1. Introduction
MXenes constitute a new group of two-dimensional (2D) transition metal carbides, nitrides, and carbonitrides with the chemical formula (Mn+1XnTx, n = 1, 2, 3) and have been successfully synthesized since 2011 [1]. Specifically, MXenes can be fabricated by selectively etching the A layer from the MAX phases (Mn+1AXn, n = 1, 2, 3) materials via hydrofluoride acid (HF) [1-3]. The "M" denotes early transition metal, "A" stands for element in IIIA or IVA group, and "X" denotes carbon or nitrogen. Given the HF-etching preparation method, the MXene surfaces are generally covered with terminal groups "Tx" such as oxygen (O), fluorine (F), and hydroxyl (OH) [4, 5]. Additionally, MXenes exhibit excellent thermodynamic stability, superior mechanical strength, and flexibility [1]. Furthermore, they exhibit long-term stability (no / low activation with their low-Z chemical composition), large specific surface areas, high ion-exchange capacity, and acid resistance [6]. The aforementioned unique properties enable the use of MXenes in various applications such as supercapacitors[7], lithium ion batteries[8] catalysis[9-10], and spent fuel treatment[11-12].
It is known that the spent fuel treatment becomes a challenging environmental concern due to the contamination of actinides (e.g., U and metallic fission products Am, Cs, Sr, and I). Adsorption and separation constitute regular and efficient approaches to clear radioactive contamination. Moreover, 2D MXenes with the aforementioned abundant active sites and high specific surface area are proposed as effective adsorbents of the radioactive metallic contaminations [11, 12]. For example, Wang et al. [11] reported the utility of Ti2CTx nanosheets in enhanced separation of U(VI) ion (the uranium species generally exist in the form of uranyl (UO22+) in practical environments) from aqueous solutions with a high capacity corresponding to 470 mg/g. Simultaneously, an enhanced adsorption capacity for the radionuclide is reached when compared to that of dry T3C2Tx via rational control of the interlayer space of hydrated T3C2Tx MXenes. It is also reported that with higher adsorption capacity, MXenes exhibit significant prospects in terms of the removal of U(VI) when compared with other nanomaterials such as graphene oxide and metal organic frameworks [12].
Currently, several MXenes including M2C (TiC, Nb2C, V2C), Ti3C2, Ta4C2, (Ti0.5, Nb0.5)2C, Ti3CN, et al., have been successfully synthesized [13-16]. Moreover, some MXenes, such as Mg2C, are theoretically predicted by using CALYPSO software [17]. Among the diverse MXenes, titanium-containing (Ti-) MXenes correspond to the most extensively studied species. Given the HF etching preparation method, the MXene surfaces are typically covered with terminal groups such as O, F, and OH [4, 5]. Recently, full Cl-terminated MXenes are synthesized via a molten chlorinated salt etching method [18]. Thus, given the copious family members of such materials, their fundamental properties vary with their chemical compositions and surface functionalization. Most MXenes exhibit metallic properties, which can be tuned via element doping, surface functionalization, defects, external electric field, and stress among others [19-21].
