Next Article in Journal
Effect of Artificial Aging on Mechanical and Tribological Properties of CAD/CAM Composite Materials Used in Dentistry
Next Article in Special Issue
Effect of Cooling Rate during Glazing on the Mechanical and Optical Properties of Monolithic Zirconia with 3 mol% Yttria Content
Previous Article in Journal
Investigating Optimal Confinement Behaviour of Low-Strength Concrete through Quantitative and Analytical Approaches
Previous Article in Special Issue
Flash Sintering of YSZ/Al2O3 Composites: Effect of Processing and Testing Conditions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

An Engineering Zirconia Ceramic Made of Baddeleyite

by
Vyacheslav V. Rodaev
*,
Andrey O. Zhigachev
,
Alexander I. Tyurin
,
Svetlana S. Razlivalova
,
Viktor V. Korenkov
and
Yuri I. Golovin
Institute for Nanotechnology and Nanomaterials, Derzhavin Tambov State University, Internatsionalnaya Str. 33, 392000 Tambov, Russia
*
Author to whom correspondence should be addressed.
Materials 2021, 14(16), 4676; https://doi.org/10.3390/ma14164676
Submission received: 15 July 2021 / Revised: 12 August 2021 / Accepted: 18 August 2021 / Published: 19 August 2021

Abstract

:
Wet high-energy milling and uniaxial pressing are used to fabricate CaO-stabilized tetragonal zirconia polycrystalline ceramic (Ca-TZP) with decent mechanical characteristics, i.e., a hardness of 11.5 GPa, Young’s modulus of 230 GPa, and fracture toughness of 13 MPa·m0.5. The effect of CaO concentration and the sintering temperature on phase composition and mechanical characteristics of CaO-stabilized zirconia ceramic made of baddeleyite is investigated.

1. Introduction

Zirconia can exist in three allotropic forms: monoclinic (m-ZrO2), tetragonal (t-ZrO2), and cubic (c-ZrO2). Pure zirconia is monoclinic from ambient temperature to 1170 °C. In the range of 1170–2370 °C, it is tetragonal, and from 2370 °C to the melting point, it is cubic [1]. Upon cooling reverse t-ZrO2→m-ZrO2, a transition occurs at 950 °C which is accompanied with an increase in volume by about 4.5% [1]. It leads to unalloyed zirconia cracking upon cooling and provides its poor mechanical properties. Various stabilizing oxides are used to prevent an undesirable t-ZrO2→m-ZrO2 transition and to keep t-ZrO2 at room temperature [2]. Stabilized zirconia, consisting of t-ZrO2 grains, are called tetragonal zirconia polycrystals (TZP). TZP exists in the certain range of the dopant concentrations.
TZP ceramics are attractive because of their excellent room-temperature mechanical properties [3]. Mainly TZP ceramics are produced through chemical processing of zircon. Baddeleyite is another natural source of zirconia containing monoclinic ZrO2 in the range of 96.5–98.5 wt%. Baddeleyite is much cheaper than zirconia synthesized from zircon, but chemical ways of stabilizing are not efficient for baddeleyite. In [4], high-energy milling was successfully applied to baddeleyite to prepare CaO-stabilized ZrO2 nanopowder. CaO was chosen as an inexpensive alternative to Y2O3 and CeO2 which commonly used for producing engineering TZP ceramics.
Mechanical properties of TZP ceramics strongly depend on its grain size [5]. Above a critical grain size, TZP ceramics are susceptible to spontaneous t-ZrO2→m-ZrO2 transformation throughout the material volume. Below a certain grain size, local t-ZrO2→m-ZrO2 transformation induced by mechanical impact, giving rise to TZP ceramics toughening, does not occur, resulting in reduced TZP ceramics fracture toughness. The increase in sintering temperature and time leads to larger grain size [5].
The aim of this work is to define CaO concentration and the sintering regime which gives rise to a Ca-TZP ceramic originating from baddeleyite which possesses competitive mechanical properties.

