Next Article in Journal
Effect of Rubber Fiber Content on the Mechanical Properties of Calcareous Sand
Previous Article in Journal
Biomechanical Comparison of Titanium and CFR-PEEK Intramedullary Nails Using Finite Element Analysis
Previous Article in Special Issue
Dynamic Mechanical Behavior of Nanosilica-Based Epoxy Composites Under LEO-like UV-C Exposure
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Production and Mechanical Performance of Tantalum Strengthened Alumina–Zirconia Composites with Graphene Addition

by
Pavel Peretyagin
1,2,
Oleg Yanushevich
2,
Natella Krikheli
2,
Yuri Pristinskiy
1,
Nestor Washington Solis Pinargote
1,2,
Anton Smirnov
1,* and
Nikita Grigoriev
1
1
Spark Plasma Sintering Research Laboratory, Moscow State University of Technology “STANKIN”, Vadkovsky per. 1, 127055 Moscow, Russia
2
Scientific Department, Federal State Budgetary Educational Institution of Higher Education Russian University of Medicine of the Ministry of Health of the Russian Federation, Dolgorukovskaya Str. 4, 127006 Moscow, Russia
*
Author to whom correspondence should be addressed.
J. Compos. Sci. 2025, 9(11), 577; https://doi.org/10.3390/jcs9110577
Submission received: 12 August 2025 / Revised: 23 October 2025 / Accepted: 24 October 2025 / Published: 27 October 2025
(This article belongs to the Special Issue Mechanical Properties of Composite Materials and Joints)

Abstract

High density alumina–zirconia–tantalum ceramic–metal composites with the addition of 0.5 vol.% of graphene oxide were fabricated via a wet processing technique followed by spark plasma sintering. Scanning electron microscopy confirmed the even distribution of metal particles in the composite matrix. The thermal reduction of graphene oxide after consolidation at 1500 °C was proved using Raman spectroscopy. The engineered materials exhibit a fracture resistance of 16 MPa∙m1/2, which is 30% greater than in the reference ZTA ceramic composites fabricated using the same technology. That increase in fracture toughness could be down to a synergetic interaction mechanism; more specifically, crack trapping, renucleation and blunting, and elongated tantalum particles bridging. In addition to the above-mentioned mechanisms, tetragonal monoclinic phase transformation in zirconia is also an additional source of increased crack resistance in the developed composites.

