Next Article in Journal
Enhanced Mechanical Properties of Ceramic Rod-Reinforced TWIP Steel Composites: Fabrication, Microstructural Analysis, and Heat Treatment Evaluation
Previous Article in Journal
Advancements in Metal Processing Additive Technologies: Selective Laser Melting (SLM)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Superconductivity in ZrB12 under High Pressure

by
Zexiao Zhang
1,
Xu Zheng
1,*,
Hanshan Luo
1,
Chan Gao
1,
Xiaowei Xue
1,
Jingcheng Zhu
1,
Ruobin Li
1,
Changqing Jin
2 and
Xiaohui Yu
2
1
Department of Physics, Chengdu University of Technology, Chengdu 610059, China
2
Beijing National Laboratory for Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China
*
Author to whom correspondence should be addressed.
Metals 2024, 14(9), 1082; https://doi.org/10.3390/met14091082
Submission received: 15 August 2024 / Revised: 14 September 2024 / Accepted: 19 September 2024 / Published: 21 September 2024

Abstract

:
Transition metal borides have emerged as pivotal players in various fields. In addition to their exceptional properties such as high hardness, a high melting point, and corrosion resistance, certain compounds exhibit remarkable characteristics including superconductivity, magnetism, electrical conductivity, and catalytic activity. Among these compounds, ZrB12 has garnered significant attention due to its unique physicochemical properties. However, previous research on ZrB12 has predominantly focused on its mechanical behavior while overlooking the electron-electron interactions of the superconducting state. In this paper, resistance characterization of ZrB12 under high-pressure conditions was conducted to further investigate its superconductivity. Our research findings indicate that ZrB12 maintains its superconductivity within a pressure range of 0 to 1.5 GPa and is classified as a type 2 superconductor. Additionally, the results confirm the anisotropic nature of ZrB12’s superconductivity. As the pressure increases, the superconducting transition temperature undergoes a gradual decrease. Remarkably, ZrB12 exhibits metallic behavior under pressures up to 31.4 GPa. The observed decline in superconductivity in ZrB12 can be ascribed to the intensified influence of Zr’s movement on phonon dispersion, ultimately leading to a reduction in carrier concentration.

1. Introduction

The rapid progress of industrialization worldwide has led to the extensive utilization of hard materials in various fields, including weaponry, aviation, machinery, construction, geological exploration, and mining [1,2]. Traditionally, hard materials are composed of compounds featuring shorter covalent bonds centered around lighter elements with smaller atomic radii, such as boron, carbon, nitrogen, and oxygen. However, due to the limitations of traditional hard materials in terms of mechanical and electrical properties, there is a growing interest in exploring novel hard materials comprising compounds formed from light elements and transition metals [3,4]. Due to the high density and large bulk modulus of transition metals, the doping of light elements can be carried out with the d-orbital of transition metals to form quasi-covalent bonds [5], significantly enhancing the hardness. Moreover, the distribution of the d-orbitals of transition metals in the outer space of the nucleus is more complex compared to the traditional superhard materials. Therefore, these materials frequently form diverse structures, exhibiting exceptional properties and potential benefits for substituting conventional hard materials in domains such as electronics, information technology, and aerospace [6].
Significant advancements have been achieved in recent years regarding the investigation of transition metal borides. In addition to high hardness, a high melting point, and strong corrosion resistance similar to traditional superhard materials, certain compounds also exhibit properties such as superconductivity [7], magnetism, electrical conductivity [8], and catalysis [9]. Transition metal borides have emerged as extensively investigated among the transition metals, owing to their finely dispersed powder form and their proclivity to readily react with metals under relatively moderate conditions. This reactivity leads to the formation of a diverse range of stable compounds. ZrB12, as a noteworthy representative of this category, has garnered significant attention due to its remarkable physicochemical properties. The experiments and first-principles calculations on ZrB12 show that the symmetrical B-B covalent network can form delocalized π-bonds, and the interaction of the 4d-orbitals of Zr atoms with these π-bonds creates electronic channels, thereby maintaining the material’s ultrahardness while also providing good electrical conductivity [10,11,12]. Notably, heavily doped ZrB12 has a calculated hardness of up to 40 GPa [13,14], while its superconducting transition temperature approaches 6 K [15]. Singularly, ZrB12 exhibits a range of behaviors that deviate significantly from the typical characteristics observed in conventional superconductors. Several theoretical models have been proposed to elucidate its superconducting properties, each offering a distinct perspective on the nature of this intriguing material. One such model characterizes ZrB12 as a traditional superconductor with amplified surface effects that could be responsible for its unique behavior [16,17]. Another model suggests that ZrB12 possesses a strongly coupled s-wave gap structure [18], while another intriguing proposal posits that it may exhibit d-wave superconductivity. The temperature-dependent behavior of the magnetic penetration depth in ZrB12 reveals a pronounced two-band structure feature [19], which implies that the interaction between these bands is relatively weak [20]. Recent band structure calculations have determined that the Fermi surface of ZrB12 consists of both an open sheet and a closed sheet [21], which is a characteristic that could underpin the observed multiband nature of its superconductivity. Additionally, specific heat and vortices measurements consistent with both type 1 and type 2 superconducting behaviors [22,23]. Zhang et al. provided further insight by confirming, through the use of X-ray diffraction techniques, that ZrB12 belongs to the category of type 1.5 superconductivity [24,25]. These results suggest that ZrB12 may represent a distinct class of superconductors, inherently capable of simultaneously exhibiting both type 1 and type 2 characteristics simultaneously. Therefore, the contribution of electron–electron interactions in ZrB12 is expected to be substantial. However, recent research on ZrB12 has predominantly focused on the superconducting type, with limited attention given to investigating the electron–electron interactions in the superconducting character. This intriguing possibility has piqued our interest and motivated us to further explore the superconducting properties of ZrB12, particularly through the application of high-pressure techniques.
Applying pressure can often effectively change the electronic structure of compounds, and even induce phase transitions, without introducing additional impurities [26,27]. In this study, we conducted resistance characterization under pressure to investigate electron–electron interactions in the superconductor ZrB12. Encouragingly, ZrB12 exhibits exceptional stability with minimal structural impact even at pressures as high as 44 GPa, thereby preserving its phase integrity [28,29]. Our research reveals that the superconductivity of ZrB12 exhibits anisotropic characteristics, which persist even under high-pressure conditions. Furthermore, under a pressure of 1.5 GPa, ZrB12 still exhibits superconductivity at approximately 5.7 K. With increasing pressure, the superconducting transition temperature decreases. ZrB12 exhibits metallic behavior under pressures up to 31.4 GPa. The diminishing superconductivity observed in ZrB12 can be attributed to the intensified effect of Zr’s movement on phonon dispersion, ultimately causing a reduction in carrier concentration.