Additionally, the mechanical properties are extremely important characters that should be considered during their potential application as irradiation-tolerant materials or spent fuel separation media. When compared with three-dimensional (3D) bulk materials, it is important to identify in-plane stiffness constants, Young’s elastic modulus, and in-plane Poisson’s ratio for the 2D MXenes. It is reported that the elastic constants of Tin+1Xn (n = 1, 2, 3) are excellent (> 500 GPa) [22] and that their Young’s modulus is in the range of 430–520 GPa [23]. Yorulmaz et al. [20] extensively investigated the dynamical and mechanical stability of M2C (M = Sc, Ti, Zr, Mo, and Hf) and their functionalized M2CT2 (T = O and F) MXenes wherein in-plane stiffness constants are predicted in a range of 92–161 N/m. The MXenes are still considerably stiff due to their stronger M-C/N bonds although they are less stiff than other 2D materials, such as graphene, single-layer h-BN monolayer, and MoS2 with stiffness constants corresponding to 340 N/m [24], 275.9 N/m [25], and 140 N/m [26], respectively. Furthermore, the stress–strain (SS) curves are an important measure to clarify the mechanical properties of MXenes. Chakraborty et al. [27] revealed that the critical strain (i.e., strain at which the stress reaches its maximal value) of Ti2C can sustain ~ 8% and ~ 13–14% under biaxial and uniaxial loads, respectively. The aforementioned values can be significantly improved up to more than 20% via surface oxygen functionalization. Notably, very few MXenes (Mg2C) exhibit negative Poisson’s ratios [28]. In a manner similar to the modification of electronic properties, their mechanical properties can also be adjusted via substitutional element doping and defects [27]. Currently, many theoretical studies were performed [20, 23, 27-30] and focus on the mechanical properties of MXenes. However, there is a paucity of studies that examine the systematical development of mechanical properties of MXenes and especially effects of metal composition, functionalized groups, and effects of vacancy on their mechanical properties.
It should be noted that the aforementioned strengths of MXenes correspond to ideal mechanical strengths that are defined as the maximum stress required to homogeneously deform its defect-free crystal lattices at zero temperature [31, 32] and is normally significantly higher than the actual stress required to produce a mechanical deformation in defective lattices. However, the ideal mechanical strength is still a fundamental mechanical parameter that characterizes the inherent bonding in the crystal, which sets clear limits on the mechanical properties that materials can achieve [32]. It is reported that the ideal strengths of graphene correspond to 110 GPa at strain of 19.4% along the zigzag direction and 121 GPa at strain of 26.6% along the armchair direction [33]. By introducing inhomogeneous lattice fluctuations, Fu et al. [34] indicated that the mechanical strength of Ti2CO2 and Mo2CO2 can be significantly shrunk when comparing to their ideal strengths. Furthermore, given the well-controlled fabrication processes, the measured mechanical properties of graphene or MoS2 can gradually approach their theoretically-predicted mechanical values [33, 35-37].
Herein, a comprehensive first-principles study was conducted to obtain the ideal mechanical properties of M2C (M = Ti, Zr, Hf) and functionalized M2CT2 (M = Ti, Zr, Hf; T = O, F, OH) MXenes. With the aid of systematical analyses on the stress–strain curves under homogenous tension in biaxial and uniaxial directions as well as the corresponding elastic parameters, tendency of mechanical properties among different metallic compositions and chemical group terminated systems were further identified. Furthermore, the effect of vacancy on the mechanical properties of MXenes is also established.
2. Computational details
All the first-principles calculations in the study were performed via the density functional theory (DFT) implemented in the VASP (Vienna ab-initio Simulation Package) package [38, 39]. The electronic interactions were described by the projector-augmented wave (PAW) method [40]. The exchange-correlation energy was treated via the Perdew-Burke-Ernzerhof (PBE) functional of the generalized gradient approximation (GGA) [41]. For the plane-wave basis, the energy cutoff was set as 520 eV to ensure good convergence.
The hexagonal unit cells of MXenes were selected to evaluate the fundamental structural properties, and two-formula-unit orthorhombic supercells were used to investigate the mechanical properties. The corresponding k-point meshes were set as 18 × 18 × 1 and 8 × 8 × 1 with the Monkhorst–Pack scheme for the two types of models, respectively. The repeating MXene slabs were separated by a vacuum space exceeding 15 Å in c axis to eliminate their interactions, and the periodic boundary conditions were set along the zigzag and armchair directions. The atomic structures of the MXenes were fully relaxed with the convergence criterions of total energies and Hellmann-Feynman forces less than 10-7 eV and 0.001 eV/Å, respectively.