1. Introduction

Alumina (Al2O3) and zirconia (ZrO2) have long been the subject of intensive research by developers of new materials [1,2,3,4,5,6,7,8,9,10,11,12,13,14]. The capacity to manufacture these inorganic compounds from readily available and cost-effective starting materials has contributed to their widespread adoption, for example, as cutting tools for high-speed operations, dental and orthopedic implants, electrical and thermal insulation, wearable components, coatings, and more. Despite the exceptional chemical, thermal, mechanical, and tribological properties of these ceramics, it is important to remember their fragility. The fracture toughness values for alumina and zirconia typically fall into a range of 3.3–6 MPa·m1/2 и 5–12 MPa·m1/2, respectively. In recent years, vast numbers of studies have explored the benefits of combining aluminum and zirconia (ZTA) oxides. The purpose of creating these materials is to leverage the high hardness of alumina while simultaneously enhancing the toughness of the composite through the martensitic transformation in zirconia [15,16]. Furthermore, to avoid the self-induced transformation of ZrO2, its concentration in the composite must be below the percolation limit, which was determined to be 16% by volume or 22% by weight [17]. However, the presence of defects in ceramic materials, which can occur during different stages of manufacturing and processing, can lead to catastrophic failure even when the stress intensity is below the strength of the material. Ceramic reinforcement is one of the ways to avoid brittle fracture. The toughening phases employed can be depicted as whiskers, filaments, flakes, and so on. These elements exert compressive forces that mitigate the stress concentration at the fracture end, thereby preventing the crack from propagating further. There are diverse methods for enhancing the fracture toughness, such as bending or pulling out, or plastic deformation of reinforcing phases. The extent to which the crack is closed is determined by the ratio of the stresses and deformations in the reinforcements, which are affected by the yield strength of the ligament, its diameter, orientation, volume ratio, and the essence of the relationship between the matrix and reinforced particles. In recent decades several investigations of zirconia–metal, alumina–metal or zirconia–alumina–metal composites have been conducted. Huang et al. investigated Al2O3–tantalum composite laminates fabricated by hot pressing [18]. The maximum value of flexural strength (1090 ± 340 MPa) of the studied composite was more than twice the bending strength of Al2O3 subjected to high-temperature pressing under identical circumstances. Recently, researchers investigated the high-temperature mechanical performance of alumina coarse-grained composites containing 11 or 21 vol. % of metals with exceptionally high melting points [19]. The compression testing conducted at temperatures ranging from 1300 °C to 1500 °C revealed that the studied composites were resistant to brittle fracture. Thompson et al. found that the average fracture toughness values of Al2O3-based nanocomposites strengthened with niobium (Nb) and/or carbon nanotubes were almost twice as high as those of alumina nanocomposites without metal phase [20]. The fracture toughness of hot-pressed zirconia–tantalum (ZrO2/Ta) ceramic metal composites was evaluated in [21]. Due to crack bridging and plastic deformations of ductile ligaments, the high toughness value (16 ± 0.6 MPa m1/2) of fabricated ceramic–metal composites was obtained. The fatigue performance of ZrO2/Ta and ZrO2/Ta ceramic–metal composites produced by spark plasma sintering was analyzed in [22]. It was reported that, in contrast to niobium, the presence of tantalum improves a material’s ability to withstand repeated stress and resistance and safely sustain defects within a zirconia matrix due to the numerous arrangements working at various dimensionalities. It also reduces topical strain because of the permanent deformation, which in turn causes internal crack resistance and also affects external factors. Bartolomé et al. studied the mechanical properties of ceramic–metal micro–nano-hybrid composites [23]. Adding 20 vol.% of metal ligaments to alumina–nano–zirconia ceramic allowed them to increase toughness from 6 to 12 MPa∙m1/2 due to the bridging effect of metal inclusions. The ability to withstand defects safely and the crack resistance in comparison to the crack length of these composites were also evaluated [24]. It was stated that a synergistic toughening mechanism leads to enhanced damage tolerance and R-curve behavior through the combined effects of crack bridging by metal particles and stress-induced phase transformation of zirconia, resulting in a fracture toughness greater than the sum of individual contributions. Zirconia–alumina–tantalum micro-nanocomposite shows high values of crack resistance (16 MPa∙m1/2) and strength (1300 MPa) due to the significant influence on the martensitic transformation of ZrO2 and to the synchronous action of different strengthening mechanisms [25]. In addition to its outstanding mechanical properties, the studied micro-nanocomposite has low susceptibility to aging. However, it should be noted that the initial mixture of ceramic powders consisted of 37.4 wt.% alumina and 52.4 62.6 wt.% hybrid ZrO2 nanopowders, which were obtained by the by CO2 laser co-vaporization method. In the current investigation, Ta was chosen as a strengthening agent for manufacturing the ZTA ceramic composites. Metal phase exhibits a level of plasticity that is not present in ceramic materials. Consequently, in metal–ceramic composites, the energy required for irreversible strain can be harnessed when a crack spreading within the base material encounters a ductile ligament. A portion of the energy required for fracture spreading is absorbed through irreversible stress, while the rest of the energy is insufficient and may prevent it spreading further. Furthermore, deformed metal particulates can bridge cracks, thus redistributing stresses and preventing further spread. In addition, Al2O3, ZrO2, and Ta have been employed in medical applications [26,27]. In vitro and in vivo biocompatibility examinations of ceramic–metal materials were performed. The studied composites had no deleterious effect on cell proliferation, extracellular matrix production, or on cell morphology, and exhibited biocompatibility and osteoconductivity [28]. The properties and behavior of zirconia–tantalum composites were evaluated in vitro and in vivo [29,30]. In vitro studies show that composites create a surface favorable for cell growth. In vivo observation demonstrates that there is no fibrous tissue or inflammatory response at the boundary between implants and newly formed bone indicates successful osseointegration. This means the bone directly bonds to the implant surface without any intervening soft tissue, which is a key characteristic of a stable and successful implant. However, in addition to the bio- and mechanical performance, the tribological performance of materials used in joint orthopedic devices must also be considered. Wear resistance is important because it can affect the following factors. Implant stability: low wear resistance can lead to loosening of the implant, which will require surgery to restore vision. Tissue reaction: wear and debris can cause adverse tissue reactions, such as inflammation and bone loss (osteolysis), and wear can contribute to metallosis, a condition involving the release of metal debris into surrounding tissues. Clinical longevity: ultimately, wear resistance significantly determines the implant’s long-term clinical performance and the need for future interventions.
Graphene (G) has the ability to reduce friction, making it a potential solid lubricant. It can also decrease the force of friction operating on the surfaces in contact at the micro- and nanolevels [31]. Recent articles confirm that the presence of graphene in ceramic-based composites can enhance their wear performance [32,33,34,35,36,37,38,39,40]. Su et al. showed that incorporation of 0.2 wt% graphene nanoplatelets led to a significant reduction in the wear rate of tungsten carbide-ZrO2-Al2O3 ceramics [41]. Llorente et al. reported that cubic yttria-stabilized ZrO2 materials containing graphene nanoplatelets (GNPs) provide outstanding triboresistance and, in particular, an excellent enhancement of the durability of more than two orders of magnitude versus composites without GNP [42]. Sun et al. [43] fabricated different ceramic matrix/graphene composites and studied their tribology properties. An unprecedentedly weak (0.06) frictional force between the tribosurfaces was recorded because of the synergistic effect between the mechanical reliability and self-lubricating performance. Cygan et al. [44] investigated the synergistic effect of mechanical performance and tribofilm formation on the wear performance of the alumina-based materials toughened with multilayered graphene and graphene oxide. The conducted analysis showed that formation of tribofilm on the rubbing surfaces was sufficient for an effective reduction in wear. Duntu et al. [45] investigated the tribological behavior of nanocomposites with an alumina matrix strengthened with ZrO2, G and a nanosized, hollow cylindrical form of carbon. Unlike the monolithic Al2O3, the produced alumina matrix composites showed improved wear resistance (~93%) due to the generation of the solid-surface thin layer formed on the rubbed surfaces. Based on the above factors, to provide steps for further investigation—namely the evaluation of the tribological performance of studied ZTA-Ta ceramic–metal composites—graphene oxide (GO) was also incorporated into this composite. Therefore, the primary goal of this research is to analyze the mechanical capabilities of sintered ZTA-20 vol.%Ta ceramic metal composites and, at the same time, an assessment of the effect of the presence of 0.5 vol.% of graphene oxide on the mechanical performance of the fabricated materials.

2. Materials and Methods

2.1. Raw Powders

The powders that are readily available in the market were utilized as the starting materials: (1) graphite powder (d50 = 3 mm, Plasmotherm, Moscow, Russia); (2) t-ZrO2 (d50 = 0.26 µm, 3Y-TZP, 3 mol% Y2O3; TZ-3YS-E, Tosoh Corp., Tokyo, Japan); (3) alumina (d50 = 0.30 µm, α-Al2O3, A16SG, Alcoa, New York, NY, USA); (4) tantalum (d50 = 58 µm, Alfa Aesar, Karslruhe, Germany).

2.2. Materials Processing and Sintering

GO was synthesized by oxidizing graphite powder based on the modified Hummers’ method [46,47]. Tantalum flakes with d50 = 42 µm were obtained after attrition milling the initial metal powder in a teflon jar for 4 h in isopropyl medium. To produce a ceramic matrix incorporating 20 vol.% Ta lamellar particles, the powders were initially mixed in distilled water medium and then subjected to spark plasma sintering (SPS) at 1500 °C (100 °C/min) for 3 min with uniaxial pressures of 80 MPa in vacuum environment. Detailed information on ceramic and metal slurry processing, as well as sintering parameters, can be found elsewhere [48]. The as-sintered samples with a 20 mm and ±3 mm of diameter and height, respectively, were fabricated. After consolidation, GO was thermally restored in situ. Hereafter, the resulting carbon material will be referred to as reduced graphene oxide (rGO).

