Conghui Meng,
Mengfei Xu,
Shiyao Lei,
Yifei Xiao,
Cheng-Feng Du*,
Linze Fan,
Weihong Qi
*,
Long Wang and
Hong Yu
*
State Key Laboratory of Solidification Processing, Center of Advanced Lubrication and Seal Materials, Northwestern Polytechnical University, Xi'an, Shaanxi 710072, P. R. China. E-mail: cfdu@nwpu.edu.cn; qiwh216@nwpu.edu.cn; yh@nwpu.edu.cn
First published on 12th August 2025
MAX phase ceramics have garnered widespread attention owing to their numerous exceptional functionalities from their emerging new phases. M-site element regulation in MAXs stands as one of the core strategies for new phase exploration and performance optimization. However, challenges lie in the compositional design and kinetic regulation to bypass the multiple intermediate phases in acquiring new MAX phases with high purity for target properties. Herein, two new MAX phases with a Ti, Nb and Hf solid solution in the M-site, namely Ti1−xNb1−xHf2xAlC (2x = 0.2 and 0.4), are synthesized using the spark plasma sintering method. Density functional theory (DFT) simulations are employed to study their electronic structures, bonding status of different M–C and M–A bonds, and the corresponding elastic properties. The calculations indicate that in the solid solution MAX phase, the bond strengths follow the order of Nb–C > Hf–C > Ti–C and Nb–Al > Hf–Al > Ti–Al. The corresponding bulk modulus, shear modulus, and Young's modulus are calculated to be 165.55 GPa, 122.76 GPa, and 295.28 GPa, respectively. The mechanical properties of the as-prepared samples are investigated on microscopic and macroscopic scales as well. Specifically, Ti0.9Nb0.9Hf0.2AlC exhibits leading compressive strength of 1574.77 ± 31.15 MPa and fracture toughness of 7.14 ± 0.05 MPa m1/2 among the reported MAX phases. This work highlights the superiority of M-site composition regulation in boosting the mechanical properties of MAX phase ceramics.
In particular, composition design for the M-site element of MAX phase ceramics, from single to double and further to multi-elements, has been demonstrated as an effective approach in property regulation. For example, substituting 20 at% Ti with V in Ti2AlC yielded the (Ti0.8V0.2)2AlC MAX phase, which showed 29%, 36%, and 45% increases in Vickers hardness, flexural strength, and shear strength, respectively.10 Similar mechanical enhancement was also observed in the Ti/Mo solid solution (Ti0.8Mo0.2)2AlC MAX phase, which exhibited enhanced Vickers hardness, flexural strength, and fracture toughness by 44%, 34%, and 136%, respectively, as compared to its Ti-containing 211 counterpart.11 Not only mechanical properties but also thermoconductivity,12 corrosion resistance,13–18 tribological performance19–22 and thermoelectric properties23,24 of MAX phase ceramics can be regulated. For example, Du et al. reported an enhancement in high-temperature tribological properties upon Mo incorporation in the Cr2TiAlC2 MAX phase via a modified interlayer binding and optimized tribo-oxide film.25 Further addition of V to produce a high entropy (TiVCrMo)3AlC2 revealed an increase in hardness and a strong sluggish diffusion effect of elements.26 Thus, M-site solid solution in MAX phase ceramics is an innovative structure correlated property manipulation approach and is imperative for the future design of multifunctional new MAX phases.
On the other hand, the synthesis of the MAX phase with high purity is difficult due to the formation of multiple intermediate phases. Transition metal carbides, metal alloys, and metal oxides commonly occur during the heat treatment process.27–30 Besides, the formation of hierarchical MAX phases also happens, especially when there are multiple transitional metals in the M-site with a synthetic temperature overlapping. For example, Nb4AlC3 413 impurity was detected in the Nb2AlC 211 sample by hot-press sintering at 1600 °C; comparatively, sintering at the same temperature (Ti0.45Nb0.55)5AlC4 514 impurity was detected in the Ti/Nb solid solution (Ti0.45Nb0.55)2AlC 211 sample.31 Therefore, although intriguing properties can be anticipated from exploring new MAX phases with M-site solid solution, the intrinsic thermodynamic instability caused by large atomic differences in M-site candidates and kinetic competition between multiple intermediate phases during synthesis has posed the fabrication of new MAX phases a challenge.
