The search for new superhard materials has usually focused on strong covalent solids. It is, however, a huge challenge to design superhard metals because of the low resistance of metallic bonds against the formation and movement of dislocations. Here, we report a microscopic mechanism of enhancing hardness by identifying highly stable thermodynamic phases and strengthening weak slip planes. Using the well-known transition-metal borides as prototypes, we demonstrate that several low borides possess unexpectedly high hardness whereas high borides exhibit an anomalous hardness reduction. Such an unusual phenomenon originates from the peculiar bonding mechanisms in these compounds. Furthermore, the low borides have close compositions, similar structures, and degenerate formation energies. This enables facile synthesis of a multiphase material that includes a large number of interfaces among different borides, and these interfaces form nanoscale interlocks that strongly suppress the glide dislocations within the metal bilayers, thereby drastically enhancing extrinsic hardness and achieving true superhard metals. Therefore, this study not only elucidates the unique mechanism responsible for the anomalous hardening in this class of borides but also offers a valid alchemy to design novel superhard metals with multiple functionalities.

Hardness is the resistance of materials against localized plastic deformation, corresponding to the formation and movement of dislocations, and thus hard and superhard materials are very useful for a wide range of applications from reducing wear to creating machining tools.1 Traditional superhard materials like diamond and cubic boron nitride (cBN) are made up of light elements. Although these pure covalent solids possess exceptional hardness, they are thermodynamically metastable phases that must be synthesized under high pressures and temperatures, making them expensive. Moreover, diamond is nonresistant to oxidation and reactive to ferrous metals at moderate temperatures. These limitations stimulate the search for new superhard phases that exhibit high stability and are easy to synthesize. A great success has been achieved with the discovery of many hard or superhard transition-metal light-element compounds.2–6 Among them, transition-metal borides (TM1-xBx) including diborides of Os and Re,7,8 triborides (originally reported as tetraborides) of W and Mo,9–20 tetraborides of Cr, Mn and Fe,21–28 and dodecaborides of Zr, Y and Sc,29,30 have currently been the focus of promising superhard materials. Implicit in such efforts is that the dense TM atoms provide high valence electron density to resist volume compression while the rich boron atoms form strong covalent network to counteract shape deformation, both of which contribute to the high hardness. As such, the hardness of TM1-xBx is enhanced by increasing the boron content in their lattices. It is not surprising that the aforementioned superhard materials are the highest borides in the respective TM-B systems, and they thus are strong covalent crystals.

Most transition metals, however, form borides with low boron content, which results in a greatly reduced level of covalent bonding and a significantly high degree of metallic bonding. Such metallic bonds are favorable for electron transport (good metallicity) but not for high hardness due to their low resistance against the formation and movement of dislocations. In this sense, it is a huge challenge to synthesize superhard metals from low borides. A typical example is that the highest boride of tungsten WB3 (nonstoichiometric WB4) is superhard (43.3-46.3 GPa under the load of 0.49 N),9,10 and this is attributed to its strong covalent network formed by hexagonal boron sheets interconnected with zigzag W-B bonds.31 In contrast, the low boride of tungsten WB is not naturally superhard due to the presence of metal bilayers.32,33 Surprisingly, Yeung et al.34 have recently synthesized W0.5Ta0.5B by substituting Ta onto the W sites of WB, and its measured hardness touches 42.8 GPa under the load of 0.49 N, marking it as the first known superhard boride with relatively low boron content. Through systematic first-principles calculations, we have found that such an anomalous hardening from WB to W0.5Ta0.5B is attributed to the relief of the fierce dd antibonding interaction.35 This exciting material has raised great expectations for a new generation of superhard metals with a good balance between hardness and ductility, but its asymptotic hardness is still less than the superhard threshold (H ≥ 40 GPa). It is therefore highly desired to develop a new strategy to further enhance hardness.

The systems of boron with the IVth to VIth group TMs have been synthesized experimentally.36–40 Many first-principles studies41–44 have been performed on their elastic properties and these calculations provide useful information for the understanding of mechanical characters. In this work, using these well-known and accessible boron-low TM1-xBx as illustrating examples, we propose a microscopic mechanism to enhance hardness by seeking highly stable thermodynamic phases and locking their weak slip planes. Of particular interest is the correlation between hardness and boron content. Our results show that the hardness of TM1-xBx (TM=V, Nb and Ta) first increases then decreases with the rise of boron content, contrasting with the common sense that the high boron content gives rise to the high hardness. Such an anomalous phenomenon originates from their peculiar electronic mechanisms. In addition, we suggest the possibility of enhancing the extrinsic hardness and achieving true superhard metals by creating a multiphase material with a large number of interfaces among different phases.

All calculations within the framework of density functional theory (DFT) are performed using the plane-wave basis set as implemented in the Vienna Ab Initio Simulation Package (VASP) code45 and the electron-ion interactions are described by means of projector augmented wave (PAW). The exchange-correlation functionals within the generalized gradient approximation (GGA)46 are employed. A large cutoff energy of 500 eV and the dense k-point meshes with a grid of 0.02 Å-1 are adopted for the considered phases to ensure the energy convergence of about 1 meV per atom.