2.3. X-Ray Diffraction (XRD) and Raman Characterization

Analysis of the phases of studied materials and the presence of m- and t-zirconia phases on the samples’ polished and fractured surfaces was performed using an Empyrean diffractometer (PANalytical, Almelo, The Netherlands) with a radiation source Cu–Kα (λ = 1.5405981 Å) operated with an intensity of 30 mA at 40 kV. The amount of m-ZrO2 and its volume fraction were estimated based on the work of Garvie and Nicholson [49] as well as calculations by Toraya et al. [50], respectively. The presence of a graphic structure in the composition before and after sintering was determined via a Raman spectroscopy analyzer (DXRTM2, Thermo Fisher Scientific, Waltham, MA, USA) using a laser with a wavelength of 532 nm and a power of 8.0 mW.

2.4. Microstructural and Mechanical Characterization

The densities of the SPSed specimens were measured by the Archimedes hydrostatic weighing procedure. The microstructure of the metalographically ground and polished samples was analyzed using scanning electron microscopy (SEM; LYRA3, Tescan, Brno, Czech Republic) and a Focused Ion Beam Scanning Electron Microscope (FIB-SEM; AURIGA 60 CrossBeam Workstation, Carl Zeiss Microscopy, Jena, Germany). Energy-dispersive spectroscopy (EDS) was performed using X-Act (Oxford Instruments, Abingdon, UK) detector. The hardness (Hv) was determined by the Vickers method. The measurements were performed using a hardness indenter (Qness, Salzburg, Austria). The hardness testing conditions per measurement (10 imprints) were 294 N loading force with 10 s loading time. To determine the mean value, the following equation was used:
Hv = 0.1891 P d 2
where P is load (N) and d is the average length of two diagonals (mm). The values of the Vickers fracture toughness (KC) were calculated using the approach described in [51]. The flexural strength (σf) was calculated through a biaxial bending test [52]. This scheme test is recognized as a reliable and preferred method for studying fragile specimens, since the highest stress occurs in the middle of the specimens, while defects on the edges of the prismatic samples used in the three-point bending test can act as stress concentration sites and may affect the test results. The tested sample was mounted on three steel balls (∅ 3 mm) arranged in a circle with a diameter of 10 mm. To determine flexural strength, 12 SPSed disks were broken using a 5 kN universal testing machine AutoGraph AG-X (Shimadzu Corp., Kyoto, Japan). The testing scheme and method used to calculate strength have been explained in a previous article [53].