2. Materials and Methods

We produced 1–5 wt% CaO-stabilized ZrO2 powders by wet high-energy co-milling of CaO powder (Sigma-Aldrich, Saint Louis, Missouri, USA) and the baddeleyite concentrate powder with zirconia content of 99.3% (5 μm, Kovdorsky mining and processing plant, Kovdor, Russia) using the planetary mill Pulverisette 7 Premium Line (Fritsch, Idar-Oberstein, Germany) in the same way as described in [4]. The size of zirconia nanoparticles in the prepared powders is less than 20 nm. We also used the powder of chemically synthesized monoclinic zirconia (99.9%, 5 μm, Sigma-Aldrich, Saint Louis, Missouri, USA) instead of the baddeleyite concentrate to produce reference samples.
The obtained powders were uniaxially pressed under 560 MPa into pellets of a 10 mm diameter and a 2 mm thickness. The fabricated pellets were sintered at the temperature range from 1100 to 1400 °C for 4 h in air atmosphere in a muffle furnace. The sintering temperature was 500 °C and the cooling rate was 5 °C/min. The samples were cooled naturally from 500 °C to room temperature.
Phase composition analysis of sintered ceramics was performed with the help of an X-ray diffractometer (XRD) D2 Phaser (Bruker AXS, Karlsruhe, Germany) at room temperature. The XRD patterns were recorded in the 20–80° 2θ range and assigned using the PDF-2 Diffraction Database File compiled by the International Centre for Diffraction. Phase content was determined from the XRD patterns by the Rietveld method in the TOPAS software (Bruker AXS, Karlsruhe, Germany). The average grain size of t-ZrO2 and c-ZrO2 was calculated in the TOPAS software (Bruker AXS, Karlsruhe, Germany) by applying the Scherrer equation to the characteristic peaks of the tetragonal and cubic phases located in the 20–80° 2θ range.
Young’s modulus of the samples was measured on a nanoindentometer G200 (MTS Nano Instruments, Oak Ridge, TN, USA) equipped with a Berkovich diamond indenter. Young’s modulus of the samples was calculated from the load-penetration depth curves obtained under peak load of 5 N using the Oliver–Pharr method [6]. Hardness of the samples was measured by Vickers indentation with a load of 19.62 N on a hardness tester Duramin A300 (Struers, Copenhagen, Denmark). Fracture toughness (KC) was calculated from length of radial cracks starting from the corners of the indents with the Anstis equation [7]:
K C = 0.016 ( E H ) 0.5 P C 1.5
where H is Vickers hardness, E is Young’s modulus, P is indentation load to produce cracks, and C is the crack length (the distance between the center of the indent and crack tip). The Vickers indentation fracture toughness tests were performed on a hardness tester Duramin A300 (Struers, Copenhagen, Denmark) using loads of 196.2 and 294.3 N. Cracks lengths were calculated using optical microscope Axio Observer.A1m (Carl Zeiss, Oberkochen, Germany). Before tests, all the samples were polished with diamond-containing slurries. All mechanical characteristics measurements were carried out at room temperature. Sintered ceramics containing m-ZrO2 are not tested because of numerous cracks already present in the samples before measurements.