The SS relations of the 2D MXenes under both biaxial and uniaxial tension were calculated to identify their intrinsic mechanical response to the external strain/stress. The biaxial tension was introduced via increasing the in-plane lattice parameters in both zigzag and armchair directions. With respect to the uniaxial tension, the strain was applied by increasing the lattice parameters in the zigzag or armchair direction. For each tensile test, the strain applied in the model increased with an interval of 0.01 strain until the fracture or deformation occurs. For each point, the supercells were relaxed in its orthogonal direction until all the components of the Hellman–Feynman stresses were less than 0.2 GPa to decrease the effect of the stress tensor in the orthogonal direction on the axial stress. Moreover, the box size in the orthogonal direction of the monolayer structure h is ambiguous under the framework of first-principles calculations. The equivalent stress was obtained via scaling the supercell stress with a factor corresponding to h/d in the c direction where h denotes the box lattice in c direction and d denotes the actual thickness of MXenes. This method was successfully employed in the previous calculations for the SS curves of 2D materials [27].
3. Results and discussion
3.1 Structural properties and SS curves
As shown in Fig. 1a, the 2D monolayer M2C MXenes are fabricated from the bulk M2AC (A = Si/Al) MAX phase materials by selectively removing the A layer (mostly aluminum) [13]. The unit cell of M2C crystal is a hexagonal lattice with carbon layer sandwiched between two M layers as identified by the black lines shown in Fig.1 b. In order to perform the tension tests, the two-formula-unit orthorhombic supercells of M2C crystal as denoted by the dashed red lines are tailored from the pristine hexagonal M2C supercells. The first nearest-neighbor (1 NN) M-C bond direction of the orthorhombic cell is identified as the zigzag boundary while the second nearest-neighbor (2 NN) M-C bond direction is identified as the armchair boundary. Furthermore, the experimentally as-synthesized MXenes are typically chemically terminated with O, F, OH groups [13]. Specifically, the M2C and functionalized M2CT2 MXenes (as shown in Fig. 1 c) are investigated for their mechanical properties. The adsorbed sites for the functionalized groups are selected to be chemi-bonded at the hollow site under which there is no carbon atom, and these sites are identified as the most energetic favorable sites in previous studies [21].
-201911/1001-8042-30-11-014/alternativeImage/1001-8042-30-11-014-F001.jpg)
The fundamental structural properties of MXenes investigated in the study are initially investigated by fully relaxing their hexagonal unit cells. Calculated lattice parameters, total energies of the optimized unit cells, bond lengths of M-C and M-T, and monolayer thickness of M2C and their functionalized M2CT2 (M = Ti, Zr, Hf; T = O, F, OH) MXenes are listed in Table 1. Among the MXenes considered in the study, the Ti2CT2 systems are the most examined species as mentioned previously. The calculated lattice parameters and total energies of Ti2C and Ti2CT2 MXenes are compared with the available experimental studies and DFT calculations via PBE method in previous studies, and they are found to be in good agreement with each other [20-22, 29, 42].
