Cannabis Sativa

Zirconium diboride

STM image of the (2×2)-reconstructed ZrB2 (0001) surface[1]
Names
IUPAC name
Zirconium diboride
Other names
ZrB2
Identifiers
ECHA InfoCard 100.031.772 Edit this at Wikidata
  • InChI=1S/B2.Zr/c1-2;
    Key: NXBOAJHBGIROOR-UHFFFAOYSA-N
Properties
ZrB2
Molar mass 112.85 g/mol
Appearance grey-black powder
Density 6.085 g/cm3
Melting point ~3246 °C
Insoluble
Structure
Hexagonal, hP3
P6/mmm, No. 191
Hazards
Occupational safety and health (OHS/OSH):
Main hazards
Uninvestigated
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).
checkY verify (what is checkY☒N ?)

Zirconium diboride (ZrB2) is a highly covalent refractory ceramic material with a hexagonal crystal structure. ZrB2 is an ultra-high temperature ceramic (UHTC) with a melting point of 3246 °C. This along with its relatively low density of ~6.09 g/cm3 (measured density may be higher due to hafnium impurities) and good high temperature strength makes it a candidate for high temperature aerospace applications such as hypersonic flight or rocket propulsion systems. It is an unusual ceramic, having relatively high thermal and electrical conductivities, properties it shares with isostructural titanium diboride and hafnium diboride.

ZrB2 parts are usually hot pressed (pressure applied to the heated powder) and then machined to shape. Sintering of ZrB2 is hindered by the material's covalent nature and presence of surface oxides which increase grain coarsening before densification during sintering. Pressureless sintering of ZrB2 is possible with sintering additives such as boron carbide and carbon which react with the surface oxides to increase the driving force for sintering but mechanical properties are degraded compared to hot pressed ZrB2.[2]

Additions of ~30 vol% SiC to ZrB2 is often added to ZrB2 to improve oxidation resistance through SiC creating a protective oxide layer - similar to aluminium's protective alumina layer.[3]

ZrB2 is used in ultra-high temperature ceramic matrix composites (UHTCMCs).[4][5][6][7][8][9][10][11]

Carbon fiber reinforced zirconium diboride composites show high toughness while silicon carbide fiber reinforced zirconium diboride composites are brittle and show a catastrophic failure.

Preparation[edit]

ZrB2 can be synthesized by stoichiometric reaction between constituent elements, in this case Zr and B. This reaction provides for precise stoichiometric control of the materials.[12] At 2000 K, the formation of ZrB2 via stoichiometric reaction is thermodynamically favorable (ΔG=−279.6 kJ mol−1) and therefore, this route can be used to produce ZrB2 by self-propagating high-temperature synthesis (SHS). This technique takes advantage of the high exothermic energy of the reaction to cause high temperature, fast combustion reactions. Advantages of SHS include higher purity of ceramic products, increased sinterability, and shorter processing times. However, the extremely rapid heating rates can result in incomplete reactions between Zr and B, the formation of stable oxides of Zr, and the retention of porosity. Stoichiometric reactions have also been carried out by reaction of attrition milled (wearing materials by grinding) Zr and B powder (and then hot pressing at 600 °C for 6 h), and nanoscale particles have been obtained by reacting attrition milled Zr and B precursor crystallites (10 nm in size).[13] Reduction of ZrO2 and HfO2 to their respective diborides can also be achieved via metallothermic reduction. Inexpensive precursor materials are used and reacted according to the reaction below:

ZrO2 + B2O3 + 5Mg → ZrB2 + 5MgO

Mg is used as a reactant to allow for acid leaching of unwanted oxide products. Stoichiometric excesses of Mg and B2O3 are often required during metallothermic reductions to consume all available ZrO2. These reactions are exothermic and can be used to produce the diborides by SHS. Production of ZrB2 from ZrO2 via SHS often leads to incomplete conversion of reactants, and therefore double SHS (DSHS) has been employed by some researchers.[14] A second SHS reaction with Mg and H3BO3 as reactants along with the ZrB2/ZrO2 mixture yields increased conversion to the diboride, and particle sizes of 25–40 nm at 800 °C. After metallothermic reduction and DSHS reactions, MgO can be separated from ZrB2 by mild acid leaching.