3. Results and Discussion

SEM images and X-ray diffractions graphs of raw metallic powder before (Figure 1A) and after (Figure 1B) attrition milling was analyzed.
For XRD examinations, a small volume of the Ta-acetone suspension was carefully applied to the silicon (510) single-crystal wafer within a specially produced support device and the suspension was dried by evaporating the acetone. The location and area under the peaks were compared with the reference (PDF 004-0788) pattern. During milling, metal particles deformed between the grinding zirconia balls, the impeller shaft and the walls of the teflon jar. Therefore, visible texturing made the (200) reflection parallel to the specimen surface and broadening occurred as a result of deformation and fundamentally altered the crystal lattice distortion. As a result, the shape of most of the initial Ta powder particles changed to a lamellar one. The distinctive X-ray patterns and Raman spectra of ZTA-Ta-GO mixture and consolidated materials are displayed in Figure 2.
According to the data provided by the ICDD, the locations and strengths of the peaks observed in the analyzed samples were consistent with the expected peaks for monoclinic (24-1165) and tetragonal (83-0113) ZrO2, aluminum oxide (81-1667), and Ta (004-0788). The data obtained only showed the presence of phases of zirconium dioxide, aluminum oxide, and tantalum, which confirms the absence of the formation of any secondary phases or contamination during processing. It is also necessary to note the complete transformation of monoclinic zirconia in the initial powder mixture into tetragonal zirconia after consolidation. The Raman spectra of ZTA-20Ta-GO powder mixture and sintered ZTA-20Ta-GO composites are shown in Figure 2B. The spectrum in Figure 2B demonstrates the wide G peak and the slight second-order area features of sp1-, sp2-, and sp3-hybridized C-C bonds in graphene [54] and the D band (∼1350 cm−1) exhibit structural defects such as lattice distortion in ZTA-20Ta-graphene oxide powder [55]. The shift for the SPSed samples (Figure 2B) shows that graphene oxide was reduced in situ (rGO) after consolidation. This is supported by the lower-intensity proportion between D- and G-bands (ID/IG). The analysis revealed that the proportion of bands in the powder mixture was 1.02, whereas in the compact samples, the ID/IG ratio decreased to 0.35, suggesting a reduced defect level in the rGOs in the heat-treated samples. Moreover, a distinct and well-defined bivariate symmetrical shift was observed at approximately 2700 cm−1. The I2D/IG ratio in sintered composites was observed to be 0.68, compared to the initial powder mixture, which had a ratio of 0.14. This also demonstrates the recovery of the graphene structure following the sintering process. Consequently, the findings demonstrate the thermal transformation of graphene oxide, including the reduction of extensive regions of sp2, during the sintering at 1500 °C.
Particles of various shapes (whiskers, fibers, flakes, etc.) can be used to harden brittle materials. The phenomenon of the bridge effect arises when ductile inclusions either absorb the strain energy during fracture (1, 9), detach from the sintered body and elongate (2, 8), or relieve the stress at the crack tip without breaking (3, 7), as illustrated in Figure 3. Furthermore, the phenomenon of crack deflection takes place when the top of the fracture encounters a particle, causing the crack to alter its trajectory of expansion (4). Conversely, these inclusions contribute to the blunting of fracture top (5) or the dissipation of stress through plastic strain in the areas where cracks are growing (6).
The scanning electron microscopy images of the metallographic and fractured surfaces of the sintered specimens subjected to the flexural test are presented in Figure 4. The dark and light phases pertain to ZTA matrix and tantalum, respectively (Figure 4A). The metal particles are evenly distributed throughout the ceramic matrix.
Examination of the fracture areas revealed that the failure process involved either full or incomplete plastic strain and the debonding of Ta inclusions. The composite mechanical characteristics are defined by a multifaceted equilibrium of linkage at different levels, and they are largely influenced by the adhesion between the ductile particles and ceramic components. If the interface is tremendous, the high level of strain it experiences will result in brittle failure of the reinforcing particles, thereby preventing a considerable gain in crack resistance. Conversely, if the bonding is poor, the ductile inclusions will simply detach from the matrix, also preventing an improvement in fracture toughness. Thus, the boundary between ceramic and ductile particles must be capable of resisting cleavage and acting as a crack arrestor, particularly when the tantalum ligaments are big or have poor ductility. The sintered composite exhibits a variety of fracture patterns in tantalum particles of diverse shapes. On the other hand, these are fragments of metal that have been shattered due to plastic deformation, either fully or partially, as a result of stretching. In such instances, the rupture of metallic lamellae takes the form of a ductile dimpled fracture, characterized by the presence of rounded depressions and protrusions on the fractured surface (Figure 4B). Conversely, there are tantalum inclusions that have detached from the ceramic, as well as those that have been fractured predominantly by cleavage (Figure 4C). Therefore, it can be inferred that an enhancement in KC will occur when the interface surfaces show moderate levels of separation, allowing the ductile ligaments to deform plastically. The tantalum inclusions within the ceramic base demonstrate elastic properties until a fracture occurs. When this happens, the particle’s surface undergoes a change in shape, but the particle itself remains unbroken, and is still attached to the surrounding material. Even after the particle is surrounded by a crack, the metal stretches due to plastic deformation, resulting in deformation. The intense plastic deformation of ductile ligaments prior to fracture ultimately results in a substantial increase in crack resistance.Figure 5 shows a scanning electron microscopy image of the fracture surface of zirconia–alumina–tantalum/rGO composite.
As can be seen in Figure 5, a very thin layer of graphene flake seems to be strongly anchored within the ceramic matrix and a large interface area between the graphene and ceramic matrix is created during the consolidation process. The development of the crack propagation within the microstructure of the specimens after the indentation has been investigated using the FIB-SEM microscope (Figure 6). These findings indicate the occurrence of plastic deformation and the formation of cracks in the Ta ligaments. Crack-tip blunting and renucleation mechanisms are also visible (Figure 6, yellow arrows). The crack can be halted at the metal particle in such a way that it must re-form in other places.
The main role in increasing fracture resistance in the samples under consideration is played by the crack bridging and irreversible strain of the ductile ligaments. The fracture can be halted at the Ta phase, such that it must reintegrate on the other side. The results of density measurements showed that a level of compaction close to the theoretical level (99.8%) was achievable. In other publications [56,57,58,59,60], it has been reported that a small quantity (0.1 to 0.5 percent by volume) of graphene may improve the mechanical characteristics of the specimens. The composites with and without graphene addition exhibited the following average values of strength 826 ± 29 MPa and 841 ± 22 MPa, respectively. No significant increase in Hv was found in the studied samples to which graphene had been added. The values of the specimens with (8.8 ± 0.3 GPa) and without (8.7 ± 0.4 GPa) rGO were virtually identical. This can be attributed to the fact that graphene, being a malleable substance, does not enhance hardness.
The phase characterization of the polished surfaces of SPSed materials demonstrated that the tetragonal zirconia phase was the most prevalent. The results proved that the volume fractions of t- and m-ZrO2 on the polished surface in sintered composites were 97% and 3%, respectively. However, on the surface of a broken material, the volume fraction of the tetragonal phase decreased to a value of 80%, while the monoclinic phase increased to a value of 20%. This can be attributed to the formation of an oxide film of tantalum oxide (Ta2O5) which covered the initial metal powder particles that came into in contact with air. The Ta2O5-ZrO2 phase diagram shows the solubility of pentavalent oxide in zirconium dioxide (approximately 2 mol.% Ta2O5) for compositions containing a high amount of zirconium dioxide at temperatures above 1400 °C. We presume that upon consolidation, the layer that formed in the metal particles was incorporated into the solid solution. The enhanced transformability can be related to the alloying effect on the tetragonality, i.e., the cell parameter ratio c/a, of stabilized t-ZrO2 [61]. The amount of tantalum oxide that enters the solid solution inside the matrix is an important factor. Kim et al. reported that solid solutions of t-ZrO2 in the ZrO2-Y2O3-Ta2O5 system, where Ta2O5 is added as a secondary phase and with c/a axial ratio smaller than 1.0203, have fracture toughness values within the range of 5–14 MPa∙m1/2 [62]. Therefore, it is very important to maintain such a concentration of Ta2O5 in order not to exceed this value and create conditions to support the material’s ability to achieve t-m transformation. The tetragonality (c/a) of the samples was determined as the coefficient between the parameters of the crystal lattice. The results obtained are shown in Table 1.
Thus, the presence of tantalum in the ceramic matrix increases the tetragonality values and, accordingly, increases the material’s predisposition to undergo tetragonal–monoclinic transformation, while at the same time this value does not exceed the established limit (1.0203). Utilizing the same technique as in previous studies [22,25,47], the molar percentage was calculated using the elemental weight percentages obtained by energy dispersion spectroscopy and the atomic masses of the elements. The proportion of tantala was determined to be less than 2 mol.%, and no other phases were observed. The occurrence of Ta2O5 enhances the tetragonal deformation of the cubic structure, which significantly increases the ability of zirconium dioxide to undergo transformation. Therefore, t→m phase transformation of zirconia in ceramic–metal composites is an additional mechanism for enhancing fracture toughness. The present findings indicate that the mutually reinforcing effect between crack bridging and the irreversible strain of the ductile tantalum inclusions and the tetragonal-to-monoclinic phase transformation of ZrO2 induces an improvement in the fracture toughness of ZTA-Ta material of up to 16.1 ± 0.3 MPa∙m1/2 compared to value (8.6 ± 0.3 MPa∙m1/2) obtained in previously studied ZTA ceramic composites [51]. The application of the Wilcoxon test to the biaxial flexure test results showed no differences between the samples (p-value > 0.05). When comparing fracture toughness results, statistically significant differences between specimens (p-value < 0.05) with and without rGO were revealed.