Lattice parameters (Å) | E (eV) | dM-C(Å) | dM-T(Å) | Thickness d (Å) | |
---|---|---|---|---|---|
Ti2C | 3.040 (3.039a, 3.01b, 3.040c, 3.040d) | -24.675 | 2.101 | — | 2.308(2.29b, 2.31c) |
Zr2C | 3.280 | -26.206 | 2.283 | — | 2.551 |
Hf2C | 3.211 | -29.170 | 2.246 | — | 2.538 |
Ti2CO2 | 3.035 (3.035a, 3.034e, 3.04c) | -44.460(-44.459e) | 2.186 | 1.979 | 4.456 (4.45c) |
Ti2CF2 | 3.060 (3.060a, 3.04c) | -39.360 | 2.104 | 2.161 | 4.777 (4.80c) |
Ti2C(OH)2 | 3.076 (3.04c) | -51.033 | 2.115 | 2.183 | 6.789 (6.78c) |
Zr2CO2 | 3.319 | -46.994 | 2.375 | 2.131 | 4.670 |
Zr2CF2 | 3.311 | -41.221 | 2.278 | 2.323 | 5.116 |
Zr2C(OH)2 | 3.319 | -52.823 | 2.286 | 2.350 | 7.155 |
Hf2CO2 | 3.265 | -50.770 | 2.333 | 2.103 | 4.613 |
Hf2CF2 | 3.263 | -44.082 | 2.232 | 2.310 | 5.067 |
Hf2C(OH)2 | 3.274 | -55.899 | 2.243 | 2.318 | 7.052 |
The SS curves are extremely important measures to evaluate mechanical properties of a material. Generally, the SS curves are drawn via recording the amount of strain and corresponding stress during their tensile deformation. The curves reveal many essential mechanical properties including Young’s modulus, Poisson’s ratio, and critical strains. Here, the SS curves of M2C and their functionalized M2CT2 MXenes under biaxial and uniaxial (along zigzag or armchair direction) tensile strains are calculated and shown in Fig. 2 based on the methods described in the computational details section. This indicates that the SS curves of MXenes begin with a quasi-linear region at small strains and is followed by a plastic region at larger strains. The critical strain here is defined as the strain value at which the stress reaches its maximal value. Young’s modulus under biaxial and uniaxial tensions can be obtained by the linear fit on the initial slope of the corresponding SS curves.
-201911/1001-8042-30-11-014/alternativeImage/1001-8042-30-11-014-F002.jpg)
In the SS curves of pristine Ti2C MXene shown in Fig. 2 a, ideal tensile strength and critical strain of pristine Ti2C under the in-plane strain are calculated as 26.5 GPa and 7%, respectively, and they are significantly lower than that under the uniaxial strain with strengths and critical strains corresponding to 38-41GPa and 12%, respectively. The ideal tensile strengths and critical strains of Zr2C/Hf2C exhibit similar trends as those of Ti2C(see Fig. 2 b-c). This indicates that the functionalized M2CT2 systems reach maximum ideal strengths when they are subjected to 18–25% biaxial strains (Fig. S2 a, d, g). In a manner relative to the non-terminated M2C MXenes, the increased critical strains for the three functionalized M2CT2 prove that the surface functionalized groups enhance the toughness of pristine M2C MXene, which allow them to sustain larger strains without plastic deformation. The results also indicate that the biaxial SS curves for all the chemical functionalized MXenes (the black lines) are exactly drawn above the uniaxial SS curves (see the blue and red lines in Fig. 2 d-l) and indicate their relatively better mechanical properties. When the chemical functionalized MXenes are uniformly stretched in a uniaxial direction, the ideal strength along the zigzag direction exceed that along the armchair direction. It is observed that this type of behavior of Ti2CO2 is consistent with the previously calculated SS relations for Ti2CO2 and the Ti2(C0.5B0.5)O2 systems [27, 34]. The critical strains for F/OH terminated M2CT2 along the armchair direction are lower than those along the zigzag direction and indicate that fracture or deformation can occur at lower stages when applying strain along the armchair direction.
With respect to a more intuitive comparison of the mechanical properties among different metal compositions and different chemical terminated groups, the SS curves of MXenes with the sequence of metal compositions (M = Ti, Zr, Hf) and functionalized groups (T = O, F, OH) are redrawn and shown in Fig. S1 and Fig. S2, respectively. As shown in Fig. S1, the tensile properties of pristine Hf2C and Hf2CT2 (T = O, F, OH) system are the most optimal among the aforementioned three systems irrespective of the loads applied in uniaxial or biaxial directions. The Ti2CT2 and Zr2CT2 systems exhibit similar tensile behavior when stretched in the uniaxial direction and tensile properties of Ti2CT2 systems exceed those of Zr2CT2 when biaxial tension is involved. It should be noted that for non-functional MXenes, the tensile properties exhibited in the SS curves vary significantly with different metal-containing systems (see Fig. S1 a-c). This indicates that the ideal tensile strength of Hf2C is 50% and 33% higher than Ti2C and Zr2C under biaxial tension, respectively, and 28–32% improvements are observed for that along the uniaxial direction. For the functionalized M2CT2, similar tensile properties of MXenes among the different transition metal compositions can arise from the similarity in the chemical properties of Ti, Zr, and Hf elements.
