Introduction

Transition metal nitrides (TMxNy), synthesized under high-pressure and high-temperature conditions, represent a prominent class of materials exhibiting extreme usefulness in a wide variety of industrial applications1,2. When it comes to their superior mechanical properties such as high hardness and elastic moduli, most of the early transition metal mononitrides (TMN), and in particular TiN and CrN are well known hard materials and are widely used in various industrial applications, such as cutting tools or wear-resistant coatings3,4. Taking advantage of high-pressure techniques, two family members of hard nitrides (Zr3N4 and Hf3N4) of the group IVB with TM3N4 stoichiometry were successfully synthesized5, opening a promising way to obtain other nitrides with N:TM > 1 under high nitrogen pressure. Compared to early transition metals, the noble metals (TM = Ru, Rh, Pd, Os, Ir, and Pt) were previously known to hardly form nitrides with high nitrogen content. Until 2004, a novel platinum nitride with ultra-high incompressibility was obtained under extreme conditions (50 GPa and 2000 K) by Gregoryanz et al.6 and was finally determined to crystallize in the pyrite structure with a stoichiometry of 1:27,8. Thereafter, there have been considerable researches to search for other transition metal dinitrides, and so far as we know, the OsN28,9, IrN27,9,10, PdN210, RhN211, and recently RuN212 have been experimentally obtained in a direct chemical reaction between platinum group elements and molecular fluid nitrogen at high pressures and temperatures. The follow-up studies of their structures and mechanical properties have stimulated significant in their potential applications. These works have been motivated by the design of intrinsic (super)hard materials proposed by Kaner et al.13,14 that the introduction of light and covalent-bond-forming elements, such as B, C, N, and O into the transition metal (TM) lattices with highly valence-electron density is expected to enhance the shear strength against plastic deformations.

More recently, a new transition metal dinitride, TiN215, was successfully synthesized at 73(3) GPa and 2400(40) K by choice of TiN and dense N2 as starting materials. The experiment revealed that this new dinitride adopts a tetragonal CuAl2-type structure at high pressure, which is in agreement with previous theoretical prediction performed by Yu et al.16. On decompression, the experiment found that this phase is recoverable to ambient conditions and possesses a high bulk modulus of 385(7) GPa comparable to those of PtN2 (372 GPa)6 and ReB2 (360 GPa)17, much larger than that of TiN (288 GPa)18. Therefore, this new tetragonal TiN2, the first synthesized high-nitride phase in early transition metal nitrides, is expected to be a candidate as a potential superhard solid for wear- and scratch-resistant materials. However, this concept for the search of novel superhard materials failed in materials such as PtN219, and ReB220, and others21,22,23, because plastic deformation occurs in shear at large strain at the atomic level, where electronic instabilities may occur upon bond breaking in the practical measurement of hardness. Meanwhile, the macroscopic behavior of a solid is strongly related to its elastic anisotropic properties, which can reveal, in some materials, an anisotropy degree decidedly non-negligible and in some cases so extreme to suggest the proximity of material instability. Accordingly, here, we have extended the mechanical behaviors of TiN2 and presented in detail the variations of the elastic moduli along the arbitrary directions. Moreover, the stress-strain relations and the underlying atomistic bond breaking processes under the applied strains were also systematically investigated to provide a deeper insight into mechanical properties and hardness of the newly discovered TiN2. We have also applied this novel tetragonal structure to other two family members ZrN2 and HfN2 to explore the possibilities and provide guidance for future experimental efforts. We hope that the present findings will encourage further theoretical and experimental works on this class of material.