4. Conclusions

For the first time, high-density ZTA-Ta-GO material using a combination of wet processing and field-assisted sintering technology was successfully fabricated. The examination of the microscopic structure revealed a consistent dispersion of tantalum particles throughout the ceramic framework. The results of the Raman analysis confirmed that the selected consolidation temperature (1500 °C) was sufficient for thermal reduction of graphene oxide. The presence of ductile metal particles of tantalum (20 vol.%) allowed us to increase the crack resistance of sintered composites to 15.5 MPa∙m1/2. In addition to crack deflection and interfacial debonding of metal ligaments, a significant role in crack growth resistance in sintered composites is played by crack bridging and plastic deformation of elongated tantalum ligaments. It should also be noted that the presence of an oxide layer on the surface of the initial tantalum particles is a strong stabilizer, which, in turn, promotes the martensitic transformation in zirconium dioxide and also helps to increase the crack resistance of the sintered composites.
Combined use of metal phase and graphene oxide (0.5 vol.%) only slightly increases the crack resistance (16.1 MPa∙m1/2) of sintered composites. The hardness values of specimens with and without GO are almost the same, since graphene oxide is a soft material and therefore does not significantly affect hardness. Based on studies of the mechanical behavior of developed ceramic metal materials with the addition of small quantities of graphene oxide, it can be concluded that in future, these materials could be excellent candidates for use in medicine. However, it is worth noting that, despite the outstanding mechanical properties, this is not enough. Tribological studies must be carried out to assess the impact of graphene on the wearability of the composite being fabricated. These studies will be described in our next article. Only after careful analysis of tribological test data will a decision be made regarding the need for in vitro and in vivo studies.

Author Contributions

Conceptualization, A.S. and P.P.; data curation, N.W.S.P. and Y.P.; formal analysis, N.W.S.P. and P.P.; funding acquisition, N.G. and O.Y.; investigation, A.S., P.P. and Y.P.; methodology, O.Y. and N.K.; project administration, N.G. and A.S.; resources, N.G. and N.K.; software, Y.P. and P.P.; supervision, A.S. and O.Y.; validation, P.P.; N.G. and N.K.; visualization, N.W.S.P., Y.P. and P.P.; writing—original draft, A.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Health of the Russian Federation, grant number 056-00041-23-00.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Acknowledgments