The SS curves of MXenes with different functionalized groups are plotted in Fig. S2. It is clearly visible that the tensile properties of MXenes with different surface functionalization increase in a sequence corresponding to OH < F < O. With respect to the biaxial tensile tests (see Fig. S2 a, d, g), the biaxial tensile strengths and critical strains for all the chemical functionalized M2CT2 MXenes are improved when compared to that of M2C (as also see Fig. 2 d-l). The biaxial tensile critical strains of all M2CT2 exceed that of graphene (~15%) [43, 44]. Furthermore, for all the chemical functionalized MXenes, the O-terminal systems exhibit the maximum biaxial ideal strength (61-68GPa) at a lower biaxial strain of ~19%, when compared with that of F-terminal MXenes (42-47GPa; 20-22%) and OH-terminal MXenes (32–35 GPa; 23–25%). For the uniaxial tensile tests along the armchair (see Fig. S2 b, e, h) and zigzag (see Fig. S2 c, f, i) directions, it is also observed in the elastic parts (at smaller strains) of the SS curves that the O-terminated MXenes exhibit similar elastic properties to the non-terminated ones, which significantly exceed that of -F and -OH terminated MXenes. The strong interaction between the transition metal element and oxygen terminated groups aid in improving the toughness of M2C and further slow the collapse of M2CO2 MXenes. This type of strong bonding arises from the hybrid state interaction between M-d and C-p orbitals. The aforementioned results indicate that the M2C and M2CO2 MXenes can exhibit good mechanical properties, such as elastic constants and Young’s Modulus, and this is subsequently discussed. Furthermore, few uniaxial SS curves for the uniaxial tension of M2CO2 appear to increase with large strain > 20%. The M-C and M-O bond lengths are also recorded during uniaxial tension to identify the atomic structure evolution. As shown in Fig. S3 and Table S1, we consider the uniaxial SS of Ti2CO2 as an example. There are two types of Ti-C bonds wherein one is relatively more robust while another is significantly elongated during the tension test. The bond length elongates to 3.02 Å, which is close to 138% when compared to that without strain for tension along the armchair direction, indicating that the atomic structure is subject to serious lattice distortion. Therefore, the uniaxial SS curves are plotted with the same strain range of biaxial tension.
3.2 In-plane mechanical properties
By comparing with the bulk 3D materials, it extremely important to identify the in-plane elastic properties for these types of 2D materials. Young’s Modulus was calculated to identify the material's response for in-plane biaxial strain. The value of Young's modulus corresponds to the rigidity of materials, which is obtained by linear fitting of the elastic deformation stage of the stress–strain curve [23, 45] as follows:
where σ and ε denote stress and strain in the elastic deformation stage, respectively. With increases in Young's modulus, the material is more resistant to deformation. In the study, the Young’s modulus calculated from the stress–strain curves under biaxial tension are mainly discussed. The properties for uniaxial tension exhibit a similar trend via comparing the initial slope of relative SS curves as shown in Fig. 2. Specifically, the strain ranges of less than ~ 6% and ~ 8% are selected as the elastic region of the M2C and M2CT2 MXenes, respectively, as shown in Fig. 2.
As the calculated Young’s modulus shown in Fig. 3(a), non-terminated M2C MXenes exhibit the highest Young’s modulus while surface functionalization decreases their values. Moreover, the Young's modulus of M2CT2 with different functionalized groups is in a descending order of O > F > OH. Among different functionalized states, the Young's modulus of Zr2CT2 is lower than that of the Ti2CT2 and Hf2CT2 as shown in Fig. 3(b).
