Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Impact of composition on the structural, electronic, and mechanical properties of M3C2T2 MXenes

Emily Sutherlanda, Benjamin Traversob and N. Aaron Deskins*b
aDepartment of Physics, Worcester Polytechnic Institute, Worcester, Massachusets, USA. E-mail: egsutherland@wpi.edu
bDepartment of Chemical Engineering, Worcester Polytechnic Institute, Worcester, Massachusets, USA. E-mail: nadeskins@wpi.edu

Received 8th August 2025 , Accepted 12th August 2025

First published on 15th August 2025


Abstract

MXenes are a family of layered 2D materials useful for a wide variety of applications. Their properties can be fine-tuned by choice of chemical composition (metal and termination), but the vast majority of published studies have focused on Ti-based MXenes with –O, –F, and –OH terminations. Furthermore, MXenes may have ABC or ABA stacking, but typical density functional theory (DFT) studies assume only ABC stacking. Thus, most modeling papers of MXenes have focused only on specific targeted MXenes. In this work we aimed to provide a comprehensive DFT study of possible MXenes in order to motivate and characterize MXenes beyond those common in the literature. We modeled 99 different M3C2T2 MXenes (including group 4, 5, and 6 metals; also including halogen, chalcogen, –OH, and –NH terminations). We made no assumptions about preferred termination site or stacking symmetry of these MXenes. 20% of the studied MXenes were found to prefer ABA stacking. In total we performed more than 2000 DFT calculations to predict the structural, electronic, and mechanical properties of these MXenes. We identified several MXenes with exceptional properties, and identified potential applications of such MXenes. We also connected the termination/metal choice to trends in their properties. Our work highlights how different properties of MXenes can be tuned based on their composition, and thus motivates further work on these materials.


1 Introduction

In recent years MXenes have attracted much attention due to their potential tunability and unique properties.1–5 They have been studied for many applications, including energy storage,6–11 photocatalysis,12–15 electrocatalysis,16–20 electronic devices,21–23 sensing,24–27 and others. MXenes are two-dimensional materials that have the chemical formula Mn+1XnTx, where M is an early transition metal, X is C or N, and T is a surface termination. The number of layers (n) is typically 1 to 4.28 The termination can have a profound role on the MXene properties,29,30 for instance making it a metal or semiconductor depending on the termination,31–34 or changing mechanical properties.35–37 Other properties, such as work function38–42 can also change depending on the constituent components (M, X, or T). Understanding and predicting MXene properties based on their composition and structure is a critical part of advancing applications of these materials.

Most commonly, MXenes are synthesized from MAX phase materials where the A layer is removed via aqueous acids (e.g., HF), resulting in –F, –O, and/or –OH terminations.43,44 Still, other terminations on MXene surfaces are possible.29,30,45–47 For instance, electrochemical etching or using Lewis acidic molten salts may produce –Cl, –Br, –I, –P, or –Sb terminations.48–52 Furthermore, through post-processing of MXenes it is possible to remove terminations (producing bare MXenes) or substitute existing terminations with a new termination, such as –S, –Se, –Te, or –NH.49,53 Despite efforts to make MXenes having various compositions, it has been estimated that around 80% of published MXene papers have focused on Ti-based MXenes.54,55 Even papers with a theoretical component are also predominantly focused on Ti-based MXenes.55 Furthermore, density functional theory (DFT) studies often tend to focus on n = 1 MXenes.31,34,35,56–59 Ignoring Ti3C2T2 and Ti2CT2 (which are the most commonly studied MXenes), papers with a theoretical component and focused on n = 1 MXenes are around 20 times more common than papers with a theoretical component and focused on n = 2 MXenes.55 Accordingly, more data and understanding of MXenes beyond those commonly studied (e.g., metals other than Ti and n > 1) is needed in order to achieve the goal of fully tuning MXenes for targeted applications.

There have been several DFT studies performed examining the impact of different surface terminations on the properties of MXenes. Much of this work however has focused on –O, –OH, or –F terminations for various metals.11,31,35,40,58–66 O-terminated MXenes generally have higher mechanical strength35,60 and are more likely to be semiconductors28,35,58 compared to those with –F or –OH terminations. Terminating a MXene with OH groups tends to result in an ultralow work function.38–40,60,65 While there are studies which model halide (F, Br, Cl, I) or chalcogen (O, S, Se, Te) terminated Ti2CTx and Ti3C2Tx MXenes,42,67–71 studies modeling terminations other than –O, –F, and –OH for MXenes with metals other than Ti- are limited.40,71–78 In addition, many MXenes have ABC stacking,79,80 but it is possible that the atomic layers may have ABA stacking.81,82 Theory predicts ABA stacking for certain MXenes.34,66,80,83–85 Despite this, the vast majority of DFT studies on MXenes assume ABC stacking, or ignore the possibility of ABA stacking. Thus far, DFT studies comparing ABA and ABC stacking have been reported for bare MXenes34,80,84,85 and for those with –O,34,66,75,80,84 –F,66,75,84 –OH,66 –S,75 –Cl,75 or –P83 terminations. Thus, there is a need for more simulation data on non-Ti MXenes, as well as identifying the right stacking (in order, for instance, to correctly model these MXenes) for MXenes with other termination and/or metal combinations.

In this work we modeled 99 carbide MXenes (i.e., M3C2T2), including 9 metals (groups 4, 5, and 6) and 10 possible terminations (halogens, chalcogens, OH, and NH). Our goal was to provide a comprehensive analysis of how metal and termination choice affects the structural stability and other properties of the MXenes. The literature on DFT modeling of MXenes has largely focused on a small subset of MXenes (typically Ti-based; or M2CTx MXenes, n = 1), or a small subset of properties. Many literature papers have also assumed ABC stacking for MXenes. Our work represents the most comprehensive DFT study of M3C2T2 (n = 2) MXenes involving a wide range of metals and terminations, and this many structural, electronic, and mechanical properties. Fig. 1 highlights the MXenes that we modeled, as well as the properties we calculated. In order to determine the correct stable structures, we modeled both ABC and ABA stacking for each MXene. We also considered four possible surface sites for terminations. We then report several properties for each of the most stable configurations, including electronic and mechanical properties. By identifying, examining, and analyzing the observed trends, we indicate how MXene properties can be controlled through their chemical composition, and which MXenes have promising properties.


image file: d5ma00874c-f1.tif
Fig. 1 Illustration of the MXenes modeled in this work, and the various calculated properties. This work contains the most comprehensive DFT study to date which has modeled M3C2T2 (n = 2) MXenes and the listed properties.

2 Methodology

We performed DFT calculations using the Vienna ab initio Simulation Package (VASP).86–89 The exchange correlation energy was calculated using the generalized gradient approximation (GGA) functional by Perdew, Burke, an Ernzerhof (PBE).90 Core electrons were described by projector augmented wave (PAW) pseudopotentials.91,92 A cutoff energy of 450 eV was used for valence electron wavefunctions. As recommended by the VASP developers, we utilized the following number of valence electrons for each metal: Ti (12), Zr (12), Hf (10), V (13), Nb (13), Ta (11), Cr (12), Mo (14), and W (14). The Brillouin zone was sampled with a Γ-centered 8 × 8 × 1 k-point mesh. Calculation convergence criteria for the energies and forces were set to 10−5 eV and 0.002 eV Å−1, respectively. We used the Gaussian smearing method with a width 0.25 eV, which was found to minimize roughness in density of states (DOS) plots while retaining principle features. One goal of examining the electronic structure was to identify any semiconductors (MXenes are predominantly metallic). The PBE functional is known to underestimate band gaps, so electronic structures were also calculated using the HSE06 functional.93 HSE06 provides a more accurate way to verify whether a MXene is a semiconductor. Since we were primarily interested in identifying semiconductors with the HSE06 functional, for these calculations we used a smearing width of 0.05 eV, as recommended by the VASP developers for semiconductors. Identifying band gaps from DOS plots can be challenging, especially since smearing in the DOS increases overlap of electronic states and band gaps will depend on the smearing width. Identifying band gaps is also ambiguous from DOS plots since band edges are arbitrarily defined in the plots. All reported band gaps are therefore taken from the electronic eigenvalues, rather than DOS plots.

Post-processing of DOS data was performed using the VASPKIT package.94 Bader charge analysis was performed using code from the Henkelman group.95 Two-dimensional elastic properties were calculated using the energy-strain method as implemented in VASPKIT.94,96 In calculating the elastic properties, we considered strain along the x and y axis to range from −0.015 to 0.015 ε in steps of 0.005 ε. We also calculated MXene d-band centers as the first moment of the projected DOS for the metal atom d-bands.97,98 Due to the small thickness of the MXenes, we included the d-bands of all metal atoms in our calculations, rather than considering surface metal atoms only. Previous literature99–101 integrated the DOS from −∞ to a finite positive value in order to capture all relevant d-band orbitals. Similarly, based on inspection of our DOS, we integrated the d-bands from −∞ to 2 eV. All structural visualizations were carried out using VESTA.102

Correlation coefficients provide a simple way to investigate relationships between MXene properties and properties of their constituent atoms. Therefore, we calculated Pearson correlation coefficients between various MXene properties and elemental properties (as reported in the Mendeleev103 database). Correlation coefficients were found using the Orange Data Mining program.104 A more extensive investigation of these relationships using statistical methods will be the focus of future work.

We modeled monolayer M3C2 and M3C2T2 MXenes with M = Ti, Zr, Hf, V, Nb, Ta, Cr, Mo, and W (groups 4, 5 and 6), and T = O, S, Se, Te, F, Cl, Br, I, OH, and NH (including chalcogens and halogens). All MXenes were modeled at full coverage, which is expected to have higher thermodynamic stability compared to lower coverages.31 To avoid interactions between neighboring monolayers, our calculations included a cell length of 22 Å in the z-direction, which gave vacuum distances of at least ≈12 Å for terminated surfaces. We modeled both trigonal (ABC) and hexagonal (ABA) stacking symmetry (shown in Fig. 2) for all the MXenes. Additionally, there are multiple high-symmetry sites where terminal groups may attach to the MXene surface. Previous literature identified several possible stable termination sites.11,31,35,58–60,78 These include hollow sites above metal atoms (hM), hollow sites above no atoms (h), hollow sites above X layer atoms (hX), mixed sites with hM or h on one surface and hX on the other (m), and directly on top of metal atoms (t). Fig. 2 shows side views of the different termination sites along with stacking configurations. In total we modeled 8 configurations for each terminated MXene (four termination sites and two stacking arrangements). Side and top views of all the different configurations are shown in Fig. S1 and S2.


image file: d5ma00874c-f2.tif
Fig. 2 Different MXene configurations modeled in this work. Blue spheres represent metal atoms, brown spheres represent carbon atoms, and red spheres represent terminating atoms. MXenes were modeled with trigonal (ABC) stacking (a)–(e) or hexagonal (ABA) stacking (f)–(j). ABC terminations are designated as follows: hollow sites above metal atoms (hM), hollow sites above carbon (hX), mixed termination sites with one surface hM and one surface hX (m), and top sites (t). ABA terminations are designated as follows: hollow sites between metal and carbon (h), hollow sites above carbon (hX), mixed termination sites with one surface h and one surface hX (m), and top sites (t). Top views of the structures can be found in Fig. S1 and S2 in the SI.