3. Results and Discussion

Figure 1 shows ceramics XRD patterns evolution with the rise in the dopant CaO concentration increase at a fixed sintering temperature of 1300 °C.
The m-ZrO2 phase dominates in the ceramic if 1-wt% CaO is used. The characteristic peaks of m-ZrO2 at 24.1°, 28.2°, 31.5°, 34.2°, 49.3°, 50.6°, and 55.5° are observed in the XRD pattern. The most intensive characteristic peak of t-ZrO2 at 30.2° is very weak compared to m-ZrO2 reflections. Phase composition of ceramic dramatically changes at 2 wt% CaO. The peaks of m-ZrO2 disappear and only the peaks of t-ZrO2 at 30.2°, 34.6°, 35.2°, 50.2°, 50.7°, 59.3°, and 60.2° are observed in the XRD pattern. This indicates that a 2-wt% CaO-ZrO2 ceramic is a TZP ceramic. A further increase in CaO concentration from 3 to 5 wt% leads to an increase in c-ZrO2 content from 7 to 23 wt% and, respectively, t-ZrO2 content reduction from 93 to 77 wt%. The intensity of the c-ZrO2 peaks increase with the rise in c-ZrO2 content. The characteristic peaks of c-ZrO2 at 35.0° and 59.7° are clearly observed in the XRD pattern of a 5-wt% CaO-ZrO2 ceramic. Other characteristic peaks of c-ZrO2 at 30.1° and 50.2° overlap with the neighboring peaks of t-ZrO2. With the rise in CaO concentration from 2 to 5 wt%, the average grain size of t-ZrO2 decreases from 93 to 63 nm and the average grain size of c-ZrO2, on the contrary, increases to 83 nm.
It is revealed that phase composition of CaO-ZrO2 ceramic is sensitive to the sintering temperature higher than 1300 °C despite CaO concentration (Figure 2). If the sintering temperature is increased to 1400 °C m-ZrO2 becomes to dominate. The t-ZrO2 is practically absent in a ceramic and c-ZrO2 content increases with the rise in CaO concentration as evidenced by an increase in the intensity of the characteristic peaks of c-ZrO2.
Figure 3 presents the effect of the sintering temperature on phase composition of the fabricated TZP ceramic doped by 2-wt% CaO.
It is found that the 2-wt% CaO-ZrO2 ceramic consists only of t-ZrO2 if sintering is performed at 1300 °C and lower temperatures. An increase in the sintering temperature induces zirconia grains growth. The average grain size of t-ZrO2 increases from 79 to 93 nm with the rise in the sintering temperature from 1100 to 1300 °C. Phase composition of the 2-wt% CaO-ZrO2 ceramic dramatically changes if sintering temperature of 1350 °C is used. The TZP ceramic transforms into one mainly composes of m-ZrO2. The radical change in phase composition of the 2-wt% CaO-ZrO2 ceramic with the rise in the sintering temperature from 1300 to 1350 °C is related to critical t-ZrO2 grain size. In [8], it was found that, in a Ca-TZP ceramic, the t-ZrO2 grain size cannot exceed 100 nm. The Ca-TZP ceramic examined in [8] was produced using uniaxial pressing from a nanopowder obtained by hydrothermal treatment of the co-precipitated calcium and zirconium hydroxides. In our case, the average grain size of t-ZrO2 is 93 nm if the Ca-TZP ceramic is sintered at 1300 °C. For comparison, in Y2O3-stabilized TZP ceramics, t-ZrO2 grains up to 1 μm in size can exist [9]. It explains higher sintering temperatures applying to Y2O3-stabilized TZP ceramics.
Peaks’ broadening in the XRD patterns of m-ZrO2 containing ceramics (Figure 1, Figure 2 and Figure 3) may be due to significant mechanical stresses resulting in visually observed ceramic cracking.
Mechanical characteristics of CaO-containing zirconia ceramics made from baddeleyite were examined. It was found that an increase in CaO concentration from 2 to 5 wt% leads to a rise in hardness by 5.5% and a decrease in fracture toughness by 29.9% (Table 1). Wherein, Young’s modulus remains unchanged within the measurement error.
It can be seen from Table 1 that a 2-wt% CaO-ZrO2 ceramic being TZP ceramic has the best combination of mechanical characteristics and the highest fracture toughness value. In terms of hardness and Young’s modulus, the fabricated 2-wt% CaO-ZrO2 ceramic corresponds to engineering Y-TZP ceramics made of chemically synthesized zirconia stabilized with Y2O3, and surpasses them in terms of fracture toughness [1]. High fracture toughness of TZP ceramics is a result of transformation toughening [10]. The stress field in the crack tip zone induces local t-ZrO2→m-ZrO2 transformation which causes volume expansion and shear strains. It applies compressive stress at the crack tip to prevent crack propagation. As a result, further crack growth is suppressed and strength of a ceramic increases. The rise in a dopant concentration leads to a decrease in the fraction of transformable t-ZrO2 in favor of c-ZrO2 incapable of transformation. It explains fracture toughness reduction in the fabricated zirconia ceramic with an increase in CaO concentration (Table 1). It can be seen that ceramic hardness increases when fracture toughness decreases. We suppose that it is due to lower hardness of the m-ZrO2 ceramic compared to the t-ZrO2 one. Indeed, hardness measurements induce t-ZrO2→m-ZrO2 transformation under an indenter tip. Thus, resulting hardness includes hardness of t-ZrO2 and m-ZrO2 phases. A decrease in the fraction of transformable t-ZrO2 negatively affects mechanically-induced t-ZrO2→m-ZrO2 transformation and respectively reduces m-ZrO2 contribution to the resulting hardness. To confirm this assumption, we tested a 1-wt% CaO-ZrO2 ceramic, containing more than 90 wt% m-ZrO2, using a nanoindentometer due to numerous cracks in the sample. Hardness was calculated from the obtained load–penetration depth curves using the Oliver–Pharr method. The prints size was significantly smaller than the analyzed region bounded by the cracks. A 5-wt% CaO-ZrO2 ceramic, mainly composed of t-ZrO2 with a reduced ability to t-ZrO2→m-ZrO2 transformation, was tested on a nanoindentometer too. A nanoindentation by a Berkovich diamond indenter with a tip of a 20-nm radius showed that hardness of a 1-wt% CaO-ZrO2 ceramic is significantly lower than the hardness of a 5-wt% CaO-ZrO2 ceramic, namely, 8.24 ± 0.37 GPa versus 12.63 ± 0.41 GPa.