Results and Discussion

The experiment has demonstrated that TiN2 crystallizes in the tetragonal CuAl2-type structure with Ti and N atoms sitting at 4a and 8 h sites in a unit cell, as shown in Fig. 1(a). Polyhedral view of this tetragonal structure (Fig. 1(b)) reveals that TiN2 consists of the TiN8 face-sharing tetragonal antiprisms connected by N-N bonds and stacked along the c-axis, in contrast to the TMN6 octahedrons in the previous synthesized noble metals pernitrides7,8,9,10,11,12. Through the full relaxations of both lattice constants and internal atomic coordination, the obtained equilibrium structure parameters for three TMN2 compounds are listed in Table 1, among which the calculated results for TiN2 compare well with the available experimental data15. For ZrN2 and HfN2, however, there are no available experimental data for comparison and the present results could provide useful information for further experimental or theoretical investigations. According to the recent experiment by Bhadram et al.15, the pressure dependences of unit cell volume and lattice constants of TiN2 were calculated and plotted in Fig. 2, along with the experimental data15 and theoretical results of ZrN2 and HfN2. First, one can see that the calculated results for TiN2 are in agreement with the experimental data under pressure, and the incompressibility of TiN2 (Fig. 2(a)) is almost identical to that of HfN2, but larger than that of ZrN2. Furthermore, from Fig. 2(b), it can be seen that the incompressibility along the a-axis is larger than that along the c axis for each TMN2 compound, indicating their clear elastic anisotropy. Second, the E-V data under pressures deduced from the Fig. 2(a) for each TMN2 are fitted to the third order Birch-Murnaghan equation of state (EOS)24. The obtained the bulk modulus (B0) and its pressure derivative (B0′) for TiN2 are 276 GPa and 4.362 (see Table 1), which are lower than those of experimental data (385 GPa and 1.45), but consistent with the theoretical values (293 GPa and 3.7) predicted by Bhadram et al. using the same approach15. The low value of B0′ related to B0 in this discrepancy has been elucidated in this experimental work. Third, the fitted B0 values of TiN2 and HfN2 are nearly equivalent but larger than that of ZrN2, which is in accord with the calculated compressibility of volume plotted in Fig. 2(a). Overall, the accuracy of the present calculations for TiN2 is made quite satisfactory with the experimental data in Table 1 and Fig. 2, which supplies the safeguard for the following studies.

Table 1 Calculated Crystal lattices (Å), Unit cell volume (Å3), Bond length (Å, dTM-N: d1; dN-N: d2), EOS fitted Bulk modulus B0 (GPa) and its pressure derivative B0′ for each tetragonal TMN2.
Figure 1
figure 1

(a) Crystal structure of tetragonal TiN2 and (b) its polyhedral view. The large and small spheres represent Ti and N atoms, respectively.

Figure 2
figure 2

(a) The calculated normalized volumes and (b) lattice parameters as a function of pressure for tetragonal TMN2 compounds.

According to synthetic conditions of TiN2 proposed by Bhadram et al.15, the thermodynamic feasibility of ZrN2 and HfN2 is evaluated through the formation enthalpy (energy) calculations. The formation enthalpy ΔHf of each TMN2 with respect to the TMN and nitrogen at ambient conditions based on the reaction route: was quantified, where the fcc TMN phase and α-N2 phase are chosen as the reference phases. As listed in Table 1, the calculated formation enthalpies of three TMN2 dinitrides are all positive values, indicating that they are all metastable at ambient conditions. It is to be noted that the calculated formation enthalpies of ZrN2 (0.372 eV/atom) and HfN2 (0.328 eV/atom) are all close to that of TiN2 (0.398 eV/atom), which has been synthesized at 73(3) GPa and 2400(40) K by choice of TiN and dense N2 as starting materials. Thus, the syntheses of the ZrN2 and HfN2 could be expected at similar high pressure and temperatures conditions. The experiment has suggested that TiN2 can be quenchable to ambient conditions, and the dynamical stabilities of ZrN2 and HfN2 at 0 GPa have been thus carefully checked by the full phonon dispersions calculations using the 2 × 2 × 2 supercell method. Figure 3(a,b) show the phonon dispersion curves which confirm the dynamic stability of ZrN2 and HfN2 as there are no imaginary modes in the whole Brillouin zone. The lower frequencies of the phonon density of states are dominated by lattice dynamics of heavy TM atoms and higher frequencies by light N atoms.

Figure 3
figure 3

Phonon dispersion curves of TMN2 at ambient pressure: (a) ZrN2 and (b) HfN2.