After having fully relaxed the studied phases with respect to their lattice constants and internal atomic positions, we calculate the formation energy defined as ΔE = E(TM1−xBx)−(1−x)E(TM)−xE(B). According to this definition, the negative formation energy means that the phase is energetically stable. The lower the formation energy is, the more stable the phase is. The elastic constants, bulk moduli, shear moduli, Young’s moduli, and Poisson’s ratios (v) are obtained by the efficient strain-energy method.47–49 Since the hardness is a relatively complex quantity to describe the resistance of materials to plastic deformation and can be dramatically affected by extrinsic conditions, we first use Chen’s model50 to estimate the hardness, then do Tian’s model51 to cross-check the overall trends.

We initially search for highly stable thermodynamic phases by calculating the formation energy. Based on the superhard W0.5Ta0.5B structure [see Fig. 1(a)], fifteen candidate phases (TiB, Ti0.5Cr0.5B, VB, V0.5Cr0.5B, and CrB in the 3d series; ZrB, Zr0.5Nb0.5B, NbB, Nb0.5Mo0.5B, and MoB in the 4d series; HfB, Hf0.5Ta0.5B, TaB, Ta0.5W0.5B, and WB in the 5d series) are chosen to examine their stability trends. Figure 2(a) shows our obtained formation energies of these TM1-xBx phases as a function of valence electron density. When the average valence electron number equals to 4 electrons per atom (e/atom), this type of monoborides (i.e., VB, NbB, and TaB) has the lowest formation energy among the 3d, 4d and 5d series, respectively.

FIG. 1.

Top and side views of crystal structures: (a) VB, (b) V5B6, (c) V3B4, (d) V2B3, (d) VB2, and (e). The large and small spheres represent the metal and boron atoms, respectively. In (a)-(d), the red and blue metal atoms are at different layers and form the metal bilayers that slip at the lowest loads, while the boron atoms are linked into single, single and double, double, and triple zigzag chains, respectively. In (e), the boron atoms are connected into the hexagonal honeycomb sheet, while the metal atoms sit directly above and below the centers of boron hexagons and form planar layers. The black solid lines denote their respective unit cells.

FIG. 1.

Top and side views of crystal structures: (a) VB, (b) V5B6, (c) V3B4, (d) V2B3, (d) VB2, and (e). The large and small spheres represent the metal and boron atoms, respectively. In (a)-(d), the red and blue metal atoms are at different layers and form the metal bilayers that slip at the lowest loads, while the boron atoms are linked into single, single and double, double, and triple zigzag chains, respectively. In (e), the boron atoms are connected into the hexagonal honeycomb sheet, while the metal atoms sit directly above and below the centers of boron hexagons and form planar layers. The black solid lines denote their respective unit cells.

Close modal
FIG. 2.

(a) Calculated formation energy ΔE versus valence electron density for a series of monoborides; (b) calculated formation energy ΔE versus boron composition x for the TM1-xBx systems. Six phases with different compositions (i.e., the ratio of metal to boron 3:2, 1:1, 5:6, 3:4, 2:3, and 1:2, respectively) are systematically considered.

FIG. 2.

(a) Calculated formation energy ΔE versus valence electron density for a series of monoborides; (b) calculated formation energy ΔE versus boron composition x for the TM1-xBx systems. Six phases with different compositions (i.e., the ratio of metal to boron 3:2, 1:1, 5:6, 3:4, 2:3, and 1:2, respectively) are systematically considered.

Close modal

After having obtained the highly stable VB, NbB, and TaB phases, we fine-tune their boron contents and broaden them to other compositions. Six composition phases (e.g., TM3B2, TMB, TM5B6, TM3B4, TM2B3 and TMB2) are considered since they are well documented experimentally.36 Figure 2(b) gives our calculated convex curves of the formation energy, which are a bit different from previous trends41 but in agreement with the trend of the V1-xBx system.44 Interestingly, the six structures are thermodynamic ground states at different boron compositions. Besides concrete values, the three systems share nearly same stability trends, but the V1-xBx phases are more thermodynamically stable than the corresponding Nb1-xBx, and Ta1-xBx ones. The formation energy quantitatively characterizes the comprehensive information of a crystal, such as strong bonding, optimal band filling, and favorable bond topology, thus the highly thermodynamic stable solid tends to own the high resistance against dislocations.52 For x from 0.5 to 0.6, the V1-xBx phases have very low formation energies, and this implies that they could be promising superhard materials.

