Novel chalcogen and halogen surface functional groups for tuning the mechanical properties of TMDs/MXenes heterojunctions

Yuqian Zhang a, Zhiwei Liu ab, Siyu Zheng c, Changyang Yu a, Siliang Yue a, Chenliang Li *a and Hui Qi *a
aCollege of Aerospace and Civil Engineering, Harbin Engineering University, Harbin, 150001, P. R. China. E-mail: lichenliang@hrbeu.edu.cn; qihui205@sina.com
bNuclear Power Institute of China, Chengdu, 610213, P. R. China
cHiwing Aerospace Materials Research Institute (Suzhou) Co., Ltd, Suzhou, 215100, P. R. China

Received 11th April 2025 , Accepted 18th August 2025

First published on 19th August 2025


Abstract

MXenes exhibit remarkable mechanical properties due to their unique structural properties and strong atomic bonding, making them highly competitive among 2D materials. Forming heterojunctions between TMDs and MXenes offers a promising strategy to enhance material performance for advanced applications. Although extensive studies have explored the electronic and chemical properties of MXenes-based heterojunctions, investigations into their mechanical properties, particularly the effects of surface functional groups, remain limited. This work systematically investigates the structural and mechanical properties of pristine MXenes and their heterojunctions with TMDs, focusing on how novel chalcogen (S and Se) and halogen (Cl and Br) surface functional groups tune mechanical behavior, anisotropy, and microscopic failure mechanisms. Our results indicate that while surface functionalization generally reduces tensile strength, it enhances the ductility of the heterojunctions. Among the functional groups studied, –Se functionalization induces the most significant improvement in flexibility, indicating potential for applications in flexible devices. Compared with conventional –O or –F terminations reported in previous studies, these newly synthesized functional groups induce distinct anisotropic mechanical responses and tunable interfacial bonding. These findings provide a deeper understanding of the mechanical tuning of TMDs/MXenes heterojunctions by surface functional groups and offer valuable insights into their design for next-generation flexible electronics devices, high-performance sensors and thin-film batteries.


1 Introduction

MAX phase materials represent a versatile family of ternary carbides and nitrides, described by the general formula Mn+1AXn (n = 1, 2, and 3), where M is an early transition metal, A is a group A element (typical IIIA or IVA), and X is either carbon or nitrogen.1,2 In 2011, Naguib et al.3 successfully synthesized the first MXenes, two-dimensional (2D) Ti3C2, by selectively etching the Al atomic layer from Ti3AlC2 (a typical member in the MAX family) using hydrofluoric acid (HF). This process led to the emergence of MXenes with the general formula Mn+1XnTx (n = 1, 2, and 3), where M denotes the transition metal, X is carbon or nitrogen, and T denotes the surface functional groups.4,5 Due to their excellent electrical,6 thermal,7 and mechanical8,9 properties, MXenes have gained significant attention across diverse applications, especially in energy storage, catalysis, adsorption, and sensing technologies.5,10–13 Among these, Ti3C2 is notable for its unique structure,14 which enables a significantly higher metal ion storage capacity than other 2D materials.

Previous studies have shown that experimentally synthesized MXenes inevitably undergo surface functionalization;15,16 as a result, extensive research has focused on the effect of functional groups in tuning the overall properties of MXenes. Notably, surface functional groups are known to have substantial impacts on the essential properties of MXenes enhancing their Li-ion capacity,17 Seebeck coefficients at low temperatures,18 metallic conductivity,19 light-harvesting efficiency20 and high photothermal-conversion performance.21 Moreover, functionalized MXenes have demonstrated great potential in biosensing and nanophotonics, enabling advanced applications such as electrochemical biosensors for SARS-CoV-2 detection22 and CRISPR/Cas12a-based plasmonic platforms for rapid virus identification.23 Furthermore, the integration of these sensors with DNA origami also enables ultrasensitive detection of circulating tumor DNA,24 highlighting their biomedical potential. The effect of conventional functional groups (–O, –F, and –OH) on the mechanical behavior of MXenes has been widely studied.25,26 Among these, the presence of –O groups has been found to stabilize the surface atomic layers, preventing collapse and enhancing mechanical resilience. Recently, Kamysbayev et al.27 reported the first successful synthesis of Ti3C2 terminated with a class of novel surface functional groups (–S, –Se, –Cl, and –Br) through substitution reactions in molten inorganic salts. These newly developed Ti3C2X2 (X = S, Se, Cl, and Br) exhibit promising potential for catalytic applications; for instance, Ti3C2S2 and Ti3C2Se2 exhibit enhanced surface activity and high catalytic performance in the hydrogen evolution reaction (HER).28,29 However, despite their recent emergence and potential, the influence of these novel functional groups on the mechanical properties of MXenes has not been systematically explored yet.

Beyond individual MXenes, the formation of heterojunctions between MXenes and other 2D materials, particularly transition metal dichalcogenides (TMDs), provides additional opportunities to enhance material properties through synergistic interactions. These heterojunctions combine the unique properties of both monolayer materials, such as the electrical conductivity and chemical stability of MXenes with the catalytic and optoelectronic properties of TMDs. As a result, MXenes/TMDs heterojunctions have gained significant attention for applications in flexible sensors,30,31 energy storage,32,33 and catalytic systems.34,35 Chandran et al.36 found that the addition of MoS2 reduced the charge transfer resistance in MoS2/MXenes heterojunctions, resulting in enhanced cycling stability and reversibility, making them suitable for supercapacitor electrodes. Similarly, Tao et al.37 successfully assembled MoSe2/Ti3C2 heterojunctions, which exhibited improved nitrogen reduction reaction activity, positioning them as effective catalysts. Liu et al.38 reported that vertical strain tunes the interfacial properties of MoS2/Ti3C2 heterojunctions, transitioning from Ohmic to Schottky contacts under tensile strain. The role of functional groups in tuning the properties of MXenes-based heterojunctions has also drawn increasing attention. For instance, altering the –O and –F functional groups allows precise tuning of the Schottky barrier height.39 However, previous studies have primarily focused on conventional –O, –F and –OH functional groups. Although Ti3C2 functionalized with novel groups (–S, –Se, –Cl, and –Br) has been successfully synthesized,27 most previous studies have continued to focus on conventional –O, –F and –OH functional groups. Consequently, the influence of these novel surface terminations on the mechanical performance and failure mechanisms of TMDs/MXenes heterojunctions remains largely unexplored and lacks systematic investigation. To fill this important gap, it is imperative to explore how these unconventional terminations tune the mechanical behavior of TMDs/MXenes heterostructures.

This study aims to clarify how novel halogen (Cl and Br) and chalcogen (S and Se) surface functional groups tune the anisotropic mechanical behavior and failure mechanisms of TMDs/MXenes heterojunctions. We have systematically calculated the elastic constants, Poisson's ratios and stress–strain behavior of TMDs/Ti3C2 and TMDs/Ti3C2X2 (S, Se, Cl, and Br) heterojunctions. Our analysis reveals that functionalization decreases the stiffness and tensile strength of heterojunctions while enhancing their ductility, with –Se functional groups providing the most significant ductility improvement. Furthermore, the mechanical properties of TMDs, Ti3C2X2 and TMDs/Ti3C2X2 heterojunctions in the zig-zag (x-) and armchair (y-) directions are similar under small deformations. According to the microscopic analysis, it can be observed that Ti–C bonds remain dominant in resisting tensile deformation in the y-direction. These findings provide new insights into the effect of functional groups on the mechanical properties of TMDs/MXenes heterojunctions, with significant implications for flexible electronic devices and thin-film batteries.