The total and projected density of states (DOS) of each TMN2 at ambient pressure was plotted to further elaborate the electronic bonding feature, as shown in Fig. 4(a–c), respectively. All TMN2 compounds show metallic bonding because of finite value of DOS at the Fermi level (EF), which originates mostly from the TM-d orbitals and the N-p orbitals. The major orbital occupancy in the energy range of −8–0 eV stems from the strong hybridized states of TM-d and N-p orbitals, as the usual cases in the most TMxNy compounds. The typical feature of the total DOS is the presence of a “pseudogap” (a sharp valley around the EF), which is supposed the borderline between the bonding and antibonding states25,26,27. For TiN2, it is noteworthy that the bonding states are completely filled with the Fermi energy located exactly at the “pseudogap”. For ZrN2 and HfN2 (see Fig. 4(b,c)), it is found that the EF shifts toward the higher energy and lies left at the pseudogap with a relative more electronic density of states [N(EF)]. It is known that for the most stable structure there is enough room to accommodate all its valence electrons into bonding states so as to bring the EF to a valley position separating bonding and antibonding states (pseudogap) favorable for structural stability. Therefore, the TiN2 is energetically more favorable compared to the ZrN2 and HfN2 in the tetragonal phase. Figure 4(d–f) offers the calculated crystal overlap Hamilton population (COHP)28 for the TM-N and the N-N bonding inside TiN2, ZrN2, and HfN2, respectively. For the TM-N combinations in all plots, there are only bonding states in the entire occupied regions, and antibonding states show up in the unoccupied crystal orbitals, well above the EF. For the N-N combinations in TMN2, the antibonding 1πg* states (starting near −3.5 eV) are almost completely occupied at the top of the conduction band, and a portion of the metallic nature can be ascribed to these states being occupied at the EF. For TiN2, this point has been addressed in a recent work by Yu et al.16. As demonstrated in previous work29, for the case of PtN2, charge transferred from Pt to N (1.05 e) results in the full filling of antibonding 1πg* states of N24− and leads to the elongation of N-N bonds. In a similar way, this mechanism is also applicable to the case of tetragonal TiN2 although antibonding states are not completely filled and there are differences in electronic and structural configurations, as suggested by Bhadram et al.15 and Yu et al.16. Consequently, a charge balance of N24− in these TMN2 is a good working hypothesis, and this leaves the TM atoms in TMN2 in a d0 configuration. In order to compare the “ionicity” of the three dinitrides, we also analyzed the charge density topology through the Bader charge analyses30. The calculated charges of the three nitrides show decreasing trends from Hf 2.13N2−2.13 to Zr1.96 N2−1.96 and Ti1.75N2−1.75, indicating the relatively lower polarity of Ti-N bond. Meanwhile, it has been demonstrated31 that the shortening of the N-N bond is ascribed to the decrease in charge transfer from TM to N (qtrans) when one monitors the pernitrides from early to late TM elements. It can be seen that as the TM element moves from Hf through Zr to Ti, as qtrans from 2.13 e through 1.96 e to 1.75 e, and as dN-N from 1.461 Å through 1.434 Å to 1.385 Å.

Figure 4
figure 4

Total and projected DOS of TMN2: (a) TiN2, (b) ZrN2, and (c) HfN2. Projected COHP curves of various bonds in TMN2: (d) TiN2, (e) ZrN2, and (f) HfN2. The Fermi level (EF) is indicated by vertical dashed lines.