We next evaluate the mechanical properties of the TM1-xBx phases. Although several previous works predicted the mechanical properties of some TM1-xBx phases,39,41,42 their results are quite scattered and in parts contradict each other due to the inconsistency in different techniques, and thus we reexamine these properties. Figure 3 shows their bulk moduli, shear moduli, and hardness as a function of the boron content. Although their mechanical properties vary widely with different boron contents, the overall trends of the three systems are very similar. This is not unexpected because V, Nb, and Ta lie in the same group on the periodic table. Most remarkable is that the changing trend of incompressibility is obviously different from that of hardness. As shown in Fig. 3(a), the bulk modulus increases with the rise of boron contents. The Ta1-xBx phases are more incompressible than the corresponding Nb1-xBx and V1-xBx ones, and TaB2 becomes the most incompressible material among them. In contrast, in Fig. 3(b)–3(c), the shear modulus and hardness ascend gradually with the boron content from x=0.4 (TM3B2) to x=0.6 (TM2B3), but then descend from x=0.6 (TM2B3) to x=2/3 (TMB2). Hence, the hardness maximum occurs at the boron contents x=0.6 (TM2B3). A positive correlation between hardness and boron content is widely expected, driving many attempts to hunt for superhard materials from the high-boron-content borides. It is rather surprising that the TM1-xBx (TM=V, Nb and Ta) systems do not follow such a monotonic correlation. A similar anomalous stress response has recently been reported in the WBx compounds.53 

FIG. 3.

Calculated bulk modulus B (a), shear modulus G (b), and theoretical hardness H (c) versus boron composition x for the V1-xBx (in red), Nb1-xBx (in blue), and Ta1-xBx (in green) systems. Six phases with different compositions (i.e., the ratio of metal to boron 3:2, 1:1, 5:6, 3:4, 2:3, and 1:2, respectively) are systematically considered.

FIG. 3.

Calculated bulk modulus B (a), shear modulus G (b), and theoretical hardness H (c) versus boron composition x for the V1-xBx (in red), Nb1-xBx (in blue), and Ta1-xBx (in green) systems. Six phases with different compositions (i.e., the ratio of metal to boron 3:2, 1:1, 5:6, 3:4, 2:3, and 1:2, respectively) are systematically considered.

Close modal

On the other hand, a large bulk modulus is often found in traditional superhard materials. For example, the hardest material, diamond, also has the largest bulk modulus. A theory of bulk modulus of covalent solids is useful for suggesting schemes to increase hardness.54 The Ta1-xBx phases possess the highest bulk modulus among the same-composition phases of the three systems, but their hardness values are not the highest. The V1-xBx phases with lower bulk moduli are much harder than the corresponding phases of the other two systems. Completely opposite to the traditional superhard materials in which the bulk modulus and hardness share the same trend, this class of TM1-xBx (TM=V, Nb and Ta) exhibits obviously different trends between the hardness and bulk modulus.

Through seeking highly stable phases and fine-tuning their boron contents, the four low borides (VB, V5B6, V3B4 and V2B3) are identified to possess the highest hardness among these materials. Table I lists their lattice parameters and mechanical properties. For comparison, the related results of W0.5Ta0.5B are also listed. Calculated elastic constants manifest that these four phases are mechanically stable. Their shear moduli (≥ 232GPa), Young’s moduli (≥ 543GPa), Poisson’s ratios (≤ 0.171), and hardness values (≥ 36.6GPa) are by far superior to the corresponding values (201 GPa, 497 GPa, 0.239, 32.6 GPa, respectively)33 of the recently synthesized W0.5Ta0.5B phase, and even match that of the high boride WB3 (252 GPa, 588 GPa, 0.168, 38.9 GPa, respectively).15,31 To the best of our knowledge, the four borides are the hardest borides with the same TM/B composition ratios.

TABLE I.

Calculated equilibrium lattice parameters a, b, c (Å), elastic constants Cij (GPa), bulk modulus B (GPa), shear modulus G (GPa), Young’s modulus E (GPa), Poisson’s ratio υ, and hardness H (GPa) of the representative VB, V5B6, V3B4, and V2B3 phases. For comparison, the available data for W0.5Ta0.5B from Ref. 33 are also listed.

a, b, cC11, C22, C33, C44, C55, C66, C12, C13, C23BGEvH
VB 3.050, 8.036, 2.969 496, 637, 616, 217, 280, 216, 132, 156, 77 275 232 543 0.171 36.6 
V5B6 21.220, 2.976, 3.049 654, 629, 508, 272, 252, 243, 93, 131, 149 280 241 571 0.164 38.8 
V3B4 13.208, 2.980, 3.042 653, 638, 501, 265, 251, 249, 100, 139, 146 282 245 572 0.162 39.1 
V2B3 3.039, 18.408, 2.982 504, 667, 655, 267, 257, 248, 133, 137, 111 284 249 579 0.161 39.9 
W0.5Ta0.53.250, 8.593, 3.129 508, 544, 571, 199, 248, 239, 212, 224, 184 310 201 497 0.239 32.6 
a, b, cC11, C22, C33, C44, C55, C66, C12, C13, C23BGEvH
VB 3.050, 8.036, 2.969 496, 637, 616, 217, 280, 216, 132, 156, 77 275 232 543 0.171 36.6 
V5B6 21.220, 2.976, 3.049 654, 629, 508, 272, 252, 243, 93, 131, 149 280 241 571 0.164 38.8 
V3B4 13.208, 2.980, 3.042 653, 638, 501, 265, 251, 249, 100, 139, 146 282 245 572 0.162 39.1 
V2B3 3.039, 18.408, 2.982 504, 667, 655, 267, 257, 248, 133, 137, 111 284 249 579 0.161 39.9 
W0.5Ta0.53.250, 8.593, 3.129 508, 544, 571, 199, 248, 239, 212, 224, 184 310 201 497 0.239 32.6 