2 Computational methods

All density functional theory (DFT) calculations were performed using the Cambridge Sequential Total Energy Package (CASTEP) code within the Material Studio (MS).40 For lattice parameters and atomic coordinate optimization, we employed the Perdew–Burke–Ernzerhof (PBE) functional within the generalized gradient approximation (GGA) as the exchange–correlation function.41 To account for the interlayer interactions in the heterojunctions, a semi-empirical dispersion correction based on the Grimme approach (DFT-D2)42 was employed. Although the DFT-D2 method was known to have limitations in accurately describing van der Waals (vdW) interactions, it has been widely employed in previous MXenes-based first-principles studies.43,44 These studies confirmed that the DFT-D2 approach is capable of accurately capturing the structural and mechanical properties of MXenes systems. Furthermore, its adoption in this work ensures methodological consistency with our earlier publications. Additionally, ultrasoft pseudopotentials were used to accurately model the interactions between ionic cores and valence electrons, facilitating the smoothing of wave functions45 and enhancing computational performance. We utilized the [Ar]3p3d4s configuration to represent the valence of Ti, the [He]2s2p configuration for C, the [Ne]3s3p configuration for S and Cl, the [Kr]4d5s configuration for Mo, and the [Ar]4p4s configuration for Br and Se. The plane-wave cutoff energy was set to 500 eV, and the Brillouin zone was sampled using a 9 × 5 × 1 Monkhorst–Pack k-point grid.46 In geometry optimization sections, the Broyden–Fletcher–Goldfarb–Shanno (BFGS) method47 was utilized with convergence criteria set at 10−7 eV per atom for total energy and 0.005 eV Å−1 for force,48 respectively. Spin polarization effects were not considered, as the local magnetic moments originating from transition metal atoms at the edges of pristine MXenes layers are significantly weakened or even entirely diminished due to interfacial interactions after forming the heterojunctions with TMDs, as reported in previous studies.49 To avoid periodic interactions between neighboring layers, a vacuum layer of 15 Å was introduced in the z-direction.50,51 The choice of this vacuum thickness is validated in previous studies on MXenes-based heterojunctions,52,53 where a 15 Å vacuum layer was shown to effectively prevent the interactions between adjacent layers. Additionally, our previous work54 on the electronic properties of TMDs/MXenes heterojunctions further confirms that this vacuum thickness is sufficient.

3 Results and discussion

3.1 Structural properties

The structural features of monolayer MoS2, MoSe2, and Ti3C2, as well as the atomic configurations of both pristine TMDs/Ti3C2 and functionalized TMDs/Ti3C2X2 (X = S, Se, Cl, and Br) heterojunctions, were comprehensively established and validated in our previous study.54 To ensure consistency and comparability, we adopt the most stable stacking configurations with the lowest total binding energies as identified in ref. 54, for all subsequent analyses in this study. Specifically, for the pristine TMDs/Ti3C2 heterojunctions, the ZM_SA configuration is adopted for MoS2/Ti3C2 and the SA_ZM configuration is adopted for MoSe2/Ti3C2, as shown in Fig. 1(a). For the functionalized heterojunctions, the functional groups are positioned at the hollow sites of the surface Ti atoms on the surface and aligned vertically to the intermediate Ti layer, as shown in Fig. 1(b). To focus on the intrinsic effect of functional group types, we adopted a fully saturated model with a fixed maximum coverage, where each surface Ti atom is terminated by a chalcogen or halogen atom, forming the Ti3C2X2 (X = S, Se, Cl, and Br) structures.
image file: d5cp01389e-f1.tif
Fig. 1 The side views of (a) TMDs/Ti3C2 heterojunctions in ZM_SA and SA_ZM configurations and (b) TMDs/Ti3C2X2 (X = S, Se, Cl, and Br) heterojunctions.54 The atomic alignments at the TMDs/MXenes interfaces are explicitly indicated by red dashed lines.

Earlier work54 demonstrated that surface functionalization increases the interlayer distance (d) and induces noticeable Ti–C bond elongation in TMDs/Ti3C2 heterojunctions, with MoSe2/Ti3C2Se2 exhibiting a zero bond length difference (d56) between Mo–S(Se)(1) and Mo–S(Se)(2). While our previous study focused on tuning the electronic properties of TMDs/MXenes heterojunctions through biaxial strain and external electric fields, the present study investigates how surface functionalization and uniaxial strains tune the mechanical performance. These structural properties provide qualitative insights into the expected mechanical behavior. Pristine TMDs/MXenes heterojunctions with a smaller interlayer distance generally exhibit stronger vdW coupling and more efficient load transfer, potentially leading to higher tensile strength but reduced ductility. In contrast, surface functionalization increases the interlayer distance, thereby weakening direct vdW interactions and facilitating interlayer sliding, which enhances ductility while reducing the ultimate strength. For the MoSe2/Ti3C2Se2 heterojunctions, the bond-length difference d56 = 0 implies negligible vertical distortion of the MoSe2 lattice and overall weak interfacial coupling, preserving the intrinsic in-plane anisotropy of MoSe2 and thus leading to the highest expected mechanical anisotropy among the systems studied.

3.2 Electronic properties

Interfacial charge coupling in TMDs/MXenes heterojunctions was examined through the work function analysis of each monolayer, as listed in Table 1. The work function quantifies the energy barrier for electron emission, with higher values indicating stronger electron confinement. Pristine MoS2 and MoSe2 exhibit work functions of 4.36 and 4.89 eV, respectively, while bare Ti3C2 presents a lower value of 3.96 eV, consistent with its metallic nature. Surface functionalization significantly tunes the electronic properties of Ti3C2. Chalcogen terminations (–S and –Se) increase the work function to 5.02 and 5.08 eV, respectively, indicating enhanced electron affinity. In contrast, halogen terminations lead to divergent trends, with Ti3C2Cl2 showing a slight increase (4.19 eV) and Ti3C2Br2 a notable decrease (3.72 eV) in work function.
Table 1 Calculated work functions (eV) of pristine TMDs, Ti3C2 and Ti3C2X2 (X = S, Se, Cl, and Br) monolayers
Formula Work function (eV)
MoS2 4.36
MoSe2 4.89
Ti3C2 3.96
Ti3C2S2 5.02
Ti3C2Se2 5.08
Ti3C2Cl2 4.19
Ti3C2Br2 3.72


Notably, the mismatch in the work function between TMDs and functionalized MXenes is a key factor governing the interfacial electron transfer and potential alignment. When the MXenes exhibits a higher work function (Ti3C2S2 and Ti3C2Se2), electrons flow from the TMDs layer toward the MXenes, enhancing interfacial the charge redistribution and strengthening vdW interactions. Conversely, in systems where the TMDs have a higher work function (Ti3C2Cl2 and Ti3C2Br2), electron transfer proceeds in the opposite direction, resulting in weakened interfacial coupling. These results confirm that surface functional groups can effectively regulate the electronic environment of Ti3C2 and thus control the direction and magnitude of charge transfer at TMDs/MXenes interfaces.

To further analyze the interlayer electron transport, the electron localization function (ELF) of MoS2/Ti3C2X2 and MoSe2/Ti3C2X2 is shown in Fig. S1(a) and (b), respectively. In the MoS2 and MoSe2 layers, high electron localization around the Mo and S (Se) atoms indicates strong covalent interactions, with an ELF value of around 0.5. In the Ti3C2 layer, electron localization is observed around C and Ti atoms due to lone-pair electrons of Ti atoms. Although these electrons do not participate directly in bonding, they contribute to localized electron density, which can interact with external substances, making them potentially valuable for catalysis or sensing.55 The introduction of functional groups reduces electron density around the metal atoms and even at the interfaces, weakening the interactions between the TMDs and Ti3C2 layers.56 This reduction is most noticeable in –Cl and –Br functionalization, indicating limited electron transport due to the lack of a continuous electron-rich region.

Given these electronic properties, it is feasible to infer that different functional groups could have a substantial influence on interfacial coupling strength and mechanical performance. As demonstrated in our previous study,54 –Cl and –Br terminations preserve the Ti-3d dominated metallicity near the Fermi level, whereas –S and –Se functionalization induces pronounced band flattening and electron localization, thereby reducing electrical conductivity. In particular, –Cl and –Br functionalization is expected to preserve strong interlayer interactions due to the sustained metallic character, which may promote efficient charge delocalization and facilitate stress transfer across the interface, thereby maintaining or potentially enhancing the ductility of the heterojunction. In contrast, the strong electronic localization induced by –S and –Se functionalization may weaken interlayer coupling, reduce stress transfer across the interface, and consequently result in increased anisotropy and reduced ductility.