Synthesis of UHTCs by boron carbide reduction is one of the most popular methods for UHTC synthesis. The precursor materials for this reaction (ZrO2/TiO2/HfO2 and B4C) are less expensive than those required by the stoichiometric and borothermic reactions. ZrB2 is prepared at greater than 1600 °C for at least 1 hour by the following reaction:

2ZrO2 + B4C + 3C → 2ZrB2 + 4CO

This method requires a slight excess of boron, as some boron is oxidized during boron carbide reduction. ZrC has also been observed as a product from the reaction, but if the reaction is carried out with 20–25% excess B4C, the ZrC phase disappears, and only ZrB2 remains. Lower synthesis temperatures (~1600 °C) produce UHTCs that exhibit finer grain sizes and better sinterability. Boron carbide must be subjected to grinding prior to the boron carbide reduction to promote oxide reduction and diffusion processes.

Boron carbide reductions can also be carried out via reactive plasma spraying if a UHTC coating is desired. Precursor or powder particles react with plasma at high temperatures (6000–15000 °C) which greatly reduces the reaction time.[15] ZrB2 and ZrO2 phases have been formed using a plasma voltage and current of 50 V and 500 A, respectively. These coating materials exhibit uniform distribution of fine particles and porous microstructures, which increased hydrogen flow rates.

Another method for the synthesis of UHTCs is the borothermic reduction of ZrO2, TiO2, or HfO2 with B.[16] At temperatures higher than 1600 °C, pure diborides can be obtained from this method. Due to the loss of some boron as boron oxide, excess boron is needed during borothermic reduction. Mechanical milling can lower the reaction temperature required during borothermic reduction. This is due to the increased particle mixing and lattice defects that result from decreased particle sizes of ZnO2 and B after milling. This method is also not very useful for industrial applications due to the loss of expensive boron as boron oxide during the reaction.

Nanocrystals of ZrB2were successfully synthesized by Zoli's reaction, a reduction of ZrO2 with NaBH4 using a molar ratio M:B of 1:4 at 700 °C for 30 min under argon flow.[17][18]

ZrO2 + 3NaBH4 → ZrB2 + 2Na(g,l) + NaBO2 + 6H2(g)

ZrB2 can be prepared from solution-based synthesis methods as well, although few substantial studies have been conducted. Solution-based methods allow for low temperature synthesis of ultrafine UHTC powders. Yan et al. have synthesized ZrB2 powders using the inorganic-organic precursors ZrOCl2•8H2O, boric acid and phenolic resin at 1500 °C.[19] The synthesized powders exhibit 200 nm crystallite size and low oxygen content (~ 1.0 wt%). ZrB2 preparation from polymeric precursors has also been recently investigated. ZrO2 and HfO2 can be dispersed in boron carbide polymeric precursors prior to reaction. Heating the reaction mixture to 1500 °C results in the in situ generation of boron carbide and carbon, and the reduction of ZrO2 to ZrB2 soon follows.[20] The polymer must be stable, processable, and contain boron and carbon to be useful for the reaction. Dinitrile polymers formed from the condensation of dinitrile with decaborane satisfy these criteria.

Chemical vapor deposition can be used to prepare zirconium diboride. Hydrogen gas is used to reduce vapors of zirconium tetrachloride and boron trichloride at substrate temperatures greater than 800 °C.[21] Recently, high-quality thin films of ZrB2 can also be prepared by physical vapor deposition.[22]

Defects and secondary phases in zirconium diboride[edit]

Zirconium diboride gains its high-temperature mechanical stability from the high atomic defect energies (i.e. the atoms do not deviate easily from their lattice sites).[23] This means that the concentration of defects will remain low, even at high temperatures, preventing failure of the material.

