Using the particle swarm optimization algorithm, we conducted an extensive search for the high-pressure stable structure of technetium diboride (TcB2) within the pressure range of 0–400 GPa. At zero pressure, the P63/mmc (hP6-TcB2) structure is considered the ground state configuration. As the pressure increases, a structural transition from hP6-TcB2 to P6/mmm (hP3-TcB2) occurs at approximately 174.9 GPa. We discuss the bonding between the two distinct phases and analyze the contribution of different atomic bonds to maintaining their structural stability. Meanwhile, the temperature–pressure phase diagram of TcB2 was successfully determined for the first time through the quasi-harmonic approximation method. It is predicted that the transition pressure from hP6-TcB2 to hP3-TcB2 can be reduced to about 164 GPa at a room temperature of 300 K. These results provide valuable insights into the behavior of TcB2 under different temperature and pressure conditions and open up new possibilities for exploring its potential applications in a variety of environments.

Superhard materials have become an indispensable component of engineering technology due to their exceptional properties, making them highly valuable in various applications, such as abrasives, cutting tools, and wear-resistant coatings.1–4 Consequently, extensive research efforts have been dedicated to the pursuit of novel hard materials. In general, the incorporation of light elements (B, C, N, and O) into electron-rich transition metals represents a reliable approach for synthesizing high hardness materials.5 The synthesis of ReB2, OsB2, RhB1.1, Re2C, TcB2, and OsN26–14 under extreme conditions has been widely reported. Among the materials mentioned, TcB2 stands out for its exceptional mechanical properties and chemical inertness, making it an attractive candidate for superhard materials.13,15–25

According to previous reports, technetium diboride has a rhenium diboride (hP6-TcB2) type structure under ambient conditions. Earlier in the experiment, Trzebiatowski and Rudzinski13 synthesized TcB2 by pressing boron and technetium powders into pellets and sintering in a vacuum. TcB2 was indexed as of hexagonal symmetry, and each conventional cell contains two units and belongs to the P63/mmc space group. Later, Wang18 evaluated the structure stability and electronic characteristics of TcB2 with P63/mmc (hP6-TcB2) and Pmmn (oP6-TcB2) by first-principles calculations under environmental conditions. The results show that hP6-TcB2 is the ground state structure of the two elastically stable structures. In addition, the calculated C33 of hP6-TcB2 is as high as 1022 GPa and the G/B value is 0.87, indicating that it is a potential superhard material. Immediately, Aydin and Simsek19 investigated the electronic structure and mechanical properties of hP6-TcB2 and hP3-TcB2 at zero pressure. It is found that hP6-TcB2 is more energetically favorable than hP3-TcB2, and both of them are non-spin polarized materials. In addition, both hP6-TcB2 and hP3-TcB2 satisfy the elastic stability criterion and are considered to be elastically stable. On the basis of the previously proposed elastically stable structures, Deligoz et al.20 calculated the phonon dispersion curves of hP6-TcB2, hP3-TcB2, and oP6-TcB2 by direct methods and confirmed that they are all dynamically stable.

Furthermore, the formation enthalpies (with reference to Tc + B) of hP6 and oP6 at 0–100 GPa were investigated by Zhong et al.21 The results show that the formation enthalpies of the considered structures under zero pressure are all negative, implying that they are thermodynamically stable. It is also found that hP6-TcB2 is structurally stable in the range of 0–100 GPa. Subsequently, Kuang et al.22 selected seven possible transition metal boride structures for probing and discussed the thermodynamic stability of these structures at zero pressure. The obtained formation enthalpies indicate that hP6-TcB2, oP6-TcB2, hR9-TcB2, and hP3-TcB2 are all thermodynamically stable and can be synthesized at ambient pressure. Moreover, they further explored the relative formation enthalpies of hP6-TcB2 and hP3-TcB2 at 0–200 GPa. The results indicate that hP6-TcB2 remains stable over a wide range of 0–147.5 GPa and transforms into hP3-TcB2 phase at around 147.5 GPa.

In summary, previous studies have speculated the structure of TcB2 based on known abundant transition metal diborides.18–26 However, the resulting phase transition sequence may not offer a complete overview and still lacks a complete understanding of the mechanical behavior of TcB2 under high pressures. In order to obtain a more precise phase transition sequence of technetium diboride under different pressures, we conducted a comprehensive exploration of the crystal structure of TcB2 across an extensive pressure range (0–400 GPa) by using the particle swarm optimization (PSO) algorithm27 in the CALYPSO code.28 This approach is effectively capable of finding stable or meta-stable structures only depending on the given chemical composition. By employing the PSO method over an extensive pressure range, we find that hP6-TcB2, hP3-TcB2, oP6-TcB2, hR9-TcB2, and the orthorhombic phase Cmcm (oC12-TcB2) all exhibit thermodynamic and dynamic stability under ambient conditions, in which the hP6-TcB2 structure remains stable from 0 to 174.9 GPa, and it will transform to the hP3-TcB2 phase with the further increase of pressure.