-201911/1001-8042-30-11-014/alternativeImage/1001-8042-30-11-014-F003.jpg)
In-plane stiffness C and Poisson’s ratio ν are other two important constants to characterize the properties of reduced dimensional MXenes. In the case of 2D materials, given the reduced dimensionality of these materials and ambiguity of the thickness of a monolayer structure h, it is more important to define the in-plane stiffness C as opposed to the classical 3D Young’s modulus as a measure of the mechanical property, as extensively used in literature for different 2D materials [46]. The in-plane stiffness constant C is calculated as follows [28, 46]:
where S0 denotes the equilibrium in-plane area of the MXenes, ε denotes biaxial strain while Es denotes the energy difference between the MXenes in their equilibrium and tensile states, and this is denoted as strain energy. The Poisson’s ratio of a 2D material is defined as the ratio of the contractile strain along one direction to the tensile strain along another direction, and it is positive for most materials [47, 48]. It is expressed as follows:
where
The calculated in-plane stiffness (C), Young’s modulus (E), ideal strength (σ), critical biaxial strain, and Poisson’s ratio (ν) for the M2C and their functionalized M2CT2 (M = Ti, Zr, Hf; T = O, F, OH) MXenes are listed in Table 2. The calculated in-plane stiffness constants of M2C (M = Ti, Zr, Hf) MXenes are in the range of 125–154 N/m. These properties are significantly improved to be in the range of 268–294 N/m for M2CO2, and this significantly exceeds that for M2CF2 and M2C(OH)2. The in-plane stiffness constants of MXenes are lower than that of graphene, single layer h-BN, and higher than MoS2 as shown in Table 2 [24-26].
C | E | σ | ν | ||
---|---|---|---|---|---|
Ti2C | 563 / 130 (125a, 142b) | 634(577b, 610c) | 26.720 | 7 (8b) | 0.300 (0.366b) |
Zr2C | 521 / 133 (129a) | 646 | 33.628 | 10 | 0.270 |
Hf2C | 651 / 165 (154a) | 771 | 44.316 | 10 | 0.263 |
Ti2CO2 | 603 / 268 (238a) | 578 (570b, 567c) | 67.946 | 20 (23b) | 0.295 (0.303b) |
Zr2CO2 | 582 / 271 (240a) | 548 | 61.181 | 19 | 0.298 |
Hf2CO2 | 679 / 294 (271a) | 594 | 67.968 | 18 | 0.250 |
Ti2CF2 | 420 / 200 (159a) | 393 | 47.358 | 20 | 0.300 |
Zr2CF2 | 402 / 205 (141a) | 346 | 42.261 | 20 | 0.294 |
Hf2CF2 | 368 / 186 (157a) | 371 | 46.096 | 22 | 0.270 |
Ti2C(OH)2 | 289 / 195 | 267 | 35.417 | 23 | 0.228 |
Zr2C(OH)2 | 266 / 190 | 245 | 32.013 | 23 | 0.228 |
Hf2C(OH)2 | 294 / 207 | 273 | 34.964 | 25 | 0.220 |
Graphene | — / 340d | ||||
h-BN | — / 275.9e | ||||
MoS2 | — / 140f |
As mentioned in Sect. 3.1, the surface functionalization with O, F, and OH groups allow M2C MXenes to sustain extended strains. The critical strains for the non-terminated M2C are calculated in a range of 7–10%, and it increases to 18–20%, 20–22%, and 23–25% for M2CT2 in the presence of surface terminal O, F, and OH groups, respectively.