The layered bonding between each layer is also very strong but means that the ceramic is highly anisotropic, having different thermal expansions in the 'z' <001> direction. Although the material has excellent high temperature properties, the ceramic has to be produced extremely carefully as any excess of either zirconium or boron will not be accommodated in the ZrB2 lattice (i.e. the material does not deviate from stoichiometry). Instead it will form extra lower melting point phases which may initiate failure under extreme conditions.[23]

Diffusion and transmutation in zirconium diboride[edit]

Zirconium diboride is also investigated as a possible material for nuclear reactor control rods due to the presence of boron. [citation needed]

10B + nth → [11B] → α + 7Li + 2.31 MeV.

The layered structure provides a plane for helium diffusion to occur. He is formed as a transmutation product of boron-10—it is the alpha particle in the above reaction—and will rapidly migrate through the lattice between the layers of zirconium and boron, however not in the 'z' direction. Of interest, the other transmutation product, lithium, is likely to be trapped in the boron vacancies that are produced by the boron-10 transmutation and not be released from the lattice.[23]

References[edit]

  1. ^ Fleurence, A.; Friedlein, R.; Ozaki, T.; Kawai, H.; Wang, Y.; Yamada-Takamura, Y. (2012). "Experimental Evidence for Epitaxial Silicene on Diboride Thin Films". Physical Review Letters. 108 (24): 245501. Bibcode:2012PhRvL.108x5501F. doi:10.1103/PhysRevLett.108.245501. PMID 23004288.
  2. ^ Zhang, S. C; Hilmas, G. E; Fahrenholtz, W. G (2006). "Pressureless Densification of Zirconium Diboride with Boron Carbide Additions". Journal of the American Ceramic Society. 89 (5): 1544–50. doi:10.1111/j.1551-2916.2006.00949.x.
  3. ^ Fahrenholtz, William G (2007). "Thermodynamic Analysis of ZrB2–SiC Oxidation: Formation of a SiC-Depleted Region". Journal of the American Ceramic Society. 90 (1): 143–8. doi:10.1111/j.1551-2916.2006.01329.x.
  4. ^ Zoli, L.; Sciti, D. (2017). "Efficacy of a ZrB 2 –SiC matrix in protecting C fibres from oxidation in novel UHTCMC materials". Materials & Design. 113: 207–213. doi:10.1016/j.matdes.2016.09.104.
  5. ^ Zoli, L.; Vinci, A.; Silvestroni, L.; Sciti, D.; Reece, M.; Grasso, S. (2017). "Rapid spark plasma sintering to produce dense UHTCs reinforced with undamaged carbon fibres". Materials & Design. 130: 1–7. doi:10.1016/j.matdes.2017.05.029.
  6. ^ Sciti, D.; Zoli, L.; Silvestroni, L.; Cecere, A.; Martino, G.D. Di; Savino, R. (2016). "Design, fabrication and high velocity oxy-fuel torch tests of a C f -ZrB 2 - fiber nozzle to evaluate its potential in rocket motors". Materials & Design. 109: 709–717. doi:10.1016/j.matdes.2016.07.090.
  7. ^ Galizia, Pietro; Failla, Simone; Zoli, Luca; Sciti, Diletta (2018). "Tough salami-inspired C f /ZrB 2 UHTCMCs produced by electrophoretic deposition". Journal of the European Ceramic Society. 38 (2): 403–409. doi:10.1016/j.jeurceramsoc.2017.09.047.
  8. ^ Vinci, Antonio; Zoli, Luca; Sciti, Diletta; Melandri, Cesare; Guicciardi, Stefano (2018). "Understanding the mechanical properties of novel UHTCMCs through random forest and regression tree analysis". Materials & Design. 145: 97–107. doi:10.1016/j.matdes.2018.02.061.
  9. ^ Zoli, L.; Medri, V.; Melandri, C.; Sciti, D. (2015). "Continuous SiC fibers-ZrB 2 composites". Journal of the European Ceramic Society. 35 (16): 4371–4376. doi:10.1016/j.jeurceramsoc.2015.08.008.