For potential engineering applications, the elastic stabilities, incompressibility, and rigidity of three TMN2 dinitrides are determined from the calculated elastic constants by applying a set of given strains with a finite variation between −0.01 and +0.01. Table 2 summarizes the calculated single-crystal elastic constants Cij and derived Hill elastic moduli as well as Poisson’s ratios of TMN2 dinitrides and compares them with those of typical hard substances TMN (TM = Ti, Zr, and Hf)32,33,34,35. The calculated six independent elastic constants of TiN2 agree well with recent theoretical results16, and the derived bulk moduli of three TMN2 dinitrides also accord well with those directly obtained from the fitting of the third-order Birch-Murnaghan EOS (see Table 1), demonstrating the reliability of the present calculations. The mechanical stabilities of three dinitrides satisfy the Born-Huang criterion36 for a tetragonal crystal [C11 > 0, C33 > 0, C44 > 0, C66 > 0, (C11 − C12) > 0, (C11 + C33 − 2C13) > 0, and 2(C11 + C12) + C33 + 4C13 > 0], indicating their mechanically stable at ambient conditions. From Table 2, the high-incompressible nature of TMN2 is disclosed by the calculated bulk modulus (TiN2: 276 GPa, ZrN2: 250 GPa, HfN2: 275 GPa), originating from the covalent TMN8 polyhedrons connected by the strong N-N covalent bonds in systems. Meanwhile, these values are comparable with the corresponding theoretical calculations and experimental data (in brackets) of typical hard transition metal mononitrides TMN, TiN: 278 GPa (288 GPa), ZrN: 250 GPa (215 GPa), HfN: 273 GPa (306 GPa). The critical values of the ratio of shear modulus G to bulk modulus B of about 0.57 separates brittle (G/B > 0.57) and ductile (G/B < 0.57) materials. For three TMN2 dinitrides, their G/B values (TiN2: 0.707, ZrN2: 0.608, HfN2: 0.626) are all larger than 0.57, implying that they are intrinsically brittle. The theoretical Vickers hardness Hv of each TMN2 was estimated by using the Chen’s empirical model37, Hv = 2(k2G)0.585 − 3. The calculated hardness value for TiN2, ZrN2, and HfN2 is 26.1 GPa, 18.1 GPa, and 23.2 GPa, respectively, making them potentially interesting for applications as hard coating materials. By using the Bader atoms-in-molecules (AIM) method, the strong covalent nature of the N-N and TM-N bonds in TMN2 were quantitatively revealed by the evidences of local charge densities at their bond critical points (BCPs) with negative Laplacian values. The obtained who can measure the bond strength related to the mechanical behaviors located at N2 dumbbells and TM-N bonds decrease in the sequence of TiN2: (2.324 e3) > ZrN2 (2.047 e3) ≈ HfN2: (2.045 e3) and TiN2: (0.457 e3) > HfN2: (0.444 e3) > ZrN2 (0.428 e3), respectively. Therefore, compared to TiN2 and HfN2, the ZrN2 exhibits the lowest moduli and hardness. Next we investigate the mechanical anisotropy of tetragonal TMN2 by calculating the orientation dependences of the Young’s modulus E and shear modulus G which can be determined from the elastic compliance constants sij38. The computational details of elastic moduli-crystal orientation dependences conducted here are presented in the Supporting information section. Figure 5 illustrates the three-dimensional surface representation showing the variation of Young’s modulus with direction for each dinitride. Clearly, all three TMN2 dinitrides exhibit a well-pronounced elastic anisotropy due to their three-dimensional pictures show a large deviation from the spherical shape, which qualifies an isotropic medium. From Fig. 5(a–c), the calculated Emax/Emin ratio of the Young’s moduli for TiN2, ZrN2, and HfN2 is 2.115, 2.543, 2.766, respectively. The Emax/Emin ratios for TiN2 and ZrN2 are much larger than those of fcc TiN (1.148) and ZrN (1.539) proposed by Brik et al.39, suggesting that the TMN2 with a larger elastic anisotropy may impose certain limitations on their possible applications. More specifically, the directional Young’s moduli along tensile axes within (001), (100), and specific planes are plotted in Fig. 6(a–c). For example, the variation of Young’s modulus in the (001) crystal plane for the quadrant of directions [uvw] between [100] (θ = 0°) and [010] (θ = 90°), the TiN2/ZrN2/HfN2 exhibits a maximum of E[110] = 734/631/725 GPa and a minimum of E[100] = E[010] = 367/250/262 GPa, respectively. From Fig. 6(a–c), the ordering of Young’s modulus as a function of direction for three TMN2 dinitrides is E[110] > E[001] > E[111] > E[011] > E[100]. Similarly, the orientation dependences of the shear modulus G were also conducted for shear on (001), (100), andplanes. From Fig. 6(d), the shear modulus of the TiN2 is independent of the shear stress from [100] to [010] directions within (001) basal plane, and the TiN2 possesses its minimum value for shear on [110] and its maximum value for shear on (100)[010] (G(100)[010] = 343 GPa). The similar cases can be also found for ZrN2 and HfN2 in Fig. 6(e,f).