3.3 Mechanical properties

Both monolayer TMDs and Ti3C2 possess a 2D hexagonal lattice structure, which presents challenges for the accurate calculation of their mechanical properties. Therefore, we redefined the lattice vectors and converted the original hexagonal lattice into a 2D orthorhombic crystal structure prior to mechanical property calculations, as illustrated in Fig. S2(a) and (b), respectively. Uniaxial strain is applied along the x- and y-directions, corresponding to the zig-zag and armchair orientations, respectively.

We first calculated the elastic properties of pristine TMDs and Ti3C2 monolayers as well as functionalized Ti3C2X2 (X = S, Se, Cl, and Br) systems, with the results listed in Table 2. In the thickness direction, the in-plane stiffness constants are expressed in units of N m−1 instead of GPa, because monolayer TMDs consist of atomic layers, and the effective thickness cannot be clearly defined. Our calculated elastic constants for monolayer MoS2 and MoSe2 are in close agreement with reported previous DFT results,57 with C11 deviations below 5.5 N m−1 and Poisson's ratio differences within 0.01. Similarly, the computed C11 value for pristine Ti3C2 is consistent with the value reported by Tang et al.58 These results further support the reliability of our computational approach for accurately describing the mechanical properties of MXenes-based systems.

Table 2 Comparison of in-plane stiffness constants (C11, C22, and C12) and Poisson's ratios (v12, v21) for pristine TMDs, Ti3C2, and functionalized Ti3C2X2 (X = S, Se, Cl, and Br)
Formula C 11 C 22 C 12 ν 12 ν 21
MoS2 129.7 N m−1 129.7 N m−1 33.7 N m−1 0.26 0.26
MoS257 124.2 N m−1 124.2 N m−1 31.1 N m−1 0.25 0.25
MoSe2 112.3 N m−1 112.3 N m−1 24.7 N m−1 0.22 0.22
MoSe257 108.1 N m−1 108.1 N m−1 24.9 N m−1 0.23 0.23
Ti3C2 506.3 GPa 491.2 GPa 91.1 GPa 0.19 0.18
Ti3C2S2 412.6 GPa 397.3 GPa 107.3 GPa 0.27 0.26
Ti3C2Se2 289.1 GPa 280.2 GPa 89.6 GPa 0.32 0.31
Ti3C2Cl2 376.4 GPa 340.2 GPa 105.4 GPa 0.31 0.28
Ti3C2Br2 387.2 GPa 373.4 GPa 104.5 GPa 0.28 0.27


Notably, monolayer MoS2 and MoSe2 exhibit nearly isotropic mechanical properties, with C11 and C22 values of 129.7 N m−1 and 112.3 N m−1, respectively. A higher Poisson ratio generally indicates greater mechanical ductility, which is consistent with recent findings.59 The calculated C11 value for MoS2 is significantly lower than that of h-BN and graphene, despite its higher Poisson's ratio, indicating that monolayer MoS2 exhibits better ductility. When converting the C11 of Ti3C2 from GPa to N m−1, we found that it corresponds to 241.8 N m−1, which exceeds that of MoS2 on the same scale; however, MoS2 shows greater stiffness per unit thickness compared to Ti3C2.

Ti3C2 exhibits a remarkably high stiffness constant, which significantly surpasses those of TMDs monolayers. However, surface functionalization leads to a notable reduction in stiffness. Among the functionalized MXenes, Ti3C2Se2 displays the lowest stiffness constants, with C11 and C22 values of 289.1 and 280.2 GPa, respectively. Additionally, Ti3C2Se2 exhibits the highest Poisson's ratio of 0.32, indicating that it experiences a greater degree of transverse expansion under uniaxial tensile stress compared to its counterparts. Furthermore, the elastic constants and Poisson's ratios reveal that the mechanical properties of Ti3C2X2 are largely isotropic, with similar behavior in both the x- and y-directions.

Furthermore, Table 3 lists the in-plane stiffness constants, Young's modulus and Poisson's ratios of the TMDs/Ti3C2 and TMDs/Ti3C2X2 heterojunctions. The in-plane stiffness of the TMDs/Ti3C2 heterojunctions is lower than that of pristine Ti3C2 but higher than that of the MoS2 and MoSe2 layers, suggesting a synergistic strengthening effect. The Poisson ratios of the TMDs/Ti3C2X2 (X = S, Cl, Br) heterojunctions exhibit nearly identical values in both the x- and y-directions, indicating isotropic mechanical behavior.

Table 3 The in-plane stiffness constants (C11, C22, and C12), Young's modulus (Y12 and Y21) and Poisson's ratios (v12, v21) of TMDs/Ti3C2 and TMDs/Ti3C2X2 (X = S, Se, Cl, and Br) heterojunctions
Formula C 11 (GPa) C 22 (GPa) C 12 (GPa) Y 1 (GPa) Y 2 (GPa) ν 12 ν 21
MoS2/Ti3C2 387.4 478.1 81.2 373.4 461.1 0.17 0.21
MoS2/Ti3C2S2 349.6 321.2 80.3 329.2 302.7 0.25 0.23
MoS2/Ti3C2Se2 260.2 327.1 101.4 228.5 287.6 0.31 0.39
MoS2/Ti3C2Cl2 321.5 309.8 89.9 295.6 284.3 0.29 0.28
MoS2/Ti3C2Br2 329.3 316.8 85.5 306.7 293.9 0.27 0.26
MoSe2/Ti3C2 444.6 412.3 115.4 411.8 382.3 0.28 0.26
MoSe2/Ti3C2S2 335.5 324.2 100.5 303.9 294.2 0.31 0.30
MoSe2/Ti3C2Se2 337.2 599.1 107.8 318.3 525.3 0.18 0.32
MoSe2/Ti3C2Cl2 326.4 350.1 94.5 301.4 322.8 0.27 0.29
MoSe2/Ti3C2Br2 351.7 321.8 77.2 332.5 304.1 0.24 0.22


It is worth noting that the MoSe2/Ti3C2Se2 heterojunctions display pronounced anisotropy, with a Poisson ratio of 0.18 in the x-direction and 0.32 in the y-direction. This anisotropy can be attributed to the larger atomic radius of –Se and the severe distortion of the –Se atomic arrangement within the MoSe2 and Ti3C2Se2 layers under uniaxial strain. This distortion disrupts the hexagonal symmetry of MoSe2 and brings the –Se atoms in closer proximity. The attraction between the upper and lower Se– atoms in this direction contributes to the observed anisotropy and increased resistance to shear deformation in the MoSe2/Ti3C2Se2 heterojunctions.

In the MoS2/Ti3C2 heterojunctions, the introduction of functional groups reduces stiffness in both the x- and y-directions. Among the functionalized heterojunctions, MoS2/Ti3C2Se2 exhibits the lowest Young's modulus in the x-direction, decreasing by 144.9 to 228.5 GPa compared to pristine Ti3C2. This significant reduction is largely attributed to –Se functionalization, as the larger atomic radius and weaker bond strength of Se lead to greater lattice distortion and weaker interlayer interactions. In the y-direction, the Young's modulus for MoS2/Ti3C2 is 461.1 GPa, while the –Cl functional groups have notable effects, resulting in a minimum Young's modulus of 284.3 GPa for MoS2/Ti3C2Cl2, indicating a considerable difference of 176.8 GPa. These results suggest that the heterojunction is more sensitive to strain in the y-direction relative to the x-direction.

In contrast, for MoSe2/Ti3C2 heterojunctions, functionalization shows a different trend in the y-direction. For MoSe2/Ti3C2Se2 heterojunctions, the Young's modulus increases to 525.3 GPa, indicating a 65.1% improvement in structural stiffness, suggesting that the –Se functional group enhances the resistance to deformation in the y-direction. This effect is possibly due to increased electron density between the –Se functional group and Ti3C2, strengthening the heterojunction's mechanical stability. While surface functionalization reduces the stiffness and Young's modulus of pristine MXenes and TMDs/MXenes heterojunctions, it plays a crucial role in tuning surface stability and electronic properties provided in our previous studies.54 Moreover, the reduction in stiffness and Young's modulus provides valuable theoretical insights into the development of flexible devices. Therefore, by further systematically investigating the effect of functional groups on the overall mechanical behavior of heterojunctions, we aim to optimize their performance for specific applications.