As widely accepted, the high hardness usually exists in strong covalent solids like high borides. It therefore is a bit puzzling that the four low borides are so hard, and several fundamental questions remain to be answered: (i) Why does VB exhibit higher stability and hardness than the newly reported W0.5Ta0.5B? (ii) Why do the hardness values of the four borides (VB, V5B6, V3B4, and V2B3) gradually enhance with the rise of boron content? (iii) What limits the hardness of VB2, the increase of boron content leading to an unexpected reduction of hardness? The answers to these questions must lie in the structural and electronic properties of this class of systems.

As illustrated in Fig. 1(a), VB is an orthorhombic structure (space group Cmcm) that is adopted by many transition-metal monoborides. Its metal sublattice consists of trigonal prisms. Each boron atom is situated in the interstitial centre of a trigonal prism and in contact with six metal atoms at the corners of the prism. The central holes of these trigonal prisms are connected to channels, thus the boron atoms are linked into zigzag chains. Since the boron concentration in VB is relatively low, its structure contains vanadium bilayers. Although these V-V bonds contribute to the good metallicity, they are mechanically weaker than the V-B and B-B bonds. As such, the metal bilayers form the weak slip planes. These easy-glide planes govern the formation and movement of dislocations and thus dominate the intrinsic hardness. Indeed, the high-pressure radial diffraction experiment of W0.5Ta0.5B34 and the calculated study of CrB38 have corroborated the presence of the weak slip planes. Similar behaviors also occur in other structural systems. For example, the Os-Os metallic bonding layers reduce the resistance of OsB2 against large shear stresses in the easy-slip directions and thus limit the hardness.55 As a result, if the slide dislocations within the metal bilayers can be prevented from moving in the crystals, the overall hardness of this type of monoborides should be able to be enhanced.

We now study the electronic structures of VB to reveal its underlying hardening mechanism. Its band structures, total density of states (DOS), and the V- and B-projected DOS are plotted in Figs. 4(a)–4(d), respectively. Several bands cross its Fermi level, indicating the electrical conductivity in VB. This good metallicity is uncommon since traditional superhard materials like diamond and cBN are insulators. We thus expect this class of superhard metals to find exciting applications (e.g., large volume press anvils, hard electromechanical coatings). The bonding mechanism of VB can be understood as follows: First, the crystal field effect splits the V-3d state into the t2g bonding states (derived from dxy, dyz, and dzx orbitals) and eg antibonding states (derived from dz2 and dx2-y2 orbitals). Then, a strong hybridization between the B-2p and t2g states forms the pd bonding states and the pd antibonding states. This bonding picture is confirmed by our calculated electronic structures. As shown in Figs. 4(a)–4(d), the lowest two bands in the range (-12, -6) eV have predominantly B-2s character that is responsible for the B-B covalent bonds of boron chains. The six bands in the range (-6, 0) eV are basically derived from the pd bonding states, while the six bands in the unoccupied high region above 2 eV are the corresponding pd antibonding part. The four bands in the range (0, 2) eV belong to the dd antibonding states. Interestingly, the dd antibonding states and the pd bonding states near the Fermi level respond oppositely to the sliding within the metal bilayers. On one hand, the dd antibonding states mainly originate from the metallic bonding within the metal bilayers and are unfavorable. Moreover, the filling of the dd antibonding states leads to the electrostatic repulsion within the metal bilayers. These are beneficial for the slide dislocations within the metal bilayers and thus provide a negative contribution to the hardness. On the other hand, the pd bonding states, which have strong directional character, are extremely resistive to shear stresses. Since each boron atom is surrounded by six neighbor metal atoms, the filling of the pd bonding states restricts the dislocations between the metal and boron layers. Moreover, each boron atom also bonds to a metal atom in the second-neighbor layer, and this pd bonding offers additional locking effect that impedes the slipping within the metal bilayers. Hence, these give rise to a positive contribution to the hardness. Importantly, VB has the optimal valence electron number to fully occupy the pd bonding states, while leaving the dd antibonding states unoccupied. This optimal band filling plays a key role in high stability and hardness of VB. For W0.5Ta0.5B, the dd antibonding states start to be populated, leading to a drop in its formation energy and hardness.35 These results not only explain why VB is much harder than W0.5Ta0.5B, but also present the main arguments that why we pursue the highly stable thermodynamic phases in the search for superhard metals.

FIG. 4.

Calculated electronic structures, total density of states (DOS), the V-projected DOS, and the B-projected DOS (from left to right) of VB (top panels) and VB2 (bottom panels). The Fermi level is set at 0 eV and indicated by a horizontal dashed line.

FIG. 4.