In this work, the crystal structure prediction was performed in the range of 0–400 GPa containing up to 4 formula units (f.u.) per simulation cell by using the PSO algorithm27 as implemented in the CALYPSO code.28 This method can predict stable or meta-stable crystal structures based on given chemical compositions under specific external conditions. More details of this structural search method can be found in Refs. 27 and 28. So far, it has been successfully applied not only to element solids, but also to binary and ternary compounds.29–32 

The ab initio optimizations and theoretical simulations are performed using ultrasoft pseudopotential, as implemented in the Cambridge Serial Total Energy Package (CASTEP).33,34 The generalized gradient approximation of Perdew–Burke–Ernzerhof (PBE)35 is used to describe the exchange-correlation function form, and the valence electron interactions with 2s22p1 and 4d55s2 are selected for the analysis of the electron configurations for B and Tc atoms, respectively. To ensure that the convergence condition is achieved, geometry optimization is performed until the energy and force differences are less than 5 × 10−6 eV/atom and 0.01 eV/Å, respectively. For the integration over the Brillouin zone, the Monkhorst–Pack k-point meshes36 are 21 × 21 × 7 for hP6-TcB2, 12 × 12 × 9 for hP3-TcB2, 8 × 13 × 9 for oP6-TcB2, 22 × 22 × 5 for hR9-TcB2, and 4 × 18 × 18 for oC12-TcB2. In addition, the phonon calculations were performed by using a supercell approach with the finite displacement method37 in the CASTEP code, and 8 × 8 × 4 and 10 × 10 × 8 k-meshes were used for the force calculation of hP6-TcB2 and hP3-TcB2 supercells, respectively.

After full geometric optimization, we present the structural parameters of different TcB2 phases in Table I, along with other theoretical data.13,19–22,25,26 As shown, the calculated a0 and c0 of hP6-TcB2 are very close to the experimental values, with a difference of only 0.03% and 0.07%, respectively. To further evaluate their thermal stability, the formation enthalpy of the identified structures at zero pressure was calculated. The formula for the ΔH of TcB2 is as follows: ΔH = H (TcB2) − H (Tc) − 2H (B), where boron in α phase and technetium in a P63/mmc state were used as the reference structures. The obtained results are listed in Table II. It is evident that hP6-TcB2, hP3-TcB2, oP6-TcB2, hR9-TcB2, and oC12-TcB2 all have negative formation enthalpies at zero pressure. Among them, hP6 has the relatively lowest formation enthalpies (−1.34 eV/f.u.) at zero pressure and is the most stable phase under ambient conditions. Meanwhile, the formation enthalpies of hP3-TcB2, oP6-TcB2, hR9-TcB2, and oC12-TcB2 are −0.18, −1.18, −1.16, and −0.86 eV/f.u., respectively, which are consistent with other theoretical data.15,18,19,21,22,25,26

TABLE I.

The predicted lattice constants and atomic coordinates for hP6-TcB2, hP3-TcB2, oP6-TcB2, hR9-TcB2, and oC12-TcB2 at different pressures, together with available experimental and theoretical data.

Space groupPearson symbolLattice constants a0, b0, c0 (Å)AtomxyzReference
P63/mmc (0 GPa) hP2.891  7.448 Tc(2c) 0.3333 0.6667 0.25 This work 
B(4f) 0.3333 0.6667 0.5476 
2.877 7.421 Tc(2c) 0.3333 0.6667 0.25 Cal.19  
B(4f) 0.3333 0.6667 0.5476 
2.896 7.428     Cal.20  
2.875 7.418 Cal.21  
2.897 7.416 Cal.26  
2.892 7.453 Exp.13  
Pmmn (0 GPa) oP4.597 2.883 4.102 Tc(2a) 0.8687 This work 
B(4f) 0.1958 0.3766 
4.576 2.869 4.082     Cal.22  
4.571 2.869 4.081 Cal.21  
4.594 2.893 4.094 Cal.20  
4.572 2.869 4.080 Cal.25  
R 3 ¯ m (0 GPa) hR2.900  11.19 Tc(3a) This work 
B(6c) 0.1978 
Cmcm (0 GPa) oC12 12.473 2.893 2.922 Tc(4c) 0.3872 0.75 This work 
B(4c) 0.2514 1.25 
B(4c) 0.5 0.4609 1.25 
P6/mmm (0 GPa) hP2.958  3.399 Tc(1b) 0.5 This work 
B(2c) 0.3333 0.6667 
2.937  3.390     Cal.22  
2.935 3.373 Cal.20  
2.940 3.384 Cal.25  
P6/mmm (180 GPa) hP2.678  3.025 Tc(1b) 0.5 This work 
B(2c) 0.3333 0.6667 
Space groupPearson symbolLattice constants a0, b0, c0 (Å)AtomxyzReference
P63/mmc (0 GPa) hP2.891  7.448 Tc(2c) 0.3333 0.6667 0.25 This work 
B(4f) 0.3333 0.6667 0.5476 
2.877 7.421 Tc(2c) 0.3333 0.6667 0.25 Cal.19  
B(4f) 0.3333 0.6667 0.5476 
2.896 7.428     Cal.20  
2.875 7.418 Cal.21  
2.897 7.416 Cal.26  
2.892 7.453 Exp.13  
Pmmn (0 GPa) oP4.597 2.883 4.102 Tc(2a) 0.8687 This work 
B(4f) 0.1958 0.3766 
4.576 2.869 4.082     Cal.22  
4.571 2.869 4.081 Cal.21  
4.594 2.893 4.094 Cal.20  
4.572 2.869 4.080 Cal.25  
R 3 ¯ m (0 GPa) hR2.900  11.19 Tc(3a) This work 
B(6c) 0.1978 
Cmcm (0 GPa) oC12 12.473 2.893 2.922 Tc(4c) 0.3872 0.75 This work 
B(4c) 0.2514 1.25 
B(4c) 0.5 0.4609 1.25 
P6/mmm (0 GPa) hP2.958  3.399 Tc(1b) 0.5 This work 
B(2c) 0.3333 0.6667 
2.937  3.390     Cal.22  
2.935 3.373 Cal.20  
2.940 3.384 Cal.25  
P6/mmm (180 GPa) hP2.678  3.025 Tc(1b) 0.5 This work 
B(2c) 0.3333 0.6667 
TABLE II.