It should be noted that an increase in CaO-ZrO2 ceramic hardness with the rise in CaO concentration cannot be explained in terms of the empirical Hall-Petch relationship [11], which describes the phenomenon whereby hardness (or strength) of materials increases with reducing the grain size:
H = H 0 + k D
where H is the measured hardness, H0 is the intrinsic hardness related to the resistance of lattice to dislocation motion, k is the material-specific strengthening coefficient, and D is the average grain size.
The observed dependence of CaO-ZrO2 ceramic hardness on the inversed square root of the effective grain size is not a line as required by the Hall–Petch relationship (Figure 4). The effective grain size of the CaO-ZrO2 ceramic was calculated for each CaO concentration from the range of 2–5 wt% knowing the content of t-ZrO2 and c-ZrO2 as well as the average grain size of t-ZrO2 and c-ZrO2 (Table 2).
It is revealed that the rise in the sintering temperature in the range of 1100–1200 °C results in simultaneous increase in hardness, fracture toughness, and Young’s modulus of a 2-wt% CaO-ZrO2 ceramic (Table 3) despite its phase composition constancy (Figure 3). It allows concluding that the observed improvement in mechanical properties is related to ceramic densification.
Indeed, the sintering temperature of 1100 °C is lower than the Tammann temperature of zirconia [12]; therefore, the fabricated ceramic is unsintered and possesses poor mechanical characteristics. On the contrary, if sintering occurs at 1200–1300 °C, a well-sintered dense ceramic with decent mechanical characteristics is obtained. The density of a 2-wt% CaO-ZrO2 ceramic sintered at temperatures of 1100–1300 °C measured by the Archimedes method with distilled water as the immersing medium is presented in Table 4.
According to the data in Table 3 and Table 4 the rise in the sintering temperature leads to the ceramic densification (porosity reduction), which improves the mechanical characteristics of the ceramic. The similar effect was observed, for example, in [13] where an increase in the sintering temperature resulted in Y2O3-stabilized TZP ceramic densification, which in turn led to the rise in Young’s modulus, bending strength and fracture toughness of the given ceramic.
Fracture toughness of a 2-wt% CaO-ZrO2 ceramic is 35.5% higher if sintering occurs at 1300 °C and not at 1200 °C. It can be related to the over-stabilization effect [2] since phase composition of the ceramic remains unchanged with the rise in the sintering temperature from 1200 to 1300 °C. For a given dopant concentration, TZP ceramic fracture toughness decreases as t-ZrO2 grains become smaller and their ability to t-ZrO2→m-ZrO2 transformation reduces. An increase in the sintering temperature induces observed t-ZrO2 grains growth and thus facilitates t-ZrO2→m-ZrO2 transformation following the introduction of a crack.
To reduce energy consumption, the effect of sintering time on mechanical characteristics of a 2-wt% CaO-ZrO2 ceramic was examined too. It was found that sintering time reduction from 4 to 1 h has no effect on hardness, fracture toughness, and Young’s modulus of a 2-wt% CaO-ZrO2 ceramic within the measurements error. The 2-wt% CaO-ZrO2 ceramic sintered at 1300 °C for 1 h is characterized by the following mechanical characteristics: hardness of 11.52 ± 0.05 GPa, Young’s modulus of 226 ± 8 GPa, and fracture toughness of 12.83 ± 0.41 MPa·m0.5.
The used baddeleyite concentrate contains impurities in the amount of 0.7 wt%. The dominating ones are SiO2, SO3, P2O5, TiO2, Fe2O3, CaO, and MgO [14]. To investigate the effect of impurities on the mechanical characteristics of zirconia ceramics, we compared hardness, fracture toughness, and Young’s modulus of CaO-ZrO2 ceramics made of the baddeleite concentrate and chemically synthesized monoclinic zirconia. According to Table 1, 2-wt% CaO-ZrO2 ceramics sintered at 1300 °C were chosen for testing. It was found that both ceramics made of chemically synthesized monoclinic zirconia and the baddeleyite concentrate are TZP ceramics. Their mechanical characteristics are given in Table 5.
It can be seen from Table 5 that the values of analyzed mechanical characteristics for both ceramics are similar. It indicates that impurities amount in the baddeleyite concentrate is too low to worsen the mechanical characteristics of resulting ceramics. Thus, zirconia ceramics with the same mechanical characteristics can be produced using the baddeleyite concentrate as a raw material instead of the more expensive chemically synthesized monoclinic zirconia. However, it should be noted that zirconia ceramic made of the baddeleyite concentrate is yellow-tinted due to impurities (Figure 5).