2. Materials and Methods

High-purity Zr (>99.95%) and B (>99.99%) powders in the molar ratio of Zr:B = 1:12 were homogeneously mixed and compacted into cylindrical pellets. Subsequently, the cylinder was placed within a BN thermal insulation layer, which was further enclosed in a graphite furnace. As pyrophyllite has low hardness, good slippage, insulation, sealing, high temperature resistance, and good heat preservation performance, it is suitable as a pressure transmission medium. The assembly was put into a cubic press where the reactants were gradually heated to 2000 °C (5 °C/min) under a pressure of 5.5 GPa for 15 min to obtain stable ZrB12.
The products were washed with distilled ethyl alcohol to remove the impurities, followed by drying in an oven at 348 K. We used a 6 × 600 t high-pressure machine and pyrophyllite as the pressure medium. The method of pressure measurement is conducted based on the variation in metal resistance during solid-to-solid phase transition under different pressures. In this paper, we used Bi (Ⅰ–Ⅱ) and Ba (Ⅱ–Ⅲ) phase transitions. The pressure of the sample corresponding to the pressure points of 2.55 GPa and 5.5 GPa can be obtained through the metals Bi and Ba. The internal diagram of the 6 × 600 t high-pressure machine is shown in Figure S1. The typical process is as follows: firstly, the pressure is elevated to the desired value; once the pressure is stable, the heating procedure is started until the required temperature is reached. Depending on specific synthesis conditions, diverse sample assembly configurations can be employed within six WC anvils.
The symmetries of different crystal orientations for a given type of crystal can be utilized to determine their respective orientations. In the case of ZrB12 single crystals, their orientation is primarily determined using the Laue method. The method requires the data acquisition plate to be in front of the sample and perpendicular to the incident X-ray, with the sample positioned 40 mm higher than the plate. Figure S2 shows the Laue diffraction of [110] (dyad symmetry) and [100] (quadruple symmetric). The single-crystal cutting method involves using a low-speed cutting machine to precisely cut along the single-crystal orientation. By aligning the crystal orientation parallel to the saw blade of the cutting machine, crystals with various orientations can be obtained.
The electronic transport properties of a [100]-oriented ZrB12 single crystal with high pressure up to 44 GPa were measured using four-probe electrical conductivity methods in a CuBe alloy diamond anvil cell (DAC). The diamond culet had a diameter of 300 μm, and Au wires with a diameter of 18 μm were used as electrodes. A T301 stainless steel gasket was compressed to a thickness of 20 μm with a 150 μm diameter hole, into which the cubic BN was pressed as an insulating layer. A small center hole of 100 μm in diameter was drilled to serve as the sample chamber, where NaCl fine powder served as a pressure-transmitting medium. A piece of ZrB12 single crystal measuring 50 μm × 50 μm × 25 μm was loaded into this chamber along with a ruby for pressure measurement. Heat capacity measurements on single crystalline samples 1 × 1 × 0.5 mm3 in size were carried out on a PPMS (QD China, Beijing, China).
High-P synchrotron X-ray diffraction experiments using a DAC were performed at the Beijing Synchrotron Radiation Facility (BSRF), Beijing, China. The obtained polycrystalline ZrB12 was ground into powders and loaded into the sample hole in a stainless steel gasket with neon as the pressure-transmitting medium. A few ruby balls were also loaded into the same sample chamber to serve as an internal pressure standard. The collected angle-dispersive diffraction data were analyzed by integrating 2D images as a function of 2θ using the program Fit2D to obtain the conventional, one-dimensional diffraction profiles.