Calculated elastic constants Cij (in GPa) and formation enthalpy ΔH (in eV/atom) of TcB2 in hP6, hP3, oP6, hR9, and oC12 phases under different pressures, together with available calculated results.

PhaseC11C12C13C22C23C33C44C55C66ΔHReference
hP6 (0 GPa) 586 146 103   943 250  220 −1.34 This work 
558 175 103   935 245   −1.32 Cal.15  
579 156 114   949 248  212  Cal.18  
595 142 96   937 251  227  Cal.19  
609 133 98   943 256  239 −1.29 Cal.21  
608 136 111   947 256    Cal.25  
557 166 97   936 249  195  Cal.26  
oP6 (0 GPa) 543 163 137 543 88 826 183 272 212 −1.18 This work 
532 183 145 549  835 199 206 231 −1.13 Cal.21  
536 173 103 545 144 821 274 183 208  Cal.18  
534 186 145 540  833 193 305 228  Cal.25  
hR9 (0 GPa) 570 136 131   884 268  217 −1.16 This work 
oC12 (0 GPa) 643 187 122 626 165 675 210 174 247 −0.86 This work 
hP3 (0 GPa) 592 186 155   529 47  203 −0.18 This work 
615 184 156   545 58  216 −0.19 Cal.22  
617 178 156   550 72    Cal.25  
hP3 (180 GPa) 1466 642 705   1333 325  412  This work 
PhaseC11C12C13C22C23C33C44C55C66ΔHReference
hP6 (0 GPa) 586 146 103   943 250  220 −1.34 This work 
558 175 103   935 245   −1.32 Cal.15  
579 156 114   949 248  212  Cal.18  
595 142 96   937 251  227  Cal.19  
609 133 98   943 256  239 −1.29 Cal.21  
608 136 111   947 256    Cal.25  
557 166 97   936 249  195  Cal.26  
oP6 (0 GPa) 543 163 137 543 88 826 183 272 212 −1.18 This work 
532 183 145 549  835 199 206 231 −1.13 Cal.21  
536 173 103 545 144 821 274 183 208  Cal.18  
534 186 145 540  833 193 305 228  Cal.25  
hR9 (0 GPa) 570 136 131   884 268  217 −1.16 This work 
oC12 (0 GPa) 643 187 122 626 165 675 210 174 247 −0.86 This work 
hP3 (0 GPa) 592 186 155   529 47  203 −0.18 This work 
615 184 156   545 58  216 −0.19 Cal.22  
617 178 156   550 72    Cal.25  
hP3 (180 GPa) 1466 642 705   1333 325  412  This work 

To determine the phase transition sequence of TcB2 at 0 K, the enthalpy differences of hP3-TcB2, oP6-TcB2, hR9-TcB2, oC12-TcB2 with respect to the hP6-TcB2 phase were obtained within the pressure range of 0–400 GPa. The calculated results are drawn in Fig. 1. As shown, hP6-TcB2 possesses the lowest formation enthalpy at zero pressure and is considered to be the ground state. Moreover, the structural stability of hP6-TcB2 endures until 174.7 GPa. As the pressure continues to rise, the hP6 phase transforms into the hP3 phase, and the latter remains structurally stable up to 400 GPa. The crystal structures of them are illustrated in Fig. 2.

FIG. 1.

Calculated enthalpy difference with respect to hP6-TcB2 under different pressures.

FIG. 1.

Calculated enthalpy difference with respect to hP6-TcB2 under different pressures.

Close modal
FIG. 2.

Top and three-dimensional views of the predicted crystal structures. (a) hP6-TcB2 at 0 GPa and (b) hP3-TcB2 at 180 GPa. The blue and red spheres represent the B and Tc atoms, respectively.

FIG. 2.

Top and three-dimensional views of the predicted crystal structures. (a) hP6-TcB2 at 0 GPa and (b) hP3-TcB2 at 180 GPa. The blue and red spheres represent the B and Tc atoms, respectively.

Close modal

Meanwhile, the dynamic stability of hP6-TcB2 and hP3-TcB2 under pressure is further discussed, and the calculated phonon dispersion curves under different pressures are presented in Fig. 3. The results show that hP6-TcB2 and hP3-TcB2 exhibit no imaginary phonon frequencies throughout the entire Brillouin zone at 0–180 and 160–400 GPa, respectively, suggesting that these phases are dynamically stable in a large range of pressures. Furthermore, Fig. 4 illustrates the pressure–volume curve of TcB2 under pressure up to 400 GPa by fitting the data energy-volume to the Vinet equation of state.38 As shown, the volume from hP6-TcB2 to hP3-TcB2 is reduced by 4.66%, which means that the structural transformation between them should be a first-order phase transition.

FIG. 3.

Phonon dispersion curves of hP6-TcB2 at (a) 0 and (b) 180 GPa and hP3-TcB2 at (c) 160 and (d) 400 GPa, respectively.

FIG. 3.

Phonon dispersion curves of hP6-TcB2 at (a) 0 and (b) 180 GPa and hP3-TcB2 at (c) 160 and (d) 400 GPa, respectively.

Close modal
FIG. 4.