Both MoS2 and MoSe2 exhibit higher tensile strengths in the y-direction relative to the x-direction, as shown in Fig. 2. Specifically, for MoS2 in the y-direction, the maximum stress reaches 22.47 GPa at a strain of 23%. In comparison, MoSe2 demonstrates lower stiffness and tensile strength, as evidenced by reduced stress values for corresponding strain percentages. A distinct step-like jump appears in the stress–strain curves of monolayer MoS2 and MoSe2 around 8% tensile strain, resulting from a structural transition caused by the relative displacement of the outer S or Se atomic layers. This behavior correlates with a well-known strain-induced semiconductor-to-metal phase transition in TMDs monolayers.60


image file: d5cp01389e-f2.tif
Fig. 2 The stress–strain curves of pristine (a) MoS2 and (b) MoSe2 under uniaxial tensile strains along the x- and y-directions. The hollow symbols indicate the peak points of the corresponding curves, representing the onset strain of mechanical instability.

At tensile strains below 5%, both MoS2 and MoSe2 display near-isotropic mechanical behavior with consistent stress–strain responses along the x- and y-directions. However, beyond the strain of 6%, the stress–strain curves reveal divergent tensile strength values along these two directions, suggesting that higher strain levels induce distortion of the orthorhombic six-membered ring structure in TMDs. This distortion disrupts the original symmetry of the material, leading to pronounced anisotropic mechanical properties. These findings emphasize the susceptibility of TMDs to structural asymmetry under large tensile strains.

As shown in Fig. 3, pristine Ti3C2 exhibits nearly isotropic mechanical behavior, with minimal variation in stress–strain response in the x- and y-directions up to a maximum strain of 24%. However, functionalized Ti3C2 demonstrates pronounced anisotropy, characterized by higher tensile strength along the x-direction compared to the y-direction. This anisotropic behavior arises from the disruption of the original Ti–C bonding caused by different surface functional groups. Among the Ti3C2X2 heterojunctions, Ti3C2Se2 shows the highest anisotropy with the largest tensile strength difference between the x- and y-directions, evident at the early stage of strain. This suggests that Ti–Se bonding is notably weaker along the y-direction, leading to a substantial reduction in the original symmetry of Ti3C2.


image file: d5cp01389e-f3.tif
Fig. 3 The stress–strain curves of pristine (a) Ti3C2 and (b)–(e) Ti3C2X2 (X = S, Se, Cl, and Br) under uniaxial tensile strains along the x- and y-directions. The hollow symbols indicate the peak points of the corresponding curves, representing the onset strain of mechanical instability.

In contrast, the tensile strength of all Ti3C2X2 is lower compared to that of pristine Ti3C2, particularly along the y-direction, suggesting that functional groups generally weaken the mechanical properties of Ti3C2. Among the functional groups studied, Ti3C2Br2 shows the greatest reduction in tensile strength along the x-direction. While Ti3C2S2 shows the greatest reduction along the y-direction, decreasing from 49.19 GPa in pristine Ti3C2 to 26.59 GPa. Moreover, the slopes of the stress–strain curves for Ti3C2S2 and Ti3C2Se2 along the y-direction are significantly smaller compared to those along the x-direction, indicating that the elastic modulus is lower in the y-direction than in the x-direction.

To assess the influence of different functional groups on the mechanical properties of TMDs/MXenes heterojunctions, we further analyzed the uniaxial tensile strains in x- and y-directions of MoS2/Ti3C2 and MoS2/Ti3C2X2 (X = S, Se, Cl, and Br). As shown in Fig. 4(a), the heterojunctions functionalized with Se and Cl in the x-direction exhibit enhanced resistance to deformation, as indicated by their higher tensile strengths compared to the bare MoS2/Ti3C2 heterojunctions (34.33 GPa), with MoS2/Ti3C2Se2 exhibiting the highest tensile strength (36.72 GPa). In contrast, as illustrated in Fig. 4(b), functionalization in the y-direction has a more pronounced weakening effect, where the bare MoS2/Ti3C2 heterojunction shows the highest tensile strength (31.81 GPa). Notably, Se functionalization significantly reduces the tensile strength to 23.87 GPa, which can be attributed to weakened Ti–Se bonding.


image file: d5cp01389e-f4.tif
Fig. 4 The stress–strain curves of MoS2/Ti3C2 and MoS2/Ti3C2X2 (X = S, Se, Cl, and Br) heterojunctions under uniaxial tensile strains along the (a) x- and (b) y-directions. The hollow symbols indicate the peak points of the corresponding curves, representing the onset strain of mechanical instability.

The trend in critical strain further reveals distinct anisotropic mechanical responses. In the x-direction, –Br, –Se, and –Cl functionalization enhances the ductility, with the bare MoS2/Ti3C2 heterojunctions exhibiting a critical strain of 21% before nonlinear deformation initiates. Notably, –Se and –Cl functionalization contributes the most significant improvements with MoS2/Ti3C2Se2 and MoS2/Ti3C2Cl2 withstanding strains of up to 29%, reflecting their excellent mechanical ductility. This enhancement indicates the redistribution of electron density across the interface, where Se functionalization strengthens Ti–Se interactions. Moreover, –Br functionalization also improves ductility with a critical strain close to 26%, whereas –S functionalization leads to a lower critical strain of 18%.

In contrast, as illustrated in Fig. 4(b), all functionalization in the y-direction decreases the ductility of the MoS2/Ti3C2 heterojunction. The bare MoS2/Ti3C2 heterojunction shows a critical strain of approximately 23%, slightly higher than in the x-direction. –Br and –Cl functionalization reduces the critical strain to 21%, while –S and –Se functionalization further decreases the critical strain to 18% and 17%, respectively. This indicates that the critical strain along the y-direction is generally lower compared to the x-direction, suggesting that the heterojunction is more easily disrupted when subjected to strain along the y-direction. This disparity arises from the structural anisotropy, where the Ti–C bonds oriented along the y-direction exhibit reduced resistance to deformation.

Similar to MoS2-based heterojunctions, the comparison of black curves in Fig. 5(a) and (b) shows that the MoSe2/Ti3C2 heterojunctions exhibit nearly identical tensile strengths and critical strains in both the x- and y-directions, indicating a high degree of mechanical isotropy. However, MoSe2/Ti3C2X2 (X = S, Se, and Br) heterojunctions show higher tensile strengths and critical strains in the x-direction than in the y-direction, reflecting anisotropic mechanical behavior.


image file: d5cp01389e-f5.tif
Fig. 5 The stress–strain curves of MoSe2/Ti3C2 and MoSe2/Ti3C2X2 (X = S, Se, Cl, and Br) heterojunctions under uniaxial tensile strains along the (a) x- and (b) y-directions. The hollow symbols indicate the peak points of the corresponding curves, representing the onset strain of mechanical instability.

In the x-direction, the MoSe2/Ti3C2 heterojunction exhibits a higher tensile strength (34.01 GPa), with only a slight enhancement in the MoSe2/Ti3C2Se2 heterojunction (36.14 GPa). While in the y-direction, the bare MoSe2/Ti3C2 heterojunction shows the highest tensile strength (36.02 GPa), but functionalization with S, Se, Cl, and Br significantly lowers the tensile strength. Additionally, the tensile strength of the MoSe2/Ti3C2Se2 heterojunction in the y-direction is less than half of its value in the x-direction, indicating that the number of Ti–C bonds resisting tensile strain in the y-direction is significantly lower than that in the x-direction. Therefore, MoSe2/Ti3C2Se2 has the largest difference in ultimate strength between the two directions, displaying the most pronounced mechanical anisotropy. These observations are in agreement with the predictions for the mechanical performance of the TMDs/MXenes heterojunctions discussed in Sections 3.1 and 3.2.