3. Results

ZrB12 crystallizes in a cubic crystal system characterized by the space group Fm-3m, which is assigned the number 225 in the International Tables for Crystallography. Within this symmetrical framework, the Zr atoms are located at specific Wyckoff positions, which are designated sites within the unit cell that can be occupied by atoms or ions. The Zr atoms are surrounded by clusters of boron atoms, forming B12 icosahedra. These B12 clusters form a rigid, three-dimensional network that contributes to the exceptional properties of ZrB12, including a high melting point and excellent thermal conductivity. The cubic arrangement with B12 clusters renders ZrB12 a material of significant interest for various high-temperature and high-strength applications [30]. The Zr atoms in ZrB12 have an electron configuration of 4s24p64d35s1 and the B atoms have 2s22p1, which serve as valence electrons. These contribute to covalent bonding, thereby fulfilling the characteristics of ultra-hard materials, including a high average valence electron count per atom, a maximal number of covalent bonds per unit volume, and a high bond energy. In order to investigate the electrical transport properties under external pressure on ZrB12, we measured temperature dependence resistance from 2 to 300 K for both [110] and [100] orientations of a ZrB12 single crystal, as shown in Figure 1a. Here, the Tconset is determined as the interception between two straight lines below and above the superconducting transition, while Tczero is defined as the zero-resistivity temperature. As the temperature decreases, the Tconset for the [100] and [110] is 5.8 K and 5.7 K, whereas the Tczero is 4.6 K and 3.4 K, respectively [31]. Moreover, the resistivity has a certain difference, and the resistivity in the [100] orientation is significantly greater than that of the [110] orientation. These findings demonstrate the anisotropic electrical behavior of ZrB12, which is consistent with prior research reports [13]. The specific heat measurements of the [100] orientation below 10 K reveal a peak heat capacity within the weak magnetic field near 0 T, as shown in Figure 1b. An exothermic process around 5.4 K was confirmed, providing further evidence for a superconducting transition around this temperature [15,32]. In ZrB12, which is classified as a two-gap superconductor exhibiting dynamic charge stripes, the interplay between these fluctuating charges and magnetic vortices can occur even under relatively low magnetic fields. This interaction induces a pronounced anisotropy in the superconducting properties of the material. Specifically, the behavior of the superconducting characteristics, such as the critical temperature and the critical current density, becomes highly dependent on the orientation of the applied magnetic field. This anisotropy is a direct consequence of the dynamic charge stripe order within the superconductor, which modulates the electronic structure and thereby influences the pinning and motion of vortices within the material. Bolotina et al. show a stepped singularity of the specific heat capacity of the [100] and [110] crystal orientations at around 5.8 K, and another hump of the [110] orientation appeared at around 5.7 K, demonstrating the band-gap anisotropy of ZrB12 single crystals [33]. Generally, the resistivity of ZrB12 will immediately approach 0 upon reaching the first superconducting temperature. Therefore, in our resistivity experiment, we can only observe the tendency of the resistivity drop to zero and cannot observe the second peak of the [110] orientation. The presence of such a pronounced anisotropy in the superconducting behavior of ZrB12 highlights the intricate relationship between the charge stripe fluctuations and the vortex dynamics, and it underscores the complexity of the superconducting state in materials with this unique electronic structure [33]. Consequently, the change in the heat capacity under the magnetic field, the magnitude of the resistivity, and the differences in the superconducting temperature collectively confirm the anisotropy of ZrB12.
Furthermore, we characterized the temperature dependence of resistance for a ZrB12 single crystal with the [100] orientation at 1.5 GPa under an external magnetic field ranging from 0 to 1 T (as shown in Figure 2). As shown in Figure 2a, the Tconset at 0.05, 0.1, 0.15, 0.2, 0.25, and 0.3 T is 5.4, 5.0, 4.7, 4.3, 3.9, and 3.5 K, while the corresponding Tczero values are 3.2, 2.8, 2.1 K, respectively. TC decreases with an increasing applied magnetic field H, and this variation is illustrated in Figure 2b. Using the Werthamer–Helfand–Hohenberg (WHH) formula,
H c 2 ( 0 ) = 0.693 T c 0 × d H c 2 d T T c 0