The study was carried out on the equipment of the Center of Collective Use “State Engineering Center” of the MSUT “STANKIN” (project 075-15-2021-695, unique id RF—2296.61321X0013). The authors would like to thank Nikita Nikitin (Spark Plasma Sintering Research Laboratory, Moscow State University of Technology “STANKIN”, Russia) for his advice on data processing.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Kaushal, S.; Zeeshan, M.D.; Ansari, M.I.; Sharma, D. Progress in tribological research of Al2O3 ceramics: A review. Mater. Today 2023, 82, 163–167. [Google Scholar]
  2. Chen, W.; Han, M.; Yang, S. Research Progress of Al2O3 Ceramic Composites. J. Mater. Eng. 2011, 3, 91–96. [Google Scholar]
  3. Wang, H.Z.; Gao, L.; Guo, J.K. The effect of nanoscale SiC particles on the microstructure of Al2O3 ceramics. Ceram. Int. 2000, 26, 391–396. [Google Scholar]
  4. Gao, L.; Hong, J.; Miyamoto, H.; Torre, S.D.D. Bending strength and microstructure of Al2O3 ceramics densified by spark plasma sintering. J. Eur. Ceram. Soc. 2000, 20, 2149–2152. [Google Scholar] [CrossRef]
  5. Shen, Z.; Johnsson, M.; Zhao, Z.; Nygren, M. Spark Plasma Sintering of Alumina. J. Eur. Ceram. Soc. 2002, 85, 1921–1927. [Google Scholar]
  6. Wang, S.W.; Chen, L.D.; Hirai, T. Densification of Al2O3 Powder Using Spark Plasma Sintering. J. Mater. Res. 2011, 15, 982–987. [Google Scholar]
  7. Piconi, C. Alumina. In Comprehensive Biomaterials; Elsevier: Amsterdam, The Netherlands, 2011; pp. 73–94. [Google Scholar]
  8. Smirnov, A.; Podrabinnik, P.A.; Babushkin, N.N.; Kuznetsova, E.V.; Pristinskiy, Y.O.; Khmyrov, R.S. Development of Al2O3 and PLA ceramic-polymer filament for 3D printing by fused deposition modelling method. AIP Conf. Proc. 2022, 2467, 020047. [Google Scholar]
  9. Tuan, W.H.; Chen, R.Z.; Wang, T.C.; Cheng, C.H.; Kuo, P.S. Mechanical properties of Al2O3/ZrO2 composites. J. Eur. Ceram. Soc. 2002, 22, 2827–2833. [Google Scholar]
  10. Meng, F.; Liu, C.; Zhang, F.; Tian, Z.; Huang, W. Densification and mechanical properties of fine-grained Al2O3–ZrO2 composites consolidated by spark plasma sintering. J. Alloys Compd. 2012, 512, 63–67. [Google Scholar] [CrossRef]
  11. Wu, Z.; Liu, W.; Wu, H.; Huang, R.; He, R.; Jiang, Q.; Wu, S. Research into the mechanical properties, sintering mechanism and microstructure evolution of Al2O3-ZrO2 composites fabricated by a stereolithography-based 3D printing method. Mater. Chem. Phys. 2018, 207, 1–10. [Google Scholar]
  12. Nettleship, I.; Stevens, R. Tetragonal zirconia polycrystal (TZP)—A review. Int. J. High Technol. Ceram. 1987, 3, 1–32. [Google Scholar] [CrossRef]
  13. Stevens, R. An introduction to Zirconia: Zirconia and Zirconia Ceramics, 2nd ed.; Magnesium Electrón Publications; Magnesium Electrum: Twickenham, NY, USA, 1986. [Google Scholar]
  14. Garvie, R.C.; Hannink, R.H.; Pascoe, R.T. Ceramic steel? Nature 1975, 258, 703. [Google Scholar] [CrossRef]
  15. Hannink, R.H.J.; Kelly, P.M.; Muddle, B.C. Transformation toughening in zirconia-containing ceramics. J. Am. Ceram. Soc. 2000, 83, 461–487. [Google Scholar] [CrossRef]
  16. 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]
  17. Pecharromán, C.; Bartolomé, J.F.; Requena, J.; Moya, J.S.; Deville, S.; Chevalier, J.; Torrecillas, R. Percolative Mechanism of Aging in Zirconia-Containing Ceramics for Medical Applications. Adv. Mater. 2003, 15, 507–511. [Google Scholar] [CrossRef]
  18. Huang, T.S.; Rahaman, M.N.; Bal, B.S. Alumina–tantalum composite for femoral head applications in total hip arthroplasty. Mater. Sci. Eng. C 2009, 29, 1935–1941. [Google Scholar] [CrossRef]
  19. Weidner, A.; Ranglack-Klemm, Y.; Zienert, T.; Aneziris, C.G.; Biermann, H. Mechanical High-Temperature Properties and Damage Behavior of Coarse-Grained Alumina Refractory Metal Composites. Materials 2019, 12, 3927. [Google Scholar] [CrossRef]
  20. Thomson, K.E.; Jiang, D.; Yao, W.; Ritchie, R.O.; Mukherjee, A.K. Characterization and mechanical testing of alumina-based nanocomposites reinforced with niobium and/or carbon nanotubes fabricated by spark plasma sintering. Acta Mater. 2012, 60, 622–632. [Google Scholar] [CrossRef]
  21. Smirnov, A.; Bartolomé, J.F. Mechanical Properties and Fatigue Life of ZrO2-Ta Composites Prepared by Hot Pressing. J. Eur. Ceram. Soc. 2012, 32, 3899–3904. [Google Scholar] [CrossRef]
  22. Smirnov, A.; Beltrán, J.I.; Rodriguez-Suarez, T.; Pecharromán, C.; Muñoz, M.C.; Moya, J.S.; Bartolomé, J.F. Unprecedented simultaneous enhancement in damage tolerance and fatigue resistance of zirconia/Ta composites. Sci. Rep. 2017, 7, 449. [Google Scholar] [CrossRef]
  23. Bartolomé, J.F.; Gutiérrez-González, C.F.; Torrecillas, R. Mechanical properties of alumina–zirconia–Nb micro–nano-hybrid composites. Compos. Sci. Technol. 2008, 68, 1392–1398. [Google Scholar] [CrossRef]
  24. Gutierrez-Gonzalez, C.F.; Bartolomé, J.F. Damage Tolerance and R-Curve Behavior of Al2O3–ZrO2–Nb Multiphase Composites with Synergistic Toughening Mechanism. J. Mater. Res. 2008, 23, 570–578. [Google Scholar] [CrossRef]
  25. Smirnov, A.; Bartolomé, J.F.; Kurland, H.D.; Grabow, J.; Müller, F.A. Design of a new zirconia–alumina–Ta micro-nanocomposite with unique mechanical properties. J. Am. Ceram. Soc. 2016, 99, 3205–3209. [Google Scholar] [CrossRef]
  26. Burger, W.; Kiefer, G. Alumina, Zirconia and Their Composite Ceramics with Properties Tailored for Medical Applications. J. Compos. Sci. 2021, 5, 306. [Google Scholar] [CrossRef]
  27. Piconi, C.; Sprio, S. Oxide Bioceramic Composites in Orthopedics and Dentistry. J. Compos. Sci. 2021, 5, 206. [Google Scholar] [CrossRef]
  28. Bartolomé, J.F.; Moya, J.S.; Couceiro, R.; Gutiérrez-González, C.F.; Guitián, F.; Martinez-Insua, A. In vitro and in vivo evaluation of a new zirconia/niobium biocermet for hard tissue replacement. Biomaterials 2016, 76, 313–320. [Google Scholar] [CrossRef]
  29. Smirnov, A.; Yanushevich, O.; Krikheli, N.; Solis Pinargote, N.W.; Peretyagin, P.; Grigoriev, S.; Alou, L.; Sevillano, D.; López-Piriz, R.; Guitian, F. 3Y-TZP/Ta Biocermet as a Dental Material: An Analysis of the In Vitro Adherence of Streptococcus Oralis Biofilm and an In Vivo Pilot Study in Dogs. Antibiotics 2024, 13, 175. [Google Scholar] [CrossRef]
  30. Smirnov, A.; Guitián, F.; Ramirez-Rico, J.; Bartolomé, J.F. A zirconia/tantalum biocermet: In vitro and in vivo evaluation for biomedical implant applications. J. Mater. Chem. B. 2024, 12, 8919–8928. [Google Scholar] [CrossRef]
  31. Sun, J.; Du, S. Application of graphene derivatives and their nanocomposites in tribology and lubrication: A review. RSC Adv. 2019, 9, 40642–40661. [Google Scholar] [CrossRef]
  32. Paul, G.; Hirani, H.; Kuila, T.; Murmu, N.C. Nanolubricants Dispersed with Graphene and its Derivatives: An Assessment and Review of the Tribological Performance. Nanoscale 2019, 11, 3458–3483. [Google Scholar] [CrossRef] [PubMed]
  33. Markandan, K.; Chin, J.K.; Tan, M.T.T. Recent progress in graphene based ceramic composites: A review. J. Mater. Res. 2016, 32, 84–106. [Google Scholar] [CrossRef]
  34. Ahmad, I.; Anwar, S.; Xu, F.; Zhu, Y. Tribological investigation of multi-layer graphene reinforced alumina ceramic nanocomposites. J. Tribol. 2018, 141, 022002. [Google Scholar] [CrossRef]
  35. Solis Pinargote, N.W.; Meleshkin, Y.; Bentseva, E.; Kuznetsova, E.; Kytmanov, A.; Kurmysheva, A.Y.; Smirnov, A. Influence of graphene oxide content on the wear resistance of zirconia toughened alumina composites consolidated by spark plasma sintering. High Temp. Mat. Proc. Int. Q. High-Technol. Plasma Proc. 2024, 28, 81–91. [Google Scholar]
  36. Sktani, Z.D.I.; Arab, A.; Mohamed, J.J.; Ahmad, Z.A. Effects of additives additions and sintering techniques on the microstructure and mechanical properties of Zirconia Toughened Alumina (ZTA): A review. Inter. J. Refrac. Met. And Hard Mater. 2022, 106, 105870. [Google Scholar] [CrossRef]
  37. Yu, Z.; Zheng, Y.; Sun, J.; Zhao, J. Recent advances in toughening of alumina-based ceramic machining tools. Inter. J. Refrac. Met. Hard Mater. 2026, 137, 107436. [Google Scholar] [CrossRef]