4. Conclusions

Optimal dopant concentration and the sintering regime to obtain Ca-TZP ceramic originating from baddeleyite with competitive mechanical characteristics are defined. The 2-wt% CaO-ZrO2 ceramic sintered at 1300 °C for 1 h possesses hardness ~11.5 GPa, Young’s modulus ~230 GPa, and fracture toughness ~13 MPa·m0.5. In terms of hardness and Young’s modulus, the fabricated 2-wt% CaO-ZrO2 ceramic corresponds to engineering Y-TZP ceramics made of chemically synthesized zirconia stabilized with Y2O3, and surpasses them in terms of fracture toughness. It is revealed that the rise in the CaO concentration from 2 to 5 wt% stimulates c-ZrO2 content increase and a Ca-TZP ceramic transforms into a ceramic, consisting of both t-ZrO2 and c-ZrO2. An increase in CaO concentration leads to the rise in hardness and a decrease in fracture toughness of a CaO-ZrO2 ceramic wherein its Young’s modulus remains unchanged. Furthermore, it is found that Ca-TZP ceramic cannot be produced if the sintering temperature of 1350 °C or higher is used.

Author Contributions

V.V.R.: investigation, writing—original draft preparation, editing; A.O.Z.: investigation; A.I.T.: investigation; S.S.R.: investigation; V.V.K.: investigation; Y.I.G.: supervision. All authors have read and agreed to the published version of the manuscript.