where Tc0 is the superconducting transition temperature under zero magnetic field. At 1.5 GPa, the upper critical field Hc2(0) was estimated to be 0.64 T. From Figure 1b and Figure 2a, it is apparent that the intensification of the magnetic field to 1 Tesla completely suppresses superconductivity, indicating that the [100] orientation of the ZrB12 single crystal belongs to the classification of type 2 superconductivity. Based on the Ginzburg–Landau theory, Abrikosov categorized superconductors into two distinct types: type 1 and type 2, according to the magnitude of the Ginzburg–Landau parameter. ZrB12 has been proposed to exhibit a two band-effect and related type 1.5 superconductivity (containing type 1 and type 2) by using X-ray diffraction and μSR studies [26,34]. The observation of type 2 superconductivity presented in Figure 2b does not contradict the conclusion of type 1.5 superconductivity, as the type 1.5 superconductivity of ZrB12 can only be observed below Tc, whereas our discussion primarily focuses on the relationship between Tc and the magnetic field H. We take into account the anisotropy along the [100] direction, which restricts our observations solely to type 2 superconductivity. Furthermore, as illustrated in Figure 2a, at relatively low magnetic fields and below Tc, the superconducting transition exhibits a remarkably sharp profile. As the magnetic field intensifies, the transition gradually broadens, which may be attributed to various vortex phase transitions [26]. Figure 2b shows the characteristic of the upper critical field Hc2(0) in type 2 superconductors, which provides insight into the microscopic mechanism of ZrB12. Generally, two independent mechanisms govern the suppression of superconductivity by magnetic fields. One involves the orbital pair breaking of Cooper pairs in the superconducting state, which is associated with screening currents that exclude the external field which is commonly referred to as the orbital limit. The other mechanism is a spin effect resulting from Zeeman splitting, which is applicable only to singlet pairings and is widely known as the Pauli paramagnetic limit. Evidently, ZrB12 aligns with the former theoretical explanation.
In order to comprehensively investigate the electrical transport properties within a high-pressure environment, it is imperative to ensure and validate the structural stability of the ZrB12 compound under compressive forces, without experiencing any detrimental alterations or deformations. This validation is crucial because the stability of ZrB12’s crystal lattice directly impacts its ability to effectively conduct electricity under such extreme conditions, thereby influencing the reliability and accuracy of measurements pertaining to electrical transport properties. A high-pressure synchrotron X-ray diffraction (λ = 0.6199 Å) study conducted in a DAC was used to investigate the phase stability of ZrB12. Figure 3a presents the XRD patterns obtained during ambient-temperature compression, revealing that the ZrB12 single crystal sustains its structural stability up to 43 GPa, without any observable phase transition. The lattice parameters of ZrB12 in Figure 3b exhibit a nearly linear decrease with increasing pressure. In Figure S3, the pressure–volume data for ZrB12 demonstrate a gradual reduction in both the lattice parameters and volume as pressure increases. This can be attributed to the compression of atomic distances under elevated pressures, resulting in enhanced repulsive interactions among atoms and consequently higher crystal incompressibility. The Zr atomic radius is significantly larger than that of B, the interplanar spacing of ZrB12 gradually decreases under pressure, and the influence on the orbital of Zr is much greater than that of B. Therefore, on the premise of ensuring the structural stability of ZrB12 under pressure, our research on the high-pressure behavior of ZrB12 primarily focuses on investigating alterations in the Zr-d atomic orbital. The third-order Birch–Murnaghan equation of state (EOS) was employed to fit the ZrB12 unit cell, resulting in a bulk modulus B = 221.1 GPa, as illustrated in Figure S3, which demonstrates its compressibility. Therefore, it can be concluded that ZrB12 maintains its structural stability under high pressure. Subsequently, we investigated the electrical transport characteristics of ZrB12 single crystals under pressures ranging from 4.4 to 44 GPa, as depicted in Figure 4. Specifically, Figure 4a exhibits the [100] orientation resistance–temperature curves from 2 to 300 K across 4.4 to 20 GPa, with an inset figure presenting these curves normalized between 2 and 10 K. The obtained data clearly demonstrate that as pressure increases, ZrB12 consistently maintains its superconducting state, while exhibiting a decrease in Tc. Furthermore, under pressures ranging from 31.4 to 44 GPa, ZrB12 exhibits metallic behavior, as illustrated in Figure 4b.
Figure 5 presents the phase diagram of the superconducting transition temperature (SC) dependence pressure of ZrB12, illustrating a region of gradual decrease in Tconset and Tczero from 0 to 44 GPa. The relationship between the superconducting transition temperature and pressure exhibits a gradual decline with a slope of approximately 0.236 K/GPa. From the phase diagram, we can easily visualize the evolution and intimated correlations between the metal and SC as a function of pressure. As pressure increases gradually, Tconset is suppressed, accompanied by an initial enhancement of Tc with a broad superconducting transition width, as displayed in Figure 4a. Interestingly, our observations under applied pressure indicate that despite the absence of structural phase transition in ZrB12 under elevated pressures, the superconducting transition temperature disappears at 20 GPa, coinciding with an increase in pressure. This phenomenon exhibits a metallic behavior, which is strongly associated with the electron–electron interaction within the single crystal.