Herein, Ti, Nb and Hf solid solution 211 MAX phase ceramics are synthesized for the first time by a spark plasma sintering (SPS) method. The variation of morphology, structural information and electronic properties along with an increasing Hf concentration are investigated. Density functional theory (DFT) simulations are employed to study the electronic structure, bonding status with respect to different M–A and M–C bonds and corresponding elastic properties. Furthermore, the mechanical properties of M-site solid solution MAXs are investigated microscopically and macroscopically with their failure behaviors elucidated. Specifically, excellent compressive strength of 1574.77 ± 31.15 MPa and fracture toughness of 7.14 ± 0.05 MPa m1/2 are achieved, implying the superiority of M-site composition regulation in boosting the mechanical properties of MAX phase ceramics.
The density of the Ti1−xNb1−xHf2xAlC series was calculated using the Archimedes method. As summarized in Table S2, TiNbAlC reveals a density of 5.21 g cm−3, which is 98.65% of its theoretical value. With increasing Hf concentration, the density increases to 5.80 g cm−3 Ti0.9Nb0.9Hf0.2AlC and further to 6.23 g cm−3 Ti0.8Nb0.8Hf0.4AlC, which are 99.85% and 97.82% of their theoretical value, respectively.
The morphology and structure of the Ti1−xNb1−xHf2xAlC (2x = 0, 0.2 and 0.4) series were characterized using backscattered electron (BSE) detection in scanning electron microscopy (SEM). The BSE image of TiNbAlC (Fig. 2a) reveals a dense and smooth grey matrix with dark micron/submicron size particles embedded in it. At a higher magnification, energy-dispersive spectroscopy (EDS) mapping was employed to distinguish these two phases (Fig. S1). On the grey matrix, Ti, Nb, Al and C are homogeneously distributed, which is assignable to the TiNbAlC MAX phase. Enrichment of Al and O observed on the dark micron/submicron particles is attributed to the Al2Ox phase, and this assignment corroborated XRD results. With Hf incorporation, besides the grey matrix and dark particles, additional dispersion of bright particles is noticed in both Ti0.9Nb0.9Hf0.2AlC (Fig. 2b) and Ti0.8Nb0.8Hf0.4AlC (Fig. 2c). EDS mapping displays a similar homogeneous distribution of Ti, Nb, Hf, Al and C on the grey matrix and overlapping of Al and O on the dark particles for both MAXs, signifying a similar inclusion of Al2Ox inside the Ti0.9Nb0.9Hf0.2AlC and Ti0.8Nb0.8Hf0.4AlC matrices (Fig. S2 and S3, respectively). Point analysis was further conducted to determine the composition of the bright particles. As revealed in Fig. S4–S7, a sharp increase in the atomic percentage of both Hf and O is observed in the bright particles as compared to the grey matrix, validating the formation of HfOx impurity in both Ti0.9Nb0.9Hf0.2AlC and Ti0.8Nb0.8Hf0.4AlC. Fig. S8–S10 depict the surface morphology of etched specimens, where random orientation of slab-like grains is observed for all three samples with grain size varying from 10 μm to 20 μm in length.