3 Results and discussion

3.1 Structural properties

3.1.1 Preferred stacking and termination sites. We first identified the most stable structures for each MXene, especially to understand how the composition affects the preferred MXene geometry. In Fig. 3, and in (Tables S1, S2 and Fig. S3, S4), we report the most stable structures and relative energy values comparing different stacking configurations and sites for each termination. Our results showed that, while the majority of M3C2 and M3C2T2 MXenes displayed an energetic preference for ABC stacking, some MXenes were more stable with ABA stacking. In total, 20% of the modeled MXenes were more stable with ABA stacking, which is not trivial. A statistical summary of the most stable configurations and terminations is found in Fig. S3.
image file: d5ma00874c-f3.tif
Fig. 3 Indication of the most stable stacking and termination sites for each MXene, as determined by calculated energies. MXenes which prefer ABC stacking are shown in blue shades, while MXenes which prefer ABA stacking are shown in green shades. 20% of the studied MXenes preferred ABA stacking.

We report energy differences between ABA and ABC stacking configurations (ΔEABA–ABC = EABAEABC) for each bare MXene in Table S1. We found that all bare MXenes containing group 4 metals (Ti, Zr, Hf) had a strong energetic preference for ABC stacking, with an average ΔEABA–ABC value of 2.72 eV. We also found that bare MXenes with group 5 metals (V, Nb, Ta) had an energetic preference for ABC stacking, with an average ΔEABA–ABC value of 1.85 eV. On the other hand, bare MXenes with group 6 metals (Cr, Mo, W) preferred ABA stacking, with negative ΔEABA–ABC values for Mo3C2 (−0.30 eV) and W3C2 (−1.13 eV). Bare Cr3C2 had a ΔEABA–ABC of 0.11 eV, which indicated that it may be stable with either ABC or ABA stacking. Gouveia et al.80 also found similar results for bare MXenes.

The most stable structures from the eight considered configurations (see Fig. 2) are reported in Fig. 3 and Fig. S4 for each M3C2T2 MXene. We found that all terminated MXenes with group 4 metals showed an energetic preference for ABC stacking, regardless of termination, and also all preferred the ABC-hM configuration. Similarly, all terminated MXenes with group 5 metals also showed an energetic preference for ABC stacking, with the majority of MXenes preferring ABC-hM. However, several MXenes preferred the ABC-hX configuration over ABC-hM (e.g., V3C2T2 with –F and –OH terminations; Nb3C2T2/Ta3C2T2 with –F, –Cl, –Br, and –OH terminations; Nb3C2 (NH)2). In the ABC-hX configuration, C and T atoms are in alignment for increased steric and/or Coulombic repulsion, while such repulsion is minimized with the ABC-hM configuration. This is typically cited as the reason ABC-hM is preferred over ABC-hX.11,34,70,105

Several MXenes involving group 6 metals were more stable with ABA stacking. Furthermore, several MXenes with group 6 metals had a near degeneracy between ABC and ABA configurations (see Fig. S4) which is in contrast to the distinct preference for ABC stacking shown by MXenes with group 4 or 5 metals. For Cr-based MXenes, ABC stacking is the dominant configuration, but some terminations induce ABA stacking (e.g., O, S, NH). A shift occurs with Mo- and W-based MXenes, where an even larger number of MXenes prefer ABA stacking. All but the largest terminations (e.g., –Br, –I) led to ABA stacking being preferred over ABC stacking. Gouveia et al. also found that M3C2O2 MXenes with group 6 metals prefer ABA stacking, but did not model terminations other than –O.80

It has been shown in previous work on other layered materials, that the electron count of the material may be predictive of the energetically preferred structure.66,106,107 Electron count is defined as the number of valence d electrons remaining on each metal atom after charge transfer to non-metal atoms.107 For example, Ti3C2F2 has three Ti atoms (each with four valence d electrons), two C atoms (each requiring four electrons to fill their outer shell), and two F atoms (each requiring one electron to fill their outer shell). The electron count per metal atom for Ti3C2F2 is then [(3)(+4) + (2)(−4) + (2)(−1)]/3 = 0.6 e per M. Electron count is given for each bare MXene in Table S1 and for each atomic terminated MXene in Table S2.

We found that MXenes with group 6 metals had the highest electron counts (3.33 e per M for bare; ≥ 2 e per M for terminated) and were the only metals which had any preference for ABA over ABC stacking. This supports the argument by Gouveia et al.80 that higher electron density (for instance as caused by more d electrons) leads to stronger ABA preference. For bare MXenes, ΔEABA–ABC values followed this trend for the group 6 metals: Cr > Mo > W. Correspondingly, the total number of electrons follows this trend: Cr < Mo < W, which further confirms the idea that more electron density may favor ABA stacking.

For terminated MXenes, we also found a relationship between the electron count and stacking/termination preferences. In addition to electron count, the relative sizes of atoms in a crystal are known to be a determining factor for structural preference in materials.106,108 For example, the Goldschmidt tolerance factor and the ionic filling fraction can be used to predict the structure of perovskites and other materials.109 We define VM/T as the ratio of atomic volume of the metal atom to atomic volume of the termination atom. Values of VM/T are provided in Table S2, utilizing atomic volume data from the Mendeleev database.103 We summarized our findings on the preferred stacking and termination sites in Fig. 4. This shows how the electron count and relative sizes of the metal to termination are key to the preferred stacking and termination site. We also indicate in this graphic the local geometry of the termination/metal sites, as exhibited in Fig. S5. Those MXenes with smaller electron count and/or smaller VM/T ratios tend to prefer ABC stacking (e.g., group 4 or 5 metals; group 6 metals with large terminations). On the other hand those MXenes with an electron count of ≥2 e per M and large VM/T ratios (e.g., group 6 metals with O or S) preferred ABA stacking. Thus, we have uncovered an important relationship for predicting the energetically preferred configuration of MXenes.


image file: d5ma00874c-f4.tif
Fig. 4 A summary of our findings on the preferred stacking, preferred termination, and preferred local geometry, as they relate to electron count and VM/T.
3.1.2 Termination binding energies. We next predicted the binding energy (EB) of the terminations, which is the energy to bind the terminations to bare MXenes. These values are useful for comparing the stability of different terminations, since one can directly compare how much energy is released (or required) to terminate a bare MXene with a given termination. The difference in binding energies of two different terminations provides the energy released or required to exchange the terminations (i.e., EB-termination1EB-termination2 < 0 implies termination 1 is more stable on the surface compared to termination 2). For atomic terminated MXenes, we calculated the binding energy according to eqn (1)
 
EB = EM3C2T2EM3C2ET2. (1)
Here EM3C2T2, EM3C2, and ET2 are, respectively, the energies of the most stable terminated MXene, the bare MXene, and the gas-phase dimer of the terminating element.31,72,110,111 The Supplementary Information describes how binding energies of –OH and –NH terminations were calculated. The calculated binding energies are given in Fig. 5 and Table S3.

image file: d5ma00874c-f5.tif
Fig. 5 Termination binding energies (eV) for terminated MXenes in their most stable configurations. Green lines are chalcogen terminated, blue lines are halogen terminated, and black lines are –NH and –OH terminated.

Our results showed that the most stable terminations were either –O or –F, having average binding energies of −9.26 and −9.32 eV, respectively. Other terminations were notably less stable compared to –O and –F. For instance, the average binding energy for –S (third most stable termination) was 3.15 eV less stable than –O and 3.21 eV less stable than –F. Generally, chalcogen terminations were more stable than halogen terminations (with the exception of –F), and the binding energies within a group weakened as the termination valence electrons occupied larger shells (i.e., moving down the periodic table). Our results showed that metals with a higher number of valence d electrons tended to have weaker binding energies (e.g., EB-Cr3C2T2-avg > EB-V3C2T2-avg > EB-Ti3C2T2-avg). We also found that within each group, binding energies tended to become more stable as the valence shell of the metal increased (e.g., EB-Ti3C2T2-avg > EB-Zr3C2T2-avg > EB-Hf3C2T2-avg). Notably, MXenes with molecular terminations, –NH and –OH, typically had weaker binding energies than the atomic terminations. Since the –F and –O terminations are so stable, terminating MXenes with other species may require non-aqueous synthesis methods, such as using molten inorganic salts.48–51 Indeed, all the considered terminations have been successfully incorporated into MXenes,49,51,112 albeit thus far only with a few select metals (e.g., Ti, Nb, Ta).

3.1.3 Lattice constants. We report, in Fig. 6 and Table S4, calculated lattice constants for bare MXenes. The lattice constant values for bare MXene structures ranged from 2.87 Å (Mo3C2 and W3C2) to 3.35 Å (Zr3C2), with an average value of 3.08 Å. We found that, with the exception of V3C2, the lattice constants of bare MXenes decreased as the number of valence d electrons increased across a given period (e.g., aZr3C2 > aNb3C2 > aMo3C2). This trend was also found in previous studies of M3C2,60,80 and in a study of M2XS2.73 Yang et al. attributed the decreased lattice constant to a decrease in length of metal bonds as the number of valence d electrons increased.73 Notably, atomic radii decreases as the number of d electrons increases within a period, and this trend follows the same pattern in our lattice constants. In fact, we found a strong correlation (+0.83) between our predicted lattice constant for bare MXenes and the atomic radii of the metal atoms. The same study by Yang et al.73 further noted that the lattice constants for those MXenes with 4d transition metals are larger than for those with 3d transition metals, due to an increase in ionic radius. They found that the lattice constants of MXenes with 4d transition metals were also larger than those with 5d transition metals, which was attributed to a reduction in metal bond lengths due to scalar-relativistic effects. We also observed that the calculated lattice constant values for bare M3C2 MXenes were largest for structures with 4d transition metals, as compared to the lattice constants of the 3d and 5d transition metals in the same group (e.g., aZr3C2 > aHf3C2 > aTi3C2).
image file: d5ma00874c-f6.tif
Fig. 6 Lattice constants (Å) for MXenes in their most stable configurations. Red circles are bare structures, green squares have terminations from chalcogens, blue triangles have terminations from halogens, while black diamonds are –OH and –NH terminations. Hollow markers indicate an ABA configuration, while solid markers indicate an ABC configuration.

The calculated lattice constant values for all configurations of the terminated MXenes are found in Table S5. Our discussion, however, is limited to the lattice constant values for the most stable configurations of the terminated MXenes, which are shown in Fig. 6. These lattice constant values ranged from 2.68 Å (Cr3C2O2) to 3.55 Å (Zr3C2I2), with an average lattice constant value of 3.17 Å. We expect the size of the terminated MXenes to be influenced by both the metal and termination. Especially when the size of both the termination and metal are larger, it is expected to lead to larger lattice constants. Indeed, our data showed that the lattice constant of atomic-terminated MXenes were moderately correlated with the atomic radii of both metal and termination atoms, with correlation coefficients of, respectively, +0.65 and +0.56.