4. Discussion

The exceptional hardness exhibited by ZrB12 can be largely attributed to the presence of strong covalent bonds between the boron atoms within its crystalline structure. These B-B covalent bonds are a direct consequence of the unique integration and interaction of boron elements within the framework of transition metals, specifically Zr. The robustness of these bonds significantly contributes to the overall hardness of the ZrB12 crystal, rendering it a material of interest in various high-performance applications where resistance to wear and deformation is crucial. The existence and positioning of Zr atoms play a crucial and indispensable role in facilitating and enhancing the electrical conductivity and superconducting capabilities of this specific material. Furthermore, under conditions of zero pressure, it is observed that there is a complete absence of energetic interaction or overlap between the optical phonon modes and the acoustic phonon modes [35]. This intriguing phenomenon can be attributed to the distinct and varying atomic masses of Zr atoms and B atoms as they are arranged within the crystal lattice unit cell. The disparity in their masses ensures that the vibrational modes associated with these atoms do not intersect energetically, thereby maintaining a clear separation between the two types of phonon modes. This separation is essential for the material to exhibit its unique electrical properties, particularly its ability to conduct electricity and transition into a superconducting state under specific conditions [35]. As pressure rises, ZrB12 exhibits remarkable resistance to compression, which is attributed to the robust B-B covalent bond. The band structure of ZrB12 remains largely unaltered, with the primary variations originating from the Zr atomic orbital. Consequently, our focus lies in exploring the disappearance of superconductivity resulting from the electronic–electron interactions in ZrB12 under pressure, along with the corresponding changes in the Zr atomic orbital. The electronic density of states (DOS) and the partial density of states (PDOS) of ZrB12 at 0 GP demonstrated that the DOS at the Fermi level is predominantly dominated by Zr-d states. This indicates that the electron carriers in ZrB12 originate primarily from the Zr-d states. In addition, the changes in cell parameters and volume with pressure derived from our experimental values (as shown in Figure 3 and Figure S3) are basically consistent with the theoretical calculated values of Li, X. et al. [35]. According to the calculation results of ZrB12 by Li, X. et al., the DOS of ZrB12 at 0, 20, 50 GPa was 1.4679 eV, 1.4001 eV, and 1.3571 eV, respectively. When the pressure is below 50 GPa, the carrier concentration gradually decreases under the action of pressure and reaches the minimum value at 50 GPa, indicating that ZrB12 exhibits limited compressibility due to its high bulk modulus. In general, in the case of internal stability of ZrB12 the change in carrier concentration and resistance are correlated. As can be seen from Figure 4a, the resistance of our sample gradually decreases under the applied pressure. However, from Figure 4b, when the pressure is further increased to 44 GPa, ZrB12 exhibits significant resistance to compression, with a resistance level similar to that observed at 31.4 GPa, indicating that the change in carrier concentration during this stage is relatively small and gradually reaches a minimum value. These results are consistent with the calculated value. As the pressure increases, the contribution of Zr movement to phonon dispersion becomes more significant [35]. With the growing influence of Zr’s movement on phonon dispersion, it may result in a reduction in carrier concentration. Therefore, within the pressure range of 0 to 44 GPa, the carrier concentration of ZrB12 steadily decreases as pressure rises. This reduction in carrier concentration leads to a decrease in its superconducting transition temperature. Lower carrier concentrations require even lower temperatures for achieving superconducting transition. Notably, at 20 GPa, the superconducting transition was undetectable, suggesting orbital pair breaking of Cooper pairs in the superconducting state of ZrB12, thereby indicating a transition to a metallic state.

5. Conclusions

In summary, previous research has primarily focused on the mechanical and structural properties of ZrB12, with limited attention given to electron–electron interactions in the superconducting character. In this paper, we discuss the electrical transport properties of ZrB12 single crystals under high-pressure conditions. Our findings reveal that ZrB12 single crystal superconductivity is anisotropic, which persists even under high pressure. Under a pressure ranging from 0 to 1.5 GPa, ZrB12 (crystal orientation of [100]) maintains its superconducting nature and is classified as a type 2 superconductor. At 1.5 GPa, the upper critical field Hc2(0) was estimated to be 0.64 T. With increasing pressure, the superconducting transition temperature decreases. At 20 GPa, no evidence of superconducting transition was observed. The metallic behavior of ZrB12 remains evident under pressures ranging from 31.4 to 44 GPa. To further delve into the superconducting properties of ZrB12, we conducted an in-depth analysis of its crystal structure and electronic interactions by a phase diagram of the superconducting transition temperature dependence pressure of ZrB12. Within the ZrB12 crystal lattice, the crucial presence of Zr atoms enables the conduction and superconductivity properties of this material, with electron carriers primarily originating from the Zr-d states. As the pressure escalates, the contribution of Zr movement to phonon dispersion becomes increasingly significant. This enhanced influence of Zr’s movement on phonon dispersion may result in a decrease in carrier concentration. Consequently, within the pressure range of 0 to 44 GPa, the carrier concentration of ZrB12 progressively diminishes as the pressure rises. This decrement in carrier concentration leads to a corresponding decline in its superconducting transition temperature, requiring even lower temperatures for the occurrence of superconducting transition. Notably, at 20 GPa, the superconducting transition becomes undetectable, indicating orbital pair breaking of Cooper pairs in the superconducting state of ZrB12, thus signifying a transition to a metallic state.
To gain a deeper understanding of the superconducting mechanisms in ZrB12, further studies are needed to investigate the electronic structure, phonon dispersions, and electron–phonon coupling under pressure. Such studies would provide valuable insights into the nature of superconductivity in ZrB12 and potentially lead to the discovery of new superconductors with enhanced properties.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/met14091082/s1, Figure S1: The internal diagram of the 6 × 600 t high- pressure machine.; Figure S2: Laue diffraction of (a) [110] and (b) [100] orientation; Figure S3: The pressure-–volume data for ZrB12 and equation of state analysis.