The electronic status of each element in Ti1−xNb1−xHf2xAlC (2x = 0, 0.2 and 0.4) series was investigated using X-ray photoelectron spectroscopy (XPS). The high-resolution Ti 2p spectrum of TiNbAlC (Fig. 2d) is deconvoluted into three 2p3/2/2p1/2 pairs assigned to Ti–C (454.48 and 459.58 eV), Ti3+ (456.36 and 460.93 eV) and Ti4+ (458.72 and 464.57 eV) species, respectively.26,29 The high-resolution Nb 3d spectrum of TiNbAlC (Fig. 2e) exhibits three deconvoluted doublets, corresponding to distinct chemical states of Nb–C (203.08 eV and 205.80 eV), Nb4+ (204.36 eV and 207.08 eV), and Nb5+ (207.40 eV and 210.12 eV) species, respectively.44,45 The Al electron core level in TiNbAlC (Fig. 2f) is resolved into two distinct peaks of Al–M (72.51 eV) and Al–O (74.46 eV) bonds, respectively.26 In the C 1s spectrum, the binding energy at 282.75 eV confirms the presence of C–M species. While the other three peaks centered at 284.75, 286.53, and 288.74 eV correspond to the adventitious contaminants29 (Fig. 2g). Similar XPS spectra and deconvoluted doublets of Ti, Nb, Al and C elements in Ti0.9Nb0.9Hf0.2AlC and Ti0.8Nb0.8Hf0.4AlC are observed as compared to TiNbAlC. In the high resolution Hf 4f spectra (Fig. S11), three similar characteristic 4f5/2/4f7/2 pairs corresponding to Hf–C, Hfx+ and Hf4+ species are noted in both Ti0.9Nb0.9Hf0.2AlC and Ti0.8Nb0.8Hf0.4AlC samples, which is consistent with the coordination environment of the 4f site in the M2C slab and HfOx impurities.
Density functional theory (DFT) was employed to examine the electronic structure of the Ti1−xNb1−xHf2xAlC solid solution system. A 2 × 2 × 1 supercell was constructed to simulate the solid solution Ti1−xNb1−xHf2xAlC structure with a special quasi-random structure (SQS) in an alloy theoretical automated toolkit (ATAT). As presented in Fig. 3a, among a total of 16 4f sites, Ti takes 7 sites, Nb takes 7 sites and Hf takes 2 sites. The total and projected densities of states (DOSs) of Ti1−xNb1−xHf2xAlC in Fig. 3b present a continuous pattern with a local minimum at the Fermi level (Ef), indicating the metallic nature and good stability of the Ti1−xNb1−xHf2xAlC structure.46 The orbital-resolved DOS reveals the orbital-wise contributions of each element. In the range of −7.5–−2.5 eV, states are mainly contributed by the hybridization of 3d orbitals of transition metal elements (M = Ti, Nb and Hf) and 2p orbitals of C element. While in the range of −2.5 – Ef, states are dominantly contributed by the M 3d–Al 2p hybridization. Furthermore, the bonding status of M–C and M–A bonds is revealed through the crystal orbital Hamilton population (COHP) analysis in Fig. 3c. All paired electrons in Ti–Al, Nb–Al and Hf–Al bonds appear as bonding interactions. Comparatively, in Ti–C, Nb–C and Hf–C bonds, the paired electrons reveal a majority of bonding interactions with some negligible antibonding interactions appearing close to the Ef. A similar antibonding feature in the M–C bond near the Ef was also observed in other 211 MAX phase materials.47
