–834. [Google Scholar] [CrossRef]
  67. Katrych, S.; Grytsiv, A.; Bondar, A.; Rogl, P.; Velikanova, T.; Bohn, M. Structural Materials: Metal–Silicon–Boron: On the Melting Behavior of Mo–Si–B Alloys. J. Alloys Compd. 2002, 347, 94–100. [Google Scholar] [CrossRef]
  68. Demirskyi, D.; Agrawal, D.; Ragulya, A. Tough Ceramics by Microwave Sintering of Nanocrystalline Titanium Diboride Ceramics. Ceram. Int. 2014, 40, 1303–1310. [Google Scholar] [CrossRef]
  69. Demirskyi, D.; Sakka, Y. High-Temperature Reaction Consolidation of TaC–TiB2 Ceramic Composites by Spark-Plasma Sintering. J. Eur. Ceram. Soc. 2015, 35, 405–410. [Google Scholar] [CrossRef]
  70. MacMillan, M.F.; Devaty, R.P.; Choyke, W.J.; Goldstein, D.R.; Spanier, J.E.; Kurtz, A.D. Infrared Reflectance of Thick P-type Porous SiC Layers. J. Appl. Phys. 1996, 80, 2412–2419. [Google Scholar] [CrossRef]
  71. Sciti, D.; Silvestroni, L.; Sans, J.-L.; Mercatelli, L.; Meucci, M.; Sani, E. Tantalum Diboride-Based Ceramics for Bulk Solar Absorbers. Sol. Energy Mater. Sol. Cells 2014, 130, 208–216. [Google Scholar] [CrossRef]
  72. Tao, Y.; Mao, Z.; Yang, Z.; Zhang, J. Preparation and Characterization of Polymer Matrix Passive Cooling Materials with Thermal Insulation and Solar Reflection Properties Based on Porous Structure. Energy Build. 2020, 225, 110361. [Google Scholar] [CrossRef]
  73. Caron, S.; Garrido, J.; Ballestrín, J.; Sutter, F.; Röger, M.; Manzano-Agugliaro, F. A Comparative Analysis of Opto-Thermal Figures of Merit for High Temperature Solar Thermal Absorber Coatings. Renew. Sustain. Energy Rev. 2022, 154, 111818. [Google Scholar] [CrossRef]
  74. G03 Committee. Tables for Reference Solar Spectral Irradiances: Direct Normal and Hemispherical on 37 Tilted Surface; ASTM International: West Conshohocken, PA, USA, 2012. [Google Scholar]
  75. Watanabe, H.; Ishii, J.; Wakabayashi, H.; Kumano, T.; Hanssen, L. Spectral Emissivity Measurements. In Experimental Methods in the Physical Sciences; Germer, T.A., Zwinkels, J.C., Tsai, B.K., Eds.; Academic Press: Cambridge, MA, USA, 2014; Volume 46, pp. 333–366. [Google Scholar]
  76. Sciti, D.; Silvestroni, L.; Mercatelli, L.; Sans, J.-L.; Sani, E. Suitability of Ultra-Refractory Diboride Ceramics as Absorbers for Solar Energy Applications. Sol. Energy Mater. Sol. Cells 2013, 109, 8–16. [Google Scholar] [CrossRef]
  77. Silvestroni, L.; Capiani, C.; Sciti, D.; Sani, E. Coloring Zirconium Oxide for Novel Energy Saving Industrial Applications. Renew. Energy 2022, 190, 223–231. [Google Scholar] [CrossRef]
  78. Arancibia-Bulnes, C.A.; Estrada, C.A.; Ruiz-Suárez, J.C. Solar Absorptance and Thermal Emittance of Cermets with Large Particles. J. Phys. D Appl. Phys. 2000, 33, 2489. [Google Scholar] [CrossRef]
Figure 1. Mean particle diameter and WC-Co amount added after 10, 20 and 30 min of HEM.
Figure 1. Mean particle diameter and WC-Co amount added after 10, 20 and 30 min of HEM.
Nanomaterials 13 02683 g001
Figure 2. Backscatter electron micrographs of TiB2 powders before and after 10, 20 and 30 min of HEM.
Figure 2. Backscatter electron micrographs of TiB2 powders before and after 10, 20 and 30 min of HEM.
Nanomaterials 13 02683 g002
Figure 3. Cumulative finer mass percent vs. diameter (a) and mass frequency vs. diameter (b) for TiB2 particles before and after 10, 20 and 30 min HEM.
Figure 3. Cumulative finer mass percent vs. diameter (a) and mass frequency vs. diameter (b) for TiB2 particles before and after 10, 20 and 30 min HEM.