  10. ^ Sciti, D.; Murri, A. Natali; Medri, V.; Zoli, L. (2015). "Continuous C fibre composites with a porous ZrB2 Matrix". Materials & Design. 85: 127–134. doi:10.1016/j.matdes.2015.06.136.
  11. ^ Sciti, D.; Pienti, L.; Murri, A. Natali; Landi, E.; Medri, V.; Zoli, L. (2014). "From random chopped to oriented continuous SiC fibers–ZrB2 composites". Materials & Design. 63: 464–470. doi:10.1016/j.matdes.2014.06.037.
  12. ^ Çamurlu, H. Erdem & Filippo Maglia. (2009). "Preparation of nano-size ZrB 2 powder by self-propagating high-temperature synthesis". Journal of the European Ceramic Society. 29 (8): 1501–1506. doi:10.1016/j.jeurceramsoc.2008.09.006.
  13. ^ Chamberlain, Adam L.; William G. Fahrenholtz; Gregory E. Hilmas (2009). "Reactive hot pressing of zirconium diboride". Journal of the European Ceramic Society. 29 (16): 3401–3408. doi:10.1016/j.jeurceramsoc.2009.07.006.
  14. ^ Nishiyama, Katsuhiro; et al. (2009). "Preparation of ultrafine boride powders by metallothermic reduction method". Journal of Physics: Conference Series. 176 (1): 012043. Bibcode:2009JPhCS.176a2043N. doi:10.1088/1742-6596/176/1/012043.
  15. ^ Karuna Purnapu Rupa, P.; et al. (2010). "Microstructure and Phase Composition of Composite Coatings Formed by Plasma Spraying of ZrO2 and B4C Powders". Journal of Thermal Spray Technology. 19 (4): 816–823. Bibcode:2010JTST...19..816K. doi:10.1007/s11666-010-9479-y. S2CID 136019792.
  16. ^ Peshev, P. & Bliznakov, G. (1968). "On the borothermic preparation of titanium, zirconium and hafnium diborides". Journal of the Less Common Metals. 14: 23–32. doi:10.1016/0022-5088(68)90199-9.
  17. ^ Zoli, Luca; Costa, Anna Luisa; Sciti, Diletta (December 2015). "Synthesis of nanosized zirconium diboride powder via oxide-borohydride solid-state reaction". Scripta Materialia. 109: 100–103. doi:10.1016/j.scriptamat.2015.07.029.
  18. ^ Zoli, Luca; Galizia, Pietro; Silvestroni, Laura; Sciti, Diletta (23 January 2018). "Synthesis of group IV and V metal diboride nanocrystals via borothermal reduction with sodium borohydride". Journal of the American Ceramic Society. 101 (6): 2627–2637. doi:10.1111/jace.15401.
  19. ^ Yan, Yongjie; et al. (2006). "New Route to Synthesize Ultra-Fine Zirconium Diboride Powders Using Inorganic–Organic Hybrid Precursors". Journal of the American Ceramic Society. 89 (11): 3585–3588. doi:10.1111/j.1551-2916.2006.01269.x.
  20. ^ Su, Kai & Sneddon, Larry G. (1993). "A polymer precursor route to metal borides". Chemistry of Materials. 5 (11): 1659–1668. doi:10.1021/cm00035a013.
  21. ^ Randich, E. (1979). "Chemical vapor deposited borides of the form (Ti,Zr)B2 and (Ta,Ti)B2". Thin Solid Films. 63 (2): 309–313. Bibcode:1979TSF....63..309R. doi:10.1016/0040-6090(79)90034-8.
  22. ^ Magnuson, Martin; Tengdelius, Lina; Greczynski, Grzegorz; Hultman, Lars; Högberg, Hans (2018). "Chemical bonding in epitaxial ZrB2 studied by X-ray spectroscopy". Thin Solid Films. 649: 89–96. arXiv:1801.08663. Bibcode:2018TSF...649...89M. doi:10.1016/j.tsf.2018.01.021. S2CID 103292992.
  23. ^ a b c Middleburgh, Simon C.; Parfitt, David C.; Blair, Paul R.; Grimes, Robin W. (2011). "Atomic Scale Modeling of Point Defects in Zirconium Diboride". Journal of the American Ceramic Society. 94 (7): 2225–2229. doi:10.1111/j.1551-2916.2010.04360.x.

Leave a Reply