Calculated electronic structures, total density of states (DOS), the V-projected DOS, and the B-projected DOS (from left to right) of VB (top panels) and VB2 (bottom panels). The Fermi level is set at 0 eV and indicated by a horizontal dashed line.

Close modal

Through optimizing the band filling, VB has the highest hardness among this class of monoborides. To further enhance the hardness, the weak metal bilayers are strengthened by the increase of boron contents. In VB, V5B6, V3B4 and V2B3 structures, their metal skeletons of trigonal prisms are nearly the same. However, the boron atoms accommodate in the central holes of different trigonal prisms, forming single boron chains in VB [Fig. 1(a)], single and double alternate boron chains in V5B6 [Fig. 1(b)], double boron chains in V3B4 [Fig. 1(c)], and triple boron chains in V2B3 [Fig. 1(d)]. It is clear that the number of metal bilayers gradually decreases with the increase of boron concentrations from VB to V2B3. Importantly, the increasing boron contents tend to form an increasing number of covalent B-B and V-B bonds. These factors continuously strengthen the mechanical properties from VB to V2B3, especially for an enhanced trend of their hardness.

When further increasing boron contents to form VB2, it becomes the AlB2 structure (space group P6/mmm), even though all trigonal prisms of metal atoms are filled by the boron atoms. As shown in Fig. 1(e), the boron atoms form rigid hexagonal sheets. The metal atoms sit directly in the interstices above and below the centers of boron hexagons and form planar metal layers. Our calculated band structures, total DOS, V-projected DOS, and B-projected DOS of VB2 are plotted in Figs. 4(e)–4(h), respectively. Its bonding mechanism is completely different from that of VB. The local environment breaks the degeneracy of V-3d orbitals, forming three types of orbitals, namely, degenerate 3dxy and 3dx2-y2, degenerate 3dyz and 3dxz, and 3dz2. Since the V-3d states lie very close in energy to the B-2sp states, there is a strong hybridization among these states. The lowest five bands in the range (-15, -1.5) eV can be viewed as the bonding states of the V-B hybridization, among which the lower three (upper two) bands are mainly derived from the degenerate 3dxy and 3dx2-y2 (3dyz and 3dxz) states and the B-2sp2 bonding (B-2pz nonbonding) states, while the region above -1.5 eV is the corresponding antibonding part. The sixth band ranging from -1.5 to 1.5 eV is mainly derived from the antibonding V-3dz2 states. This analysis naturally leads to the conclusion that this structure shows the highest hardness when the five bonding bands are completely filled. That is to say, the hardness maximum occurs at 10 electrons per formula unit. For VB2, each formula unit has 11 electrons and the antibonding states start to be populated. It is this strong antibonding interaction that leads to the anomalous hardness reduction of VB2. This bonding mechanism responsible for the unusual hardness reduction is transferable to other systems or properties. As a matter of fact, the hardness maximum of cubic TMCxN1-x at 8.4 electrons per formula unit2 and the highest stability of hexagonal TMB2 at 10 electrons per formula unit56 have already confirmed this.

By applying the above atom-scale schemes (i.e., optimizing band filling and tuning boron content), the four borides (VB, V5B6, V3B4 and V2B3) become the competitive superhard metals, but their asymptotic hardness is still less than 40GPa. We further propose a nanoscale way of enhancing extrinsic hardness by creating a multiphase material. As discussed above, their compositions are adjacent and structures are similar. One common structural feature is the presence of the metal bilayers, which limits the hardness enhancement. We thus envisage that one creates a multiphase solid-solution where the easy-slip directions of several borides are essentially random. This particular material not only includes a large number of interfaces among the different phases but also introduces some lattice strains. These interfaces form nanoscale interlocks that strongly restrict the glide dislocations within the metal bilayers, thus drastically enhancing the extrinsic hardness and achieving the true superhard metal. In fact, the recently synthesized nanotwin diamond and cBN have well illustrated that the creation of nanoscale interfaces indeed increase the extrinsic hardness,57,58 and the nature of strengthening is the effective blockage of dislocation motion by numerous grain boundaries.

According to the above calculations, VB has the lowest formation energy (-0.849 eV/atom), and the formation energies of V5B6 (-0.830 eV/atom), V3B4 (-0.819 eV/atom), and V2B3 (-0.798 eV/atom) also are extremely low. Hence, the four phases are thermodynamically viable. It is interesting that their energy differences are very small, and the maximum span is as low as 0.051 eV/atom. Considering the temperature effect on the formation energy, such a small energy separation indicates that the four borides are energetically degenerate. Also, they share close compositions and similar structures. These factors thus provide a useful indication that this multiphase material should be able to synthesizable under appropriate conditions. One good example is the recently successful synthesis of the superhard and metallic W0.5Ta0.5B nanowires through flux growth.59 We therefore expect future efforts to synthesize the particular multiphase material of V1-xBx. If realized, this should be record making true superhard metals from low borides.