Funding

The reported study was funded by Russian Foundation for Basic Research (RFBR) according to the research project No. 18-29-17047 and partially supported by the Ministry of Science and Higher Education of the Russian Federation in the frame work of agreement No. 075-15-2021-709.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data included in this study are available upon request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gautam, C.; Joyner, J.; Gautam, A.; Rao, J.; Vajtai, R. Zirconia based dental ceramics: Structure, mechanical properties, biocompatibility and applications. Dalton Trans. 2016, 45, 19194–19215. [Google Scholar] [CrossRef] [PubMed]
  2. Kelly, J.R.; Denry, I. Stabilized zirconia as a structural ceramic: An overview. Dent. Mater. 2008, 24, 289–298. [Google Scholar] [CrossRef] [PubMed]
  3. Zhang, X.; Wu, X.; Shi, J. Additive manufacturing of zirconia ceramics: A state-of-the-art review. J. Mater. Res. Technol. 2020, 9, 9029–9048. [Google Scholar] [CrossRef]
  4. Zhigachev, A.O.; Umrikhin, A.V.; Golovin, Y.I.; Farber, B.Y. Preparation of nanocrystalline calcia-stabilized tetragonal zirconia by high-energy milling of baddeleyite. Int. J. Appl. Ceram. Technol. 2015, 12, E82–E89. [Google Scholar] [CrossRef]
  5. El-Ghany, O.S.A.; Sherief, A.H. Zirconia based ceramics, some clinical and biological aspects: Review. Future Dent. J. 2016, 2, 55–64. [Google Scholar] [CrossRef]
  6. Fischer-Cripps, A.C. Nanoindentation, 3rd ed.; Springer: New York, NY, USA, 2011; p. 282. [Google Scholar]
  7. Anstis, G.R.; Chantikul, P.; Lawn, B.R.; Marshall, D.B. A critical evaluation of indentation techniques for measuring fracture toughness: I, direct crack measurements. J. Am. Ceram. Soc. 1981, 64, 533–538. [Google Scholar] [CrossRef]
  8. Pyda, W.; Haberko, K. CaO-containing tetragonal ZrO2 polycrystals (Ca-TZP). Ceram. Int. 1987, 13, 113–118. [Google Scholar] [CrossRef]
  9. Bannister, M. Science and Technology of Zirconia V, 1st ed.; CRC Press: Boca Raton, FL, USA, 1993; p. 880. [Google Scholar]
  10. Chevalier, J.; Gremillard, L.; Virkar, A.V.; Clarke, D.R. The tetragonal-monoclinic transformation in zirconia: Lessons learned and future trends. J. Am. Ceram. Soc. 2009, 92, 1901–1920. [Google Scholar] [CrossRef]
  11. Trunec, M. Effect of grain size on mechanical properties of 3Y-TZP ceramics. Ceram. Silikáty 2008, 52, 165–171. [Google Scholar]
  12. Xu, Y.; Luo, C.; Zheng, Y.; Ding, H.; Wang, Q.; Shen, Q.; Li, X.; Zhang, L. Characteristics and performance of CaO-based high temperature CO2 sorbents derived from a sol–gel process with different supports. RSC Adv. 2016, 6, 79285–79296. [Google Scholar] [CrossRef]
  13. Deng, Z.-Y.; Yang, J.-F.; Beppu, Y.; Ando, M.; Ohji, T. Effect of agglomeration on mechanical properties of porous zirconia fabricated by partial sintering. J. Am. Ceram. Soc. 2002, 85, 1961–1965. [Google Scholar] [CrossRef]
  14. Lokshin, É.P.; Lebedev, V.N.; Lyakhov, V.P.; Kampel’, F.B.; Popovich, V.F. Zirconium-containing materials for ceramics and refractories manufactured from baddeleyite-containing products of the Kovdorskii mining-and-dressing works joint-stock co. Refract. Ind. Ceram. 2002, 43, 353–358. [Google Scholar] [CrossRef]
Figure 1. The XRD patterns of a CaO-ZrO2 ceramic sintered at 1300 °C containing different dopant amount; m, t, c—the most intense characteristic peaks of m-ZrO2, t-ZrO2 and c-ZrO2, respectively.
Figure 1. The XRD patterns of a CaO-ZrO2 ceramic sintered at 1300 °C containing different dopant amount; m, t, c—the most intense characteristic peaks of m-ZrO2, t-ZrO2 and c-ZrO2, respectively.
Materials 14 04676 g001
Figure 2. The XRD patterns of (a) 2-wt% CaO-ZrO2 and (b) 3-wt% CaO-ZrO2 ceramics sintered at temperatures in the range of 1200–1400 °C; m, t, c—the most intense characteristic peaks of m-ZrO2, t-ZrO2 and c-ZrO2, respectively.
Figure 2. The XRD patterns of (a) 2-wt% CaO-ZrO2 and (b) 3-wt% CaO-ZrO2 ceramics sintered at temperatures in the range of 1200–1400 °C; m, t, c—the most intense characteristic peaks of m-ZrO2, t-ZrO2 and c-ZrO2, respectively.
Materials 14 04676 g002
Figure 3. The XRD patterns of 2-wt% CaO-ZrO2 ceramic sintered at different temperatures; m, t—the most intense characteristic peaks of m-ZrO2 and t-ZrO2, respectively.