The volume of hP6-TcB2 and hP3-TcB2 as a function of pressure.

FIG. 4.

The volume of hP6-TcB2 and hP3-TcB2 as a function of pressure.

Close modal

The elastic constants of solids and their associated properties can provide essential insights into understanding the mechanical performance of materials. Thus, we initially calculated the elastic constants of hP6-TcB2, hP3-TcB2, oP6-TcB2, hR9-TcB2, and oC12-TcB2 at 0 GPa by a stress–strain method.39,40 The obtained results, together with the calculated elastic constants of hP3-TcB2 at 180 GPa and other relevant theoretical data,15,18,19,21,22,25,26 are presented in Table II. As shown, our obtained results at 0 GPa are consistent well with the available theoretical data. Moreover, it is well known that the mechanical stability of different crystal structures can be assessed using the Born–Huang stability criteria.41 After careful comparison, hP6-TcB2, oP6-TcB2, hR9-TcB2, and oC12-TcB2 all meet their mechanical stability criteria, indicating they should be mechanically stable at 0 GPa. Furthermore, the high-pressure phase hP3-TcB2 is mechanically stable at 180 GPa because its elastic constants also satisfy the mechanical stability criteria at this pressure.

In addition, it is worth noting that the C33 of hP6-TcB2, oP6-TcB2, and hR9-TcB2 are 943, 826, and 884 GPa, respectively, which is comparable to the experimentally synthesized hard material ReB2,10 implying that they may also be hard materials. To further evaluate their application potential in the field of superhard materials, the hardness characteristics of different TcB2 phases were assessed using the empirical model proposed by Chen et al.42 It can be expressed as
(1)
where B and G stand for the inherent resistance to changes in bulk and shear deformation, which can be calculated from the elastic constants Cij. The evaluation results, along with other theoretical data,18,19,21,22 are presented in Table III. At zero pressure, the Vickers hardness of hP6-TcB2 is 38.4 GPa, followed by hR9-TcB2 (37.7 GPa), oP6-TcB2 (32.8 GPa), oC12-TcB2 (27.5 GPa), and hP3-TcB2 (7.3 GPa). Generally, the Vickers hardness of superhard materials should be above 40 GPa. Therefore, hP6-TcB2, hR9-TcB2, oP6-TcB2, and oC12-TcB2 can be used as candidates for hard materials rather than superhard materials. Unfortunately, even at the pressure of 180 GPa, the Vickers hardness of the high-pressure phase hP3-TcB2 is only 20 GPa, which is not suitable for application in the field of superhard materials. In addition, a B/G ratio greater than 1.75 indicates that the material is ductile, while a ratio below this threshold signifies it is brittle.43 As can be seen from Table III, hP6-TcB2, hR9-TcB2, oP6-TcB2, and oC12-TcB2 exhibits a brittle nature at zero pressure, while hP3-TcB2 at 180 GPa shows ductility with a B/G ratio of 2.61.
TABLE III.

Calculated bulk modulus B (in GPa), shear modulus G (in GPa), Young's modulus E (in GPa), Poisson's ratio σ, and Vickers's hardness HV (in GPa) of TcB2 at different pressures, together with other theoretical results.

PhaseBGEB/GσHVReference
hP6 (0 GPa) 309 258 606 1.20 0.17 38.4 This work 
314 227 704 1.38 0.21  Cal.18  
306 260 608  0.17 37.0 Cal.19  
309 269 623  0.16 34.0 Cal.22  
oP6 (0 GPa) 294 288 544 1.02 0.19 32.8 This work 
302 240 569 1.26 0.19  Cal.21  
hR9 (0 GPa) 303 256 599 1.18 0.17 37.7 This work 
oC12 (0 GPa) 318 212 521 1.5 0.23 27.5 This work 
hP3 (0 GPa) 300 114 303 2.63 0.33 7.3 This work 
307 128 336 2.40 0.32  Cal.22  
hP3 (180 GPa) 930 357 950 2.61 0.33 20.0 This work 
PhaseBGEB/GσHVReference
hP6 (0 GPa) 309 258 606 1.20 0.17 38.4 This work 
314 227 704 1.38 0.21  Cal.18  
306 260 608  0.17 37.0 Cal.19  
309 269 623  0.16 34.0 Cal.22  
oP6 (0 GPa) 294 288 544 1.02 0.19 32.8 This work 
302 240 569 1.26 0.19  Cal.21  
hR9 (0 GPa) 303 256 599 1.18 0.17 37.7 This work 
oC12 (0 GPa) 318 212 521 1.5 0.23 27.5 This work 
hP3 (0 GPa) 300 114 303 2.63 0.33 7.3 This work 
307 128 336 2.40 0.32  Cal.22  
hP3 (180 GPa) 930 357 950 2.61 0.33 20.0 This work 
Furthermore, according to the obtained B and G, Young's modulus and Poisson's ratio can be calculated as follows:
(2)

The obtained Young's modulus E and Poisson's ratio σ are also listed in Table III. Generally, the higher Young's modulus E, the stronger the resistance to compression and tension within the elastic deformation. Therefore, the order of stiffness for TcB2 at 0 GPa should be hP6 > hR9 > oP6 > oC12 > hP3. In addition, Poisson's ratio can be used to evaluate the degree of covalent bond directionality, with larger Poisson's ratio, the weaker directionality of the covalent bond. Thus, the degree of TcB2 covalent bond directionality at 0 GPa is hR9 > hP6 > oP6 > oC12 > hP3. In order to more intuitively describe the elastic properties of different TcB2 phases, the correlations between the bulk modulus B, the shear modulus G, Young's modulus E, and hardness HV under different structures are presented in Fig. 5. Moreover, at the pressure of 180 GPa, Young's modulus of hP3-TcB2 is 950 GPa and Poisson's ratio is 0.33.