In summary, we have proposed an atom- and nanoscale method of hardness enhancement in low borides by pursuing high stable thermodynamic phases and restricting weak slip planes. Using the typical TM1-xBx systems as illustrative cases, we find that the four low borides (VB, V5B6, V3B4 and V2B3) are harder than superhard W0.5Ta0.5B and exhibit the electrical conductivity. The four borides share close compositions, similar structures, and small energy differences. We thus suggest that one produces a multiphase material with a large number of interfaces among different phases. These interfaces form nanoscale interlocks that strongly hinder the glide dislocations within the metal bilayers of each boride, accordingly enhancing the extrinsic hardness and achieving the true superhard metal.

Furthermore, our results show that the low borides have high hardness whereas the high borides exhibit an unusual reduction of hardness, which brings an exception to the conventional knowledge that the high boron content improves the hardness. Such an anomalous hardness phenomenon stems from the peculiar bonding mechanisms in these borides. Therefore, the present work not only reveals the unique mechanism responsible for the anomalous hardness trend of this class of TM1-xBx, but also provides a multiscale avenue to creating novel superhard metals with rich functionalities.

We thank Prof. Changfeng Chen for his valuable suggestions. We acknowledge the financial support from the National Natural Science Foundation of China (No. 51671126). Chun Tang acknowledges the support from Jiangsu University start-up funding.