In most cases, chalcogen terminations resulted in smaller lattice constants compared to the halogen terminations with the same valence shell (e.g., aM3C2S2-Avg < aM3C2Cl2-Avg). This corresponded with our finding in Section 3.1.2 that chalcogen terminations tended to result in stronger binding energies compared to the halogen terminations with the same valence shell. Other work60,67 has also reported larger lattice constants for halogen terminations compared to their corresponding chalcogen terminations. Additionally, we noted that the average lattice constant of terminated structures decreased as the number of valence d electrons of the metal increased (e.g., aZr3C2T2-avg > aNb3C2T2-avg > aMo3C2T2-avg), which was also found to be the case for bare MXenes. Those MXenes with –NH and –OH terminations have lattice constants similar to those with –O and –F, respectively, likely due to electronic similarities.31,60,113,114

For these MXenes that show a preference for ABA stacking, our calculated ABA lattice parameter values were 0.7–5.9% smaller than our calculated values for the corresponding termination site with ABC stacking (see Table S5). This is in agreement with previous studies which found that ABA stacking leads to lattice parameters which are reduced by up to 8% compared to ABC stacking.34,66,80 This reduction was attributed to reduced steric and/or coulombic repulsion between metal atoms in ABA stacking compared to ABC stacking.34

3.2 Electronic properties

3.2.1 Atomic charges. Atomic charges provide insight on electron transfer between the components of the MXenes, including the nature of bonding between atoms, and the oxidation state of the individual elements. We report Bader charges of the different atoms in Tables S7–S9. Fig. S6 also provides the charges of surface metal atoms and the terminations for the most stable MXenes.

Unsurprisingly, all metals were positive (cationic) in the MXenes, while the terminations were negative (anionic). The charges of surface layer metal atoms in bare MXenes ranged from 0.60 (Mo3C2) to 1.20 (Hf3C2), with an average value of 0.97. On the other hand, surface metals for terminated MXenes ranged from 0.94 (W3C2Te2) to 2.23 (Zr3C2O2), with an average value of 1.56. The terminations thus induce the metals to become more oxidized. Termination charges on the other hand had smaller absolute values, and ranged from −1.26 (Zr3C2I2) to −0.16 (W3C2O2), with an average value of −0.67. Average charges of –NH (QT-M3C2(NH)2-Avg = −0.98) and –OH (QT-M3C2(OH)2-Avg = −0.75) were comparable to, respectively, –O and –F terminations (QT-M3C2O2-Avg = −1.08, QT-M3C2F2-Avg = −0.73). This observation may be explained by electronic similarities between the respective atomic (–O/–F) and molecular terminations (–NH/–OH).31,60,113,114 Our data showed that more stable terminations were more anionic. We found a correlation coefficient of +0.85 between termination Bader charge and termination binding energy for the atomic species, confirming the relationship between charge and stability. Thus, we conclude that more oxidation/reduction results when stronger interactions occur between the metal and termination.

Furthermore, we found a relationship between Bader charge values and atomic electronegativity. Our predicted decrease in metal charges corresponded to increased electronegativity of the metal atom. For bare MXenes, there was a correlation coefficient of −0.90 between the Bader charges of surface metal atoms and electronegativity of the metal species. For terminated MXenes, the correlation coefficient was −0.69. Thus, metal atoms with larger electronegativity are less likely to be oxidized since they tend to attract electrons. In contrast, the termination charges became more negative with increased electronegativity of the termination atom, since atoms with larger electronegativity will attract more electrons and become more negative. The correlation coefficient between termination electronegativity and termination Bader charge was −0.60.

3.2.2 Work function. All work function values are reported in Fig. 7 and Table S10. The work function for the bare structures ranged from 4.11 eV (Zr3C2) to 4.92 eV (W3C2), with an average value of 4.53 eV. This was a very narrow range despite the variety of metals in the MXenes, indicating that it may be difficult to tune the work function of bare MXenes. On the other hand, terminated MXenes had a much wider range, from 1.67 eV (Cr3C2(OH)2) to 7.80 eV (Cr3C2O2). Generally halogen terminations had lower work functions than chalcogen terminations, as the average work function for halogen terminated MXenes (4.27 eV) was found to be nearly 1 eV lower than the average work function of chalcogen terminated MXenes (5.25 eV). Particularly, –O terminations resulted in the largest work functions for all MXenes with group 5 and 6 metals (ϕM3C2O2-avg being 5.75 and 7.46 eV with, respectively, group 5 and 6 metals), while –S terminations produced the largest work functions for those with group 4 metals (ϕM3C2S2-avg was 6.33 eV with group 4 metals). On the other hand, –Te and –I terminated MXenes resulted in the lowest average work functions (ϕM3C2Te2-avg = 4.09 eV, ϕM3C2I2-avg = 3.61 eV) for atomic terminations.
image file: d5ma00874c-f7.tif
Fig. 7 Work functions (eV) for MXenes in their most stable configurations. Red lines are bare MXenes, green lines are chalcogen terminated, blue lines are halogen terminated, and black lines are OH and NH terminated.

Work functions for MXenes terminated by OH groups, however, ranged from 1.67 eV (Cr3C2(OH)2) to 3.05 eV (W3C2(OH)2), with an average value of 2.34 eV. These results for –OH agree with literature, where it has been well established that –OH terminated MXenes display ultralow work functions.38–40,60,65 The work function values of –NH terminated MXenes ranged from 2.28 eV (Ta3C2(NH)2) to 3.31 eV (Nb3C2(NH)2), and had an average value of 2.82 eV, again being ultralow work functions. Thus, termination selection is crucial for engineering MXene work function while choice of metal may be less important.

Because the work function is such a critical electronic property, there have been a number of attempts to explain the work functions of materials, including dependence on various material properties.34,38,39,42,65,67,68,115–121 Although these studies report connections between work function and various properties, quantitative analysis connecting these properties to the work function is generally lacking. Thus, we calculated the correlation coefficients between the MXene work function and various atomic properties, as well as between the work function and other MXene properties (see Table S11). Atomic properties of the metal and terminating atoms were taken from the Mendeleev database.103 MXene properties (e.g., termination binding energy, M–T bond distances, Bader charges, p-band center) were from our calculations, and are reported in various tables in the SI.

The highest correlations occurred with termination atomic volume (−0.74), termination p-band center (+0.70), termination specific heat capacity (+0.67), M–T bond distance (−0.66), termination Mendeleev number (−0.65), termination atomic number (−0.64), termination atomic weight (−0.63), termination double covalent radius (−0.63), termination period number (−0.63), and termination triple covalent radius (−0.61). These correlations indicate that the termination, rather than the metal, is key to the MXene work functions. It is worth noting that a number of these properties, namely atomic volume, specific heat capacity, M–T bond distance, atomic weight, period number, and covalent radii, are all interrelated. In particular, we found that each of these properties were highly correlated with the atomic number of the corresponding termination atom (correlation coefficients being ≥|0.78|).

3.2.3 Conducting nature. We calculated the density of states and band gaps to establish the conducting nature of the MXenes (e.g., semiconducting or metallic). The PBE functional is widely utilized for many calculations, but is well known to underestimate band gaps,122–124 so we also calculated the DOS and band gap values using the HSE06 functional. We must acknowledge that the HSE06 functional may not be accurate for describing properties of metals125–127 and, in some cases, has been found to overestimate the band gap.122,123,127–129

Our results for MXene electronic structure using the PBE functional indicated that all 99 M3C2/M3C2T2 MXenes were metallic. DOS plots calculated using the PBE functional can be found in Fig. S11–S19. When using the HSE06 functional, however, we predicted 21 MXenes to be narrow-gap semiconductors. Band gap values (calculated with HSE06) are given in Table 1, and the corresponding DOS plots are given in Fig. S8–S10. In accordance with the definition of semiconductors presented by Sheng S. Li,130 we considered only those with Eg > 0.1 eV to be semiconducting. These MXenes included three Ti-based MXenes, Cr3C2, Zr3C2(NH)2, six Nb-based MXenes, Mo3C2, two Hf-based MXenes, three Ta-based MXenes, and four W-based MXenes. Every termination, except Te, when combined with the right metal, formed a semiconductor MXene. The band gap values for these structures ranged from 0.12 eV (Nb3C2(OH)2) to 0.65 eV (Nb3C2F2). Literature has largely focused on predicting band gaps for M2CTx MXenes,33,42,131–136 leaving questions about M3C2Tx MXenes. Indeed, we are not aware of previous electronic structure calculations using the HSE06 functional for most MXenes we predicted (n = 2 MXenes) to be semiconducting. In the case of MXenes which have previous electronic structure calculations using the HSE06 functional, there is disagreement throughout literature. Hf3C2O2 has been found to be a semiconductor with reported band gaps of 0.16 eV,137 (similar to our value of 0.18 eV), 0.43 eV,138 and 0.95 eV.139 We report a band gap of 0.14 eV for Ti3C2O2, while other papers140 found it to be metallic. Literature also predicts M2CO2 (M = Hf, Zr, Ti) MXenes to be semiconductors,33 while we also predict several O-terminated MXenes to be semiconductors.

Table 1 Band gap values (Eg) for MXenes which were identified as semiconductors using the HSE06 functional. Band gap values are given in units of eV. Corresponding DOS plots for each semiconducting MXene are given in Fig. S8–S10
MXene Eg MXene Eg
Ti3C2O2 0.14 Mo3C2 0.23
Ti3C2I2 0.15 Hf3C2O2 0.18
Ti3C2(NH)2 0.15 Hf3C2(NH)2 0.30
Cr3C2 0.48 Ta3C2Se2 0.24
Zr3C2(NH)2 0.26 Ta3C2Cl2 0.54
Nb3C2S2 0.24 Ta3C2Br2 0.47
Nb3C2Se2 0.18 W3C2 0.20
Nb3C2F2 0.65 W3C2O2 0.32
Nb3C2Cl2 0.51 W3C2S2 0.19
Nb3C2Br2 0.47 W3C2(NH)2 0.25
Nb3C2(OH)2 0.12    


Despite DFT predictions of semiconductors, experimental studies identifying semiconductor MXenes are limited.132,141,142 The discrepancy between theoretical predictions and experimental results regarding electronic properties of MXenes has been attributed to defects143 and mixed terminations132 which may be introduced during synthesis. Thus, overcoming current synthesis challenges may be the key to experimental realization of semiconducting MXenes. Our work may further provide direction on what compositions may be worth pursuing to create semiconductor MXenes.

Electrical conductivity is another important property of MXenes, especially for applications such as catalysis, spintronics, batteries, LEDs, photodetectors, and others. However, calculating electrical conductivity can be difficult since it depends on many factors, such as charge carrier concentration (which in turn depends on donor state concentration) and charge mobility. Rather, we examined the DOS values at the Fermi level, N(EF), which has been used as a proxy for electrical conductivity in metal MXenes.140,144–147 Larger DOS at the EF should in principle correspond to larger carrier concentration since more free electrons would be available. We report, in Fig. S7 and Table S12, N(EF) values as calculated with the PBE functional. We do not report N(EF) values using the HSE06 functional, since this functional may have low accuracy for predicting properties of metals.125–127 The N(EF) values spanned 0.49 states per eV (Hf3C2O2) to 6.08 states per eV (Cr3C2Te2). The metals Ti, Zr, and Hf had some of the lowest values of N(EF), indicating potentially low electrical conductivity for these MXenes. This is noteworthy, considering that M2CO2 with M = Ti, Zr, and Hf are the MXenes most frequently predicted to be semiconducting.31,33–35,58,59,148 The highest values of N(EF) were found to be with Cr- and V-based MXenes, having average N(EF) values of 1.33 and 1.22 states per eV respectively, compared to an overall average value for all MXenes of 0.87 states per eV. This suggests that such MXenes may have larger electrical conductivity.