FIG. 5.

Correlations between bulk modulus B, shear modulus G, Young's modulus E, and hardness HV of TcB2 in different structures.

FIG. 5.

Correlations between bulk modulus B, shear modulus G, Young's modulus E, and hardness HV of TcB2 in different structures.

Close modal

To enhance our understanding of the electronic structure of TcB2, we depicted the density of states distribution for the hP6 phase at 160 GPa and the hP3 phase at 180 GPa in Fig. 6. As shown, the two high-pressure phases of TcB2 show electron accumulation near the Fermi level, highlighting their inherent metallic characteristics. The notable electron density of states (DOS) values at the Fermi energy level in both phases can be attributed to the contribution of Tc-3d and B-2d orbitals. When focusing on the energy interval under consideration, the total DOS of the two phases exhibits a similar basic composition, almost both being Tc-3d with a B-2p orbital superposition shape. This indicates that both phases possess strong Tc-3d and B-2p hybridization, forming strong covalent bonds.

FIG. 6.

The PDOS and –COHP of TcB2: (a) PDOS (hP6) and –COHP (hP6) at 160 GPa and (b) PDOS (hP3) and –COHP (hP3) at 180 GPa.

FIG. 6.

The PDOS and –COHP of TcB2: (a) PDOS (hP6) and –COHP (hP6) at 160 GPa and (b) PDOS (hP3) and –COHP (hP3) at 180 GPa.

Close modal

The crystal orbital Hamilton population (COHP) curve enables the assessment of the relative contribution of a specific bond to the overall bonding energy of the system, thereby facilitating the analysis of the interaction strength between two distinct types of bonds, namely, B–B bonds and Tc–B bonds. In addition, integrated crystal orbital Hamilton population (ICOHP) provides energy values after integration up to the Fermi level, which can be used to visually quantify the strength of covalent bonds. Using the LOBSTER post-processing program,44,45 we calculated of the COHP and ICOHP of the B–B and B–Tc bonds in the hP6 and hP3 phases. The obtained results are also illustrated in Fig. 6. As shown, two distinct categories of B–B bonds and two distinctive types of B–Tc bonds can be distinguished in the hP6 phase. Among them, the B–B bond with a length of 1.60 Å (ICOHP = 8.79 eV) constitutes most of the total B–B bonds. Another type of B–B bond with a length of 2.81 Å, which involves B atoms closer between two folded boron layers, forms a small amount of covalent bonding (ICOHP = 0.38 eV). A similar trend is observed for the B–Tc bonds, where there are two types of Tc–B bonds with lengths of 2.05 and 2.06 Å. Although there is only a small difference of 0.01 Å between them, their covalent bond strengths are almost twice as different (ICOHP = 4.37 eV and ICOHP = 2.46 eV, respectively).

In comparison, the high-pressure phase hP3-TcB2 exhibits a structurally more concise arrangement. The flat hexagonal boron layers overlap with the technetium layers, with technetium atoms positioned directly below the hexagonal rings [see Fig. 6(b)], resulting in simpler bonding characteristics within the crystal. Notably, the neighboring B–B bonds showcase robust covalent bonding traits, featuring an ICOHP value of 8.65 eV. However, the covalent bonding strength of the Tc–B bonds that connect boron layers with the transition metal layers is comparably weaker, with an ICOHP value of 3.78 eV. For both hP6-TcB2 and hP3-TcB2, the strong covalent bond formed by B–B bonds serves as the primary factor for maintaining the structural stability. However, the close connection between B atoms constitutes the compact B layer, so covalent bonding strength in B–Tc bonds is comparatively weaker, yet not negligible when compared to B–B bonds. Moreover, it is noteworthy that the hardness of the hP6 phase is significantly greater than that of hP3, which may be attributed to the hP6 phase having a zigzag configuration of the B atomic framework. However, under pressure, the zigzag B framework in the hP6 phase is replaced by a flatter B layer, resulting in enhanced interlayer sliding between B and Tc layers. This characteristic is also reflected in the hP3 phase, which exhibits a smaller shear modulus of 114 GPa.

In order to further comprehensively investigate the bonding characteristics between the two high-pressure phases, the electron localizability can be quantified through the Pauli principle, so as to more intuitively understand the electron distribution in distinct regions.46 The electron localization density of hP6-TcB2 and hP3-TcB2 is plotted in Fig. 7, which can visualize the bonding states between different atoms. In general, high Electron Localization Function (ELF) values are considered to form a stronger localization feature in the region, while low ELF values correspond to gaps in an atomic orbital. As can be seen in Fig. 7, the B–B bonds in both different phases of TcB2 have high ELF values, indicating the existence of strong covalent bonds between the immediately neighboring B atoms. In addition, the ELF values between the adjacent Tc–B are around 0.5, indicating that there are partial covalent bonds in Tc–B, which is consistent with our analysis from the density of states and COHP.

FIG. 7.

Contours of ELF of (a) hP6-TcB2 on the planes (110) at 160 GPa and (b) hP3-TcB2 on the planes (110) at 180 GPa.

FIG. 7.

Contours of ELF of (a) hP6-TcB2 on the planes (110) at 160 GPa and (b) hP3-TcB2 on the planes (110) at 180 GPa.