Author Contributions

Conceptualization, X.Z. and X.Y.; Methodology, Z.Z.; Validation, Z.Z., H.L., X.X., J.Z. and R.L.; Investigation, Z.Z., H.L., X.X., J.Z. and R.L.; Resources, X.Z., C.G., C.J. and X.Y.; Data curation, Z.Z.; Writing—original draft preparation, Z.Z.; Writing—review & editing, X.Z., C.G., C.J. and X.Y.; Visualization, X.Z., C.G., C.J. and X.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This study were partially supported by Sichuan Science and Technology Program (No. 2023NSFSC1371)and Chengdu University of Technology 2023 Young and Middle-aged Backbone Teachers Development Funding Program (No. 10912-JXGG2023-09014).

Data Availability Statement

The original contributions presented in the study are included in the article/supplementary material, further inquiries can be directed to the corresponding author/s.

Acknowledgments

The authors would like to express their gratitude to the Synergic Extreme Condition User Facility (SECUF) for its invaluable support in several laboratory experiments.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Haines, J.; Léger, J.M.; Bocquillon, G. Synthesis and design of superhard materials. Annu. Rev. Mater. Res. 2001, 31, 1–23. [Google Scholar] [CrossRef]
  2. Xu, B.; Tian, Y. Ultrahardness: Measurement and enhancement. J. Phys. Chem. C 2015, 119, 5633–5638. [Google Scholar] [CrossRef]
  3. Qiao, Z.; Ma, X.; Zhao, W.; Tang, H. High-pressure sintering the novel light (W0.25Al0.75)C hard material without binder phase. J. Alloys Compd. 2009, 472, 211–213. [Google Scholar] [CrossRef]
  4. Zhang, S.F.; Zhou, C.; Sun, G.Q.; Wang, X.; Bao, K.; Zhu, P.W.; Zhu, J.M.; Wang, Z.Q.; Zhao, X.B.; Tao, Q.; et al. Electronic structure and hardness of Mn3N2 synthesized under high temperature and high pressure. Metals 2022, 12, 2164. [Google Scholar] [CrossRef]
  5. Prosviryakov, A.; Mondoloni, B.; Churyumov, A.; Pozdniakov, A. Microstructure and hot deformation behaviour of a novel Zr-alloyed high-boron steel. Metals 2019, 9, 218. [Google Scholar] [CrossRef]
  6. Xu, B.; Tian, Y. Superhard materials: Recent research progress and prospects. Sci. China Mater. 2015, 58, 132–142. [Google Scholar] [CrossRef]
  7. Nagamatsu, J.; Nakagawa, N.; Muranaka, A.; Znan, Y.; Akmsu, J. Superconductivity at 39 K in magnesium diboride. Nature 2001, 410, 63–64. [Google Scholar] [CrossRef]
  8. Kayhan, M.; Hildebrandt, E.; Frotscher, M.; Senyshyn, A.; Albert, B. Neutron diffraction and observation of superconductivity for tungsten borides, WB and W2B4. Solid State Sci. 2012, 14, 1656–1659. [Google Scholar] [CrossRef]
  9. Zhu, J.; Cannizzaro, F.; Liu, L.; Zhang, H.; Kosinov, N.; Filot, I.; Rabeah, J.; Brückner, A.; Hensen, E. Ni–In synergy in CO2 hydrogenation to methanol. ACS Catal. 2021, 11, 11371–11384. [Google Scholar] [CrossRef]
  10. Gasparov, V.A.; Sidorov, N.S.; Zver’kova, I.I.; Khassanov, S.S.; Kulakov, M.P. Electron transport, penetration depth, and the upper critical magnetic field in ZrB12 and MgB2. J. Exp. Theor. Phys. 2005, 101, 98–106. [Google Scholar] [CrossRef]
  11. Paderno, Y.B.; Adamovskii, A.A.; Lyashchenko, A.B.; Paderno, V.N.; Fillipov, V.B.; Naydich, Y.V. Zirconium dodecaboride-based cutting material. Powder Metall. Met. C 2004, 43, 546–548. [Google Scholar] [CrossRef]
  12. Post, B.; Glaser, F.W. Crystal structure of ZrB12. JOM 1952, 4, 631–632. [Google Scholar] [CrossRef]
  13. Ma, T.; Li, H.; Zheng, X.; Wang, S.; Wang, X.; Zhao, H.; Han, S.; Liu, J.; Zhang, R.; Zhu, P.; et al. Ultrastrong boron frameworks in ZrB12: A highway for electron conducting. Adv. Mater. 2017, 29, 1604003. [Google Scholar] [CrossRef]