3.3 Vacancy effect on the mechanical properties of MXenes
Defects and especially vacancy defects are unavoidable during the preparation or application of MXenes and can deteriorate material performance. On another hand, defects can tailor the properties of a material to a significant extent. Therefore, it is highly desirable to evaluate the vacancy effect on the mechanical properties of MXenes. The relative stability of vacancy is evaluated by the defect formation energy (Ef), which is defined as follows [49-51]:
where Edef denotes the total energy of the defective supercell, Eper denotes the energy of the defect-free supercell, and μ0 denotes the chemical potential of the vacancy species. Three possible types of vacancy defects can exist, namely metal vacancy (VM), carbon vacancy (VC), and functional group vacancy (VT) in MXenes. The titanium vacancies were experimentally obtained [52] while the carbon vacancies intrinsically exist in MAX phase materials [53]. We consider Ti2CO2 MXene as an example, and the formation energy of typical vacancies in Ti2CO2 MXene were systematically investigated as a function of vacancy concentrations [21]. The chemical potentials of Ti/Zr/Hf and C refer to the energy of their pure solids, and the chemical potential of O corresponds to the half energy of an O2 molecule. From the point of vacancy formation energy, this indicates that the carbon vacancy exhibits the lowest formation energy for various concentration when compared to that of titanium and oxygen vacancies, indicating that carbon vacancies are most likely to be formed [54]. Conversely, the atomic structures of Ti2CO2 MXene with titanium vacancy are quite unstable for high vacancy concentration, and it becomes stable in lower defect concentration supercells [52]. Specifically, in order to prevent the structure from losing stability due to excessive vacancy defect concentration, single carbon vacancy in the two-formula-unit orthorhombic supercells of M2CT2 (i.e., the mostly stable type of defect in MXenes) is selected to identify the defect effect on the mechanical properties of functionalized MXenes. The carbon vacancy exhibits stability in a comparable small cell of MXene lattice, which also saves computational resources in terms of a systematical calculation of their mechanical properties.
The calculated formation energies of carbon vacancy in M2CT2 (M = Ti, Zr, Hf; T = O, F, OH) MXenes are listed in Table 3. It is reported that the carbon formation energy for MXenes depends on the carbon vacancy concentration. Specifically, the formation energy of carbon vacancy in Ti2CO2 is calculated as 0.0015 eV, which is slightly smaller when compared to the predicted value in previous studies [21, 54] due to the difference in the shape of super lattices. For all the MXenes with different terminated groups, the tendency for carbon formation energies increases in the following order corresponding to Ti, Zr, and Hf. While for the MXene with specific metal compositions, it exhibits an increasing trend similar to the terminated O, OH, and F groups. The carbon vacancy in the case of Ti2CO2 MXene is identified with the lowest formation energy, indicating that carbon defect is most likely to be formed in all the cases that are considered.
Vacancy formation energy (eV) | |||
---|---|---|---|
Ti | Zr | Hf | |
M2CO2 | 0.0015 (0.385a) | 1.5808 | 2.3617 |
M2CF2 | 2.5718 (2.7b) | 3.1134 | 3.5381 |
M2C(OH)2 | 2.2903 | 2.8672 | 3.3730 |
The effect of carbon vacancy on the mechanical properties of M2CO2 (M = Ti, Zr, Hf) systems are systematically investigated by exploring their SS curves as shown in Fig. 4. It is noted that the stress components along the armchair and zigzag directions for the biaxial tension are slightly different from each other, and we consider the smaller component of the stress obtained along the orthogonal directions. The results indicate that the mechanical properties of the materials are significantly weakened due to the effect of carbon vacancy. The in-plane strength and critical strain of Ti2CO2 with single carbon vacancy for the biaxial tensile correspond to 34.1 GPa and 9%, respectively, which are approximately half that of defect-free structures. The values are also lower than the results (~ 44.4 GPa, 15%) calculated in previous studies by introducing lattice fluctuations [34]. Moreover, the strength of Zr2CO2 and Hf2CO2 also decreased by ~ 40%. The elastic parameters are also calculated as listed in Table 4, and Young’s modulus and in-plane stiffness of each system are also reduced by approximately 20–27% and 17–21%, respectively. The results indicate that the toughness of the material is significantly reduced, and the defective M2CO2 may deform under lower strain than those without defects. Subsequently, we compare the tensile behavior of the M2CO2 (M = Ti, Zr, Hf) MXene systems with carbon vacancy under biaxial tension (Fig. 5(a)). After the carbon vacancy is introduced, the critical strain of Zr2CO2 system is derived as ~ 10%, which indicates that the Zr2CO2 can deform under ~ 10% biaxial strain while Ti2CO2 and Hf2CO2 can withstand up to 12% biaxial strain. As previously mentioned, Zr2CO2 exhibits the lowest Young's modulus, and thus the material is most prone to deformation.