The termination had an important impact on the N(EF) values, with strong interplay between the metal and termination. For example, we calculated a correlation coefficient between the electron count (which depends on the metal and termination; see Section 3.1.1) and N(EF) values to be +0.46, indicating at least a modest correlation between the two. Furthermore, we found that N(EF) values of terminated MXenes were correlated more strongly (−0.64) with the atomic volume ratio, VM/T, than with any elemental properties of the metal or termination atom alone. For MXenes terminated by chalcogens, at least for groups 4 and 6, the N(EF) values tended to increase with increasing atomic number of the chalcogen termination. Particularly, –Te terminations resulted in some of the MXenes with the highest N(EF) values. On the other hand, group 5 metals showed the opposite trend where N(EF) values tended to decrease with increasing atomic number of the chalcogen termination. Halogen terminations gave low N(EF) values with group 4 metals, larger N(EF) values with group 5 metals that tended to increase with termination atomic number, and high N(EF) values with group 6 metals. The –OH and –NH terminations yielded moderate N(EF) values with group 4 metals, relatively high values with group 5 metals, and both low (NH) and high (OH) N(EF) values with group 6 metals. Interestingly, we found that the lowest values of N(EF) for MXenes with group 6 metals were those with the ABA-hX configuration (i.e., Cr3C2T2 with T = O, S, and NH, Mo3C2T2 with T = O, S, Se, and NH, and W3C2T2 with T = O, S, Se, Te and OH). Thus, our results show that there are strong interactions between the metal and termination that may affect the conductivity, and these interactions are complex.

3.2.4 d-band center. The d-band center (εd) of a metal has been established as a good descriptor for predicting adsorption energies and surface reactivity.97,98,149 The d-band center of MXenes has been linked to their performance as supercapacitors,150 gas sensors,151 and catalysts.152–154 We report, in Fig. 8 and Table S13, the d-band center of the MXenes in their most stable configurations. The d-band centers of bare MXenes decreased (i.e., shifted further from EF) as the number of valence d electrons increased (e.g., εd-Ti3C2 > εd-V3C2 > εd-Cr3C2), similar to what was reported by Hammer and Nørskov.149 A similar trend is that d-band centers become more negative with larger valence shells (e.g., εd-Ti3C2 > εd-Zr3C2 > εd-Hf3C2).
image file: d5ma00874c-f8.tif
Fig. 8 Calculated d-band center values (eV) of MXenes in their most stable configurations. Energies are relative to the Fermi energy (0 eV). Results are using the PBE functional. Red lines are bare MXenes, green lines are chalcogen terminated, blue lines are halogen terminated, and black lines are OH and NH terminated.

Our calculated values of the d-band center for terminated MXenes ranged from −3.49 eV (W3C2O2) to −0.61 eV (Ti3C2I2). The d-band center trends of terminated MXenes were similar to the trends of bare MXenes. Regardless of termination species, the d-band center of M3C2T2 MXenes decreased as the metal valence shell increased (e.g., εd-Ti3C2T2 > εd-Zr3C2T2 > εd-Hf3C2T2). In general, the d-band center of terminated MXenes also decreased as the number of metal valence d electrons increased (e.g., εd-Zr3C2T2-Avg > εd-Nb3C2T2-Avg > εd-Mo3C2T2-Avg).

Indeed, we found a correlation of −0.78 between metal atomic number and our calculated values of metal d-band centers for terminated MXenes, indicating the strong dependence of the d-band center on the metal choice. In addition to the strong effect of the metal on d-band center, the termination also affected the d-band center significantly. For example, Hf3C2O2 had a d-band center of −3.48 eV, while Hf3C2I2 had a d-band center of −1.71 eV, a difference of 1.77 eV. Generally chalcogen-terminated MXenes had more negative d-band centers compared to halogen-terminated MXenes within the same row of the periodic table. We also found that –NH/–O and –OH/–F terminations resulted in similar values of εd, each pair having electronic similarities.31,60,113,114

3.3 Mechanical properties

Materials with desirable mechanical properties are key for many applications, such as biomedical devices, strain sensors, flexible electronics, microfluidic devices, and composite materials. We calculated 2D mechanical properties, including elastic constants (C11 and C12), Young's modulus (E), Poisson's ratio (ν), shear modulus (G), and bulk modulus (K). Calculated mechanical properties are reported in Table S15. All MXenes were found to be mechanically stable according to the elastic stability criteria (C11 > 0 and C11 > |C12|).96,155

Young's modulus functions as a measure of material limitations, influencing the operating conditions which will result in deformation, as well as typifying material stiffness. It should be noted that we found the elastic constant C11, shear modulus, and bulk modulus (Table S15) to be highly correlated to Youngs modulus (correlation coefficient ≥0.90). Elastic constant C12 and Poisson's ratio were also correlated with each other (0.74). We therefore focused our discussion on Youngs modulus and Poisson's ratio, since these two are representative mechanical properties.

3.3.1 Young's modulus. We show in Fig. 9 and Table S15 calculated Young's modulus values. We found that Young's modulus values ranged from 93 N m−1 (Ti3C2Te2) to 535 N m−1 (W3C2O2), with an average value of 308 N m−1. For a given valence shell, we found that average Young's modulus values were larger for chalcogen terminated MXenes as compared to those with the corresponding halogen terminations (e.g., EM3C2O2-avg > EM3C2F2-avg). Also, MXenes with chalcogen terminations typically showed greater variability in Young's modulus values compared to those with halogen terminations. The range of Young's modulus values for chalcogen-terminated MXenes was 442 N m−1, with a standard deviation of 113 N m−1, while the range for halogen-terminated MXenes was 364 N m−1 with a standard deviation of 89 N m−1.
image file: d5ma00874c-f9.tif
Fig. 9 Young's modulus (N m−1) for MXenes in their most stable configurations. Red circles are bare structures, green squares have terminations from chalcogens, blue triangles have terminations from halogens, while black diamonds are –OH and –NH terminations. Hollow markers indicate an ABA configuration, while solid markers indicate an ABC configuration.

MXenes with –O and –F terminations produced the highest values of Youngs modulus for atomic-terminated structures, with average values of 414 N m−1 and 339 N m−1, respectively. In comparison, the 2D Young's modulus for graphene is around 340 N m−1.67,156 MXenes with –NH terminations also had large Young's modulus values, having an average of 400 N m−1. These results can be explained, at least in part, by considering our calculated Bader charges in Section 3.2.1 (Fig. S6 and Tables S7, S8), which showed that –O and –NH terminations resulted in the highest M–T charge transfer (i.e., large positive charges on the metal and large negative charges on the termination). Increased charge transfer corresponds to stronger bonds and, therefore, to stiffer materials. This is also corroborated by the large binding energies of –O.

We also considered the impact of the metal on MXene stiffness. On average, Zr-based MXenes had the lowest values of Young's modulus (EZr3C2T2-avg = 241 N m−1), while W-based MXenes had the highest values (EW3C2T2-avg = 426 N m−1). Our results showed that metals with more valence d electrons tended to have larger Youngs modulus values (e.g., EZr3C2T2-avg < ENb3C2T2-avg < EMo3C2T2-avg). Within a given metal group, Youngs modulus values tended to increase as the valence shell increased (e.g., EV3C2T2-avg < ENb3C2T2-avg < ETa3C2T2-avg).

It is well known that Youngs modulus is affected by the crystal structure, and MXenes are no exception. Our results showed that ABA stacking resulted in larger Young's modulus values compared to ABC stacking. We found that Youngs modulus for MXenes with ABC stacking ranged from 93 N m−1 (Ti3C2Te2) to 404 N m−1 (Ta3C2Cl2), with an average value of 271 N m−1, while the Youngs modulus for MXenes with ABA stacking (shown with hollow markers in Fig. 9) ranged from 338 N m−1 (Mo3C2) to 535 N m−1 (W3C2O2), with an average value of 455 N m−1. A comparison between ABC-hM and ABC-hX indicated the ABC-hX configuration resulted in some of the highest values of Youngs modulus. For those MXenes which preferred the ABC-hX configuration (e.g., M3C2T2 with M = V, Nb, Ta and T = F, OH) Young's modulus values ranged from 357 N m−1 (Nb3C2Br2) to 404 N m−1 (Ta3C2Cl2), with an average value of 380 N m−1. Kazemi et al. also reported larger Youngs modulus values for ABC-hX configurations compared to ABC-hM.68

3.3.2 Poisson's ratio. Poisson's ratio (ν) is an indicator of material flexibility. Specifically, it is a measure of the deformation perpendicular to an applied force. We found the Poisson's ratio values, reported in Table S15, to have a range from −0.40 (W3C2I2) to 0.75 (Ti3C2Te2), with an average value of 0.26. A box plot of the values for bare and terminated MXenes is shown in Fig. S20. In fact, although the range of values was 1.15, the standard deviation was only 0.15, indicating most values were close to the average of 0.26. Indeed, the 25th percentile was 0.20, while the 75th percentile was 0.29. Fig. S20 shows the distribution of values, and indicates that, other than a few outliers, there is very little variation in our data, regardless of metal or termination. Some MXenes are rather unique, and are discussed in the following paragraph.

We found eight MXenes with Poisson's ratio values >0.5, which is typically indicative of instability under stress.157,158 These included M3C2Te2 with M = Ti, Zr, Hf, and Cr, and M3C2 (OH)2 with M = Ti, and Zr. On the other extreme, we found three MXenes with a negative Poisson's ratio (NPR). This occurs when a material expands, rather than contracts, perpendicular to applied strain (i.e., it becomes wider when stretched). This phenomena is rare, but has been predicted for a number of 2D materials.159–163 We found NPR values for Cr3C2 (−0.15), W3C2Br2 (−0.02), and W3C2I2 (−0.40). These materials are therefore expected to exhibit enhanced toughness and vibrational damping compared to a MXene with a positive Poisson's ratio.162,163 Our results are corroborated by experimental studies which reported NPR values for some MXenes with group 6 metals.159,160 For example, Wu et al.159 reported a NPR of −0.43 for W2C and Mortazavi et al.160 reported an NPR of −0.15 for Mo2C.