Close modal
In Sec. III A, we established the phase transition sequence of hP6-TcB2 → hP3-TcB2 under pressure at zero Kelvin, and the calculation of the phonon dispersion curves shows that the two phases are dynamically stable in a large range of pressures. For a more precise depiction of the material's phase transition behavior under high-pressure and high-temperature conditions, we established the phase diagram of TcB2 by the quasi-harmonic approximation (QHA).47 The expression for Helmholtz free energy within the quasi-harmonic method is as follows:
(3)
where E(V) is the total energy and Fvib(V,T) is the nonequilibrium vibrational Helmholtz energy, which can be expressed as
(4)
where g(ω,V) is the phonon density of states and serves as a critical parameter for investigating the physical and thermal properties of solids. In addition, Ggibbs is expressed as
(5)

In general, when the Gibbs free energy reaches its minimum value, the material exists in a stable phase. Thus, through computation and analysis of the Gibbs free energy, one can anticipate the phase transition behavior depicted in the phase diagram. The high-pressure and high-temperature phase diagram of TcB2 are shown in Fig. 8. It is obvious that temperature has a significant effect on the structural stability of TcB2. With the increase of temperature, the transition pressure of hP6-TcB2 → hP3-TcB2 gradually decreases. At a room temperature of 300 K, the phase transformation from hP6-TcB2 to hP3-TcB2 occurs at about 164 GPa.

FIG. 8.

The phase diagram of TcB2 at high pressure and finite temperature.

FIG. 8.

The phase diagram of TcB2 at high pressure and finite temperature.

Close modal

By employing a combination of particle swarm optimization for structure prediction and density functional theory calculations, we successfully predicted the stable phases of TcB2 under both ambient and high pressures and investigated their structural stability over a wide range of pressures. The phase transition sequence of hP6-TcB2 → hP3-TcB2 under pressure was established, with a transition pressure of approximately 174.9 GPa. The calculation of elastic constants and phonon dispersion curves indicates that both hP6-TcB2 and hP3-TcB2 exhibit mechanical and dynamic stability in a large range of pressures. Moreover, the hP6-TcB2 demonstrates a higher bulk modulus and lower Poisson's ratio, suggesting its potential for low-pressure compression applications. For hP3-TcB2, the transition from the folded B layer to a flatter B layer under pressure results in an increased ease of sliding between the B layer and the transition metal layer. This is also reflected in its low hardness and shear modulus. Electronic structure calculations show that both hP6-TcB2 and hP3-TcB2 exhibit characteristics of metallic behavior. Furthermore, the B–B bonds in these two phases contribute to the majority of the covalent bonds in the system, which is crucial for its structural stability. Finally, the high-pressure and high-temperature phase diagram of TcB2 was constructed for the first time using the QHA method. The results show that the transition pressure of hP6-TcB2 → hP3-TcB2 gradually decreases with rising temperature.

This work was supported by the Natural Science Basic Research Program of Shaanxi Province under Grant No. 2024JC-YBQN-0044, the National Natural Science Foundation of China under Grant No. 11904282, and the Doctoral Scientific Research Foundation of the Xi’an University of Science and Technology under Grant No. 2018QDJ029.

The authors have no conflicts to disclose.