  38. Wang, R.; Zhang, F.; Yang, K.; Xiao, N.; Tang, J.; Xiong, Y.; Zhang, G.; Duan, M.; Chen, H. Important contributions of carbon materials in tribology: From lubrication abilities to wear mechanisms. J. Alloys Compd. 2024, 979, 173454. [Google Scholar] [CrossRef]
  39. Goswami, S.; Ghosh, R.; Hirani, H.; Mandal, N. Mechano-tribological performance of Graphene/CNT reinforced alumina nanocomposites–Review and quantitative insights. Ceram. Inter. 2022, 48, 11879–11908. [Google Scholar] [CrossRef]
  40. Yang, K.; Zhang, F.; Chen, Y.; Zhang, H.; Xiong, B.; Chen, H. Recent progress on carbon-based composites in multidimensional applications. Composites Part A Appl. Sci. Manuf. 2022, 157, 106906. [Google Scholar] [CrossRef]
  41. Su, Y.; Zhu, T.; Cheng, Y.; Sun, N.; Wang, H.; Li, Y.; Hu, F.; Xie, Z. Tribological properties and wear mechanisms of WC-ZrO2-Al2O3-GNPs ceramics against typical counterparts. Wear 2025, 564, 205705. [Google Scholar] [CrossRef]
  42. Llorente, J.; Belmonte, M. Rolled and twisted graphene flakes as self-lubricant and wear protecting fillers into ceramic composites. Carbon 2020, 159, 45–50. [Google Scholar] [CrossRef]
  43. Sun, C.; Huang, Y.; Shen, Q.; Wang, W.; Pan, W.; Zong, P.A.; Yang, L.; Xing, Y.; Wan, C. Embedding two-dimensional graphene array in ceramic matrix. Sci. adv. 2020, 6, eabb1338. [Google Scholar] [CrossRef] [PubMed]
  44. Cygan, T.; Petrus, M.; Wozniak, J.; Cygan, S.; Teklińska, D.; Kostecki, M.; Jaworska, L.; Olszyna, A. Mechanical properties and tribological performance of alumina matrix composites reinforced with graphene-family materials. Ceram. Inter. 2020, 46, 7170–7177. [Google Scholar] [CrossRef]
  45. Duntu, S.H.; Eliasu, A.; Ahmad, I.; Islam, M.; Boakye-Yiadom, S. Synergistic effect of graphene and carbon nanotubes on wear behaviour of alumina-zirconia nanocomposites. Mater. Characteriz. 2021, 175, 111056. [Google Scholar] [CrossRef]
  46. Hummers, W.S., Jr.; Offeman, R.E. Preparation of graphitic oxide. J. Am. Chem. Soc. 1958, 80, 1339. [Google Scholar] [CrossRef]
  47. Smirnov, A.; Solis Pinargote, N.W.; Peretyagin, N.; Pristinskiy, Y.; Peretyagin, P.; Bartolomé, J.F. Zirconia Reduced Graphene Oxide Nano-Hybrid Structure Fabricated by the Hydrothermal Reaction Method. Materials 2020, 13, 687. [Google Scholar] [CrossRef]
  48. Smirnov, A.; Peretyagin, P.; Bartolomé, J.F. Processing and mechanical properties of new hierarchical metal-graphene flakes reinforced ceramic matrix composites. J. Eur. Ceram. Soc. 2019, 39, 3491–3497. [Google Scholar] [CrossRef]
  49. Garvie, R.C.; Nicholson, P.S. Phase analysis in zirconia systems. J Amer. Ceram.Soc. 1972, 55, 303–305. [Google Scholar] [CrossRef]
  50. Toraya, H.; Yoshimura, M.; Somiya, S. Calibration curve for quantitative analysis of the monoclinic-tetragonal ZrO2 system by X-ray difraction. J. Am. Ceram. Soc. 1984, 67, 119–121. [Google Scholar] [CrossRef]
  51. ISO 6872:2015; Dentistry—Ceramic Materials. International Organization for Standardization: Geneva, Switzerland, 2015.
  52. Miranzo, P.; Moya, J.S. Elastic/plastic indentation in ceramics: A fracture toughness determination method. Ceram. Int. 1984, 10, 147–152. [Google Scholar] [CrossRef]
  53. Grigoriev, S.; Smirnov, A.; Pinargote, N.W.S.; Yanushevich, O.; Kriheli, N.; Kramar, O.; Pristinskiy, Y.; Peretyagin, P. Evaluation of Mechanical and Electrical Performance of Aging Resistance ZTA Composites Reinforced with Graphene Oxide Consolidated by SPS. Materials 2022, 15, 2419. [Google Scholar] [CrossRef] [PubMed]
  54. Tomanik, E.; Christinelli, W.; Souza, R.M.; Oliveira, V.L.; Ferreira, F.; Zhmud, B. Review of Graphene-Based Materials for Tribological Engineering Applications. Eng 2023, 4, 2764–2811. [Google Scholar] [CrossRef]
  55. Shahriary, L.; Athawale, A.A. Graphene oxide synthesized by using modified Hummers approach. Int. J. Renew. Energy Environ. Eng. 2014, 2, 58–63. [Google Scholar]
  56. Miranzo, P.; Ramírez, C.; Román-Manso, B.; Garzón, L.; Gutiérrez, H.R.; Terrones, M.; Ocal, C.; Osendi, M.I.; Belmonte, M. In situ processing of electrically conducting graphene/SiC nanocomposites. J. Eur. Ceram. Soc. 2013, 33, 1665. [Google Scholar] [CrossRef]
  57. Yazdani, B.; Xia, Y.; Ahmad, I.; Zhu, Y. Graphene and carbon nanotube (GNT)-reinforced alumina nanocomposites. J. Eur. Ceram. Soc. 2015, 35, 179. [Google Scholar] [CrossRef]
  58. Lahiri, D.; Khaleghi, E.; Bakshi, S.R.; Li, W.; Olevsky, E.A.; Agarwal, A. Graphene-induced strengthening in spark plasma sintered tantalum carbide–nanotube composite. Scr. Mater. 2013, 68, 285. [Google Scholar] [CrossRef]
  59. Walker, L.S.; Marotto, V.R.; Rafiee, M.A.; Koratkar, N.; Corral, E.L. Toughening in graphene ceramic composites. ACS Nano 2011, 5, 3182. [Google Scholar] [CrossRef]
  60. Grigoriev, S.N.; Vereschaka, A.A.; Vereschaka, A.S.; Kutin, A.A. Cutting tools made of layered composite ceramics with nano-scale multilayered coatings. Procedia CIRP 2012, 1, 301–306. [Google Scholar] [CrossRef]
  61. Loganathan, A.; Gandhi, A.S. Effect of phase transformations on the fracture toughness of yttria stabilized zirconia. Mater. Sci. Eng. A. 2012, 556, 927–935. [Google Scholar] [CrossRef]
  62. Kim, D.-J.; Tien, T.-Y. Phase Stability and Physical Properties of Cubic and Tetragonal ZrO2 in the System ZrO2–Y2O3–Ta2O5. J. Am. Ceram. Soc. 1991, 74, 3061–3065. [Google Scholar] [CrossRef]
Figure 1. XRD analysis and SEM images of raw (A) and milled (B) Ta powder.
Figure 1. XRD analysis and SEM images of raw (A) and milled (B) Ta powder.
Jcs 09 00577 g001
Figure 2. (A) X-Ray measurements of ZTA-20Ta-GO mixture (1) and sintered (2) ZTA-Ta-reduced graphene oxide samples. “α”, “t”, “m” and Ta indicate alumina, tetragonal and monoclinic ZrO2, and tantalum, respectively. (B) Raman shifts in ZTA-20Ta-GO mixture and sintered ZTA-20Ta-rGO composite.
Figure 2. (A) X-Ray measurements of ZTA-20Ta-GO mixture (1) and sintered (2) ZTA-Ta-reduced graphene oxide samples. “α”, “t”, “m” and Ta indicate alumina, tetragonal and monoclinic ZrO2, and tantalum, respectively. (B) Raman shifts in ZTA-20Ta-GO mixture and sintered ZTA-20Ta-rGO composite.
Jcs 09 00577 g002
Figure 3. Strengthening techniques in ceramics with ductile particles of diverse forms.
Figure 3. Strengthening techniques in ceramics with ductile particles of diverse forms.
Jcs 09 00577 g003
Figure 4. SEM micrographs showing the polished (A) and fractured (B,C) surface of zirconia–alumina–tantalum/rGO composite. Yellow arrows demonstrate the crack bridging and plastic deformation of ductile ligaments. Red arrows indicate the separation between the matrix and tantalum ligaments. Blue arrows indicate the cleavage of metal particles. (D) Zirconia–alumina–tantalum/rGO sample with the crack generated after indentations.
Figure 4. SEM micrographs showing the polished (A) and fractured (B,C) surface of zirconia–alumina–tantalum/rGO composite. Yellow arrows demonstrate the crack bridging and plastic deformation of ductile ligaments. Red arrows indicate the separation between the matrix and tantalum ligaments. Blue arrows indicate the cleavage of metal particles. (D) Zirconia–alumina–tantalum/rGO sample with the crack generated after indentations.
Jcs 09 00577 g004
Figure 5. SEM image of the surface of the broken zirconia–alumina–tantalum/rGO composite.
Figure 5. SEM image of the surface of the broken zirconia–alumina–tantalum/rGO composite.
Jcs 09 00577 g005
Figure 6. FIB-SEM micrographs of crack spreading in the ZTA-Ta specimens with reduced graphene oxide. The red arrow shows crack bridging and irreversible deformation of the ductile particle.
Figure 6. FIB-SEM micrographs of crack spreading in the ZTA-Ta specimens with reduced graphene oxide. The red arrow shows crack bridging and irreversible deformation of the ductile particle.
Jcs 09 00577 g006
Table 1. Lattice parameters and tetragonality of studied materials.
Table 1. Lattice parameters and tetragonality of studied materials.
Materialscac/a
Matrix5.17355.09971.014
ZTA-Ta5.18275.08651.0189
ZTA-Ta-rGO5.18275.08651.0189
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