3.4 Summarizing the composition–property relationships

Throughout this work, we calculated various structural, electronic, and mechanical properties of M3C2/M3C2T2 MXenes. Identifying how these properties are influenced by metal choice and termination species is important for optimizing MXenes for targeted applications. In Fig. 10, we summarize the trends in how the various properties depend on the group and period of the metals and terminations. For each property, arrows point in the direction of increasing values. For example, it can be seen that termination binding energies, lattice constants, and d-band centers all increase by selecting a halogen termination, rather than a chalcogen termination. On the other hand, work function and Young's modulus values would tend to increase for chalcogens compared to halogens. Similar relationships are shown for how termination period affects different properties (e.g., termination binding energy increases with increased period or atomic number). Fig. 10 also shows how metal period and group affects the various calculated properties. Our work indicates which metal and termination choices should be selected to obtain desired properties.
image file: d5ma00874c-f10.tif
Fig. 10 A summary of our findings regarding the general impact of metal and termination species on the preference for ABA stacking (a), termination binding energy (b), lattice constant (c), the magnitude of Bader charges for termination and surface layer metal atoms (d), work function (e), DOS at Fermi energy (f), metal d-band center (g), and Young's modulus (h). Arrows point in the direction of generally increasing values. Arrows are only shown for discernible trends, as a clear relationship was not always present (e.g., lattice constant and metal period). Poisson's ratio is not shown, as we found no distinct trends in the data.

3.5 Notable MXenes

In addition to identifying the general relationships between MXene properties and their composition, we also identified some promising MXenes for various applications. For example, when combined with –O terminations, group 6 metals resulted in ϕ > 7 eV (ϕCr3C2O2 = 7.80 eV, ϕMo3C2O2 = 7.44 eV, ϕW3C2O2 = 7.15 eV), which are considered ultra-high work functions.164 Ultra-high work functions may be desirable for applications involving electron capture and storage, such as in LEDs,165–167 transistors,168–170 and electron capture in solar cells.171–173 Several MXenes were identified as having low work functions (<3 eV), predominantly those terminated with NH or OH groups. These include all the metals (except W) with –OH terminations. These also include M = V, Cr, Mo, Ta, and W with –NH terminations. MXenes with ultra-low values may be ideal for facilitating charge transport in applications including electron emission devices,174–176 hydrogen production,177–179 and thermionic energy conversion.65,180,181

Furthermore, we identified several potential semiconductors, as presented in Table 1. These MXenes may be useful in electronic, photocatalysis, sensor, and thermoelectric applications.32,182,183 As discussed in Section 3.2.3, several MXenes were identified as having low conductivity (e.g., Ti3C2O2, Zr3C2O2, and Hf3C2O2), while other MXenes were identified as having high conductivity (e.g., Cr3C2Te2 and Mo3C2Br2). High conductivity may be desirable for energy storage, sensing applications, electronic devices, thermoelectrics, and other applications.22,183 Finally, we note that higher d-band center values may lead to better catalytic reactivity,154 increased capacitance in supercapacitors,150 or improved gas sensing ability.151 Metals with heavier halogens (such as I and Br) tended to have the highest d-band centers, especially when paired with Ti, V, or Cr.

Identifying MXenes with desirable mechanical properties may also be important. High stiffness for –O terminated MXenes, as compared to other terminations, has been widely reported in literature.35,67,68,70,78,184 In addition to –O, we found that –NH terminated MXenes also exhibited superior stiffness compared to other terminations, with an average Youngs modulus value of 400 N m−1. The highest Young's modulus values occurred with Mo- and W-based MXenes (e.g., 535 N m−1 for W3C2O2, 533 N m−1 for W3C2 (NH)2, and 505 N m−1 for Mo3C2O2), which incidentally preferred ABA stacking. These MXenes may be useful for applications where high stiffness is desired, such as in protective coatings,185–187 biomedical applications,188–190 or composite materials.191–193 On the other hand, some MXenes had low Young's modulus values (≈100 N m−1), such as Ti3C2Te2, Cr3C2, and Mo3C2Br2. Due to our predicted combination of low Young's modulus and high electrical conductivity for those three MXenes, they may be of particular interest in flexible, wearable sensors.194–196

We noted in Section 3.3.2 that most MXenes had Poisson's ratio values near 0.26. Several MXenes (M3C2Te2 with M = Ti, Zr, Hf, and Cr, and M3C2 (OH)2 with M = Ti, and Zr) had high values (>0.5), indicating their instability under stress. A few MXenes (Cr3C2, W3C2Br2, W3C2I2) had negative Poisson's ratios, which may make them potentially suitable for use in sensors, biomedical devices, clothing, and other applications.162,197

As noted here and in Sections 3.2 and 3.3, several MXenes with non-traditional terminations (–O, –OH, –F) have interesting properties. For instance, –Te and –I terminated MXenes have a unique combination of properties, including large lattice constants (i.e., large surface to volume ratio), low work function, high electrical conductivity, high d-band center values, and low Young's modulus, which make them strong candidates for a variety of applications. However, these terminations have the lowest binding energies. Therefore, the continued development of synthesis methods which overcome thermodynamics issues, and produce MXenes with heavy terminations may be beneficial to the field. For example, molten salt etching,49 electrochemical etching,51 and chemical vapor deposition198 have all been shown to provide significant control over the resulting termination species and coverage.

4 Conclusions

In this study, we modeled 99 different M3C2/M3C2T2 MXenes in order investigate the impact of chemical composition on their structural, electronic, and mechanical properties. We considered an extended compositional space compared to most previous MXene studies. This included M = Ti, V, Cr, Zr, Nb, Mo, Hf, Ta, and W with T = bare, O, S, Se, Te, F, Cl, Br, I, NH, and OH. We first predicted the most stable configuration for each MXene (i.e., stacking and termination site), which required over 700 simulations. Our calculations showed that most MXenes had an energetic preference for ABC stacking, but some group 6 metals preferred ABA stacking (20% of the studied MXenes). We found that stacking and termination site preference were related to electron count and metal/termination atomic volume ratio. The MXene lattice constants were found to be highly dependent on the atomic configurations, and was found to be proportional to metal and termination atom sizes. We also calculated termination binding energies and found that the common –O and –F terminations were significantly more stable than other terminations, presenting a challenge for synthesizing MXenes with other terminations.

We determined a number other properties, including electronic and mechanical properties. Several properties strongly depended on the metal and termination, and the trends for such properties were established (see Fig. 10). We identified several notable MXenes, with interesting calculated properties. For example, MXenes terminated with –NH and –OH generally had ultra low work functions, while M3C2O2 MXenes with M = Cr, Mo, W had ultra high work functions. We found the termination had a much greater effect than metal choice on the work function. Using the HSE06 functional, we classified several MXenes as narrow-gap semiconductors (see Table 1). Every termination could form a semiconducting MXene when paired with the right metal, with the exception of –Te. Some MXenes (typically halogen terminations with period 4 metals) had higher d-band centers, which may be useful for catalysis, energy storage, or gas sensing. Our calculations of mechanical properties showed that –O and –NH terminations resulted in very stiff MXenes. ABA stacking also led to stiffer MXenes, compared to ABC stacking. Most Poisson's ratio values were near the average of 0.26, but three MXenes had a negative Poisson's ratio. To summarize, we have shown how termination and metal choices affect various properties of MXenes. We identified the trends of such properties, and our work provides guidance and motivation to synthesize and study MXenes beyond typical MXenes, e.g., beyond Ti3C2O2 or Ti3C2F2.

Author contributions

ES performed the DFT simulations, analyzed the results, and contributed to the writing of the paper. BT performed DFT simulations. NAD oversaw the project and contributed to the writing.

Conflicts of interest

There are no conflicts to declare.

Data availability

Data from the various tables and figures for this article have been included in the SI. See DOI: https://doi.org/10.1039/d5ma00874c

Simulation files, including geometries, can be found at https://github.com/Deskins-group/Structure-Files/tree/master/MXene-structure-properties.

Acknowledgements

This research was supported in part by the National Science Foundation under grant NRT-HDR-2021871. This research was performed using computational resources supported by the Academic & Research Computing group at Worcester Polytechnic Institute.