It is found that the critical strain in the y-direction is significantly lower compared to the x-direction, indicating fewer Ti–C bonds resisting deformation along the y-direction. The bare MoSe2/Ti3C2 heterojunction exhibits the smallest critical strains in the x-direction. For MoSe2/Ti3C2Se2, the tensile strain in the x-direction reaches over 26% before yielding, while yielding occurs at only 11% tensile strain in the y-direction. A similar trend is observed for MoSe2/Ti3C2S2, where the yield limit in the x-direction is approximately double that in the y-direction. The introduction of –Br functionalization allows the heterojunction to withstand up to 28% critical strain in the x-direction, but this reduces to 20% in the y-direction. Overall, –Br, –S, Se, and Cl functional groups generally enhance the ductility of the MoSe2/Ti3C2 heterojunction in the x-direction, while only –Cl and –Br improve ductility in the y-direction.

To further quantify the anisotropic mechanical behavior of the investigated systems, Table 4 summarizes the critical strain and ultimate strength of pristine TMDs, Ti3C2X2 (X = S, Se, Cl, and Br), and their heterojunctions under uniaxial tensile strain. Monolayer MoS2 and MoSe2 exhibit higher strength and critical strain along the y-direction, reflecting their intrinsic anisotropy, while pristine Ti3C2 shows nearly isotropic behavior. After surface functionalization, all Ti3C2X2 systems exhibit reduced ultimate strength in the y-direction. However, the x-direction consistently shows higher strength and ductility for most functionalized systems.

Table 4 Summary of the critical strain (%) and ultimate strength (GPa) of pristine TMDs, Ti3C2, Ti3C2X2 (X = S, Se, Cl, and Br) and their heterojunctions under uniaxial tensile strains along the x- and y-directions
Formula x-direction y-direction
Critical strain (%) Ultimate strength (GPa) Critical strain (%) Ultimate strength (GPa)
MoS2 19 15.66 23 22.47
MoSe2 19 12.76 26 20.42
Ti3C2 24 51.11 20 49.19
Ti3C2S2 23 40.66 17 26.59
Ti3C2Se2 25 43.08 18 26.83
Ti3C2Cl2 25 43.58 13 30.10
Ti3C2Br2 30 39.99 18 29.60
MoS2/Ti3C2 21 34.33 23 31.81
MoS2/Ti3C2S2 18 31.04 18 26.95
MoS2/Ti3C2Se2 29 36.72 17 23.87
MoS2/Ti3C2Cl2 29 36.01 21 31.22
MoS2/Ti3C2Br2 26 33.57 21 28.78
MoSe2/Ti3C2 12 34.01 13 36.02
MoSe2/Ti3C2S2 24 32.34 14 23.99
MoSe2/Ti3C2Se2 26 36.14 11 18.51
MoSe2/Ti3C2Cl2 22 31.76 24 29.13
MoSe2/Ti3C2Br2 28 33.57 20 28.62


When TMDs are combined with Ti3C2 and Ti3C2X2 to form heterojunctions, a synergistic mechanical enhancement is observed. MoS2/Ti3C2Cl2 and MoS2/Ti3C2Se2 heterojunctions exhibit superior performance in the x-direction, with critical strains of up to 29% and ultimate strengths above 36 GPa. In the y-direction, MoSe2/Ti3C2 shows the highest ultimate strength (36.02 GPa) but with limited ductility. Among all studied heterojunctions, MoSe2/Ti3C2Se2 displays the greatest mechanical anisotropy, with significantly lower critical strain in the y-direction due to weakened interlayer bonding from –Se functionalization. Overall, –Br and –Cl terminations enhance strength and ductility along the x-direction, while –Se improves stretchability but at the cost of reduced strength, especially in the y-direction. These findings highlight the importance of surface functionalization and strain orientation in tuning mechanical properties for flexible electronics.

The strain energy (Es) inflection point marks the transition from the elastic to the plastic region, where irreversible structural changes begin to occur.61 In the linear elastic region, as strain energy increases, the structure retains its honeycomb configuration, allowing it to return to its original undeformed size upon the release of tension. The DFT-D2 scheme provides a computationally efficient correction for vdW interactions; however, it may introduce systematic errors in absolute energy values. Nevertheless, it remains relatively consistent in predicting energy trends and structural changes.62 Therefore, our focus is placed on the overall energy trends and critical inflection points, rather than on the absolute energy values. As shown in Fig. 6(a), the Es of all heterojunctions increases almost linearly with strains of up to around 30%. Notably, the MoS2/Ti3C2Br2 and MoS2/Ti3C2Cl2 heterojunctions exhibit the highest Es, indicating that these systems require more energy to maintain the given strain, which implies stronger bonding interactions. In contrast, the MoS2/Ti3C2Se2 heterojunction shows a sharp increase in Es, but requires the least amount of Es overall, suggesting better ductility but weaker mechanical strength. A similar trend is observed in Fig. 6(b), while the MoSe2 system displays slightly lower absolute Es values, which indicates that MoSe2-based heterojunctions are generally more flexible than their MoS2 counterparts. Moreover, –Br and –Cl functional groups terminated heterojunctions exhibit higher Es, indicating stronger bonds that may enhance Ti3C2-based heterojunctions through improved interfacial interactions. In contrast, –S and –Se functional groups weaken the structures, with the TMDs/Ti3C2Se2 heterojunctions exhibiting the highest ductility.


image file: d5cp01389e-f6.tif
Fig. 6 The strain energy (Es) of (a) MoS2/Ti3C2X2 and (b) MoSe2/Ti3C2X2 (X = S, Se, Cl, and Br) heterojunctions under uniaxial tensile strains. The blue/pink areas mark the plastic phase of the heterojunction with the highest Es.

Therefore, –Br and –Cl functionalization enhances the Es of the heterojunction, increasing its resistance to deformation and stiffness, which is primarily attributed to the strengthened interfacial interactions induced by their high electronegativity. –Br and –Cl functionalization induces significant charge redistribution when terminating the Ti3C2X2 surface, thereby strengthening the bonding between Ti and X (X = Br and Cl). This enhanced bonding increases the vdW interactions between the Ti3C2 and TMDs layers, resulting in improved interfacial stability. As a result, the TMDs layer experiences greater constraints, leading to increased stiffness in the heterojunction. Moreover, the stronger bonding between the Ti3C2X2 and TMDs layers effectively restrains interfacial sliding, which is particularly advantageous for flexible electronic applications, as they reduce the risk of delamination or peeling under sustained mechanical stress.

3.4 Microscopic mechanism

The bond length behavior at the atomic scale offers a clearer understanding of the stress–strain curves. Therefore, we further analyzed the bond lengths along the y-direction for both the bare MoS2/Ti3C2 and representative MoS2/Ti3C2Se2 heterojunctions. As shown in Fig. 7(a), the bond lengths of bonds (2) and (3), located in the middle of the Ti3C2 layer, are longer than those of bonds (1) and (4), which are positioned on the top and bottom surfaces. This variation is due to the higher bond energies of bonds (2) and (3), which make them more resistant to elongation. Additionally, interlayer interaction results in bond (1) on the top surface being slightly longer than bond (4) on the bottom surface.
image file: d5cp01389e-f7.tif
Fig. 7 The Ti3C2 bond length of (a) MoS2/Ti3C2 and (b) MoS2/Ti3C2Se2 heterojunctions under uniaxial tensile strains along the y-direction.

At early stages of tensile strain, the lengths of the Ti–C bonds increase linearly, reflecting the elastic deformation of the Ti3C2 layer. The synchronous elongation of bonds (2) and (3) with bonds (1) and (4) is due to uniform stretching of interatomic interactions under tensile load. However, the extension of these Ti–C bonds is limited by the effective radius of the outer electrons of Ti and C atoms, preventing the bonds from stretching indefinitely. As the tensile strain increases, bond lengths (1), (2), and (3) begin to rapidly decrease around 24% strain, indicating bond fracture, which corresponds directly to the material failure observed in Fig. 4(b). Meanwhile, the Mo–S bond within the MoS2 layer fractures once the tensile strain exceeds 23%, disrupting its interlayer interaction with the Ti–C bonds in the Ti3C2 layer. This results in a substantial decrease in the bond lengths of the already fractured Ti–C bonds (1), (2), and (3).