Yi X. Wang: Conceptualization (equal); Data curation (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Writing – original draft (equal). H. Wu: Data curation (equal); Formal analysis (equal); Methodology (equal); Software (equal); Writing – original draft (equal). Wu N. Xie: Investigation (equal); Methodology (equal); Visualization (equal). Xiao F. Wang: Investigation (equal); Methodology (equal); Software (equal). Shao W. Sun: Investigation (equal); Methodology (equal); Software (equal); Writing – review & editing (equal). Jian B. Gu: Formal analysis (equal); Investigation (equal); Methodology (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
H.-Y.
Chung
,
M. B.
Weinberger
,
J. B.
Levine
,
A.
Kavner
,
J. M.
Yang
,
S. H.
Tolbert
, and
R. B.
Kaner
, “
Synthesis of ultra-incompressible superhard rhenium diboride at ambient pressure
,”
Science
316
(
5823
),
436
439
(
2007
).
2.
J. B.
Levine
,
S. H.
Tolbert
, and
R. B.
Kaner
, “
Advancements in the search for superhard ultra-incompressible metal borides
,”
Adv. Funct. Mater.
19
(
22
),
3519
3533
(
2009
).
3.
H.
Sumiya
,
N.
Toda
, and
S.
Satoh
, “
Mechanical properties of synthetic type IIa diamond crystal
,”
Diam. Relat. Mater.
6
(
12
),
1841
1846
(
1997
).
4.
M.
Magnuson
,
L.
Hultman
, and
H.
Högberg
, “
Review of transition-metal diboride thin films
,”
Vacuum
196
,
110567
(
2022
).
5.
R. B.
Kaner
,
J. J.
Gilman
, and
S. H.
Tolbert
, “
Designing superhard materials
,”
Science
308
(
5726
),
1268
1269
(
2005
).
6.
Y. X.
Wang
,
Z. X.
Yan
,
W.
Liu
, and
G. L.
Zhou
, “
Novel high pressure structures and mechanical properties of rhenium diboride
,”
J. Appl. Phys.
126
(
13
),
135901
(
2019
).
7.
R. W.
Cumberland
,
M. B.
Weinberger
,
J. J.
Gilman
,
S. M.
Clark
,
S. H.
Tolbert
, and
R. B.
Kaner
, “
Osmium diboride, an ultra-incompressible, hard material
,”
J. Am. Chem. Soc.
127
(
20
),
7264
7265
(
2005
).
8.
J. V.
Rau
and
A.
Latini
, “
New hard and superhard materials: RhB1.1 and IrB1.35
,”
Chem. Mater.
21
(
8
),
1407
1409
(
2009
).
9.
A. F.
Young
,
C.
Sanloup
,
E.
Gregoryanz
,
S.
Scandolo
,
R. J.
Hemley
, and
H.
Mao
, “
Synthesis of novel transition metal nitrides IrN2 and OsN2
,”
Phys. Rev. Lett.
96
(
15
),
155501
(
2006
).
10.
J.
Qin
,
D.
He
,
J.
Wang
,
L.
Fang
,
L.
Lei
,
Y.
Li
,
J.
Hu
,
Z.
Kou
, and
Y.
Bi
, “
Is rhenium diboride a superhard material?
,”
Adv. Mater.
20
(
24
),
4780
4783
(
2008
).
11.
A.
Latini
,
J. V.
Rau
,
R.
Teghil
,
A.
Generosi
, and
V. R.
Albertini
, “
Superhard properties of rhodium and iridium boride films
,”
ACS Appl. Mater. Interfaces
2
(
2
),
581
587
(
2010
).
12.
J.
Yang
,
H.
Sun
, and
C.
Chen
, “
Is osmium diboride an ultra-hard material?
,”
J. Am. Chem. Soc.
130
(
23
),
7200
7201
(
2008
).
13.
W.
Trzebiatowski
and
J.
Rudzinski
, “
The composition and structure of technetium nitride and technetium borides
,”
J. Common Met.
6
(
3
),
244
245
(
1964
).
14.
E. A.
Juarez-Arellano
,
B.
Winkler
,
A.
Friedrich
,
L.
Bayarjargal
,
V.
Milman
,
J.
Yan
, and
S. M.
Clark
, “
Stability field of the high-(P, T) Re2C phase and properties of an analogous osmium carbide phase
,”
J. Alloys Compd.
481
(
1
),
577
581
(
2009
).
15.
J. H.
Wu
and
G.
Yang
, “
Phase stability and physical properties of technetium borides: A first-principles study
,”
Comput. Mater. Sci.
82
,
86
91
(
2014
).
16.
D. R.
Armstrong
, “
The electronic structure of ReB2 and TcB2
,”
J. Common Met.
67
(
1
),
191
203
(
1979
).
17.
A.
Pallas
and
K.
Larsson
, “
Structure determination of the 4d metal diborides:  A quantum mechanical study
,”
J. Phys. Chem. B
110
(
11
),
5367
5371
(
2006
).
18.
Y. X.
Wang
, “
Elastic and electronic properties of TcB2 and superhard ReB2: First-principles calculations
,”
Appl. Phys. Lett.
91
(
10
),
101904
(
2007
).
19.
S.
Aydin
and
M.
Simsek
, “
First-principles calculations of MnB2, TcB2, and ReB2 within the ReB2-type structure
,”
Phys. Rev. B
80
(
13
),
134107
(
2009
).
20.
E.
Deligoz
,
K.
Colakoglu
,
H. B.
Ozisik
, and
Y. O.
Ciftci
, “
Lattice dynamical properties of TcB2 compound
,”
Solid State Sci.
14
(
7
),
794
800
(
2012
).
21.
M. M.
Zhong
,
X. Y.
Kuang
,
Z. H.
Wang
,
P.
Shao
,
L. P.
Ding
, and
X. F.
Huang
, “
Phase stability, physical properties, and hardness of transition-metal diborides MB2(M = Tc, W, Re, and Os): First-principles investigations
,”
J. Phys. Chem. C
117
(
20
),
10643
10652
(
2013
).
22.
F. G.
Kuang
,
X. Y.
Kuang
,
S. Y.
Kang
, and
X. F.
Huang
, “
Theoretical investigation of pressure-induced effects on mechanical characteristics of TcB2: A potential hard material
,”
Curr. Inorg. Chem.
5
(
2
),
143
150
(
2015
).
23.
C.
Ying
,
E.
Zhao
,
L.
Lin
, and
Q.