Table 2 Calculated Elastic constants Cij, Bulk modulus B, Shear modulus G, and Young’s modulus E (in units of GPa) for each tetragonal TMN2.
Figure 5
figure 5

Three-dimensional surface representations of the Young’s modulus E for TMN2: (a) TiN2, (b) ZrN2, and (c) HfN2.

Figure 6
figure 6

Orientation dependences of the Young’s modulus E for TMN2: (a) TiN2, (b) ZrN2, (c) and HfN2. Orientation dependence of the Shear modulus G for TMN2: (d) TiN2, (e) ZrN2, and (f) HfN2.

To determine the electronic and structural stabilities as well as the ideal strengths of three TMN2 compounds, the stress-strain relations upon tension and shear for tetragonal TMN2 phase are calculated in some main crystallographic directions through projection of a 12-atom unit cell onto the corresponding crystal axes with one axis parallel to the strain direction for tension deformation, or with one axis parallel to the slip direction and another axis perpendicular to the slip plane for shear deformation. The schematic of tensile/shear deformation and the ideal strengths deduced from the stress-strain curves for three TMN2 compounds are shown in Fig. 7. From Fig. 7(a–c), one can see that the calculated tensile strengths show a similar anisotropy for all three compounds. It shows that all three TMN2 have strong stress responses in the [110] directions (TiN2: 74.14 GPa, ZrN2: 64.95 GPa, HfN2: 69.59 GPa) that accord well with their largest directional Young’s moduli (see Fig. 6), which measure the resistance against uniaxial tensions. However, the weakest tensile strength along [011] with the peak tensile stresses below 20 GPa for TMN2 (TiN2: 19.15 GPa, ZrN2: 14.88 GPa, HfN2: 16.53 GPa) is much lower than those of 40 GPa for PtN219 and 31.1 GPa for TiN40 along the [100] directions. The anisotropy ratio of tensile strength (σmax:σmin) for TiN2 (3.87) is smaller than those of ZrN2 (4.36) and HfN2 (4.21). Meanwhile, the shear strengths upon large strains for three TMN2 are presented in Fig. 7(d–f) in order to further examine the shear deformation where plastic deformation proceeds irreversibly on the atomic scale. First, the highest shear strength for TMN2 is found under the (100)[010] direction compare well with their largest shear modulus orientation in the (100) principal shear plane shown in Fig. 6(d–f). Second, the values of the ideal shear strength τ and shear strain γ of the weakest system is TiN2: (, γ = 0.166), ZrN2: (, γ = 0.188), and HfN2: (, γ = 0.177), which is basically lower than that41 of TiN: (, γ = 0.21), ZrN: (, γ = 0.17), and HfN: (, γ = 0.15), respectively, showing their lower shear resistance or hardness than these known hard wear-resistant materials. Third, the lowest shear strength of TMN2 is lower than the lowest tensile strength. This means the failure mode in tetragonal TMN2 phase is dominated by the shear type.

Figure 7
figure 7

Calculated tensile stress-strain relations for TMN2: (a) TiN2, (b) ZrN2, (c) and HfN2. Calculated shear stress-strain relations for TMN2: (d) TiN2, (e) ZrN2, and (f) HfN2.