In contrast, as shown in Fig. 7(b), the bond lengths of (5)(6), (2)(3), and (1)(4) exhibit synchronous changes under strain due to the reduced electrical activation of Ti atoms by Se atoms on the Ti3C2Se2 surface. Both the (2)(3) and (1)(4) bonds increase almost linearly with strain, with the (2)(3) bonds displaying a steadier linear increase corresponding to the stress increase before the critical strain. As the strain approaches 17%, the sharp elongation of the (1)(4) bonds show the structural distortion, marking the material's critical point. Compared to the dramatic shortening of the Ti–C bond length at large strains in Fig. 7(a), the Ti–Se bonds exhibit less contraction, which suggests that the Se-functional group increases ductility of the MoS2/Ti3C2 heterojunction. The Ti–Se bonds located on the upper and lower surfaces exhibit longer bond lengths than the Ti–C bonds. Notably, the (5)(6) bond lengths remain unchanged under increasing tensile strain until stress yielding occurs, while the Ti–C bonds show more significant elongation compared to the Ti–Se bonds. This behavior suggests that Ti–C bonds are more resistant to tensile deformation than Ti–Se bonds.

Fig. 8(a) illustrates the variation in Ti–C bond lengths within the MoSe2/Ti3C2 heterojunctions under uniaxial tensile strain along the y-direction. Notably, the (2) and (3) bonds located in the middle of the Ti3C2 layer is approximately 0.15 Å longer than the (1) and (4) bonds located on the surface. Furthermore, the (1) bond near the MoSe2 layer and the (4) bond at the bottom of Ti3C2 exhibit nearly identical lengths, leading to overlapping curves in the bond length distribution. This overlap indicates that the vdW interactions between the MoSe2 and Ti3C2 layers are relatively weaker compared to the interlayer coupling. As the tensile strain increases to 15%, the (1) and (4) bonds elongate synchronously, corresponding to the structural failure of the bare MoSe2/Ti3C2 heterojunction, as shown in Fig. 5(a). As a result, the Ti1 atoms are further attracted to the MoSe2 layer, causing the elongation of the (1) bond. Meanwhile, the elongation of the (4) bond occurs as the MoSe2 layer shifts, which weakens the vdW attraction and initiating repulsive forces. This dynamic between bond elongation and vdW interactions is essential in determining the mechanical failure of the heterojunction at critical strain levels.


image file: d5cp01389e-f8.tif
Fig. 8 The Ti3C2 bond length of (a) MoSe2/Ti3C2 and (b) MoSe2/Ti3C2Se2 heterojunctions under uniaxial tensile strains along the y-direction.

As tensile strain is applied, the decoupling of electronic interactions between the MoSe2 and Ti3C2Se2 layers results in synchronous changes in the bond lengths of bonds (2)(3) and (1)(4), while the bond lengths of bonds (5) and (6) remain nearly constant before fracture, as shown in Fig. 8(b). As the tensile strain increases to 7%, the bond lengths of (2) and (3) in the middle of the Ti3C2Se layer reach 2.35 Å and 2.36 Å, respectively, at marking the beginning of failure. When tensile strain increases to 10%, the bond lengths of bonds (5) and (6) start to decrease, indicating that the Mo–Se bond in the upper MoSe2 layer has fractured, thereby reducing the tensile resistance of the Ti3C2Se2 layer. Continued tensile strain results in an increase in the bond lengths of bonds (5) and (6), indicating the complete failure of the Ti–Se bonds. The Ti–Se bonds exhibit greater ductility and elongate earlier than the Ti–C bonds, suggesting that the Se atoms are particularly sensitive to deformation. Consequently, the Ti–Se bonds in the MoSe2/Ti3C2Se2 heterojunction are the weakest in resisting tensile deformation, followed by the Ti–C bonds located in the middle of the Ti3C2Se2 layer, and lastly, the Ti–C bonds at the surface. This finding indicates that the presence of Ti–Se bonds reduces the ultimate strength of the MoSe2/Ti3C2 heterojunction.

We also compared the evolution of interlayer distance d in TMDs/Ti3C2 and TMDs/Ti3C2Se2 heterojunctions under uniaxial strain along the y-direction, as shown in Fig. 9. The variation in layer spacing reflects the interfacial interaction strength under strain and provides insight into strain-induced charge transfer. At lower strain levels, all heterostructures exhibit a slight increase in interlayer distance due to atomic rearrangement and weak vdW relaxation. With increasing strain, the Poisson effect leads to a significant reduction in interlayer distance, indicating enhanced interfacial charge coupling.


image file: d5cp01389e-f9.tif
Fig. 9 The interlayer distance d (Å) of MoS2/Ti3C2, MoS2/Ti3C2Se2, MoSe2/Ti3C2, and MoSe2/Ti3C2Se2 heterojunctions under uniaxial tensile strains along the y-direction.

Notably, MoSe2/Ti3C2Se2 exhibit the largest initial interlayer distance (2.96 Å) and the most pronounced reduction under critical strain, which suggests that Se-induced charge redistribution enhances flexibility but weakens interfacial adhesion. For MoS2/Ti3C2 and MoSe2/Ti3C2 heterojunctions, the relatively stable and moderate reduction in the interlayer distance under strain suggests that strong Ti–C and Mo–S (or Mo–Se) bonds effectively mediate interfacial stress transfer. In contrast, MoS2/Ti3C2Se2 and MoSe2/Ti3C2Se2 exhibit a larger initial d and more pronounced contraction, indicating that Se functionalization weakens Ti–X and Mo–Se bonding due to the reduced orbital overlap and enhanced charge delocalization. These results highlight the crucial role of interfacial bond strength in determining the mechanical robustness of TMDs/MXenes heterostructures under tensile deformation.

To clarify the atomic-level fracture mechanisms, Mulliken charge populations of key interfacial bonds were analyzed under both strain-free and critical strain conditions, as summarized in Tables S1–S4. In the MoS2/Ti3C2 heterojunction, Ti atoms exhibit a positive charge at the strain-free stage. Upon reaching the critical strain, a noticeable increase in the net charge on Ti and Mo atoms along with enhanced localization on S atoms indicates a charge redistribution from MoS2 to the Ti3C2. These results indicate a weakening of the Mo–S bond and enhanced interfacial charge coupling, which is consistent with the observed interlayer contraction and mechanical failure. A more pronounced charge redistribution is observed in the MoSe2/Ti3C2 heterojunction. Under the critical strain, the Mo atoms exhibit a significant increase in the positive charge, while the Ti atoms show a minimal variation, suggesting that charge transfer predominantly occurs within the TMDs layer. This internal charge redistribution from Mo to Se atoms, indicative of Mo–Se bond weakening, likely contributes to the pronounced mechanical anisotropy and the reduced critical strain observed along the y-direction.

In the MoSe2/Ti3C2Se2 heterojunction, the Ti atoms exhibit a notable increase in the positive charge, while the surface –Se functional groups show pronounced positive polarization. In contrast, the Mo atoms maintain the negative charge, indicating that electron transfer is predominantly directed from the MoSe2 layer toward the Ti3C2Se2 interface. This redistribution weakens the Ti–Se interfacial bonding; however, partial Mo–Se bond integrity is preserved due to the localized electron density on Mo atoms prior to mechanical failure. A comparable trend is observed in the MoS2/Ti3C2Se2 heterojunction, where charge polarization across Ti–X and Mo–S bonds under strain indicates progressive interfacial debonding. Overall, these findings highlight the critical role of interfacial electron redistribution in tuning the bond strength and fracture behavior.