References

  1. Y. Gogotsi and B. Anasori, ACS Nano, 2019, 13, 8491–8494 CrossRef CAS PubMed.
  2. M. Sokol, V. Natu, S. Kota and M. W. Barsoum, Trends Chem., 2019, 1, 210–223 CrossRef CAS.
  3. L. Verger, V. Natu, M. Carey and M. W. Barsoum, Trends Chem., 2019, 1, 656–669 CrossRef CAS.
  4. X. Jiang, A. V. Kuklin, A. Baev, Y. Ge, H. Ågren, H. Zhang and P. N. Prasad, Phys. Rep., 2020, 848, 1–58 CrossRef CAS.
  5. A. VahidMohammadi, J. Rosen and Y. Gogotsi, Science, 2021, 372, eabf1581 CrossRef CAS PubMed.
  6. B. Anasori, M. R. Lukatskaya and Y. Gogotsi, Nat. Rev. Mater., 2017, 2, 1–17 Search PubMed.
  7. S. Sun, C. Liao, A. M. Hafez, H. Zhu and S. Wu, Chem. Eng. J., 2018, 338, 27–45 CrossRef CAS.
  8. J. Pang, R. G. Mendes, A. Bachmatiuk, L. Zhao, H. Q. Ta, T. Gemming, H. Liu, Z. Liu and M. H. Rummeli, Chem. Soc. Rev., 2019, 48, 72–133 RSC.
  9. W. Meng, X. Liu, H. Song, Y. Xie, X. Shi, M. Dargusch, Z.-G. Chen, Z. Tang and S. Lu, Nano Today, 2021, 40, 101273 CrossRef CAS.
  10. P. Lokhande, A. Pakdel, H. Pathan, D. Kumar, D.-V. N. Vo, A. Al-Gheethi, A. Sharma, S. Goel, P. P. Singh and B.-K. Lee, Chemosphere, 2022, 297, 134225 CrossRef CAS PubMed.
  11. Q. Tang, Z. Zhou and P. Shen, J. Am. Chem. Soc., 2012, 134, 16909–16916 CrossRef CAS PubMed.
  12. P. Kuang, J. Low, B. Cheng, J. Yu and J. Fan, J. Mater. Sci. Technol., 2020, 56, 18–44 CrossRef CAS.
  13. X. Xie and N. Zhang, Adv. Funct. Mater., 2020, 30, 2002528 CrossRef CAS.
  14. Y. Zhao, M. Que, J. Chen and C. Yang, J. Mater. Chem. C, 2020, 8, 16258–16281 RSC.
  15. J. K. Im, E. J. Sohn, S. Kim, M. Jang, A. Son, K. D. Zoh and Y. Yoon, Chemosphere, 2021, 270, 129478 CrossRef CAS PubMed.
  16. A. Liu, X. Liang, X. Ren, W. Guan, M. Gao, Y. Yang, Q. Yang, L. Gao, Y. Li and T. Ma, Adv. Funct. Mater., 2020, 30, 2003437 CrossRef CAS.
  17. H. Wang and J.-M. Lee, J. Mater. Chem. A, 2020, 8, 10604–10624 RSC.
  18. Y. Wang, Y. Nian, A. N. Biswas, W. Li, Y. Han and J. G. Chen, Adv. Energy Mater., 2021, 11, 2002967 CrossRef CAS.
  19. Z. Kang, M. A. Khan, Y. Gong, R. Javed, Y. Xu, D. Ye, H. Zhao and J. Zhang, J. Mater. Chem. A, 2021, 9, 6089–6108 RSC.
  20. L.-N. Shi, L.-T. Cui, Y.-R. Ji, Y. Xie, Y.-R. Zhu and T.-F. Yi, Coord. Chem. Rev., 2022, 469, 214668 CrossRef CAS.
  21. H. Kim, Z. Wang and H. N. Alshareef, Nano Energy, 2019, 60, 179–197 CrossRef CAS.
  22. F. Shahzad, A. Iqbal, H. Kim and C. M. Koo, Adv. Mater., 2020, 32, 2002159 CrossRef CAS PubMed.
  23. A. Ahmed, S. Sharma, B. Adak, M. M. Hossain, A. M. LaChance, S. Mukhopadhyay and L. Sun, InfoMat, 2022, 4, e12295 CrossRef CAS.
  24. A. Sinha, Dhanjai, H. Zhao, Y. Huang, X. Lu, J. Chen and R. Jain, TrAC, Trends Anal. Chem., 2018, 105, 424–435 CrossRef CAS.
  25. K. Deshmukh, T. Koväřïk and S. P. Khadheer, Coord. Chem. Rev., 2020, 424, 213514 CrossRef CAS.
  26. D. H. Ho, Y. Y. Choi, S. B. Jo, J.-M. Myoung and J. H. Cho, Adv. Mater., 2021, 33, 2005846 CrossRef CAS PubMed.
  27. Y. Pei, X. Zhang, Z. Hui, J. Zhou, X. Huang, G. Sun and W. Huang, ACS Nano, 2021, 15, 3996–4017 CrossRef CAS PubMed.
  28. Y. Gogotsi and Q. Huang, ACS Nano, 2021, 15, 5775–5780 CrossRef CAS PubMed.
  29. M. Mozafari and M. Soroush, Mater. Adv., 2021, 2, 7277–7307 RSC.
  30. T. Su, X. Ma, J. Tong, H. Ji, Z. Qin and Z. Wu, J. Mater. Chem. A, 2022, 10, 10265–10296 RSC.
  31. 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.
  32. X.-H. Zha, Q. Huang, J. He, H. He, J. Zhai, J. S. Francisco and S. Du, Sci. Rep., 2016, 6, 27971 CrossRef PubMed.
  33. Y. Zhang, W. Xia, Y. Wu and P. Zhang, Nanoscale, 2019, 11, 3993–4000 RSC.
  34. D. Ontiveros, F. Viñes and C. Sousa, J. Mater. Chem. A, 2023, 11, 13754 RSC.
  35. X. H. Zha, K. Luo, Q. Li, Q. Huang, J. He, X. Wen and S. Du, EPL, 2015, 111, 26007 CrossRef.
  36. Z. Guo, J. Zhou, C. Si and Z. Sun, Phys. Chem. Chem. Phys., 2015, 17, 15348–15354 RSC.
  37. Y. Bai, K. Zhou, N. Srikanth, J. H. Pang, X. He and R. Wang, RSC Adv., 2016, 6, 35731–35739 RSC.
  38. M. Khazaei, M. Arai, T. Sasaki, A. Ranjbar, Y. Liang and S. Yunoki, Phys. Rev. B: Condens. Matter Mater. Phys., 2015, 92, 075411 CrossRef.
  39. Y. Liu, H. Xiao and W. A. I. Goddard, J. Am. Chem. Soc., 2016, 138, 15853–15856 CrossRef CAS PubMed.
  40. N. M. Caffrey, Nanoscale, 2018, 10, 13520–13530 RSC.
  41. S. W. Koh, L. Rekhi, Arramel, M. D. Birowosuto, Q. T. Trinh, J. Ge, W. Yu, A. T. S. Wee, T. S. Choksi and H. Li, ACS Appl. Mater. Interfaces, 2024, 16, 66826–66836 CrossRef CAS PubMed.
  42. K. Yusupov, J. Björk and J. Rosen, Nanoscale Adv., 2023, 5, 3976–3984 RSC.
  43. M. Naguib, M. W. Barsoum and Y. Gogotsi, Adv. Mater., 2021, 33, 2103393 CrossRef CAS PubMed.
  44. K. R. G. Lim, M. Shekhirev, B. C. Wyatt, B. Anasori, Y. Gogotsi and Z. W. Seh, Nat. Synth., 2022, 1, 601–614 CrossRef CAS.
  45. H. Tang, R. Wang, L. Shi, E. Sheremet, R. D. Rodriguez and J. Sun, Chem. Eng. J., 2021, 425, 131472 CrossRef CAS.
  46. T. Zhang, L. Chang and X. Xiao, Small Methods, 2023, 2201530 CrossRef CAS PubMed.
  47. H. Shi, P. Zhang, Z. Liu, S. Park, M. R. Lohe, Y. Wu, A. Shaygan Nia, S. Yang and X. Feng, Angew. Chem., Int. Ed., 2021, 60, 8689–8693 CrossRef CAS PubMed.
  48. M. Li, J. Lu, K. Luo, Y. Li, K. Chang, K. Chen, J. Zhou, J. Rosen, L. Hultman, P. Eklund, P. O. A. Persson, S. Du, Z. Chai, Z. Huang and Q. Huang, J. Am. Chem. Soc., 2019, 141, 4730–4737 CrossRef CAS PubMed.
  49. 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.
  50. T. Zhang, L. Chang, X. Zhang, H. Wan, N. Liu, L. Zhou and X. Xiao, Nat. Commun., 2022, 13, 6731 CrossRef CAS PubMed.
  51. H. Ding, Y. Li, M. Li, K. Chen, K. Liang, G. Chen, J. Lu, J. Palisaitis, P. O. A. Persson, P. Eklund, L. Hultman, S. Du, Z. Chai, Y. Gogotsi and Q. Huang, Science, 2023, 379, 1130–1135 CrossRef CAS PubMed.
  52. M. Shen, W. Jiang, K. Liang, S. Zhao, R. Tang, L. Zhang and J. Wang, Angew. Chem., Int. Ed., 2021, 60, 27013–27018 CrossRef CAS PubMed.
  53. I. Persson, J. Halim, H. Lind, T. W. Hansen, J. B. Wagner, L.-Å. Näslund, V. Darakchieva, J. Palisaitis, J. Rosen and P. O. Å. Persson, Adv. Mater., 2019, 31, 1805472 CrossRef PubMed.
  54. S. Venkateshalu, M. Shariq, B. Kim, M. Patel, K. S. Mahabari, S.-I. Choi, N. K. Chaudhari, A. N. Grace and K. Lee, J. Mater. Chem. A, 2023, 11, 13107–13132 RSC.
  55. Y. Georgantas, F. P. Moissinac and M. Bissett, Graphene 2D Mater., 2024, 9, 5–26 CrossRef.
  56. A. Champagne and J.-C. Charlier, J. Phys. Mater., 2021, 3, 032006 CrossRef.
  57. M. Khazaei, A. Ranjbar, M. Arai, T. Sasaki and S. Yunoki, J. Mater. Chem. C, 2017, 5, 2488–2503 Search PubMed.
  58. U. Yorulmaz, A. Özden, N. K. Perkgöz, F. Ay and C. Sevik, Nanotechnology, 2016, 27, 335702 CrossRef PubMed.
  59. H. Zhang, G. Yang, X. Zuo, H. Tang, Q. Yang and G. Li, J. Mater. Chem. A, 2016, 4, 12913–12920 Search PubMed.
  60. K. Luo, X.-H. Zha, Y. Zhou, Q. Huang, S. Zhou and S. Du, Int. J. Quantum Chem., 2020, 120, e26409 Search PubMed.
  61. T. Schultz, N. C. Frey, K. Hantanasirisakul, S. Park, S. J. May, V. B. Shenoy, Y. Gogotsi and N. Koch, Chem. Mater., 2019, 31, 6590–6597 CrossRef CAS.
  62. M. López, K. S. Exner, F. Viñes and F. Illas, Adv. Theory Simul., 2022, 2200217, 1–9 Search PubMed.
  63. R. Ibragimova, P. Erhart, P. Rinke and H.-P. P. Komsa, J. Phys. Chem. Lett., 2021, 12, 2377–2384 CrossRef CAS PubMed.
  64. R. Ibragimova, M. J. Puska and H.-P. Komsa, ACS Nano, 2019, 13, 9171–9181 CrossRef CAS PubMed.
  65. H. A. Tahini, X. Tan and S. C. Smith, Nanoscale, 2017, 9, 7016–7020 RSC.
  66. S. Bera and H. Kumar, J. Phys. Chem. C, 2023, 127, 20734–20741 CrossRef CAS.
  67. M. Faraji, A. Bafekry, M. M. Fadlallah, F. Molaei, N. N. Hieu, P. Qian, M. Ghergherehchi and D. Gogova, Phys. Chem. Chem. Phys., 2021, 23, 15319–15328 RSC.
  68. S. A. Kazemi, S. A. Ogunkunle, O. Allen, W. Wen, A. W.-C. Liew, S. Yin and Y. Wang, J. Phys. Mater., 2023, 6, 035004 CrossRef CAS.