The detailed atomic electronic interactions concerning respective coordination environments of Ti1−xNb1−xHf2xAlC with respect to their bond strength were further studied by Bader charge (Fig. S12) and integrated –COHP (–ICOHP) analyses. From the structural point of view, each M2C layer is interleaved between two Al layers, thus the M–C and M–A bonds are intertwined with each other. As revealed in Fig. 3d–i, the Hf incorporation induces higher electron donation towards the Al layer and a larger interlayer distance between Ti or Nb and the Al layer, which leads to weakened Ti–Al and Nb–Al bonds. For Ti(6) located within the same layer as Hf, the average bond strength for Ti(6)–Al(18, 19, 20) is integrated to be 1.82 eV bond−1. Comparatively, for Ti(3) in a different layer as Hf, the average bond strength of Ti(3)–Al(22, 23, 24) is higher at 1.95 eV bond−1. Similarly, the average bond strength of Nb(12)–Al(17, 19, 20) is 2.26 eV bond−1, which is lower than that of Nb(14)–Al(21, 22, 23) at 2.42 eV bond−1. For Hf, the average bond strength of Hf(16)–Al(17, 18, 20) (2.17 eV per bond) is stronger than that of Hf(15)–Al(17, 18, 19) (2.01 eV bond−1). On the other hand, as Hf, Ti and Nb reveal a decreasing electronegativity difference with C from Hf/C (1.25) to Ti/C (1.01) and further to Nb/C (0.95), the ionic feature of these M–C bond decreases and thus the covalent feature increases in the order of Hf–C bond to the Ti–C bond and further to the Nb–C bond. This change in ionicity of these M–C bonds can be further validated by the electron localization function (ELF) analysis, which provides localization probability information of the same-spin electrons. Fig. S13 displays the plane intercepts of all three types of Ti–C, Nb–C and Hf–C bonds. There is a highly localized electron density around the C atom. Notably, varied electron distributions are observed on the metal atoms. Compared to the almost electron-deficient region around Hf and Ti atoms, electrons with low density are found around the Nb atom, which indicates an increased covalency of the Nb–C bond. Thus, the Hf incorporation would cause varied impacts on the M–C bond strength, which is closely related to the respective distribution of paired electrons and the coordination environment of the M atom. As observed, Ti(6)–C(25, 27, 28) delivers a higher bond strength of 3.97 eV bond−1 compared to that of Ti(3)–C(25, 26, 27) at 3.88 eV bond−1. This increased Ti–C bond strength is possibly ascribed to electron delocalization on the coordinated C atoms from the adjustment of other bonding M-site atoms, as well as a higher electron accumulation at the coordinated Al atoms. An opposite bonding strength trend is observed for the covalent Nb–C bond when electron delocalization on the coordinated C atoms and electron enrichment on the coordinated Al atoms occur. A lower bond strength of 4.62 eV bond−1 is displayed for Nb(12)–C(30, 31, 32) compared to that of 4.77 eV bond−1 for Nb(14)–C(26, 27, 28). As for the Hf–C bond, the trade-off of electron distribution on C and Al gives a slightly stronger Hf(16)–C(25, 26, 27) bond as compared to the Hf(15)–C(29, 30, 32) bond. As summarized in Table S3, the average bond strength of M–C and M–Al in the cell was calculated to be 4.19 eV bond−1 and 2.12 eV bond−1, respectively, with the Ti–C/Ti–Al and Hf–C/Hf–Al bonds locating at the lower strength end and the Nb–C/Nb–Al bond located at the higher strength end. The bond strength ratio of R = M–C/M–A increases in the order of RNb (1.958) < RHf (1.962) < RTi (2.01).
Given the anisotropic bonding properties derived from the solid solution Ti1−xNb1−xHf2xAlC structure, intriguing mechanical properties are expected. Through theoretical calculations, the elastic properties of the Ti1−xNb1−xHf2xAlC structure were first calculated with the second-order elastic constants cij, and the values are summarized in Table S4. Satisfying the Born stability criteria for hexagonal crystals: c44 > 0, c11 > |c12|, (c11 + c12)c33 > 2c132,48 the Ti1−xNb1−xHf2xAlC structure reveals good mechanical stability upon applied strain. Similar to the previously reported MAX phase, the Ti1−xNb1−xHf2xAlC structure possesses anisotropic elastic constants.46 Specifically, linear deformation resistance along crystallographic directions is characterized by elastic constants: c11 for the a-axis and c33 for the c-axis.49 A higher value of c11 than c33 is displayed by Ti1−xNb1−xHf2xAlC, indicating a greater resistance along the a axis. Compared to the 211 phase with a single Ti, Nb, or Hf M-site element, both the c11 and c33 values are higher in the Ti1−xNb1−xHf2xAlC structure, suggesting enhanced elastic resistance in both directions.46 c44, a key hardness parameter and shear resistance indicator along the basal plane,49 is smaller in Ti0.875Nb0.875Hf0.25AlC than in Ti2AlC and Nb2AlC, suggesting lower hardness and shear strength in the solid solution Ti1−xNb1−xHf2xAlC. The corresponding bulk modulus (B), shear modulus (G) and Young's modulus (E) of Ti1−xNb1−xHf2xAlC were approximated by Voigt–Reuss–Hill average, which yielded 165.55, 122.76 and 295.28 GPa for these moduli, respectively (Table S5). Poisson's ratio (ν) was calculated to be 0.20. The brittleness of the structure was reviewed by the B/G ratio, where B/G >1.75 indicates ductility and lower values suggest brittleness.50 Although the obtained B/G ratio of 1.35 of Ti1−xNb1−xHf2xAlC does not fall within the ductile region, it is still higher than those of Hf2AlC,51 Ti3AlC2 (ref. 52) and ZrB12-based high-temperature ceramics.53
The mechanical properties of the as-prepared Ti1−xNb1−xHf2xAlC (2x = 0, 0.2 and 0.4) sample series were further evaluated experimentally. The nanoindentation test was utilized to determine the micromechanical properties with high spatial resolution and minimal interference from the impurity phase. As the indentation size of approximately 0.4 μm2 is within one to several grain diameters in all three samples, the mechanical properties measured by the nanoindentation test intrinsically reflect the mechanical properties of these 211 grains. Fig. 4a–c depict the displacement–load curves of the three samples with similar loading, holding, and unloading stages, indicating similar elastoplastic deformation, creep and recovery response to a consistent loading. Specifically, during the loading stage, the displacement–load curves of all three samples almost overlap with each other with low slopes, suggesting low elastic moduli. In the holding stage, displacement increases with time under constant load, indicating creep deformation. Steep initial unloading segments imply high contact stiffness. After unloading, only elastic recovery occurs, leaving displacement changes from both plastic deformation and creep processes. According to the Oliver–Pharr model, the reduced elastic modulus (E) and nanoindentation hardness (H) for the three 211 MAX phase ceramics were calculated (Fig. 4d).54 The E and H values for TiNbAlC are 214.28 and 13.40 GPa, respectively. Compared to TiNbAlC, the solid solution Ti1−xNb1−xHf2xAlC (2x = 0.2 and 0.4) samples show slightly reduced E and H.
Macroscopically, the hardness, compressive stress and fracture toughness of the 211 series were analyzed as well. The Vickers hardness (Fig. S14a) of TiNbAlC, Ti0.9Nb0.9Hf0.2AlC, and Ti0.8Nb0.8Hf0.4AlC were measured to be 5.68, 6.09, and 6.69 GPa, respectively. These values are comparable to the previously reported Ti and Nb containing single and double transition metal 211 MAX phases, i.e. Ti2AlC at 4.2–5.7 GPa,55 Nb2AlC at 6.1 GPa (ref. 30) and TiNbAlC at 5.8 GPa.56 Unlike the conventional ceramics, such as Si3N4 (ref. 57) and SiC58 with crack propagation upon indentation, the typical indentation morphology of TiNbAlC (Fig. S14b) reveal high toughness without observable cracks, which is similar to MAX phases. As the size of the indention is around 70 μm, the contribution of the grain boundary and oxide impurities to the measured hardness cannot be excluded, which may possibly be the reason for the varied trend in hardness between microscopic and macroscopic scales.
The compressive stress–strain curves of all three MAX phases depict brittle behaviour (Fig. S15). Except for the toe region, the curves of the three 211 MAXs display linear correlation between stress and strain, featuring elastic deformation upon loading. As summarized in Fig. 4e, the compressive strengths of TiNbAlC, Ti0.9Nb0.9Hf0.2AlC, and Ti0.8Nb0.8Hf0.4AlC are 1360.67 ± 19.01, 1574.77 ± 31.15, and 1493.67 ± 35.56 MPa, respectively. A similar trend in the fracture toughness of the three MAXs are observed as well (Fig. 4f). The fracture toughness values of TiNbAlC, Ti0.9Nb0.9Hf0.2AlC, and Ti0.8Nb0.8Hf0.4AlC were measured to be 5.52 ± 0.35, 7.14 ± 0.05 and 6.62 ± 0.22 MPa m1/2, respectively. These values are much higher than the current reported MAX phase ceramics, including 211, 312 and 413 phases and even the common engineering ceramic Al2O3 (Table S6). Except for the intrinsic mechanical property differences among these 211 grains, the presence of Al2O3 and HfO2 impurities would also impact their compressive strength and fracture toughness by revising crack propagation.59 For example, Chen et al. revealed improved toughness and compressive strength in Ti3AlC2 with 5% Al2O3 addition.60 Zhu et al. discovered that the compressive strength of Ti2AlC with 12% Al2O3 increased by 248%.61 Tzenov et al. found that Ti3AlC2 containing 4% Al2O3 exhibited higher mechanical properties compared to monolithic Ti3AlC2.62 As for our case, the oxide impurity increases from the 6.9 wt% in TiNbAlC to 7.8 wt% in Ti0.9Nb0.9Hf0.2AlC and further to 14.5 wt% in Ti0.8Nb0.8Hf0.4AlC, where the impact from the high content second phase cannot be ignored. Besides the strengthening effect by the oxide second phase, the existence of possible defects from insufficient density of 211 phase would, on the other hand, compromise the strength and toughness. Thus, the enhanced compressive strengths of the solid solution Ti1−xNb1−xHf2xAlC compared to TiNbAlC is likely due to the synergistic effect of the varied mechanical properties of M-site solid solution and the presence of high strength impurity phases and defects.
To better understand the mechanical behaviour in relation to the structure, the surface morphology of the fractured particles from the three MAXs was examined. Fig. S16 presents SEM images of fractured surfaces from the compressive test. As observed, the lamellar structures in all three MAXs are randomly oriented, which is beneficial for a delayed crack propagation and enhanced strength.63 These lamellar structures have a dimension of 2–5 μm in thickness and 10–20 μm in length, which are similar to their original structures. Two distinct failure modes are identified in the three MAXs, i.e., (i) the delamination of lamellar structure along the a–b plane and (ii) the transgranular fracture along the c axis. The delamination along the ab plane is facilitated by the weak A-layer/MX-layer interfacial bonding under the shear force. The transgranular fracture along the c axis features the breakage of both strong M–C bonds and weak M–A bonds. Furthermore, TiNbAlC displays a sharp cross-section surface on the fractured grains, indicating simultaneous breakage of the M–C and M–A bonds within the grain (Fig. 5a). Comparatively, both Ti0.9Nb0.9Hf0.2AlC and Ti0.8Nb0.8Hf0.4AlC depict irregular ruptured surfaces cross-sectionally along the layers (Fig. 5b and c). This zig-zag propagation of the cracks from layer to layer in Ti0.9Nb0.9Hf0.2AlC and Ti0.8Nb0.8Hf0.4AlC indicates a higher energy absorption, which is consistent with their higher compressive strengths. Similar two distinct failure surfaces as characteristics of two failure modes are observed after the fracture toughness test (Fig S17). Submicron-sized pores are observed in all three MAXs. Notably, under the BSE detector (Fig. 5d–f and S18–S21), the distribution of Al2Ox and HfOx in between grains is evident, which may play a role in impeding the crack propagation via deflection or bridging mechanisms.64
The data supporting this article have been included as part of the SI. See DOI: https://doi.org/10.1039/d5ta05278e.
This journal is © The Royal Society of Chemistry 2025 |