Hou
, “
Structural determination and physical properties of 4d transitional metal diborides by first-principles calculations
,”
Mod. Phys. Lett. B
28
(
27
),
1450213
(
2014
).
24.
C.
Ying
,
T.
Liu
,
L.
Lin
,
E.
Zhao
, and
Q.
Hou
, “
New predicted ground state and high pressure phases of TcB3 and TcB4: First-principles
,”
Comput. Mater. Sci.
144
,
154
160
(
2018
).
25.
W.
Chen
and
J. Z.
Jiang
, “
Elastic and electronic properties of low compressible 4d transition metal diborides: First principles calculations
,”
Solid State Commun.
150
(
43–44
),
2093
2096
(
2010
).
26.
X.
Miao
,
W.
Xing
,
F.
Meng
, and
R.
Yu
, “
Prediction on technetium triboride from first-principles calculations
,”
Solid State Commun.
252
,
40
45
(
2017
).
27.
Y.
Wang
,
J.
Lv
,
L.
Zhu
, and
Y.
Ma
, “
Crystal structure prediction via particle-swarm optimization
,”
Phys. Rev. B
82
(
9
),
094116
(
2010
).
28.
Y.
Wang
,
J.
Lv
,
L.
Zhu
, and
Y.
Ma
, “
CALYPSO: A method for crystal structure prediction
,”
Comput. Phys. Commun.
183
(
10
),
2063
2070
(
2012
).
29.
L.
Zhu
,
H.
Wang
,
Y.
Wang
,
J.
Lv
,
Y.
Ma
,
Q.
Cui
,
Y.
Ma
, and
G.
Zou
, “
Substitutional alloy of Bi and Te at high pressure
,”
Phys. Rev. Lett.
106
(
14
),
145501
(
2011
).
30.
X.
Li
and
F.
Peng
, “
Superconductivity of pressure-stabilized vanadium hydrides
,”
Inorg. Chem.
56
(
22
),
13759
13765
(
2017
).
31.
S.
Zhang
,
J.
Lin
,
Y.
Wang
,
G.
Yang
,
A.
Bergara
, and
Y.
Ma
, “
Nonmetallic FeH6 under high pressure
,”
J. Phys. Chem. C
122
(
22
),
12022
12028
(
2018
).
32.
Y. X.
Wang
,
Y. Y.
Liu
,
Z. X.
Yan
,
W.
Liu
,
G. L.
Zhou
, and
K. Z.
Xiong
, “
Crystal structures and mechanical properties of osmium diboride at high pressure
,”
Sci. Rep.
11
(
1
),
5754
(
2021
).
33.
V.
Milman
,
B.
Winkler
,
J. A.
White
,
C. J.
Pickard
,
M. C.
Payne
,
E. V.
Akhmatskaya
, and
R. H.
Nobes
, “
Electronic structure, properties, and phase stability of inorganic crystals: A pseudopotential plane-wave study
,”
Int. J. Quantum Chem.
77
(
5
),
895
910
(
2000
).
34.
M. C.
Payne
,
M. P.
Teter
,
D. C.
Allan
,
T. A.
Arias
, and
J. D.
Joannopoulos
, “
Iterative minimization techniques for ab initio total-energy calculations: Molecular dynamics and conjugate gradients
,”
Rev. Mod. Phys.
64
(
4
),
1045
1097
(
1992
).
35.
J. P.
Perdew
,
K.
Burke
, and
M.
Ernzerhof
, “
Generalized gradient approximation made simple
,”
Phys. Rev. Lett.
77
(
18
),
3865
3868
(
1996
).
36.
D. J.
Chadi
, “
Special points for Brillouin-zone integrations
,”
Phys. Rev. B
16
(
4
),
1746
1747
(
1977
).
37.
K.
Parlinski
,
Z. Q.
Li
, and
Y.
Kawazoe
, “
First-principles determination of the soft mode in cubic ZrO2
,”
Phys. Rev. Lett.
78
(
21
),
4063
4066
(
1997
).
38.
P.
Vinet
,
J. H.
Rose
,
J.
Ferrante
, and
J. R.
Smith
, “
Universal features of the equation of state of solids
,”
J. Phys.: Condens. Matter
1
(
11
),
1941
(
1989
).
39.
Y. X.
Wang
,
Q.
Wu
,
X. R.
Chen
, and
H. Y.
Geng
, “
Stability of rhombohedral phases in vanadium at high-pressure and high-temperature: First-principles investigations
,”
Sci. Rep.
6
,
32419
(
2016
).
40.
Y. X.
Wang
,
H.
Wu
,
Y. Y.
Liu
,
H.
Wang
,
X. R.
Chen
, and
H. Y.
Geng
, “
Recent progress in phase stability and elastic anomalies of group VB transition metals
,”
Crystals
12
(
12
),
1762
(
2022
).
41.
Z.
Wu
,
E.
Zhao
,
H.
Xiang
,
X.
Hao
,
X.
Liu
, and
J.
Meng
, “
Crystal structures and elastic properties of superhard IrN2 and IrN3 from first principles
,”
Phys. Rev. B
76
(
5
),
054115
(
2007
).
42.
X. Q.
Chen
,
H.
Niu
,
D.
Li
, and
Y.
Li
, “
Modeling hardness of polycrystalline materials and bulk metallic glasses
,”
Intermetallics
19
(
9
),
1275
1281
(
2011
).
43.
R.
Hill
, “
The elastic behaviour of a crystalline aggregate
,”
Proc. Phys. Soc. A
65
(
5
),
349
(
1952
).
44.
S.
Steinberg
and
R.
Dronskowski
, “
The crystal orbital Hamilton population (COHP) method as a tool to visualize and analyze chemical bonding in intermetallic compounds
,”
Crystals
8
(
5
),
225
(
2018
).
45.
J.
Hempelmann
,
P. C.
Müller
,
P. M.
Konze
,
R. P.
Stoffel
,
S.
Steinberg
, and
R.
Dronskowski
, “
Long-range forces in rock-salt-type tellurides and how they mirror the underlying chemical bonding
,”
Adv. Mater.
33
(
37
),
2100163
(
2021
).
46.
A.
Savin
,
A. D.
Becke
,
J.
Flad
,
R.
Nesper
,
H.
Preuss
, and
H. G.
von Schnering
, “
A new look at electron localization
,”
Angew. Chem. Int. Ed. Engl.
30
(
4
),
409
412
(
1991
).
47.
A.
Otero-de-la-Roza
,
D.
Abbasi-Pérez
, and
V.
Luaña
, “
Gibbs2: A new version of the quasiharmonic model code: II: Models for solid-state thermodynamics, features and implementation
,”
Comput. Phys. Commun.
182
(
10
),
2232
2248
(
2011
).