To further illustrate the atomistic deformation mechanism and the origin of the intriguing bond-breaking pattern of such novel materials in engineering applications, take TiN2 for example, we further investigate the variations of bond lengths and electronic structures as a function of applied strain along (110)directions. As presented in Fig. 8 where there are two types bond lengths [the Ti-N (2.201 Å) and N-N (1.385 Å) bond length is denoted as d1 and d2, respectively] in TiN2 at equilibrium state. Under increasing shear strains, the N-N lengths denoted as d2 remain nearly invariant (d2 = 1.385 Å at γ = 0 and d2 = 1.380 Å at the critical shear strain of γ = 0.166). The Ti-N length indicated as d1 in TiN8 building block is split from one bond distance to eight different bond distances denoted as ln (n = 1, 2, …) (see the inset in Fig. 8). In Fig. 8, the Ti-N bond lengths indicated as l1, l4, l5, l7, and l8 decrease in the whole studied shear strain range, on the contrary, the l2, l3, and l6 bonds in TiN8 polyhedrons increase conformably at each strain. Especially, the stretched Ti-N bonds denoted as l6 increases sharply and breaks at the critical shear strain of γ = 0.166, which limits the achievable strengths of TiN2. Such a bond-breaking can also be clearly seen from the selected Electronic Localization Function (ELF)42,43 distributions of TiN2 on (110) plane before and after shear instability. At equilibrium state (γ = 0, see Fig. 9(a)), a certain electron localization can be seen in the region between adjacent N and Ti atoms indicative of ionic bonding, whereas the electron localization located between N-Ti (l6 bonds) atoms decreases gradually upon the incremental shear strains [(110)direction] from Fig. 9(c,d). For ELF at strain of γ = 0.188 presented in Fig. 9(d), where no electron localized at l6 bonds and results in the breaking of this bond. Therefore, the shear-induced structural deformation for tetragonal TiN2 can be attributed to the collapse of TiN8 polyhedrons by simultaneously breaking of l6 bonds, and this is also the case for other two family members of ZrN2 and HfN2.

Figure 8
figure 8

Calculated bond lengths of Ti-N (ln) as a function of strain along (110) shear directions.

The dashed line represents the shear induced structural deformation firstly occurrence.

Figure 9
figure 9

The development of ELF distributions between N-Ti (l6 bond) on (110) plane in tetragonal TiN2 at selected shear strains (a) γ = 0.0, (b) γ = 0.145, (c) γ = 0.166, and (d) γ = 0.188.

Conclusions

To conclude, the structural, electronic, and mechanical properties as well as the ideal strengths of the recent synthesized tetragonal TiN2 and two family members, yet-to-be-synthesized ZrN2 and HfN2 have been systematically studied by using first-principles calculations. Phonon dispersion and formation enthalpies calculations suggest that three tetragonal TMN2 are all dynamically stable at ambient condition and can be synthesizable at readily attainable pressures. The high-incompressible of TMN2 is associated with the strong N-N covalent bonding in N2 dumbbells and polar covalent bonding between TM and N atoms in TMN8 building blocks. However, as compared with known fcc TMN, all these tetragonal TMN2 exhibit a much larger elastic anisotropy and substantially lower shear strength, which may impose certain limitations on their possible applications. Detailed analyses of the deformed atomic structures under shear strain reveal that the lattice instability of TMN2 is due to the collapse of TMN8 polyhedrons by simultaneously breaking of TM-N bonds which limits their achievable strength.

Methods

All first-principles plane wave calculations were performed using the VASP code44 in the framework of density functional theory with the generalized-gradient approximation (GGA) proposed by Perdew-Burke-Ernzerhof (PBE) exchange-correlation functional45,46. The electron-ion interaction was described by the frozen-core all-electron projector augmented wave (PAW) method47, which called for a d-electron as valence states. The integration in the Brillouin zone for all transition metals dinitrides was employed using the Monkhorst-Pack scheme48 (8 × 8 × 6), an energy cutoff of 600 eV for the plane-wave expansions, and a tetrahedron method with Blöch corrections for energy calculations and Gaussian smearing for the stress calculations. The conjugate gradient method was used for the relaxation of structural parameters. Phonon frequencies were calculated using direct supercell49, which uses the forces obtained by the Hellmann-Feynaman theorem. Chemical bonding analyses were performed by means of the crystal orbital Hamilton population (COHP) method as implemented in the LOBSTER code50,51. The independent elastic constants were determined from evaluation of stress tensor generated small strain and bulk modulus, shear modulus, Young’s modulus, and Poisson’s ratio were thus estimated by the Voigt-Reuss-Hill approximation. The stress-strain relationships were calculated by incrementally deforming the model cell in the direction of the applied strain, and simultaneously relaxing the cell basis vectors conjugated to the applied strain, as well as the positions of atoms inside the cell, at each step.

Additional Information

How to cite this article: Zhang, M. et al. Electronic bonding analyses and mechanical strengths of incompressible tetragonal transition metal dinitrides TMN2 (TM = Ti, Zr, and Hf). Sci. Rep. 6, 36911; doi: 10.1038/srep36911 (2016).

Publisher's note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.