1.
H. W.
Hugosson
,
U.
Jansson
,
B.
Johansson
, and
O.
Eriksson
,
Science
293
,
2434
(
2001
).
2.
S. H.
Jhi
,
J.
Ihm
,
S. G.
Louie
, and
M. L.
Cohen
,
Nature
399
,
132
(
1999
).
3.
L. S.
Dubrovinsky
,
N. A.
Dubrovinskaia
,
V.
Swamy
,
J.
Muscat
,
N. M.
Harrison
,
R.
Ahuja
,
B.
Holm
, and
B.
Johansson
,
Nature
410
,
653
(
2001
).
4.
R. B.
Kaner
,
J. J.
Gilman
, and
S. H.
Tolbert
,
Science
308
,
1268
(
2005
).
5.
J. C.
Crowhurst
,
A. F.
Goncharov
,
B.
Sadigh
,
C. L.
Evans
,
P. G.
Morrall
,
J. L.
Ferreira
, and
A. L.
Nelson
,
Science
311
,
1275
(
2006
).
6.
C.
Lu
,
Q.
Li
,
Y.
Ma
, and
C.
Chen
,
Phys. Rev. Lett.
119
,
115503
(
2007
).
7.
R. W.
Cumberland
,
M. B.
Weinberger
,
J. J.
Gilman
,
S. M.
Clark
,
S. H.
Tolbert
, and
R. B.
Kaner
,
J. Am. Chem. Soc.
127
,
7264
(
2005
).
8.
H. Y.
Chung
,
M. B.
Weinberger
,
J. B.
Levine
,
A.
Kavner
,
J. M.
Yang
,
S. H.
Tolbert
, and
R. B.
Kaner
,
Science
316
,
436
(
2007
).
9.
Q.
Gu
,
G.
Krauss
, and
W.
Steurer
,
Adv. Mater.
20
,
3620
(
2008
).
10.
R.
Mohammadi
,
A. T.
Lech
,
M.
Xie
,
B. E.
Weaver
,
M. T.
Yeung
,
S. H.
Tolbert
, and
R. B.
Kaner
,
Proc. Natl. Acad. Sci. USA
108
,
10958
(
2011
).
11.
Y.
Liang
,
X.
Yuan
, and
W.
Zhang
,
Phys. Rev. B
83
,
220102(R)
(
2011
).
12.
R. F.
Zhang
,
D.
Legut
,
Z. J.
Lin
,
Y. S.
Zhao
,
H. K.
Mao
, and
S.
Veprek
,
Phys. Rev. Lett.
108
,
255502
(
2012
).
13.
X.
Cheng
,
W.
Zhang
,
X. Q.
Chen
,
H.
Niu
,
P.
Liu
,
K.
Du
,
G.
Liu
,
D.
Li
,
H. M.
Cheng
,
H.
Ye
, and
Y.
Li
,
Appl. Phys. Lett.
103
,
171903
(
2013
).
14.
Q.
Li
,
D.
Zhou
,
W.
Zheng
,
Y.
Ma
, and
C.
Chen
,
Phys. Rev. Lett.
110
,
136403
(
2013
).
15.
Y.
Liang
,
J.
Yang
,
X.
Yuan
,
W.
Qiu
,
Z.
Zhong
,
J.
Yang
, and
W.
Zhang
,
Sci. Rep.
4
,
5063
(
2014
).
16.
I.
Zeiringer
,
P.
Rogl
,
A.
Grytsiv
,
J.
Polt
,
E.
Bauer
, and
G. J.
Giester
,
J. Phase Equilib. Diff.
35
,
384
(
2014
).
17.
Q.
Tao
,
D.
Zheng
,
X.
Zhao
,
Y.
Chen
,
Q.
Li
,
Q.
Li
,
C.
Wang
,
T.
Cui
,
Y.
Ma
,
X.
Wang
, and
P.
Zhu
,
Chem. Mater.
26
,
5297
(
2014
).
18.
E.
Zhao
,
J.
Meng
,
Y.
Ma
, and
Z.
Wu
,
Phys. Chem. Chem. Phys.
12
,
13158
(
2010
).
19.
A. G.
Van Der Geest
and
A. N.
Kolmogorov
,
Calphad
46
,
184
(
2014
).
20.
Y.
Liang
,
X.
Yuan
,
Z.
Fu
,
Y.
Li
, and
Z.
Zhong
,
Appl. Phys. Lett.
101
,
181908
(
2012
).
21.
H.
Niu
,
J.
Wang
,
X. Q.
Chen
,
D.
Li
,
Y.
Li
,
P.
Lazar
,
R.
Podloucky
, and
A. N.
Kolmogorov
,
Phys. Rev. B
85
,
144116
(
2012
).
22.
A.
Knappschneider
,
C.
Litterscheid
,
D.
Dzivenko
,
J.
Kurzman
,
R.
Seshadri
,
N.
Wagner
,
J.
Beck
,
R.
Riedel
, and
B.
Albert
,
Inorg. Chem.
52
,
540
(
2013
).
23.
H.
Gou
,
A. A.
Tsirlin
,
E.
Bykova
,
A. M.
Abakumov
,
G.
Van Tendeloo
,
A.
Richter
,
S. V.
Ovsyannikov
,
A. V.
Kurnosov
,
D. M.
Trots
,
Z.
Konopkova
,
H. P.
Liermann
,
L.
Dubrovinsky
, and
N.
Dubrovinskaia
,
Phys. Rev. B
89
,
064108
(
2014
).
24.
A.
Knappschneider
,
C.
Litterscheid
,
N. C.
George
,
J.
Brgoch
,
N.
Wagner
,
J.
Beck
,
J. A.
Kurzman
,
R.
Seshadri
, and
B.
Albert
,
Angew. Chem., Int. Ed.
53
,
1684
(
2014
).
25.
Y.
Liang
,
X.
Yuan
,
Y.
Gao
,
W.
Zhang
, and
P.
Zhang
,
Phys. Rev. Lett.
113
,
176401
(
2014
).
26.
H.
Gou
,
N.
Dubrovinskaia
,
E.
Bykova
,
A. A.
Tsirlin
,
D.
Kasinathan
,
W.
Schnelle
,
A.
Richter
,
M.
Merlini
,
M.
Hanfland
,
A. M.
Abakumov
,
D.
Batuk
,
G.
Van Tendeloo
,
Y.
Nakajima
,
A. N.
Kolmogorov
, and
L.
Dubrovinsky
,
Phys. Rev. Lett.
111
,
157002
(
2013
).
27.
R. F.
Zhang
,
X. D.
Wen
,
D.
Legut
,
Z. H.
Fu
,
S.
Veprek
,
E.
Zurek
, and
H. K.
Mao
,
Sci. Rep.
6
,
23088
(
2016
).
28.
Y.
Liang
,
Z.
Wu
, and
S.
Wang
,
AIP Adv.
5
,
117208
(
2015
).
29.
G.
Akopov
,
M. T.
Yeung
,
Z. C.
Sobell
,
C. L.
Turner
,
C. W.
Lin
, and
R. B.
Kaner
,
Chem. Mater.
28
,
6605
(
2016
).
30.
T.
Ma
,
H.
Li
,
S.
Wang
,
X.
Zheng
,
X.
Wang
,
H.
Zhao
,
S.
Han
,
J.
Liu
,
R.
Zhang
,
P.
Zhu
,
Y.
Long
,
J.
Cheng
,
Y.
Ma
,
Y.
Zhao
,
C.
Jin
, and
X.
Yu
,
Adv. Mater.
29
,
1604003
(
2017
).
31.
Y.
Liang
,
F.
Fu
,
X.
Yuan
,
S.
Wang
,
Z.
Zhong
, and
W.
Zhang
,
Europhys. Lett.
98
,
66004
(
2012
).
32.
H.
Gou
,
Z.
Li
,
J.
Zhang
,
H.
Niu
,
F.
Gao
,
R. C.
Ewing
, and
J.
Lian
,
Comput. Mater. Sci.
53
,
460
(
2012
).
33.
Y.
Liang
,
Y.
Gou
,
X.
Yuan
,
Z.
Zhong
, and
W.
Zhang
,
Chem. Phys. Lett.
580
,
48
(
2013
).
34.
M. T.
Yeung
,
J.
Lei
,
R.
Mohammadi
,
C. L.
Turner
,
Y.
Wang
,
S. H.
Tolbert
, and
R. B.
Kaner
,
Adv. Mater.
28
,
6993
(
2016
).
35.
Y.
Liang
,
Z.
Gao
,
P.
Qin
,
L.
Gao
, and
C.
Tang
,
Nanoscale
9
,
9112
(
2017
).
36.
K. E.
Spear
,
P. K.
Liao
, and
J. F.
Smith
,
Bull. Alloy Phase Diagr.
8
,
447
(
1987
).
37.
S.
Okada
and
T.
Lundstrom
,
J. Cryst. Growth
129
,
543
(
1993
).
38.
L.
Han
,
S.
Wang
,
J.
Zhu
,
S.
Han
,
W.
Li
,
B.
Chen
,
X.
Wang
,
X.
Yu
,
B.
Liu
,
R.
Zhang
,
Y.
Long
,
J.
Cheng
,
J.
Zhang
,
Y.
Zhao
, and
C.
Jin
,
Appl. Phys. Lett.
106
,
221902
(
2015
).
39.
P. J.
Robinson
,
G.
Liu
,
S.
Ciborowski
,
C.
Martinez-Martinez
,
J. R.
Charmorro
,
X.
Zhang
,
T. M.
McQueen
,
K. H.
Bowen
, and
A. N.
Alexandrova
,
Chem. Mater.
29
,
9892
(
2017
).
40.
P.
Wang
,
R.
Kumar
,
E. M.
Sankaran
,
X.
Qi
,
X.
Zhang
,
D.
Popov
,
A. L.
Cornelius
,
B.
Li
,
Y.
Zhao
, and
L.
Wang
,
Inorg. Chem.
57
,
1096
(
2018
).
41.
T.
Yao
,
Y.
Wang
,
H.
Li
,
J.
Lian
,
J.
Zhang
, and
H.
Gou
,
Comput. Mater. Sci.
65
,
302
(
2012
).
42.
Y.
Liang
,
X.
Yuan
,
Z.
Fu
,
Y.
Li
, and
Z.
Zhong
,
Appl. Phys. Lett.
101
,
181908
(
2012
).
43.
X.
Xu
,
K.
Fu
,
L.
Li
,
Z.
Lu
,
X.
Zhang
,
Y.
Fan
,
J.
Lin
,
G.
Liu
,
H.
Luo
, and
C.
Tang
,
Physica B
419
,
105
(
2013
).
44.
Y.
Pan
,
Y. H.
Lin
,
J. M.
Guo
, and
M.
Wen
,
RSC Adv.
4
,
47377
(
2014
).
45.
J. P.
Perdew
,
K.
Burke
, and
M.
Ernzerhof
,
Phys. Rev. Lett.
77
,
3865
(
1996
).
46.
G.
Kresse
and
J.
Furthmuller
,
Phys. Rev. B
54
,
11169
(
1996
).
47.
R.
Hill
,
Proc. Phys. Soc., London, Sect. A
65
,
349
(
1952
).
48.
Y.
Liang
,
C.
Li
,
W.
Guo
, and
W.
Zhang
,
Phys. Rev. B
79
,
024111
(
2009
).
49.
Y.
Liang
,
A.
Li
,
J.
Zhao
, and
W.
Zhang
,
Mod. Phys. Lett. B
23
,
1281
(
2009
).
50.
X. Q.
Chen
,
H.
Niu
,
D.
Li
, and
Y.
Li
,
Intermetallics
19
,
1275
(
2011
).
51.
Y.
Tian
,
B.
Xu
, and
Z.
Zhao
,
Int. J. Refract. Met. Hard Mater.
33
,
93
(
2012
).
52.
Y.
Liang
,
Z.
Zhong
, and
W.
Zhang
,
Comput. Mater. Sci.
68
,
222
(
2013
).
53.
Q.
Li
,
D.
Zhou
,
W.
Zheng
,
Y.
Ma
, and
C.
Chen
,
Phys. Rev. Lett.
115
,
185502
(
2015
).
54.
A. Y.
Liu
and
M. L.
Cohen
,
Science
245
,
841
(
1989
).
55.
J.
Yang
,
H.
Sun
, and
C.
Chen
,
J. Am. Chem. Soc.
130
,
7200
(
2008
).
56.
Y.
Liang
,
Z.
Wu
,
X.
Yuan
,
W.
Zhang
, and
P.
Zhang
,
Nanoscale
8
,
1055
(
2016
).
57.
Y.
Tian
,
B.
Xu
,
D.
Yu
,
Y.
Ma
,
Y.
Wang
,
Y.
Jiang
,
W.
Hu
,
C.
Tang
,
Y.
Gao
,
K.
Luo
,
Z.
Zhao
,
L. M.
Wang
,
B.
Wen
,
J.
He
, and
Z.
Liu
,
Nature
493
,
385
(
2013
).
58.
Q.
Huang
,
D.
Yu
,
B.
Xu
,
W.
Hu
,
Y.
Ma
,
Y.
Wang
,
Z.
Zhao
,
B.
Wen
,
J.
He
,
Z.
Liu
, and
Y.
Tian
,
Nature
510
,
250
(
2014
).
59.
M. T.
Yeung
,
G.
Akopov
,
C. W.
Lin
,
D. J.
King
,
R. L.
Li
,
Z. C.
Socbell
,
R.
Mohammadi
, and
R. B.
Kaner
,
Appl. Phys. Lett.
109
,
203107
(
2016
).