  14. Pan, Y.; Chen, S. Influence of alloying elements on the mechanical and thermodynamic properties of ZrB2 boride. Vacuum 2022, 198, 110898. [Google Scholar] [CrossRef]
  15. Papaconstantopoulos, D.A.; Mehl, M.J. First-principles study of superconductivity in high-pressure boron. Phys. Rev. B 2002, 65, 172510. [Google Scholar] [CrossRef]
  16. Tsindlekht, M.I.; Leviev, G.I.; Asulin, I.; Sharoni, A.; Millo, O.; Felner, I.; Paderno, Y.B.; Filippov, V.B.; Belogolovskii, M.A. Tunneling and magnetic characteristics of superconducting ZrB12 single crystals. Phys. Rev. B 2004, 69, 212508. [Google Scholar] [CrossRef]
  17. Khasanov, R.; Castro, D.D.; Belogolovskii, M.; Paderno, Y.; Keller, H. Anomalous electron-phonon coupling probed on the surface of superconductor ZrB12. Phys. Rev. B 2005, 72, 224509. [Google Scholar] [CrossRef]
  18. Daghero, D.; Gonnelli, R.S.; Ummarino, G.A.; Calzolari, A.; Dellarocca, V.; Stepanov, V.A.; Filippov, V.B.; Paderno, Y.B. Andreev-reflection spectroscopy in ZrB12 single crystals. Supercond. Sci. Technol. 2004, 17, S250. [Google Scholar] [CrossRef]
  19. Gasparov, V.A.; Sidorov, N.S.; Zver’kova, I.I. Two-gap superconductivity in ZrB12: Temperature dependence of critical magnetic fields in single crystals. Phys. Rev. B 2006, 73, 94510. [Google Scholar] [CrossRef]
  20. Nicol, E.J.; Carbotte, J.P. Properties of the superconducting state in a two-band model. Phys. Rev. B 2005, 71, 054501. [Google Scholar] [CrossRef]
  21. Gasparov, V.; Sheikin, I.; Levy, F.; Teyssier, J.; Santi, G. The de Haas van Alphen effect study of the Fermi surface of ZrB12. J. Phys. Conf. Ser. 2009, 150, 052059. [Google Scholar] [CrossRef]
  22. Wang, Y.; Lortz, R.; Paderno, Y.; Filippov, V.; Abe, S. Specific heat and magnetization of a ZrB12 single crystal: Characterization of a type-II/1 superconductor. Phys. Rev. B 2005, 72, 24548. [Google Scholar] [CrossRef]
  23. Ge, J.Y.; Gutierrez, J.; Lyashchenko, A.; Filipov, V.; Li, J.; Moshchalkov, V.V. Direct visualization of vortex pattern transition in ZrB12 with Ginzburg-Landau parameter close to the dual point. Phys. Rev. B 2014, 90, 184511. [Google Scholar] [CrossRef]
  24. Huebener, R.P. The path to type-II superconductivity. Metals 2019, 9, 682. [Google Scholar] [CrossRef]
  25. Zhang, A.L.; Gao, L.X.; He, J.Y.; Filipov, V.B.; Cao, S.; Xiao, Q.L.; Ge, J.Y. Experimental evidence for type-1.5 superconductivity in ZrB12 single crystal. Sci. China Phys. Mech. Astron. 2022, 65, 297412. [Google Scholar] [CrossRef]
  26. Tang, L.; Zhang, J.; Jin, Y.; Kong, P.; Li, S.; Huo, D.; Zhang, C.; Kuang, F. Pressure-induced novel phases with the high-Tc superconductivity in zirconium dihydride. Mater. Today Commun. 2024, 40, 109516. [Google Scholar] [CrossRef]
  27. Li, Y.; Gao, Y.; Han, Y.; Liu, C.; Peng, G.; Wang, Q.; Feng, K.; Ma, Y.; Gao, C. Metallization and hall-effect of Mg2Ge under high pressure. Appl. Phys. Lett. 2015, 107, 142103. [Google Scholar] [CrossRef]
  28. Thakur, S.; Biswas, D.; Sahadev, N.; Biswas, P.K.; Balakrishnan, G.; Maiti, K. Complex spectral evolution in a BCS superconductor, ZrB12. Sci. Rep. 2013, 3, 3342. [Google Scholar] [CrossRef]
  29. Thakur, S.; Maiti, K. Unusual correlation physics in a BCS superconductor, ZrB12. Solid State Commun. 2014, 193, 45–50. [Google Scholar] [CrossRef]
  30. Escamilla, R.; Romero, M.; Morales, F. Elastic properties, Debye temperature, density of states and electron–phonon coupling of ZrB12 under pressure. Solid State Commun. 2012, 152, 249–252. [Google Scholar] [CrossRef]
  31. Yamaura, K.; Huang, Q.; Akaishi, M.; Takayama-Muromachi, E. Superconductivity in the hexagonal-layered molybdenum carbide η-Mo3C2. Phys. Rev. B 2006, 74, 184510. [Google Scholar] [CrossRef]