Figure 3. The XRD patterns of 2-wt% CaO-ZrO2 ceramic sintered at different temperatures; m, t—the most intense characteristic peaks of m-ZrO2 and t-ZrO2, respectively.
Materials 14 04676 g003
Figure 4. The dependence of hardness of the CaO-ZrO2 ceramic sintered at 1300 °C containing a different CaO amount on the inversed square root of the effective grain size. The dots are connected by lines for clarity.
Figure 4. The dependence of hardness of the CaO-ZrO2 ceramic sintered at 1300 °C containing a different CaO amount on the inversed square root of the effective grain size. The dots are connected by lines for clarity.
Materials 14 04676 g004
Figure 5. The photo of 2-wt% CaO-ZrO2 ceramics made of the baddeleyite concentrate (left) and chemically synthesized monoclinic zirconia (right). The samples are prepared for mechanical tests.
Figure 5. The photo of 2-wt% CaO-ZrO2 ceramics made of the baddeleyite concentrate (left) and chemically synthesized monoclinic zirconia (right). The samples are prepared for mechanical tests.
Materials 14 04676 g005
Table 1. Mechanical characteristics of CaO-stabilized zirconia ceramics of a baddeleyite origin sintered at 1300 °C containing different dopant amount.
Table 1. Mechanical characteristics of CaO-stabilized zirconia ceramics of a baddeleyite origin sintered at 1300 °C containing different dopant amount.
CaO Concentration,
wt%
Hardness,
GPa
Fracture Toughness,
MPa·m0.5
Young’s Modulus,
GPa
211.57 ± 0.1013.14 ± 0.49228 ± 10
311.64 ± 0.0611.89 ± 0.27227 ± 5
411.96 ± 0.0910.44 ± 0.41223 ± 6
512.21 ± 0.069.21 ± 0.35225 ± 9
Table 2. The content of t-ZrO2 and c-ZrO2 phases and the average grain size of t-ZrO2 and c-ZrO2 phases in CaO-ZrO2 ceramic sintered at 1300 °C containing a different CaO amount.
Table 2. The content of t-ZrO2 and c-ZrO2 phases and the average grain size of t-ZrO2 and c-ZrO2 phases in CaO-ZrO2 ceramic sintered at 1300 °C containing a different CaO amount.
CaO Concentration,
wt%
t-ZrO2c-ZrO2
Content, wt%Average Grain Size, nmContent, wt%Average Grain Size, nm
299931-
39378757
486691467
577632383
Table 3. Mechanical characteristics of a 2-wt% CaO-ZrO2 ceramic of a baddeleyite origin sintered at different temperatures.
Table 3. Mechanical characteristics of a 2-wt% CaO-ZrO2 ceramic of a baddeleyite origin sintered at different temperatures.
Sintering Temperature,
°C
Hardness,
GPa
Fracture Toughness,
MPa·m0.5
Young’s Modulus,
GPa
11006.34 ± 0.094.96 ± 0.39131 ± 11
120011.93 ± 0.079.70 ± 0.34227 ± 11
130011.57 ± 0.1013.14 ± 0.49228 ± 10
Table 4. The density and the relative density of 2 wt% CaO-ZrO2 ceramic of a baddeleyite origin sintered at different temperatures.
Table 4. The density and the relative density of 2 wt% CaO-ZrO2 ceramic of a baddeleyite origin sintered at different temperatures.
Sintering Temperature, °CDensity, cm3/gRelative Density, %
11004.3772.8
12005.8597.5
13005.9699.3
Table 5. Mechanical characteristics of 2-wt% CaO-ZrO2 ceramics made of the baddeleyite concentrate and chemically synthesized monoclinic zirconia. The sintering temperature is 1300 °C.
Table 5. Mechanical characteristics of 2-wt% CaO-ZrO2 ceramics made of the baddeleyite concentrate and chemically synthesized monoclinic zirconia. The sintering temperature is 1300 °C.
Raw Material of 2 wt% CaO-ZrO2 CeramicHardness,
GPa
Fracture Toughness,
MPa·m0.5
Young’s Modulus,
GPa
Baddeleite concentrate11.57 ± 0.1013.14 ± 0.49228 ± 10
Chemically synthesized monoclinic zirconia 11.97 ± 0.1212.86 ± 0.43221 ± 11
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rodaev, V.V.; Zhigachev, A.O.; Tyurin, A.I.; Razlivalova, S.S.; Korenkov, V.V.; Golovin, Y.I. An Engineering Zirconia Ceramic Made of Baddeleyite. Materials 2021, 14, 4676. https://doi.org/10.3390/ma14164676

AMA Style

Rodaev VV, Zhigachev AO, Tyurin AI, Razlivalova SS, Korenkov VV, Golovin YI. An Engineering Zirconia Ceramic Made of Baddeleyite. Materials. 2021; 14(16):4676. https://doi.org/10.3390/ma14164676

Chicago/Turabian Style

Rodaev, Vyacheslav V., Andrey O. Zhigachev, Alexander I. Tyurin, Svetlana S. Razlivalova, Viktor V. Korenkov, and Yuri I. Golovin. 2021. "An Engineering Zirconia Ceramic Made of Baddeleyite" Materials 14, no. 16: 4676. https://doi.org/10.3390/ma14164676

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