  69. X. Nie, Y. Ji, Y.-M. Ding and Y. Li, Nanotechnology, 2023, 34, 105704 Search PubMed.
  70. D. Li, X. Chen, P. Xiang, H. Du and B. Xiao, Appl. Surf. Sci., 2020, 501, 144221 CrossRef CAS.
  71. E. Rems, Y.-J. Hu, Y. Gogotsi and R. Dominko, Chem. Mater., 2024, 36, 10295–10306 CrossRef CAS PubMed.
  72. J. Björk and J. Rosen, Chem. Mater., 2021, 33, 9108–9118 CrossRef.
  73. J. Yang, A. Wang, S. Zhang, H. Wu and L. Chen, Comput. Mater. Sci., 2018, 153, 303–308 CrossRef CAS.
  74. L. Zhou, Y. Zhang, Z. Zhuo, A. J. Neukirch and S. Tretiak, J. Phys. Chem. Lett., 2018, 9, 6915–6920 CrossRef CAS PubMed.
  75. J. Kalmár and F. Karlický, J. Appl. Phys., 2024, 135, 244302 CrossRef.
  76. E. Heidari Semiromi, Z. Khorasani Baghini and A. Mostafaei, Solid State Commun., 2024, 381, 115451 CrossRef CAS.
  77. M.-L. Qin, S.-Y. Wu, T.-H. Guo, M.-Q. Wu, Q.-S. Zhu and M.-Q. Kuang, Surf. Interfaces, 2024, 46, 104073 CrossRef CAS.
  78. M. Dahlqvist and J. Rosen, npj 2D Mater. Appl., 2024, 8, 65 CrossRef CAS.
  79. H. Alnoor, A. Elśukova, J. Palisaitis, E. N. Tseng, J. Lu and L. Hultman, Mater. Today Adv., 2021, 9, 100123 CrossRef CAS.
  80. J. D. Gouveia, F. Viñes, F. Illas and J. R. B. Gomes, Phys. Rev. Mater., 2020, 4, 054003 CrossRef CAS.
  81. P. Urbankowski, B. Anasori, T. Makaryan, D. Er, S. Kota, P. L. Walsh, M. Zhao, V. B. Shenoy, M. W. Barsoum and Y. Gogotsi, Nanoscale, 2016, 8, 11385–11391 RSC.
  82. E. B. Deeva, A. Kurlov, P. M. Abdala, D. Lebedev, S. M. Kim, C. P. Gordon, A. Tsoukalou, A. Fedorov and C. R. Müller, Chem. Mater., 2019, 31, 4505–4513 CrossRef CAS.
  83. Y. Kong, S. Yan, J. Feng, S. Wang and H. Pan, J. Mater. Chem. A, 2021, 9, 597–606 RSC.
  84. J. D. Gouveia and J. R. B. Gomes, Phys. Rev. Mater., 2022, 6, 024004 CrossRef CAS.
  85. A. Jurado, Á. Morales-García, F. Viñes and F. Illas, J. Phys. Chem. C, 2021, 125, 26808–26813 CrossRef CAS PubMed.
  86. G. Kresse and J. Hafner, Phys. Rev. B: Condens. Matter Mater. Phys., 1994, 49, 14251–14269 CrossRef CAS PubMed.
  87. G. Kresse and J. Hafner, Phys. Rev. B: Condens. Matter Mater. Phys., 1993, 47, 558–561 CrossRef CAS PubMed.
  88. G. Kresse and J. Furthmüller, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169–11186 CrossRef CAS PubMed.
  89. G. Kresse and J. Furthmüller, Comput. Mater. Sci., 1996, 6, 15–50 CrossRef CAS.
  90. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef CAS PubMed.
  91. P. E. Blöchl, Phys. Rev. B: Condens. Matter Mater. Phys., 1994, 50, 17953–17979 CrossRef PubMed.
  92. G. Kresse and D. Joubert, Phys. Rev. B: Condens. Matter Mater. Phys., 1999, 59, 1758–1775 CrossRef CAS.
  93. J. Heyd, G. E. Scuseria and M. Ernzerhof, J. Chem. Phys., 2003, 118, 8207–8215 CrossRef CAS.
  94. V. Wang, N. Xu, J.-C. Liu, G. Tang and W.-T. Geng, Comput. Phys. Commun., 2021, 267, 108033 CrossRef CAS.
  95. G. Henkelman, A. Arnaldsson and H. Jónsson, Comput. Mater. Sci., 2006, 36, 354–360 CrossRef.
  96. V. Wang, G. Tang, Y. C. Liu, R. T. Wang, H. Mizuseki, Y. Kawazoe, J. Nara and W. T. Geng, J. Phys. Chem. Lett., 2022, 13, 11581–11594 CrossRef CAS PubMed.
  97. B. Hammer and J. K. Nørskov, Surf. Sci., 1995, 343, 211–220 CrossRef CAS.
  98. J. R. Kitchin, J. K. Nørskov, M. A. Barteau and J. G. Chen, J. Chem. Phys., 2004, 120, 10240–10246 CrossRef CAS PubMed.
  99. M. Gajdo, A. Eichler and J. Hafner, J. Phys.: Condens. Matter, 2004, 16, 1141–1164 CrossRef.
  100. B. S. Mun, M. Watanabe, M. Rossi, V. Stamenkovic, N. M. Markovic and P. N. Ross, J. Chem. Phys., 2005, 123, 204717 CrossRef PubMed.
  101. T. Hofmann, T. H. Yu, M. Folse, L. Weinhardt, M. Bär, Y. Zhang, B. V. Merinov, D. J. Myers, W. A. Goddard and C. Heske, J. Phys. Chem. C, 2012, 116, 24016–24026 CrossRef CAS.
  102. K. Momma and F. Izumi, J. Appl. Crystallogr., 2008, 41, 653–658 CrossRef CAS.
  103. mendeleev – A Python resource for properties of chemical elements, ions and isotopes, ver. 1.0.0, 2014, https://github.com/lmmentel/mendeleev Search PubMed.
  104. J. Demšar, T. Curk, A. Erjavec, Črt Gorup, T. Hočevar, M. Milutinovič, M. Možina, M. Polajnar, M. Toplak, A. Starič, M. Štajdohar, L. Umek, L. Žagar, J. Žbontar, M. Žitnik and B. Zupan, J. Mach. Learn. Res., 2013, 14, 2349–2353 Search PubMed.
  105. Y. Li, H. Shao, Z. Lin, J. Lu, L. Liu, B. Duployer, P. O. A. Persson, P. Eklund, L. Hultman, M. Li, K. Chen, X.-H. Zha, S. Du, P. Rozier, Z. Chai, E. Raymundo-Pñero, P.-L. Taberna, P. Simon and Q. Huang, Nat. Mater., 2020, 19, 894–899 CrossRef CAS PubMed.
  106. M. Kertesz and R. Hoffmann, J. Am. Chem. Soc., 1984, 106, 3453–3460 CrossRef CAS.
  107. A. Miura, K. Tadanaga, E. Magome, C. Moriyoshi, Y. Kuroiwa, T. Takahiro and N. Kumada, J. Solid State Chem., 2015, 229, 272–277 CrossRef CAS.
  108. D. R. Askeland and W. Wright, The science and engineering of materials, CENGAGE Learning Custom Publishing, Mason, OH, 7th edn, 2015 Search PubMed.
  109. T. Sato, S. Takagi, S. Deledda, B. C. Hauback and S.-I. Orimo, Sci. Rep., 2016, 6, 23592 CrossRef CAS PubMed.
  110. Z. Yang, Y. Zheng, W. Li and J. Zhang, J. Comput. Chem., 2019, 40, 1352–1359 CrossRef CAS PubMed.
  111. Y. Shao, F. Zhang, X. Shi and H. Pan, Phys. Chem. Chem. Phys., 2017, 19, 28710–28717 RSC.
  112. 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.
  113. T. Hu, Z. Li, M. Hu, J. Wang, Q. Hu, Q. Li and X. Wang, J. Phys. Chem. C, 2017, 121, 19254–19261 CrossRef CAS.
  114. N. Zhang, Y. Hong, S. Yazdanparast and M. Asle Zaeem, 2D Mater., 2018, 5, 045004 CrossRef CAS.
  115. Y. Xin and Y.-X. Yu, Mater. Des., 2017, 130, 512–520 CrossRef CAS.
  116. P. Roy, L. Rekhi, S. W. Koh, H. Li and T. S. Choksi, JPhys Energy, 2023, 5, 034005 CrossRef.
  117. Q. Zhu and S.-q Wang, J. Electrochem. Soc., 2016, 163, H796–H808 CrossRef CAS.
  118. F. Gossenberger, T. Roman, K. Forster-Tonigold and A. Groß, Beilstein J. Nanotechnol., 2014, 5, 152–161 CrossRef PubMed.
  119. F. Gossenberger, T. Roman and A. Groß, Surf. Sci., 2015, 631, 17–22 CrossRef CAS.
  120. R. Jacobs, J. Booske and D. Morgan, Adv. Funct. Mater., 2016, 26, 5471–5482 CrossRef CAS.
  121. T. C. Leung, C. L. Kao, W. S. Su, Y. J. Feng and C. T. Chan, Phys. Rev. B: Condens. Matter Mater. Phys., 2003, 68, 68–73 CrossRef.
  122. J. M. Crowley, J. Tahir-Kheli and W. A. Goddard, J. Phys. Chem. Lett., 2016, 7, 1198–1203 Search PubMed.
  123. H.-D. Saßnick and C. Cocchi, Electron. Struct., 2021, 3, 027001 CrossRef.
  124. E. M. Flores, M. L. Moreira and M. J. Piotrowski, J. Phys. Chem. A, 2020, 124, 3778–3785 CrossRef CAS PubMed.
  125. Y. X. Wang, H. Y. Geng, Q. Wu and X. R. Chen, J. Chem. Phys., 2020, 152, 024118 Search PubMed.
  126. T. M. Henderson, J. Paier and G. E. Scuseria, Phys. Status Solidi B, 2010, 248, 767–774 CrossRef.
  127. W. Gao, T. A. Abtew, T. Cai, Y.-Y. Sun, S. Zhang and P. Zhang, Solid State Commun., 2016, 234–235, 10–13 CrossRef CAS.
  128. K. Pokharel, C. Lane, J. W. Furness, R. Zhang, J. Ning, B. Barbiellini, R. S. Markiewicz, Y. Zhang, A. Bansil and J. Sun, npj Comput. Mater., 2022, 8, 31 CrossRef CAS.
  129. Y. Meng, X.-W. Liu, C.-F. Huo, W.-P. Guo, D.-B. Cao, Q. Peng, A. Dearden, X. Gonze, Y. Yang, J. Wang, H. Jiao, Y. Li and X.-D. Wen, J. Chem. Theory Comput., 2016, 12, 5132–5144 CrossRef CAS PubMed.
  130. S. S. Li, Semiconductor physical electronics, Springer Science & Business Media, 2012 Search PubMed.
  131. A. C. Rajan, A. Mishra, S. Satsangi, R. Vaish, H. Mizuseki, K.-R. R. Lee and A. K. Singh, Chem. Mater., 2018, 30, 4031–4038 CrossRef CAS.
  132. K. Hantanasirisakul and Y. Gogotsi, Adv. Mater., 2018, 30, 1–30 CrossRef PubMed.
  133. B. Ul Haq, S.-H. Kim, I. Khadka, R. Ahmed, S. AlFaify, A. R. Chaudhry and Y. A. Obaidat, Emergent Mater., 2023, 6, 1991–1999 CrossRef CAS.
  134. Y.-m Ding, X. Nie, H. Dong, N. Rujisamphan and Y. Li, Nanoscale Adv., 2020, 2, 2471–2477 RSC.