  32. Silaev, M.; Babaev, E. Microscopic theory of type-1.5 superconductivity in multiband systems. Phys. Rev. B 2011, 84, 094515. [Google Scholar] [CrossRef]
  33. Bolotina, N.B.; Khrykina, O.N.; Azarevich, A.N.; Shitsevalova, N.; Filipov, V.B.; Gavrilkin, S.Y.; Mitsen, K.V.; Voronov, V.V.; Sluchanko, N.E. Checkerboard patterns of charge stripes in the two-gap superconductor ZrB12. Phys. Rev. B 2022, 105, 054511. [Google Scholar] [CrossRef]
  34. Biswas, P.K.; Rybakov, F.N.; Singh, R.P.; Mukherjee, S.; Paul, D.M.K. Coexistence of type-I and type-II superconductivity signatures in ZrB12 probed by muon spin rotation measurements. Phys. Rev. B 2020, 102, 144523. [Google Scholar] [CrossRef]
  35. Li, X.; Yong, Y.; Cui, H.; Zhang, R. Mechanical behavior, electronic and phonon properties of ZrB12 under pressure. J. Phys. Chem. Solids 2018, 117, 173–179. [Google Scholar] [CrossRef]
Figure 1. (a) Temperature dependence of [100] and [110] orientation resistivity at 2 to 300 K under 0 GPa. Inset figure is superconducting temperature of ZrB12 single crystal. (b) Specific heat capacity of ZrB12 from 2 to 9 K under magnetic fields of 0 T, 1 T, and 3 T.
Figure 1. (a) Temperature dependence of [100] and [110] orientation resistivity at 2 to 300 K under 0 GPa. Inset figure is superconducting temperature of ZrB12 single crystal. (b) Specific heat capacity of ZrB12 from 2 to 9 K under magnetic fields of 0 T, 1 T, and 3 T.
Metals 14 01082 g001
Figure 2. (a) The [100] orientation superconducting transition temperature dependence of the magnetic field at 0 T, 0.05 T, 0.1 T, 0.15 T, 0.2 T, 0.25 T, 0.3 T under 1.5 GPa. (b) The relationship between the [100] orientation superconducting transition temperature and the magnetic field under 1.5 GPa.
Figure 2. (a) The [100] orientation superconducting transition temperature dependence of the magnetic field at 0 T, 0.05 T, 0.1 T, 0.15 T, 0.2 T, 0.25 T, 0.3 T under 1.5 GPa. (b) The relationship between the [100] orientation superconducting transition temperature and the magnetic field under 1.5 GPa.
Metals 14 01082 g002
Figure 3. (a) High-pressure XRD patterns collected at room temperature. (b) Lattice parameters of ZrB12 under high pressure.
Figure 3. (a) High-pressure XRD patterns collected at room temperature. (b) Lattice parameters of ZrB12 under high pressure.
Metals 14 01082 g003
Figure 4. (a) The [100] orientation resistance–temperature curves from 2 to 300 K under 4.4 to 20 GPa. (b) The [100] orientation resistance–temperature curve at pressures ranging from 31.4 to 44 GPa.
Figure 4. (a) The [100] orientation resistance–temperature curves from 2 to 300 K under 4.4 to 20 GPa. (b) The [100] orientation resistance–temperature curve at pressures ranging from 31.4 to 44 GPa.
Metals 14 01082 g004
Figure 5. The phase diagram of superconducting transition temperature (SC) dependence pressure of ZrB12.
Figure 5. The phase diagram of superconducting transition temperature (SC) dependence pressure of ZrB12.
Metals 14 01082 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, Z.; Zheng, X.; Luo, H.; Gao, C.; Xue, X.; Zhu, J.; Li, R.; Jin, C.; Yu, X. Superconductivity in ZrB12 under High Pressure. Metals 2024, 14, 1082. https://doi.org/10.3390/met14091082

AMA Style

Zhang Z, Zheng X, Luo H, Gao C, Xue X, Zhu J, Li R, Jin C, Yu X. Superconductivity in ZrB12 under High Pressure. Metals. 2024; 14(9):1082. https://doi.org/10.3390/met14091082

Chicago/Turabian Style

Zhang, Zexiao, Xu Zheng, Hanshan Luo, Chan Gao, Xiaowei Xue, Jingcheng Zhu, Ruobin Li, Changqing Jin, and Xiaohui Yu. 2024. "Superconductivity in ZrB12 under High Pressure" Metals 14, no. 9: 1082. https://doi.org/10.3390/met14091082

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop