Abstract
High-entropy materials have been proposed for applications in nuclear systems recently due to their outstanding properties in extreme environments. Chemical complexity in these materials plays an important role in irradiation tolerance since it significantly affects energy dissipation and defect behaviors under particle bombardment. Indeed, better resistance to irradiation-induced amorphization was observed in the high-entropy MAX (HE-MAX) phase (Ti, M)2SnC (Mâ=âV, Nb, Zr, Hf). However, in this work, we report an opposite trend in another series of HE-MAX phases (Ti, M)2AlC (Mâ=âNb, Ta, V, Zr). It is demonstrated that the amorphization resistance is sequentially reduced as the number of components increases from single-component Ti2AlC to (TiNbTa)2AlC and (TiNbTaVZr)2AlC. These phenomena are verified through AIMD simulations and interpreted by analyzing the underlying properties combining lattice distortion and bonding characteristics through the first-principle calculation. By developing a machine-learning (ML) model, we can directly predict lattice distortion to screen HE-MAX phases with excellent resistance to irradiation-induced amorphization. We highlight that the elemental species plays a more crucial role in the irradiation tolerance of these MAX phases than the number of constituent elements. Knowledge gained from this study will enable an improved understanding of the irradiation tolerance in HE-MAX phases and other multi-elemental ceramics.
Similar content being viewed by others
Introduction
Mn+1AXn phases (or simplified as MAX phases) are a class of layered ternary carbides or nitrides1, where M denotes an early transition metallic element, A is a main group element and X is carbon and/or nitrogen, nâ=â1, 2, or 3. MAX phases are hexagonal (space group P63/mmc, No.194) and nanolaminated with Mn+1Xn layers interleaved with A atom layers, typically possessing weak metallic bonds between M and A atoms and strong covalent bonds within the Mn+1Xn blocks1,2,3,4. Due to such specific structures and mixing rules of bonding characteristics, they exhibit a unique set of properties combining both ionic-covalent and metallic materials1,2,4,5,6. These remarkable properties include high elastic stiffness and strength, high electrical and thermal conductivity, high thermal stability, exceptional resistance to corrosion and oxidation, easy machinability, and excellent damage tolerance2,4,5,6,7,8,9,10. Moreover, some species of MAX phases are irradiation tolerant8,11,12,13, and are considered promising candidate coating materials for accident tolerance fuel (ATF) in advanced nuclear systems14,15,16.
Ion-beam irradiation is a common approach to studying the irradiation effects of materials since it can generate collision cascade and analogous radiation damage as neutron irradiation17,18,19. The accumulation of irradiation-induced defects may also trigger structural evolution or amorphization in ceramics like complex oxides20,21,22,23,24 and MAX phases8,25,26, significantly influencing the long-term durability of materials due to the property mutations and performance deterioration. Therefore, there is a great impetus to investigate the responses of MAX phases to irradiation and elucidate the mechanism for developing high amorphization-resistance materials.
The introduction of chemical complexity or the so-called âhigh entropyâ is a newly developed strategy for property tailoring27,28,29,30, and can help design enhanced irradiation-tolerance materials31,32. Previous studies have demonstrated that chemical complexity plays an important role in the irradiation tolerance of both single-phase concentrated solid solution alloys (CSAs)31,32,33 and high entropy ceramics (HECs)34. It is suggested that the effects of lattice distortion, sluggish diffusion, and chemical disorder impressively affect the generation, migration, and evolution of irradiation-induced defects. These effects lead to slower energy dissipation, enhancing defect recombination and consequently resulting in slower damage accumulation under irradiation33. A reduced antisite defect formation energy has been manifested in the high-entropy MAX (HE-MAX) phase (Ti, M)2SnC (Mâ=âV, Nb, Zr, Hf)26 such that the chemically complex (TiVNbZrHf)2SnC is prone to accommodate more point defects and maintain the crystalline lattice reflected by its better irradiation-induced amorphization resistance.
However, in this work, we report an opposite phenomenon in another series of HE-MAX phases, (Ti, M)2AlC (Mâ=âNb, Ta, V, Zr). The schematic atomic models of (TiNbTa)2AlC and (TiNbTaVZr)2AlC are shown in Fig. 1. The resistance to irradiation-induced amorphization is sequentially reduced as the number of components increases from single-component Ti2AlC to (TiNbTa)2AlC and (TiNbTaVZr)2AlC. The chemical complexity seems to play a negative role in the resistance to irradiation-induced amorphization, contrary to our previous perceptions26. The ab initio molecular dynamics (AIMD) results verified this trend and elucidated that the atomic displacement of (TiNbTaVZr)2AlC invariably remains larger than that of Ti2AlC with the increasing simulated damage dose. Through the first-principle calculation, we found that lower antisite defect formation energy is not found in this series of Al-based HE-MAX phases. We further explained the experimental phenomena through the analysis of the underlying properties combining the lattice distortion and bonding characteristics. The elemental species plays a more crucial role in the irradiation tolerance of MAX phases than the number of constituent elements. This effect is attributed to the stronger interatomic bonding characteristics in ceramics substantially associated with the elemental components. Additionally, we developed a machine learning model to predict the lattice distortion of the γ-phases by the properties and number of constituent elements.
Results and discussions
Irradiation-induced phase transformation and amorphization resistance in (Ti, M)2AlC (Mâ=âNb, Ta, V, Zr)
Selected area electron diffraction (SAED) is an efficient technique to determine the lattice or phase structure, thereby we obtained the SAED patterns of the samples during ion irradiation to record the entire process of phase transformation and amorphization. Figure 2 shows the series of in situ SAED patterns of Ti2AlC, (TiNbTa)2AlC, and (TiNbTaVZr)2AlC along the [\(11\bar{2}0\)] zone axis during the Kr2+ irradiation at room temperature (RT). All SAED patterns for each sample were obtained from the central part of one grain such that the effect of grain boundary could be ignored.
As illustrated in Fig. 2aâg for Ti2AlC, Fig. 2hâm for (TiNbTa)2AlC, and Fig. 2oât for (TiNbTaVZr)2SnC, the diffraction spots of (000l) (lââ â6n), and (\(\bar{1}10l\)) (lâ=â6n, 6nâ±â1) in the \(\left\{000l,\,\bar{1}10l\right\}\) reflections gradually attenuate and disappear with the increased dose owing to the increase of the lattice symmetry. Further irradiation transforms the intermediate γ-phase into a twinned face-centered cubic (fcc) structure, as the two spots designated by yellow arrows in Fig. 2e represent \((\bar{1}1\bar{1})\) and \((00\bar{2})\) diffraction spots of the nano-twined fcc structure. This similar process of structural transformation in Ti2AlC has been observed in previous work35. No amorphization was observed in Ti2AlC even at the irradiation dose up to 61.6âdpa. Nevertheless, both (TiNbTa)2AlC and (TiNbTaVZr)2AlC show evidence of an amorphous component with a diffraction ring in the SAED patterns, as indicated by the white arrows in Fig. 2j, q at a dose of 0.48 and 0.15âdpa, respectively. This indicates that the intermediate γ-phase starts to convert into an amorphous phase as the irradiation dose increases. Moreover, the five-component (TiNbTaVZr)2AlC exhibits a worse irradiation resistance than (TiNbTa)2AlC as its critical irradiation dose (~0.6âdpa) of complete amorphization (when diffraction spots disappear) is far less than that in (TiNbTa)2AlC (about 21.6âdpa), as shown in Fig. 3n, u. Therefore, a decrease of irradiation-induced amorphization resistance with the elemental components increase is demonstrated in this series of MAX phases, indicating that the chemical complexity plays a negative role in the irradiation resistance. This compositional trend is opposite to our previous perceptions in the HEAs33,36 and HE-MAX phases (Ti, M)2SnC (Mâ=âV, Nb, Zr, Hf)26. It should be noted that the results were checked and confirmed in various grains in each material to show a similar trend for avoiding occasionality.
Figure 3 shows the high-resolution TEM (HRTEM) images of the (TiNbTa)2AlC and (TiNbTaVZr)2AlC along the \(\left[11\bar{2}0\right]\) direction, before and after 800âkeV Kr2+ irradiation at a series of doses. These atomic scale TEM images demonstrate the structural evolution follows a multi-stage α-γ-fcc-amorphous sequence (we defined the pristine hexagonal phase as α-phase), which is in good agreement with the evolution of SAED patterns in Fig. 2.
Compared to our previous work on the irradiation-induced phase transformation in (Ti, M)2SnC (Mâ=âV, Nb, Zr, Hf)37, the fcc-phase appears between the stages of γ-phase and amorphous phase under the 800âkeV Kr2+ irradiation, as shown in Fig. 3d, h. We performed a phase-contrast simulation to further confirm the phase transition of the intermediate γ-phase to fcc-phase as proposed above. It is revealed that the simulation result is consistent with the experimental HRTEM micrograph (as shown in Supplementary Fig. 1).
Combining the results of SAED, HRTEM, and phase-contrast simulation, we thereby confirmed the multi-stage phase transformation of hex-γ-fcc in (TiNbTa)2AlC and (TiNbTaVZr)2AlC. Continuous ion bombardment will further induce a lattice disorder when the defects and chemical disorder accumulate to a critical value, triggering the crystalline phase to transform into an amorphous phase eventually.
AIMD simulations of the lattice evolution during point defect accumulation
Furthermore, ab initio molecular dynamics (AIMD) simulations were performed in Ti2AlC and (TiNbTaVZr)2AlC to probe the irradiation-induced crystalline to amorphous phase transition at the atomic scale. The radiation process is simulated by continuously introducing Frenkel pairs38,39,40,41,42. The irradiation dose for AIMD simulations is equal to the ratio of the number of Frenkel pairs introduced to the number of atoms in the simulated crystal cell. The pair correlation function of the system at an interval of 0.1âdpa was obtained to judge whether the system was fully amorphous. It is demonstrated that long-range order is still retained in Ti2AlC at 0.6âdpa, as shown in Fig. 4a. In contrast, (TiNbTaVZr)2AlC completely loses its long-range order at a dose of 0.6âdpa, indicating full amorphization (Fig. 4b). Figure 4dâg shows the atomic projections of Ti2AlC and (TiNbTaVZr)2AlC at 0âdpa and 0.6âdpa, respectively. Likewise, all the crystalline planes and rows of (TiNbTaVZr)2AlC are completely lost, indicating full amorphization, while the lattice of Ti2AlC remains stable. In addition, there are numerous cation antisite defects as well as C interstitials in the otherwise empty cation octahedral interstitial sites in Ti2AlC at 0.6âdpa, suggesting that an ordered-to-disordered phase transition process has occurred.
To quantify the fully amorphous dose in AIMD simulations, the crystallinity was defined as the percentage of hexagonal close-packed (HCP) structures in the cation sublattice identified by the Polyhedral template matching method43. As shown in Fig. 4c, with the continuous introduction of Frenkel pairs, the cation sublattice in Ti2AlC remains crystalline with a crystallinity of around 0.55. However, for (TiNbTaVZr)2AlC, it gradually loses its crystallinity, and its crystallinity decreases to 0 at about 0.56âdpa, indicating complete amorphization. Therefore, the trend of the amorphization resistance obtained from the AIMD simulation is consistent with the ion irradiation experimental results.
DFT calculations of the antisite defects formation energy
The formation energy of antisite defect was proposed to be an important indication of irradiation tolerance in complex ceramics like pyrochlores20,21, since the formation of cation antisite defects provides a recovery mechanism to accommodate lattice point defects induced by the displacement cascades and maintains its lattice structure. That means complex ceramics with a lower formation energy of antisite defect are susceptive to irradiation-induced order-to-disorder (O-D) transition and generally possess better resistance to irradiation-induced amorphization. However, Uberuaga et al. found an opposite case in spinels that have cation vacancies instead of anion vacancies like in pyrochlores23, and demonstrated that the correlation between cation disordering and amorphization resistance depends on the lattice structure (vacancies exist on the cation or anion sublattice). Similar to pyrochlore and the defective fluorite structures, ion irradiation drives the transformation of MAX phases into a structure being close-packed in the cation sublattice but with C/N vacancies in the anion sublattice25,35,37,44,45,46, and generally, the M-A antisite defect manifests almost the lowest formation energy among all point defect types8,46,47,48,49,50. Therefore, this criterion of correlation between DFT calculated M-A antisite defect formation energy (\({E}_{{form}}^{\,M-A\,{antisite}}\)) and amorphization resistance, is consistent with most radiation experiments of the common MAX phases (lower \({E}_{{form}}^{\,M-A\,{antisite}}\) indicates higher amorphization resistance)8,47,48. In our recent work, we also suggested that the enhanced irradiation resistance as the chemical complexity increases is attributed to the reduced formation energy of M-A antisite defect (\({E}_{{form}}^{\,M-A\,{antisite}}\)) in (TiVNbZrHf)2SnC, compared to corresponding single-component M2SnC (Mâ=âTi, V, Nb, Zr, Hf)37.
Nevertheless, this criterion fails in the case of Cr2AlC. For instance, Cr2AlC possesses almost the lowest \({E}_{{form}}^{M-A\,{antisite}}\) in the traditional MAX phases family47,51, yet its amorphization resistance under ion irradiation is worse than V2AlC and Ti2AlC35,52. In this work, this criterion of antisite defect formation energy fails to explain this experimental phenomenon, as shown in Fig. 5. Unlike our previous work that exhibits a decreased M-Sn (Mâ=âTi, V, Nb, Zr, Hf) antisite defect formation energy after the introduction of chemical complexity37, the \({E}_{{form}}^{M-{Al\; antisite}}\) (Mâ=âTi, Nb, Ta, V, Zr) in M2AlC, (TiNbTa)2AlC, (TiNbTaVZr)2AlC shows an irregular trend with the introduction of chemical complexity. For instance, compared to the corresponding single-component counterpart, the formation energy of Ti/Zr-Al antisite defect decreases in (TiNbTa)2AlC or (TiNbTaVZr)2AlC, while Ta/V-Al antisite defect increases instead. Therefore, the reduction of \({E}_{{form}}^{M-A\,{antisite}}\) with the introduction of chemical complexity in M2SnC (Mâ=âTi, V, Nb, Zr, Hf) is not a general rule in ME/HE-MAX phases.
It is worth mentioning that, all the \({E}_{{form}}^{M-A\,{antisite}}\) calculations are based on the pristine hexagonal phases, indicating that the criterion provides only an indication at the early stage under irradiation. More attention should be paid to the intermediate γ-phase or fcc-phase which plays a more important role in the resistance to irradiation-induced amorphization.
Structural properties of the intermediate γ-phase and fcc-phase
It has been demonstrated that the γ-phase is a necessary stage when the MAX phases undergo the complex phase transformation process during ion irradiation8,25,35,44,45,46. Additionally, an γ-to-fcc phase transformation was also observed in these three materials, as shown in Fig. 2. The properties of the γ-phase or fcc-phase may play an important role in the resistance of the MAX phases to amorphization. Therefore, the properties of lattice distortion and electronic structure of the intermediate γ-phase and fcc-phase are focused in this work since they greatly influence the accumulation of defects and lattice disorder.
The ab initio calculation was based on the intermediate γ-phase and fcc-phase with chemically disordered structures. A DFT relaxation together with a charge density topology analysis was applied to obtain the degree of lattice distortion and bonding characteristics in the solid-solution γ-(Ti, M)2AlC (Mâ=âNb, Ta, V, Zr) phases. Figure 6a illustrates one of the DFT-relaxed supercell models of γ-(TiNbTaVZr)2AlC phase. The dense yellow spheres indicate the charge density in the bond critical points (BCPs, a BCP is the first order saddle points of electron density lies on a bond path which defines the bond between two atoms53) in the lattice. Figure 6b, c shows the statistical results of the lattice distortion \(\Delta d\) and the total average charge density among all BCPs of M/Al-C (Mâ=âNb, Ta, V, Zr) bonds \(\bar{\rho }\left({r}_{b}\right)\) over three 4âÃâ4âÃâ1 SQS supercells of the pristine phase, γ-phase and corresponding converted fcc-phase in Ti2AlC, (TiNbTa)2AlC, and (TiNbTaVZr)2AlC, respectively. The lattice distortion \(\Delta d\) is described by the atomic deviation from the ideal regular lattice sites as the following calculation formula54
where (\({x}_{i},\,{y}_{i},\,{z}_{i}\)) and (\({x}_{i}^{{\prime} }\), \({y}_{i}^{{\prime} }\), \({z}_{i}^{{\prime} }\)) are the coordinates of the unrelaxed and relaxed positions of atom i, respectively. It is illustrated that the (TiNbTaVZr)2AlC possesses the largest atomic displacement of approximately 0.170 \(\mathring{\rm A}\) and 0.201 \(\mathring{\rm A}\) in the γ-phase and fcc-phase respectively, followed by (TiNbTa)2AlC and then Ti2AlC. A larger atomic displacement means that the lattice is more unstable and more inclined to undergo a lattice disorder induced by the continuous ion bombardment. Therefore, the (TiNbTaVZr)2AlC with the largest atomic displacement in the γ-phase exhibits the lowest resistance to amorphization, while γ-Ti2AlC instead shows the best irradiation-induced amorphization resistance due to its smallest lattice distortion.
Meanwhile, as shown in Fig. 6c, a similar trend of the total average charge density over all the BCPs of M/Al-C bonds \(\bar{\rho }\left({r}_{b}\right)\) is exhibited in both the original γ-phase and fcc phase. \(\bar{\rho }\left({r}_{b}\right)\) are about 0.542 \(e/{\mathring{\rm A} }^{3}\) and 0.560 \(e/{\mathring{\rm A} }^{3}\) in the γ-phase and fcc-phase of (TiNbTa)2AlC respectively, 0.547 \(e/{\mathring{\rm A} }^{3}\) and 0.568 \(e/{\mathring{\rm A} }^{3}\) in the γ-phase and fcc-phase of (TiNbTaVZr)2AlC respectively. These values are larger than that in the intermediate phases of Ti2AlC (only about 0.501 \(e/{\mathring{\rm A} }^{3}\) and 0.513 \(e/{\mathring{\rm A} }^{3}\), respectively), which indicates stronger interatomic bonding characteristics48,53 in the intermediate phase of (TiNbTa)2AlC and (TiNbTaVZr)2AlC. Figure 6d shows the corresponding BCP charge density at each M-C (Mâ=âTi, Nb, Ta, V, Zr) bonds \(\rho \left({r}_{b}\right)\) in the γ-phases of Ti2AlC, (TiNbTa)2AlC and (TiNbTaVZr)2AlC. It is illustrated that the involvement of Nb-C, Ta-C, and V-C bonds in (TiNbTa)2AlC and (TiNbTaVZr)2AlC enhances the level of charge density at the BCPs which indicates a stronger interatomic bonding characteristics53 in the intermediate γ-phase (while in fcc-phase the scenario is similar and not shown here). As proposed by Ogata et al.55, if the crystalline system possesses a stronger covalent or directionally bonding characteristic, it is expected to be more frustrated and less accommodating, exhibited by its less tolerance under irradiation. Therefore, the sequentially reduced resistance to irradiation-induced amorphization in Ti2AlC, (TiNbTa)2AlC, and (TiNbTaVZr)2AlC is interpreted as the analysis combining the lattice distortion and bonding characteristics. In addition, the pristine phase C-M bonds of (TiNbTa)2AlC is stronger than that of (TiNbTaVZr)2AlC, further indicating that it is not sufficient to analyze the amorphization resistance of the MAX phase based on the pristine phase properties alone.
Furthermore, the atomic displacements or lattice distortion of Ti2AlC and (TiNbTaVZr)2AlC are counted as a function of the simulated dose via the AIMD simulations. As shown in Fig. 7, the atomic displacements of (TiNbTaVZr)2AlC are always larger than that of Ti2AlC, in agreement with the DFT calculated results of γ-phases and fcc-phases (Fig. 6b). The atomic displacements of (TiNbTaVZr)2AlC increase more rapidly with increasing doses than Ti2AlC. When atomic displacements accumulate to a certain extent, it may be difficult for (TiNbTaVZr)2AlC to maintain crystalline structure and subsequently become amorphized. Note that the magnitude of the lattice distortion obtained from AIMD simulations is larger than that from DFT calculations. This may be due to the thermal vibrations and loss of defects recombination in the AIMD simulations.
All these results above manifest that the structural properties, especially the atomic displacement or lattice distortion during the phase transition (e.g., the intermediate γ-phase or fcc-phase) play an important role in the amorphization resistance of these materials. As shown in Fig. 8, the critical amorphization dose (damage dose for full amorphization) and corresponding lattice distortion of the γ-phase illustrate a negative correlation, for both (Ti, M)2SnC (Mâ=âV, Nb, Zr, Hf) and (Ti, M)2AlC (Mâ=âNb, Ta, V, Zr) HE-MAX phases. i.e., the material with a lower lattice distortion of the intermediate γ-phase has a higher amorphization resistance to irradiation. Furthermore, when comparing these two series of HE-MAX phases, it is shown that their lattice distortion strongly depends on the species, instead of the number of constituent elements of MAX phases. Therefore, we can infer that the strategy of âhigh entropyâ does not always work in the MAX phases family. It plays a subdominant role compared to the species of the elements in these materials. This may arise from the stronger interatomic bonding characteristics in ceramics substantially associated with the elemental components.
Prediction of γ-phase lattice distortion by machine learning model
The HE-MAX phases with excellent resistance to irradiation amorphization can be screened by calculating the γ-phase lattice distortion. However, considering the vast composition space of HE-MAX phases, using first-principle methods to calculate lattice distortion is highly time-consuming. Machine learning has shown powerful potential to address complex problems in material science56. By learning from high-dimensional input data (descriptors), machine learning methods can directly predict lattice distortion in HE-MAX phases, offering a more efficient alternative.
To train machine learning model, the training datasets were first collected from 60 HE-MAX phases with Al at the A site, C at the X site, and all binary, ternary, quaternary, and quinary combinations of Ti, V, Nb, Zr, Hf, Ta. In the collection of lattice distortions, three supercells with 128 atoms in different atomic distribution states were used to ensure statistical validity.
Subsequently, after careful investigation of the contributions to the lattice distortion of the solid solution system, 11 descriptors were chosen for the machine learning model, as summarized in Table 1. These descriptors can be calculated from the properties and numbers of the constituent elements of the HE-MAX phases, without any other additional calculations and experiments. To reduce possible overfitting introduced by the strongly correlated features and increase fitting efficiency, we removed those features with Pearson coefficients higher than 0.95. The Pearson coefficients between descriptors d3 and d7, d4 and d8, as well as d10 and d11 are greater than 0.95. Thus, the descriptors d7, d8, d11 are excluded in the following training procedure.
The parity plots comparing the calculated lattice distortion with the ML prediction are shown in Fig. 9. Eighty percent of the dataset was used for training with the remaining for testing the model performance. It can be seen from the parity plots that the machine learning model can reasonably predict γ-phase lattice distortion over the entire dataset. Visually, it can be seen that the model prediction performance is comparable for the training and test sets, indicating the robustness of the model. The machine learning model exhibits great predictive power by simply considering the constituent elementsâ basic properties. This machine learning model can accelerate the discovery of HE-MAX phases with excellent resistance to irradiation amorphization within the vast phase space.
In summary, we have combined in situ ion irradiation with TEM analysis, first-principle calculations, and AIMD simulation techniques to probe the phase transformation process and amorphization resistance of the HE-MAX phases (Ti, M)2AlC (Mâ=âNb, Ta, V, Zr). In contrast to the previous view that the introduction of chemical complexity or âhigh entropyâ can enhance the irradiation tolerance of materials, we found that the amorphization resistance is sequentially reduced as the number of constituent elements increases from single-component Ti2AlC to (TiNbTaVZr)2AlC. This trend is verified through the AIMD simulations via point defect accumulation. Through the first-principle calculation, we found that lower antisite defect formation energy is not represented in this series of HE-MAX phases, and we further explained these experimental phenomena through the analysis of the underlying structural properties combining the lattice distortion and bonding characteristics. The poor amorphization resistance of (TiNbTaVZr)2AlC arises from its large lattice distortion and strong bond covalency of the intermediate phases during irradiation. Additionally, a negative correlation between critical amorphization dose and lattice distortion of the intermediate phase was built for both series of HE-MAX phases. Using machine learning methods, we can predict the γ-phase lattice distortion in the HE-MAX phases. We highlight the elemental species, instead of the number that plays a more crucial role in the irradiation tolerance of these MAX phases. These results overturn the conventional perception and provide a new viewpoint for the design of amorphization-resistant and radiation-resistant materials in the MAX-phases family, and related multi-elemental ceramics systems.
Methods
Materials preparation
The bulk samples for irradiation in this work were synthesized at the Ningbo Institute of Materials Technology and Engineering (NIMTE), Ningbo, China. Specifically, Ti2AlC samples were prepared by the in-situ hot pressing/solid-liquid reaction process, and corresponding elemental powders were mixed in stoichiometric proportions, pressed in a graphite die, and subsequently hot-pressed in a flowing Ar atmosphere. More details about the synthesis process have been published elsewhere57,58,59. The (TiNbTa)2AlC and (TiNbTaVZr)2AlC powders were synthesized via the molt salt method and then the as-prepared powders of (TiNbTa)2AlC or (TiNbTaVZr)2AlC were filled in a graphite die for hot-press sintering to prepare its bulk sample, more details in terms of the preparation process and characterization of element composition can be found in ref. 60
Subsequently, the as-prepared bulk samples above were cut into square specimens and then mechanically thinned to ~20âμm with physical polish. The polished thin foil (with glue around) was then glued to a \({\rm{\phi }}\)3 copper ring, followed by an argon ion milling process from 4.5âkeV down to about 2âkeV (corresponding milling angle set from 4° to 2°) utilizing a Gatan PIPS 691 ion miller, thus preparing the sample with a thin wedge for in-situ irradiation and transmission electron microscope (TEM) characterization.
Ion irradiation
The in situ ion irradiation experiments in this work were performed under a 400âkV ion implanter coupling with an FEI Tecnai F30 (with a field emission gun operating at 300âkV) transmission electron microscope at Xiamen Multiple Ion Beam In-situ TEM Analysis Facility, Xiamen University61. The ion bombardment was performed with an 800âkeV Kr2+ beam at room temperature (RT). During irradiation, the selected area electron diffraction (SAED) images were recorded to observe the near real-time phase states and transformations.
The radiation damage level (displacement per atom, dpa) and the penetration of Kr ions along the ionsâ incident direction were calculated using the SRIM-2008 program (as shown in Fig. 10). The quick Kinchin-Pease mode was adopted62, and the threshold displacement energies for each element were set as 25â28âeV. The Kr2+ peak range exceeds the thickness of the TEM sample foil such that the effect of the penetrated Kr ions can be avoided. The average dpa values over the thickness (assumed to be 100ânm as in previous work63,64,65) of the TEM foils were used to estimate the damage level.
Characterization techniques
TEM observations were obtained in a 200âkV Tecnai F20 transmission electron microscope (FEI, Hillsboro, OR) with a point resolution of 0.24ânm and a line resolution of 0.102ânm at the Electron Microscopy Laboratory of Peking University. The phase contrast simulation was performed via the QSTEM program66.
Theoretical calculations
First-principles calculations based on density functional theory (DFT) were conducted in this work applying the Vienna Ab-initio Simulation Package (VASP)67. The projector augmented-wave (PAW) method68 and the generalized gradient approximation (GGA) by Perdew, Burke, and Ernzerhof (PBE)69 were employed for the electron-ion interactions and exchange-correlation function, respectively. The disordered supercells were constructed using the alloy theoretic automated toolkit (ATAT)70 based on the special quasi-random structure (SQS) method71. The gamma-centered MonkhorstâPack72 k-point of 4âÃâ4âÃâ2 is used to sample the Brillouin zone. The total energy and forces converged to better than 10â6âeV and 0.01âeV/à with a plane wave cutoff of 450âeV.
For each material of (Ti, M)2AlC (Mâ=âNb, Ta, V, Zr), three samples of 4âÃâ4âÃâ1 SQS supercells each containing 128 atoms with different atomic disordered configurations were constructed and relaxed to perform a charge density topology analysis based on the Quantum Theory of Atoms In Molecules (QTAIM)53, where the bond critical points (BCPs) were implemented by the Critic2 program73.
The antisite defect formation energy \({E}\,_{{form}}^{{antisite}}\) is defined as the following formula:
where \({E}_{T}\left({\rm{antisite}}\right)\) and \({E}_{T}\left({\rm{perfect}}\right)\) are the total energy of a supercell with one antisite defect and a perfect supercell, respectively. For (TiNbTa)2AlC and (TiNbTaVZr)2AlC, the \({E}\,_{{form}}^{{antisite}}\) of M-Al antisite defect (Mâ=âTi, Nb, Ta, V, Zr) is statistically derived from DFT calculations over 20 configurations to sample the variation of the local atomic environment.
Ab initio molecular dynamics (AIMD) simulations are also performed on the VASP code67. Starting from a perfect 200-atom (5âÃâ5âÃâ1) cell, Frenkel pairs are continuously introduced by randomly selecting any atom displaced in any direction by a distance of 5âà . Such distance prevents defects from directly recombination. Meanwhile, the displaced atoms are farther away from other atoms than 1âà , preventing close interactions. Between the two displacement events, the cell relaxes at a fixed volume at 300âk for 0.5âps. The time step is set to 1âfs and the Nose-Hoover thermostat74 is used. A single gamma point is used to sample the Brillouin zone.
Machine learning model
According to the data type and quantity in the original data set, the supervised learning model was used for training. The regression machine learning algorithm Artificial Neural Network (ANN) was selected. To reduce potential overfitting introduced by strongly correlated features, the Pearson correlation coefficient (r) between features was calculated as75:
where xi and yi are the i-th value of two different input features, respectively; \(\bar{x}\) and \(\bar{y}\) are the expectations of the two input features. Moreover, to achieve the same magnitude level of all features, each feature was normalized by:
where \({x}_{i}^{{norm}}\) and xi are the i-th normalized value and original value of the input feature x, respectively; \(\bar{x}\) and \({\sigma }_{x}\) are the expectation and standard deviation of the input feature x, respectively.
Data availability
The authors declare that the data supporting the findings of this study are available within the article and its supplementary information files or from the corresponding authors on reasonable request.
Code availability
The code used to train machine learning model is available from https://github.com/jzhang-github/elasticnet/.
References
Barsoum, M. W. The MN+1AXN phases: A new class of solids: Thermodynamically stable nanolaminates. Prog. Solid state Chem. 28, 201â281 (2000).
Radovic M & Barsoum M. W. MAX phases: Bridging the gap between metals and ceramics. Am Cream Soc Bull, 92(3): 20-27, (2013).
Sokol, M., Natu, V., Kota, S. & Barsoum, M. W. On the Chemical Diversity of the MAX Phases. Trends Chem. 1, 210â223, (2019).
Barsoum, M., Tamer, E.-R. The MAX phases-Unique new carbide and nitride materials (American Scientist, 2001).
Barsoum, M. Physical properties of the MAX phases, Encyclopedia of Materials: Science and Technology. Elsevier Amst. 1, 1.6 (2006).
Wang, J. & Zhou, Y. Recent Progress in Theoretical Prediction, Preparation, and Characterization of Layered Ternary Transition-Metal Carbides. Annu. Rev. Mater. Res. 39, 415â443 (2009).
Barsoum, M. W. & Radovic, M. Elastic and Mechanical Properties of the MAX Phases. Annu. Rev. Mater. Res. 41, 195â227 (2011).
Wang, C., Tracy, C. L. & Ewing, R. C. Radiation effects in Mn+1AXn phases. Appl. Phys. Rev. 7, https://doi.org/10.1063/5.0019284 (2020).
Sun, Z. M. Progress in research and development on MAX phases: a family of layered ternary compounds. Int. Mater. Rev. 56, 143â166 (2011).
Sauceda, D. et al. High-throughput reaction engineering to assess the oxidation stability of MAX phases. npj Comput. Mater. 7, https://doi.org/10.1038/s41524-020-00464-7 (2021).
Tunes, M. A. et al. Accelerated radiation tolerance testing of Ti-based MAX phases. Mater. Today Energy 30, https://doi.org/10.1016/j.mtener.2022.101186 (2022).
Tunes, M. A., Harrison, R. W., Donnelly, S. E. & Edmondson, P. D. A Transmission Electron Microscopy study of the neutron-irradiation response of Ti-based MAX phases at high temperatures. Acta Materialia 169, 237â247, (2019).
Tunes, M. A., Imtyazuddin, M., Kainz, C., Pogatscher, S. & Vishnyakov, V. M. Deviating from the pure MAX phase concept: Radiation-tolerant nanostructured dual-phase Cr2AlC. Sci. Adv. 7, eabf6771 (2021).
Younker, I. & Fratoni, M. Neutronic evaluation of coating and cladding materials for accident tolerant fuels. Prog. Nucl. Energy 88, 10â18 (2016).
Garcia-Diaz, B. et al. MAX phase coatings for accident tolerant nuclear fuel. Trans. Am. Nucl. Soc. 110, 994â996 (2014).
Pint, B. A., Terrani, K. A., Yamamoto, Y. & Snead, L. L. Material Selection for Accident Tolerant Fuel Cladding. Metall. Mater. Trans. E 2, 190â196 (2015).
Was, G. et al. Emulation of reactor irradiation damage using ion beams. Scr. Materialia 88, 33â36 (2014).
Zinkle, S. J. & Snead, L. L. Opportunities and limitations for ion beams in radiation effects studies: Bridging critical gaps between charged particle and neutron irradiations. Scr. Materialia 143, 154â160 (2018).
Was, G. S. & Averback, R. S. in Comprehensive Nuclear Materials 195â221, Oak Ridge National Lab.(ORNL), Oak Ridge, TN (United States) (2012).
Sickafus, K. E. et al. Radiation Tolerance of Complex Oxides. Science 289, 748â751 (2000).
Sickafus, K. E. et al. Radiation-induced amorphization resistance and radiation tolerance in structurally related oxides. Nat. Mater. 6, 217â223 (2007).
Zhang, J. et al. Ion-irradiation-induced structural transitions in orthorhombic Ln2TiO5. Acta Materialia 61, 4191â4199 (2013).
Uberuaga, B. P. et al. Opposite correlations between cation disordering and amorphization resistance in spinels versus pyrochlores. Nat. Commun. 6, 8750 (2015).
Zhao, J. et al. Complex Ga2O3 polymorphs explored by accurate and general-purpose machine-learning interatomic potentials. npj Comput. Mater. 9, https://doi.org/10.1038/s41524-023-01117-1 (2023).
Wang, C. et al. Disorder in Mn+1AXn phases at the atomic scale. Nat. Commun. 10, 622, https://doi.org/10.1038/s41467-019-08588-1 (2019).
Zhao, S. et al. Phase Transformation and Amorphization Resistance in High-entropy MAX Phase M2SnC (M=Ti, V, Nb, Zr, Hf) under In-situ Ion Irradiation. Acta Materialia 238, https://doi.org/10.1016/j.actamat.2022.118222 (2022).
George, E. P., Raabe, D. & Ritchie, R. O. High-entropy alloys. Nat. Rev. Mater. 4, 515â534 https://doi.org/10.1038/s41578-019-0121-4 (2019).
Ding, Q. et al. Tuning element distribution, structure and properties by composition in high-entropy alloys. Nature 574, 223â227 (2019).
Yin, B. & Curtin, W. A. First-principles-based prediction of yield strength in the RhIrPdPtNiCu high-entropy alloy. npj Comput. Mater. 5, https://doi.org/10.1038/s41524-019-0151-x (2019).
Giles, S. A., Sengupta, D., Broderick, S. R. & Rajan, K. Machine-learning-based intelligent framework for discovering refractory high-entropy alloys with improved high-temperature yield strength. npj Comput. Mater. 8, https://doi.org/10.1038/s41524-022-00926-0 (2022).
Zhang, Y., Osetsky, Y. N. & Weber, W. J. Tunable Chemical Disorder in Concentrated Alloys: Defect Physics and Radiation Performance. Chem. Rev. https://doi.org/10.1021/acs.chemrev.1c00387 (2021).
Zhang, Y., Egami, T. & Weber, W. J. Dissipation of radiation energy in concentrated solid-solution alloys: Unique defect properties and microstructural evolution. MRS Bull. 44, 798â811 (2019).
Zhang, Y. et al. Influence of chemical disorder on energy dissipation and defect evolution in concentrated solid solution alloys. Nat. Commun. 6, 8736 (2015).
Xiang, H. et al. High-entropy ceramics: Present status, challenges, and a look forward. J. Adv. Ceramics https://doi.org/10.1007/s40145-021-0477-y (2021).
Wang, C. et al. Irradiation-induced structural transitions in Ti2AlC. Acta Materialia 98, 197â205 (2015).
Jin, K. et al. Effects of compositional complexity on the ion-irradiation induced swelling and hardening in Ni-containing equiatomic alloys. Scr. Materialia 119, 65â70 (2016).
Zhao, S. et al. Phase transformation and amorphization resistance in high-entropy MAX phase M2SnC (M= Ti, V, Nb, Zr, Hf) under in-situ ion irradiation. Acta Materialia 238, 118222 (2022).
Jiang, C., Zheng, M.-J., Morgan, D. & Szlufarska, I. Amorphization Driven by Defect-Induced Mechanical Instability. Phys. Rev. Lett. 111, https://doi.org/10.1103/PhysRevLett.111.155501 (2013).
Devanathan, R., Gao, F. & Weber, W. J. Amorphization of silicon carbide by carbon displacement. Appl. Phys. Lett. 84, 3909â3911 (2004).
Chartier, A., Meis, C., Crocombette, J.-P., Corrales, L. R. & Weber, W. J. Atomistic modeling of displacement cascades inLa2Zr2O7pyrochlore. Phys. Rev. B 67, https://doi.org/10.1103/PhysRevB.67.174102 (2003).
Chartier, A., Catillon, G. & Crocombette, J.-P. Key role of the cation interstitial structure in the radiation resistance of pyrochlores. Phys. Rev. Lett. 102, 155503 (2009).
Liu, C., Li, Y., Shi, T., Peng, Q. & Gao, F. Oxygen defects stabilize the crystal structure of MgAl2O4 spinel under irradiation. J. Nucl. Mater. 527, 151830 (2019).
Larsen, P. M., Schmidt, S. & Schiøtz, J. Robust structural identification via polyhedral template matching. Model. Simul. Mater. Sci. Engin. 24, https://doi.org/10.1088/0965-0393/24/5/055007 (2016).
Yang, T. et al. The structural transitions of Ti3AlC2 induced by ion irradiation. Acta Materialia 65, 351â359 (2014).
Wang, C. et al. Ion-irradiation-induced structural evolution in Ti4AlN3. Scr. Materialia 133, 19â23 (2017).
Huang, Q. et al. Saturation of ion irradiation effects in MAX phase Cr2AlC. Acta Materialia 110, 1â7 (2016).
Xiao, J. et al. Investigations on Radiation Tolerance of Mn+1AXn Phases: Study of Ti3SiC2, Ti3AlC2, Cr2AlC, Cr2GeC, Ti2AlC, and Ti2AlN. J. Am. Ceram. Soc. 98 1323â1331 (2015).
Shah, S. & Bristowe, P. Point defect formation in M2AlC (M= Zr, Cr) MAX phases and their tendency to disorder and amorphize. Sci. Rep. 7, 1â8 (2017).
Zhao, S., Xue, J., Wang, Y. & Huang, Q. Ab initio study of irradiation tolerance for different Mn+1AXn phases: Ti3SiC2 and Ti3AlC2. J. Appl. Phys. 115, https://doi.org/10.1063/1.4861384 (2014).
Wang, J., Liu, B., Wang, J. & Zhou, Y. Theoretical investigation of thermodynamic stability and mobility of the intrinsic point defects in Ti3AC2 (A = Si, Al). Phys. Chem. Chem. Phys. 17, 8927â8934 (2015).
Han, H. et al. A first-principles study on the defective properties of MAX phase Cr2AlC: the magnetic ordering and strong correlation effect. RSC Adv. 6, 84262â84268 (2016).
Wang, C. et al. Structural Transitions Induced by Ion Irradiation in V2AlC and Cr2AlC. J. Am. Ceram. Soc. 99, 1769â1777 (2016).
Bader, R. Atoms in molecules: a quantum theory (Oxford University Press, 1990).
Song, H. et al. Local lattice distortion in high-entropy alloys. Phys. Rev. Mater. 1, https://doi.org/10.1103/PhysRevMaterials.1.023404 (2017).
Ogata, S., Li, J. & Yip, S. Ideal pure shear strength of aluminum and copper. Science 298, 807â811 (2002).
Schmidt, J., Marques, M. R. G., Botti, S. & Marques, M. A. L. Recent advances and applications of machine learning in solid-state materials science. npj Comput. Mater. 5, https://doi.org/10.1038/s41524-019-0221-0 (2019).
Tian, W. et al. Synthesis and thermal and electrical properties of bulk Cr2AlC. Scr. Materialia 54, 841â846 (2006).
Hu, C. et al. In SituReaction Synthesis and Mechanical Properties of V2AlC. J. Am. Ceram. Soc. 91, 4029â4035 (2008).
Lin, Z. J., Zhuo, M. J., Zhou, Y. C., Li, M. S. & Wang, J. Y. Microstructural characterization of layered ternary Ti2AlC. Acta Materialia 54, 1009â1015 (2006).
Chen, L. et al. Multiprincipal Element M2FeC (M = Ti,V,Nb,Ta,Zr) MAX Phases with Synergistic Effect of Dielectric and Magnetic Loss. Adv. Sci. 10, https://doi.org/10.1002/advs.202206877 (2023).
Huang, M., Li, Y., Ran, G., Yang, Z. & Wang, P. Cr-coated Zr-4 alloy prepared by electroplating and its in situ He+ irradiation behavior. J. Nucl. Mater. 538, https://doi.org/10.1016/j.jnucmat.2020.152240 (2020).
Stoller, R. E. et al. On the use of SRIM for computing radiation damage exposure. Nucl. Instrum. Methods Phys. Res. Sect. B. 310, 75â80, (2013).
Bugnet, M., Mauchamp, V., Oliviero, E., Jaouen, M. & Cabiocâh, T. Chemically sensitive amorphization process in the nanolaminated Cr2AC (A=Al or Ge) system from TEM in situ irradiation. J. Nucl. Mater. 441, 133â137 (2013).
Yang, T. et al. Formation of nano-twinned structure in Ti3AlC2 induced by ion-irradiation. Acta Materialia 128, 1â11 (2017).
Whittle, K. R. et al. Radiation tolerance of Mn+1AXn phases, Ti3AlC2 and Ti3SiC2. Acta Materialia 58, 4362â4368 (2010).
Koch, C. T. Determination of core structure periodicity and point defect density along dislocations (Arizona State University, 2002).
Kresse, G. & Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 54, 11169â11186 (1996).
Kresse, G. & Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 59, 1758 (1999).
John, P., Perdew, K. B. & Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 77, 3865â3868 (1996).
Van De Walle, A., Asta, M. & Ceder, G. The alloy theoretic automated toolkit: A user guide. Calphad 26, 539â553 (2002).
Zunger, A., Wei, S., Ferreira, L. G. & Bernard, J. E. Special quasirandom structures. Phys. Rev. Lett. 65, 353â356, (1990).
Pack, J. D. & Monkhorst, H. J. Special points for Brillouin-zone integrationsââa reply. Phys. Rev. B 16, 1748â1749 (1977).
Otero-de-la-Roza, A., Johnson, E. R. & Luaña, V. Critic2: A program for real-space analysis of quantum chemical interactions in solids. Comput. Phys. Commun. 185, 1007â1018 (2014).
Hoover, W. G. Canonical dynamics: Equilibrium phase-space distributions. Phys. Rev. A 31, 1695 (1985).
Zhang, J. et al. Design high-entropy carbide ceramics from machine learning. npj Comput. Mater. 8, https://doi.org/10.1038/s41524-021-00678-3 (2022).
Ziegler, J. F., Ziegler, M. D. & Biersack, J. P. SRIM â The stopping and range of ions in matter (2010). Nucl. Instrum. Methods Phys. Res. Sect. B. 268, 1818â1823 (2010).
Acknowledgements
We also thank the support of Electron Microscope Laboratory (EML) in Peking University for TEM characterization and the High Performance Computing Platform of the Center for Life Sciences (Peking University) for the first-principle calculation. This work was supported by the National Natural Science Foundation of China (Grant No. 12275009, 12192280, and 11935004) and National Magnetic Confinement Fusion Energy Research Project 2021YFE031100.
Author information
Authors and Affiliations
Contributions
H.X. performed the DFT calculations and drafted the initial manuscript. Shu.Z. and L.C. carried out the irradiation experiments. J.Z. and Shi.Z. trained the machine learning model. Y.L., K.C. and Q.H. prepared the materials. C.W. and Y.W. supervised the project, reviewed, and edited the manuscript. All authors contributed to the manuscript writing and discussion.
Corresponding author
Ethics declarations
Competing interests
The authors declare no competing interests. The author Shijun Zhao is a guest editor of this journal.
Additional information
Publisherâs note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the articleâs Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the articleâs Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
About this article
Cite this article
Xiao, H., Zhao, S., Zhang, J. et al. Distinct amorphization resistance in high-entropy MAX-phases (Ti, M)2AlC (M=Nb, Ta, V, Zr) under in situ irradiation. npj Comput Mater 10, 196 (2024). https://doi.org/10.1038/s41524-024-01370-y
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41524-024-01370-y
This article is cited by
-
Functional Applications and Data-Driven Design of High-Entropy Ceramics
High Entropy Alloys & Materials (2024)