4 Conclusions

In this study, we systematically investigated the structural and mechanical properties of TMDs/MXenes heterojunctions, focusing on the influence of functionalization and the orientation dependence. Our first-principles calculations demonstrate that the halogen (Cl and Br) and chalcogen (S and Se) surface functionalization leads to a significant reduction in in-plane stiffness and a notable increase in Poisson's ratio. Notably, MoSe2/Ti3C2Br2 and MoS2/Ti3C2Se2 exhibit the highest critical strain in the x-direction among all functionalized systems, which makes them promising candidates for flexible devices. Although the overall in-plane stiffness decreases compared to pristine Ti3C2, most functionalized heterojunctions retain higher elastic moduli than the constituent TMDs monolayers, indicating a synergistic strengthening effect within the heterojunctions.

While most TMDs/Ti3C2X2 heterojunctions exhibit nearly isotropic mechanical responses, MoSe2/Ti3C2Se2 displays marked anisotropy under large strain, primarily due to the larger atomic radius of Se that affects bonding energy and induces lattice distortion. Mechanistically, bond-level analysis highlights that Ti–C bonds bear the primary tensile load, while Ti–X bonds contribute little to this resistance. The higher bond energy of Ti–C bonds leads to their preferential breakage, which ultimately causes the strength limit of TMDs/Ti3C2X2 heterojunctions to be lower than that of TMDs/Ti3C2. Among the functionalized heterostructures, –Cl termination demonstrates the most balanced performance, combining strong interfacial charge coupling with enhanced mechanical flexibility. It maintains metallic character and supports efficient stress transfer under tensile strain, making it particularly suitable for flexible electronic applications. In contrast, –S and –Se terminations induce charge localization and weaken interfacial bonding, leading to diminished tensile strength and greater mechanical anisotropy. ELF analyses further confirm that functional group-induced charge redistribution influences interfacial bonding strength. These findings highlight the impact of functionalization on the mechanical properties of TMDs/MXenes heterojunctions, offering new insights for next-generation flexible devices, sensors, and thin-film batteries.

Author contributions

Yuqian Zhang: data curation, formal analysis, writing – original draft, and visualization. Zhiwei Liu: data curation, validation, and visualization. Siyu Zheng: data curation, methodology and visualization. Changyang Yu: formal analysis, writing – original draft and visualization. Siliang Yue: data curation, formal analysis and visualization. Chenliang Li: conceptualization, writing – review and editing, funding acquisition, software and supervision. Hui Qi: conceptualization, methodology and supervision.

Conflicts of interest

The authors declare that they have no conflicts of interest.

Data availability

The data supporting this article are described in detail in Section 2, Computational methods.

The supplementary information provides ELF plots, structural transformation diagrams, and detailed Mulliken charge population data for various TMDs/MXenes heterostructures under both strain-free and critical strain conditions. See DOI: https://doi.org/10.1039/d5cp01389e

Acknowledgements

This work was supported by the National Natural Science Foundation of China (12172097) and the Natural Science Foundation of Heilongjiang Province, China (LH2021A006).