Nanomaterials 13 02683 g003
Figure 4. Comparison of densification curves for soft homogenized (a) and high-energy-milled (b) HP sintered samples.
Figure 4. Comparison of densification curves for soft homogenized (a) and high-energy-milled (b) HP sintered samples.
Nanomaterials 13 02683 g004
Figure 5. X-ray diffraction patterns for the soft homogenized (line below) and high-energy-milled (line above) sample of TiB2 without additives (a), with 5 vol.% of B4C (b), 5 vol.% of Si3N4 (c) and 5 vol.% of MoSi2 (d).
Figure 5. X-ray diffraction patterns for the soft homogenized (line below) and high-energy-milled (line above) sample of TiB2 without additives (a), with 5 vol.% of B4C (b), 5 vol.% of Si3N4 (c) and 5 vol.% of MoSi2 (d).
Nanomaterials 13 02683 g005
Figure 6. SEM micrograph of T (left column) and T-h (right column) samples’ fractured and polished sections and EDS spectrum of T-h sample’s polished section (third row).
Figure 6. SEM micrograph of T (left column) and T-h (right column) samples’ fractured and polished sections and EDS spectrum of T-h sample’s polished section (third row).
Nanomaterials 13 02683 g006
Figure 7. SEM micrograph of TiB2 with 5 vol.% B4C soft homogenized (TC, left column) and high-energy-milled (TC-h, right column) samples’ fractured and polished sections and EDS spectrum of polished sections.
Figure 7. SEM micrograph of TiB2 with 5 vol.% B4C soft homogenized (TC, left column) and high-energy-milled (TC-h, right column) samples’ fractured and polished sections and EDS spectrum of polished sections.
Nanomaterials 13 02683 g007
Figure 8. SEM micrograph of TiB2 with 5 vol.% Si3N4 soft homogenized (TS, left column) and high-energy-milled (TS-h, right column) samples’ fractured and polished sections and EDS spectrum of polished sections.
Figure 8. SEM micrograph of TiB2 with 5 vol.% Si3N4 soft homogenized (TS, left column) and high-energy-milled (TS-h, right column) samples’ fractured and polished sections and EDS spectrum of polished sections.
Nanomaterials 13 02683 g008
Figure 9. SEM micrograph of TiB2 with 5 vol.% MoSi2 soft homogenized (TC, left column) and high-energy-milled (TC-h, right column) samples’ fractured and polished sections and EDS spectrum of polished sections.
Figure 9. SEM micrograph of TiB2 with 5 vol.% MoSi2 soft homogenized (TC, left column) and high-energy-milled (TC-h, right column) samples’ fractured and polished sections and EDS spectrum of polished sections.
Nanomaterials 13 02683 g009
Figure 10. Experimental spectral hemispherical reflectance of the sintered sample of TiB2 as received (blue line), high-energy-milled (black line) and with 5 vol.% of Si3N4 (red line) or B4C (green line).
Figure 10. Experimental spectral hemispherical reflectance of the sintered sample of TiB2 as received (blue line), high-energy-milled (black line) and with 5 vol.% of Si3N4 (red line) or B4C (green line).
Nanomaterials 13 02683 g010
Figure 11. Calculated hemispherical emittance (a) and spectral selectivity (b) of the sintered sample of TiB2 as received (blue line/symbols), high-energy-milled (black line/symbols) and with 5 vol.% of Si3N4 (red line/symbols) or B4C (green line/symbols).
Figure 11. Calculated hemispherical emittance (a) and spectral selectivity (b) of the sintered sample of TiB2 as received (blue line/symbols), high-energy-milled (black line/symbols) and with 5 vol.% of Si3N4 (red line/symbols) or B4C (green line/symbols).