  135. A. Patra, S. Jana, P. Samal, F. Tran, L. Kalantari, J. Doumont and P. Blaha, J. Phys. Chem. C, 2021, 125, 11206–11215 CrossRef CAS PubMed.
  136. Y. Zhang, M. Wu, Z. Wang, N. Zhang and C. Ge, J. Mater. Res., 2021, 36, 1678–1685 CrossRef CAS.
  137. X.-H. Zha, J. Zhou, K. Luo, J. Lang, Q. Huang, X. Zhou, J. S. Francisco, J. He and S. Du, J. Phys.: Condens. Matter, 2017, 29, 165701 CrossRef PubMed.
  138. N. Kumar, M. Kolos, S. Bhattacharya and F. Karlický, J. Chem. Phys., 2024, 160, 124707 CrossRef CAS PubMed.
  139. S. Qiao, Y. Zhang, S. Li, L. Wei, H. Wu and F. Li, Phys. Chem. Chem. Phys., 2025, 27, 513–519 RSC.
  140. Y. Xie and P. R. C. Kent, Phys. Rev. B: Condens. Matter Mater. Phys., 2013, 87, 235441 CrossRef.
  141. K. A. Papadopoulou, A. Chroneos, D. Parfitt and S.-R. G. Christopoulos, J. Appl. Phys., 2020, 128, 170902 CrossRef CAS.
  142. W. Zheng, B. Sun, D. Li, S. M. Gali, H. Zhang, S. Fu, L. Di Virgilio, Z. Li, S. Yang, S. Zhou, D. Beljonne, M. Yu, X. Feng, H. I. Wang and M. Bonn, Nat. Phys., 2022, 18, 544–550 Search PubMed.
  143. P. Eliášová, B. Šmíd, J. Vejpravová, S. Li, F. Brivio, M. Mazur, D. N. Rainer, M. I. H. Mohideen, R. E. Morris and P. Nachtigall, J. Mater. Chem. C, 2024, 12, 5431–5441 RSC.
  144. J. Hart, K. Hantanasirisakul, A. C. Lang, B. Anasori, D. Pinto, Y. Pivak, J. Tijn van Omme, S. J. May, Y. Gogotsi and M. L. Taheri, Nat. Commun., 2019, 10, 1–10 Search PubMed.
  145. R. Khanal and S. Irle, J. Chem. Phys., 2023, 158, 194701 CrossRef CAS PubMed.
  146. J. Fatima, M. Tahir, A. Rehman, M. Sagir, M. Rafique, M. A. Assiri, M. Imran and M. Alzaid, Mater. Sci. Eng. B, 2023, 289, 116230 CrossRef CAS.
  147. N. Zhang, Y. Hong, S. Yazdanparast and M. Asle Zaeem, 2D Mater., 2018, 5, 045004 Search PubMed.
  148. G. Bharathi and S. Hong, Materials, 2024, 18, 104 Search PubMed.
  149. B. Hammer and J. Nørskov, Impact of Surf. Sci. on Catalysis, Academic Press, 2000, vol. 45, pp. 71–129 Search PubMed.
  150. Y. Guan, Y. Cong, R. Zhao, K. Li, X. Li, H. Zhu, Q. Zhang, Z. Dong and N. Yang, Small, 2023, 19, 2301276 Search PubMed.
  151. Y. Wang, J. Fu, H. Hu and D. Ho, ACS Appl. Mater. Interfaces, 2023, 15, 40846–40854 Search PubMed.
  152. Y. Zhang, Y. Wang, N. Ma, B. Liang, Y. Xiong and J. Fan, Small, 2024, 20, 2306840 CrossRef CAS PubMed.
  153. P. Liu, H. Xu, X. Wang, G. Tian, X. Wen, C. Wang, C. Zeng, S. Wang, F. Fan and T. Zeng, et al., J. Colloid Interface Sci., 2024, 655, 364–370 CrossRef CAS PubMed.
  154. Z. Wang, Z. Dong, B. Wu, Z. Wang, Z. Qiu, D. Wang, Q. Zeng, X. Liu, K. Nam Hui, Z. Liu and Y. Zhang, J. Colloid Interface Sci., 2024, 676, 368–377 Search PubMed.
  155. F. Mouhat and F.-X. Coudert, Phys. Rev. B: Condens. Matter Mater. Phys., 2014, 90, 224104 CrossRef.
  156. K. Liu and J. Wu, J. Mater. Res., 2015, 31, 832–844 CrossRef.
  157. A. V. Dyskin, E. Pasternak and Y. Xu, Phys. Status Solidi B, 2017, 254, 1600851 CrossRef.
  158. S. Xinchun and R. S. Lakes, Phys. Status Solidi B, 2007, 244, 1008–1026 CrossRef.
  159. D. Wu, S. Wang, S. Zhang, J. Yuan, B. Yang and H. Chen, Phys. Chem. Chem. Phys., 2018, 20, 18924–18930 RSC.
  160. B. Mortazavi, M. Shahrokhi, M. Makaremi and T. Rabczuk, Nanotechnology, 2017, 28, 115705 CrossRef PubMed.
  161. X. Du, J. Zhao, J. Wang, X. Liu, Z. Ye, L. Fang and M. Zhou, J. Chem. Phys., 2025, 162, 014703 CrossRef CAS PubMed.
  162. C. Huang and L. Chen, Adv. Mater., 2016, 28, 8079–8096 CrossRef CAS PubMed.
  163. G. Qin and Z. Qin, Npj Comput. Mater., 2020, 6, 51 CrossRef CAS.
  164. P. Schindler, E. R. Antoniuk, G. Cheon, Y. Zhu and E. J. Reed, Adv. Funct. Mater., 2024, 34, 2401764 CrossRef CAS.
  165. Y. Li, F. Liu, Z. Sui and C. Jiang, Thin Solid Films, 2022, 741, 139038 CrossRef CAS.
  166. S. Ahn, T. Han, K. Maleski, J. Song, Y. Kim, M. Park, H. Zhou, S. Yoo, Y. Gogotsi and T. Lee, Adv. Mater., 2020, 32, 2000919 CrossRef CAS PubMed.
  167. P. Lu, J. Wu, X. Shen, X. Gao, Z. Shi, M. Lu, W. W. Yu and Y. Zhang, Adv. Sci., 2020, 7, 2001562 CrossRef CAS PubMed.
  168. S. Chuang, C. Battaglia, A. Azcatl, S. McDonnell, J. S. Kang, X. Yin, M. Tosun, R. Kapadia, H. Fang, R. M. Wallace and A. Javey, Nano Lett., 2014, 14, 1337–1342 CrossRef CAS PubMed.
  169. B. Lyu, M. Kim, H. Jing, J. Kang, C. Qian, S. Lee and J. H. Cho, ACS Nano, 2019, 13, 11392–11400 CrossRef CAS PubMed.
  170. J. You, C. Si, J. Zhou and Z. Sun, J. Phys. Chem. C, 2019, 123, 3719–3726 CrossRef CAS.
  171. S. Qamar, K. Fatima, N. Ullah, Z. Akhter, A. Waseem and M. Sultan, Nanoscale, 2022, 14, 13018–13039 RSC.
  172. J. Zhang, C. Huang and H. Yu, Appl. Phys. Lett., 2021, 119, 033506 CrossRef CAS.
  173. Z. Yu, W. Feng, W. Lu, B. Li, H. Yao, K. Zeng and J. Ouyang, J. Mater. Chem. A, 2019, 7, 11160–11169 RSC.
  174. N. U. Kiran, A. B. Deore, M. A. More, D. J. Late, C. S. Rout, P. Mane, B. Chakraborty, L. Besra and S. Chatterjee, ACS Appl. Electron. Mater., 2022, 4, 2656–2666 CrossRef CAS.
  175. J. Chen, B. Yang, Y. D. Lim, W. Duan, Y. Zhao, B. K. Tay and X. Yan, Nanotechnology, 2020, 31, 285701 CrossRef CAS PubMed.
  176. S. Yamamoto, Appl. Surf. Sci., 2005, 251, 4–13 CrossRef CAS.
  177. C. Peng, P. Wei, X. Li, Y. Liu, Y. Cao, H. Wang, H. Yu, F. Peng, L. Zhang, B. Zhang and K. Lv, Nano Energy, 2018, 53, 97–107 CrossRef CAS.
  178. M. Li, B. Cai, R. Tian, X. Yu, M. B. Breese, X. Chu, Z. Han, S. Li, R. Joshi, A. Vinu, T. Wan, Z. Ao, J. Yi and D. Chu, Chem. Eng. J., 2021, 409, 128158 CrossRef CAS.
  179. A. Sherryna and M. Tahir, ACS Appl. Energy Mater., 2021, 4, 11982–12006 CrossRef CAS.
  180. R. Jacobs, D. Morgan and J. Booske, APL Mater., 2017, 5, 116105 CrossRef.
  181. T. Ma, R. Jacobs, J. Booske and D. Morgan, J. Mater. Chem. C, 2021, 9, 12778–12790 RSC.
  182. M. Khazaei, M. Arai, T. Sasaki, M. Estili and Y. Sakka, Phys. Chem. Chem. Phys., 2014, 16, 7841–7849 RSC.
  183. M. P. Bilibana, AdSEM, 2023, 2, 100080 Search PubMed.
  184. B. Miao, T. Bashir, H. Zhang, T. Ali, S. Raza, D. He, Y. Liu and J. Bai, Renewable Sustainable Energy Rev., 2024, 199, 114506 CrossRef CAS.
  185. Y. Huang, S. Jiang, R. Liang, P. Sun, Y. Hai and L. Zhang, Chem. Eng. J., 2020, 391, 123621 CrossRef CAS.
  186. W. Chen, X. Song, X. He, Y. Su, S.-K. Oh, S. Chen and Q. Sun, Prog. Org. Coat., 2024, 186, 108021 CrossRef CAS.
  187. D. An, Z. Wang, L. Qin, Y. Wu, S. Lu, H. Yang, Z. Ma, F. J. Mawignon, J. Liu, L. Hao and G. Li, Prog. Org. Coat., 2023, 183, 107779 CrossRef CAS.
  188. A. Maleki, M. Ghomi, N. Nikfarjam, M. Akbari, E. Sharifi, M. Shahbazi, M. Kermanian, M. Seyedhamzeh, E. Nazarzadeh Zare, M. Mehrali, O. Moradi, F. Sefat, V. Mattoli, P. Makvandi and Y. Chen, Adv. Funct. Mater., 2022, 32, 2203430 CrossRef CAS.
  189. H. Huang, C. Dong, W. Feng, Y. Wang, B. Huang and Y. Chen, Adv. Drug Delivery Rev., 2022, 184, 114178 CrossRef CAS PubMed.
  190. S. Iravani, E. Nazarzadeh Zare and P. Makvandi, ACS Biomater. Sci. Eng., 2024, 10, 1892–1909 CrossRef CAS PubMed.
  191. M. Malaki and R. S. Varma, Adv. Mater., 2020, 32, 2003154 CrossRef CAS PubMed.
  192. Y. Sliozberg, J. Andzelm, C. B. Hatter, B. Anasori, Y. Gogotsi and A. Hall, Compos. Sci. Technol., 2020, 192, 108124 CrossRef CAS.
  193. J. Hu, S. Li, J. Zhang, Q. Chang, W. Yu and Y. Zhou, Chin. Chem. Lett., 2020, 31, 996–999 CrossRef CAS.
  194. R. Chen, X. Xu, D. Yu, C. Xiao, M. Liu, J. Huang, T. Mao, C. Zheng, Z. Wang and X. Wu, J. Mater. Chem. C, 2018, 6, 11193–11201 RSC.
  195. D. Yu, Y. Teng, N. Zhou, Y. Xu, X. Wang, X. Lin, Q. Wang and C. Xue, Colloids Surf., A, 2023, 659, 130752 CrossRef CAS.
  196. M. Xin, J. Li, Z. Ma, L. Pan and Y. Shi, Front. Chem., 2020, 8 DOI:10.3389/fchem.2020.00297.
  197. K. K. Saxena, R. Das and E. P. Calius, Adv. Eng. Mater., 2016, 18, 1847–1870 CrossRef CAS.
  198. D. Wang, C. Zhou, A. S. Filatov, W. Cho, F. Lagunas, M. Wang, S. Vaikuntanathan, C. Liu, R. F. Klie and D. V. Talapin, Science, 2023, 379, 1242–1247 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.