References

  1. M. Dahlqvist, M. W. Barsoum and J. Rosen, Mater. Today, 2024, 72, 1–24 CrossRef.
  2. A. Zhou, Y. Liu, S. Li, X. Wang, G. Ying, Q. Xia and P. Zhang, J. Adv. Ceram., 2021, 10, 1194–1242 CrossRef CAS.
  3. M. Naguib, M. Kurtoglu, V. Presser, J. Lu, J. Niu, M. Heon, L. Hultman, Y. Gogotsi and M. W. Barsoum, Adv. Mater., 2011, 23, 4248–4253 CrossRef CAS PubMed.
  4. B. C. Wyatt, A. Rosenkranz and B. Anasori, Adv. Mater., 2021, 33, 2007973 CrossRef CAS PubMed.
  5. Y. Wei, P. Zhang, R. A. Soomro, Q. Zhu and B. Xu, Adv. Mater., 2021, 33, 2103148 CrossRef CAS PubMed.
  6. F. Shahzad, A. Iqbal, H. Kim and C. M. Koo, Adv. Mater., 2020, 32, 2002159 CrossRef CAS PubMed.
  7. D. Xu, Z. Li, L. Li and J. Wang, Adv. Funct. Mater., 2020, 30, 2000712 CrossRef CAS.
  8. Z. Fu, N. Wang, D. Legut, C. Si, Q. Zhang, S. Du, T. C. Germann, J. S. Francisco and R. Zhang, Chem. Rev., 2019, 119, 11980–12031 CrossRef CAS PubMed.
  9. A. Lipatov, H. Lu, M. Alhabeb, B. Anasori, A. Gruverman, Y. Gogotsi and A. Sinitskii, Sci. Adv., 2018, 4, eaat0491 CrossRef PubMed.
  10. H. Qi, S. L. Yue, C. L. Li, J. Guo, F. Q. Chu, Z. Wang and Y. Q. Zhang, Mater. Today Commun., 2022, 33, 104721 CrossRef CAS.
  11. J. Guo, Y. Zhang, S. Yue, C. Li and Z. Wang, J. Phys. Chem. Solids, 2023, 181, 111531 CrossRef CAS.
  12. G. R. Bhimanapati, Z. Lin, V. Meunier, Y. Jung, J. Cha, S. Das, D. Xiao, Y. Son, M. S. Strano, V. R. Cooper, L. Liang, S. G. Louie, E. Ringe, W. Zhou, S. S. Kim, R. R. Naik, B. G. Sumpter, H. Terrones, F. Xia, Y. Wang, J. Zhu, D. Akinwande, N. Alem, J. A. Schuller, R. E. Schaak, M. Terrones and J. A. Robinson, ACS Nano, 2015, 9, 11509–11539 CrossRef CAS PubMed.
  13. H. Qi, S. Yue, C. Li, J. Guo, Z. Fan, H. Wu and Z. Wang, Diamond Relat. Mater, 2024, 144, 110973 CrossRef CAS.
  14. C. Shi, M. Beidaghi, M. Naguib, O. Mashtalir, Y. Gogotsi and S. J. Billinge, Phys. Rev. Lett., 2014, 112, 125501 CrossRef PubMed.
  15. C. Zhang, X. Wang, W. Wei, X. Hu, Y. Wu, N. Lv, S. Dong and L. Shen, ChemElectroChem, 2021, 8, 3804–3826 CrossRef CAS.
  16. X. Hui, X. Ge, R. Zhao, Z. Li and L. Yin, Adv. Funct. Mater., 2020, 30, 2005190 CrossRef CAS.
  17. O. Mashtalir, M. Naguib, V. N. Mochalin, Y. Dall'Agnese, M. Heon, M. W. Barsoum and Y. Gogotsi, Nat. Commun., 2013, 4, 1716 CrossRef PubMed.
  18. M. Khazaei, M. Arai, T. Sasaki, C. Y. Chung, N. S. Venkataramanan, M. Estili, Y. Sakka and Y. Kawazoe, Adv. Funct. Mater., 2013, 23, 2185–2192 CrossRef CAS.
  19. S. J. Kim, H. J. Koh, C. E. Ren, O. Kwon, K. Maleski, S. Y. Cho, B. Anasori, C. K. Kim, Y. K. Choi, J. Kim, Y. Gogotsi and H. T. Jung, ACS Nano, 2018, 12, 986–993 CrossRef CAS PubMed.
  20. L. Zhou, Y. Zhang, Z. Zhuo, A. J. Neukirch and S. Tretiak, J. Phys. Chem. Lett., 2018, 9, 6915–6920 CrossRef CAS PubMed.
  21. W. Y. Chen, S. N. Lai, C. C. Yen, X. F. Jiang, D. Peroulis and L. A. Stanciu, ACS Nano, 2020, 14, 11490–11501 CrossRef CAS PubMed.
  22. Z. Chen, C. Wu, Y. Yuan, Z. Xie, T. Li, H. Huang, S. Li, J. Deng, H. Lin, Z. Shi, C. Li, Y. Hao, Y. Tang, Y. You, O. A. Al-Hartomy, S. Wageh, A. G. Al-Sehemi, R. Lu, L. Zhang, X. Lin, Y. He, G. Zhao, D. Li and H. Zhang, J. Nanobiotechnol., 2023, 21, 141 CrossRef CAS PubMed.
  23. Z. Chen, J. Li, T. Li, T. Fan, C. Meng, C. Li, J. Kang, L. Chai, Y. Hao, Y. Tang, O. A. Al-Hartomy, S. Wageh, A. G. Al-Sehemi, Z. Luo, J. Yu, Y. Shao, D. Li, S. Feng, W. J. Liu, Y. He, X. Ma, Z. Xie and H. Zhang, Natl. Sci. Rev., 2022, 9, nwac104 CrossRef CAS PubMed.
  24. Z. Chen, C. Meng, X. Wang, J. Chen, J. Deng, T. Fan, L. Wang, H. Lin, H. Huang, S. Li, S. Sun, J. Qu, D. Fan, X. Zhang, Y. Liu, Y. Shao and H. Zhang, Laser Photonics Rev., 2024, 18, 2400035 CrossRef CAS.
  25. Z. H. Fu, Q. F. Zhang, D. Legut, C. Si, T. C. Germann, T. Lookman, S. Y. Du, J. S. Francisco and R. F. Zhang, Phys. Rev. B: Condens. Matter Mater. Phys., 2016, 94, 104103 CrossRef.
  26. K. Luo, X.-H. Zha, Y. Zhou, Q. Huang, S. Zhou and S. Du, Int. J. Quantum Chem., 2020, 120, e26409 CrossRef CAS.
  27. V. Kamysbayev, A. S. Filatov, H. Hu, X. Rui, F. Lagunas, D. Wang, R. F. Klie and D. V. Talapin, Science, 2020, 369, 979–983 CrossRef CAS PubMed.
  28. D. Cao, P. Shen, H. Liu, J. Li, Z. Zhang and Y. Li, Int. J. Quantum Chem., 2023, 123, e27143 CrossRef CAS.
  29. L. Zhang, W. Zhang, X. Ma, X. Zhang and J. Wen, Phys. Chem. Chem. Phys., 2022, 24, 8913–8922 RSC.
  30. M. Mathew, P. V. Shinde, R. Samal and C. S. Rout, J. Mater. Sci., 2021, 56, 9575–9604 CrossRef CAS.
  31. S. Kim, H. Shin, J. Lee, C. Park, Y. Ahn, H.-J. Cho, S. Yuk, J. Kim, D. Lee and I.-D. Kim, ACS Nano, 2023, 17, 19387–19397 CrossRef CAS PubMed.
  32. K. Hagedorn, C. Forgacs, S. Collins and S. Maldonado, J. Phys. Chem. C, 2010, 114, 12010–12017 CrossRef CAS.
  33. S. Radhakrishnan, S. K. A. Raj, S. R. Kumar, P. Johari and C. S. Rout, Nanotechnology, 2021, 32, 155403 CrossRef CAS PubMed.
  34. J. Su, G.-D. Li, X.-H. Li and J.-S. Chen, Adv. Sci., 2019, 6, 1801702 CrossRef PubMed.
  35. Y. Li, Z. Yin, G. Ji, Z. Liang, Y. Xue, Y. Guo, J. Tian, X. Wang and H. Cui, Appl. Catal., B, 2019, 246, 12–20 CrossRef CAS.
  36. M. Chandran, A. Thomas, A. Raveendran, M. Vinoba and M. Bhagiyalakshmi, J. Energy Storage, 2020, 30, 101446 CrossRef.
  37. L. Tao, L. Huang, W. Qin, K. Pang, M. Zhang, L. Duan, Y. He, G. Zhu, C. Wen, C. Yu and H. Ji, ACS Appl. Nano Mater., 2023, 6, 21068–21078 CrossRef CAS.
  38. Y. Liu, W. Zhang, B. Lv, Y. Ge, R. Zhang, B. Wang, Z. Chen, Q. Zhang and S. Sang, Surf. Interfaces, 2022, 30, 101823 CrossRef CAS.
  39. C. Li, K. Lv, X. Ding, L. Feng, X. Lv and D. Ma, Surf. Interfaces, 2024, 44, 103661 CrossRef.
  40. S. J. Clark, M. D. Segall, C. J. Pickard, P. J. Hasnip, M. J. Probert, K. Refson and M. C. Payne, Z. Kristallogr., 2005, 220, 567–570 CAS.
  41. R. Van Leeuwen and E. Baerends, Phys. Rev. A: At., Mol., Opt. Phys., 1994, 49, 2421 CrossRef CAS PubMed.
  42. E. R. McNellis, J. Meyer and K. Reuter, Phys. Rev. B: Condens. Matter Mater. Phys., 2009, 80, 205414 CrossRef.
  43. G. R. Berdiyorov, AIP Adv., 2016, 6, 055105 CrossRef.
  44. P. H. Nha, C. V. Nguyen, N. N. Hieu, H. V. Phuc and C. Q. Nguyen, Nanoscale Adv., 2024, 6, 1193–1201 RSC.
  45. G. Kresse and D. Joubert, Phys. Rev. B: Condens. Matter Mater. Phys., 1999, 59, 1758 CrossRef CAS.
  46. S. Froyen, Phys. Rev. B: Condens. Matter Mater. Phys., 1989, 39, 3168 CrossRef PubMed.
  47. J. A. Thomas and H. Fischer, J. Phys. Chem., 1992, 96, 9768–9774 CrossRef.
  48. Y.-X. Yu, Phys. Chem. Chem. Phys., 2013, 15, 16819–16827 RSC.
  49. B. Li, H. Guo, Y. Wang, W. Zhang, Q. Zhang, L. Chen, X. Fan, W. Zhang, Y. Li and W.-M. Lau, npj Comput. Mater., 2019, 5, 16 CrossRef.
  50. L. Zeng, D. Wu, J. Jie, X. Ren, X. Hu, S. P. Lau, Y. Chai and Y. H. Tsang, Adv. Mater., 2020, 32, 2004412 CrossRef PubMed.
  51. M. Hajhosseini and Z. Zeinalizadeh, Acta Mech. Sin., 2023, 39, 522463 CrossRef.
  52. R. Zhou, H. Tong, Q. Xu, R. Zhang and G. Hao, Micro Nanostruct., 2024, 192, 207880 CrossRef CAS.
  53. N. R. Hemanth, T. Kim, B. Kim, A. H. Jadhav, K. Lee and N. K. Chaudhari, Mater. Chem. Front., 2021, 5, 3298–3321 RSC.
  54. S. Zheng, C. Li, C. Wang, D. Ma and B. Wang, Nanomaterials, 2023, 13, 1218 CrossRef CAS PubMed.
  55. A. Scemama, P. Chaquin and M. Caffarel, J. Chem. Phys., 2004, 121, 1725–1735 CrossRef CAS PubMed.
  56. D. Wu, S. Wang, S. Zhang, J. Yuan, B. Yang and H. Chen, Phys. Chem. Chem. Phys., 2018, 20, 18924–18930 RSC.
  57. D. Çakır, F. M. Peeters and C. Sevik, Appl. Phys. Lett., 2014, 104, 203110 CrossRef.
  58. Q. Tang, Z. Zhou and P. Shen, J. Am. Chem. Soc., 2012, 134, 16909–16916 CrossRef CAS PubMed.
  59. Z. Guo, J. Zhou, C. Si and Z. Sun, Phys. Chem. Chem. Phys., 2015, 17, 15348–15354 RSC.
  60. J.-W. Jiang and H. S. Park, Appl. Phys. Lett., 2014, 105, 033108 CrossRef.
  61. M. Topsakal, S. Cahangirov and S. Ciraci, Appl. Phys. Lett., 2010, 96, 091912 CrossRef.
  62. S. Grimme, J. Antony, S. Ehrlich and H. Krieg, J. Chem. Phys., 2010, 132, 154104 CrossRef PubMed.

This journal is © the Owner Societies 2025
Click here to see how this site uses Cookies. View our privacy policy here.