Nanomaterials 13 02683 g011
Table 1. Sintering parameters, microstructural features and mechanical properties of hot pressed and pressureless sintered TiB2 composites.
Table 1. Sintering parameters, microstructural features and mechanical properties of hot pressed and pressureless sintered TiB2 composites.
LabelSintering AidHEMMax T/TimeBulk DensityRelative DensityCrystalline PhasesMean Grain SizeHV 1.0KIc
vol.%No°C/ming/cm3% TiB2, μmGPaMPa m1/2
TNoNo1950, 203.45≈76TiB23.2 ± 0.1--
TC5 B4CNo1900, 153.71≈78TiB22.4 ± 0.1--
TS5 Si3N4No1900, 154.2885.6 * (>98)TiB2, TiN3.1 ± 0.122.2 ± 0.55.4 ± 0.4
TM5 MoSi2No1900, 104.4395.1 * (>98)TiB2, MoSi21.8 ± 0.124.4 ± 0.45.4 ± 0.2
T-hNoWC media1900, 104.6093.4TiB2, WB, W2CoB21.0 ± 0.124.5 ± 0.25.0 ± 0.5
TC-h5 B4CWC media1900, 104.2287.9TiB2, WB1.3 ± 0.119.6 ± 0.34.1 ± 0.1
TS-h5 Si3N4WC media1830, 104.5492.4 * (>98)TiB2, TiN, WB0.9 ± 0.123.2 ± 0.23.6 ± 0.1
TM-h5 MoSi2WC media1700, 104.6794.7 * (>98)TiB2, TiC, TiSi, MoB, WB0.8 ± 0.124.4 ± 0.23.5 ± 0.1
* The sample has pockets of silica that appear as porosity (the actual density is reported in brackets).
Table 2. Sintering parameters, microstructural characteristics and mechanical properties of TiB2 samples sintered with different sintering methods and different amounts of additives. Notes: HP—hot pressing, MWS—microwave sintering, SPS—spark plasma sintering.
Table 2. Sintering parameters, microstructural characteristics and mechanical properties of TiB2 samples sintered with different sintering methods and different amounts of additives. Notes: HP—hot pressing, MWS—microwave sintering, SPS—spark plasma sintering.
Sample CompositionProcessing Details: Method,
Temperature, Dwell Time
Relative DensityMean Grain SizeHV 1.0KIcReference
wt%°C/min%TiB2, μmGPaMPa m1/2
TiB2HP, 1800, 6090-245.8Park et al. [64]
TiB2MWS, 1700, 1095.52.0245.6Demirskyi et al. [68]
TiB2SPS, 2000, 5966–1023.55.5Demirskyi and Sakka [69]
TiB2HEM, HP, 1900, 1093.41.024.55.0Present work
TiB2 + 2.5Si3N4HP, 1800, 6099-275.1Park et al. [64]
TiB2 + 10Si3N4HP, 1800, 6096-205.4Park et al. [64]
TiB2 + 3.6Si3N4HP, 1900, 15>98 *3.122.25.4Present work
TiB2 + 3.6Si3N4HEM, HP, 1830, 15>98 *0.923.23.6Present work
TiB2 + 6.8MoSi2HP, 1900, 10>98 *1.824.45.4Present work
TiB2 + 6.8MoSi2HEM, HP, 1700, 10>98 *0.824.43.5Present work
TiB2 + 10MoSi2HP, 1700, 6099.31–2274.0Ch. Murthy et al. [7]
TiB2 + 20MoSi2HP, 1700, 6098.71–2255.0Ch. Murthy et al. [7]
TiB2 + 10TaCSPS, 2000, 5>982–428.85.9Demirskyi and Sakka [69]
* The sample has pockets of silica.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Taraborelli, S.; Failla, S.; Sani, E.; Sciti, D. Sintering, Mechanical and Optical Properties of TiB2 Composites with and without High-Energy Milling. Nanomaterials 2023, 13, 2683. https://doi.org/10.3390/nano13192683

AMA Style

Taraborelli S, Failla S, Sani E, Sciti D. Sintering, Mechanical and Optical Properties of TiB2 Composites with and without High-Energy Milling. Nanomaterials. 2023; 13(19):2683. https://doi.org/10.3390/nano13192683

Chicago/Turabian Style

Taraborelli, Simone, Simone Failla, Elisa Sani, and Diletta Sciti. 2023. "Sintering, Mechanical and Optical Properties of TiB2 Composites with and without High-Energy Milling" Nanomaterials 13, no. 19: 2683. https://doi.org/10.3390/nano13192683

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