Peretyagin, P.; Yanushevich, O.; Krikheli, N.; Pristinskiy, Y.; Solis Pinargote, N.W.; Smirnov, A.; Grigoriev, N. Production and Mechanical Performance of Tantalum Strengthened Alumina–Zirconia Composites with Graphene Addition. J. Compos. Sci. 2025, 9, 577. https://doi.org/10.3390/jcs9110577

AMA Style

Peretyagin P, Yanushevich O, Krikheli N, Pristinskiy Y, Solis Pinargote NW, Smirnov A, Grigoriev N. Production and Mechanical Performance of Tantalum Strengthened Alumina–Zirconia Composites with Graphene Addition. Journal of Composites Science. 2025; 9(11):577. https://doi.org/10.3390/jcs9110577

Chicago/Turabian Style

Peretyagin, Pavel, Oleg Yanushevich, Natella Krikheli, Yuri Pristinskiy, Nestor Washington Solis Pinargote, Anton Smirnov, and Nikita Grigoriev. 2025. "Production and Mechanical Performance of Tantalum Strengthened Alumina–Zirconia Composites with Graphene Addition" Journal of Composites Science 9, no. 11: 577. https://doi.org/10.3390/jcs9110577

APA Style

Peretyagin, P., Yanushevich, O., Krikheli, N., Pristinskiy, Y., Solis Pinargote, N. W., Smirnov, A., & Grigoriev, N. (2025). Production and Mechanical Performance of Tantalum Strengthened Alumina–Zirconia Composites with Graphene Addition. Journal of Composites Science, 9(11), 577. https://doi.org/10.3390/jcs9110577

Article Metrics

Back to TopTop