VC-doped M2CO2 | C | E | σ | ν | |
---|---|---|---|---|---|
Ti2CO2-V | 479 / 210 | 420 | 34.108 | 9 | 0.300 |
Zr2CO2-V | 478 / 221 | 428 | 37.085 | 10 | 0.330 |
Hf2CO2-V | 533 / 244 | 475 | 40.721 | 12 | 0.300 |
-201911/1001-8042-30-11-014/alternativeImage/1001-8042-30-11-014-F004.jpg)
-201911/1001-8042-30-11-014/alternativeImage/1001-8042-30-11-014-F005.jpg)
For the biaxial tension of the defect-free MXenes, the derived stress components along both armchair and zigzag directions are identical to each other, and there exists anisotropy between the unidirectional tension along armchair and zigzag directions as shown in Fig. 2. For the defective MXenes, the stress components along the zigzag and armchair directions under biaxial tension are plotted as solid and dashed lines in Fig. 5, respectively. This indicates that the two stress components along the armchair and zigzag directions are different from each other after the carbon vacancy is introduced. The anisotropy is mainly attributed to the difference in the linear defect densities along armchair and zigzag directions. In order to further explore the underlying mechanism for the phenomenon, the bonding conditions between the adjacent carbon and surrounding titanium atoms are analyzed. As shown in Fig. 5(b), there are two types of bonds, namely "ba" and "bz", which are quasi-parallel to the armchair and zigzag directions, respectively. The average length of Ti-C bonds, denoted as "ba" and "bz", during biaxial tension are plotted in Fig. 5(c). When the biaxial strain increases, the bond length of "ba" and "bz" exhibits a trendency of separation. The length of Ti-C bond elongates in the zigzag direction although it shrinks in the orthogonal armchair direction with increases in strain. Thus, the stresses in the two orthogonal directions are inconsistent and exhibit anisotropic manners even for the biaxial tension loads that are applied on the in-plane surface. Furthermore, increases in stress applied to the model cause the bond lengths to exhibit a more serious separation during their elastic stages. Subsequently, atomic structures can be deformed and reconstructed in the following plastic stages.
4. Conclusion
The structural and mechanical properties of MXenes are crucial to their potential applications. In this study, the mechanical properties of M2CT2 (M = Ti, Zr, Hf; T=O, F, OH) MXenes were systematically investigated via a first-principles method. The SS curves of MXenes under homogenous tension in their biaxial and uniaxial directions and elastic parameters including Young’s modulus, in-plane stiffness, and Poisson’s ratios were calculated and analyzed. The main findings are summarized as follows:
First, ideal strength and critical strain of M2C were in a range of 26–44 GPa and 7–10%, respectively. With significantly higher ideal strength and extended critical strains, the M2CO2 MXenes exhibited optimal flexibility when compared with that of M2C, M2CF2, and M2C(OH)2. The Young’s modulus, in-plane stiffness, and Poisson’s ratio of MXenes with different surface functionalization decrease in a sequence corresponding to O > F > OH.
Second, for the M2CT2 with different metal compositions, the Hf2CT2 exhibited the best tensile performance under uniaxial and biaxial tension, and similarity in the tensile properties of MXenes among the different transition metal compositions can arise from the similarity in the chemical properties of Ti, Zr, and Hf elements. Furthermore, the anisotropy between biaxial and uniaxial tension was identified for each MXene.
Third, the effect of defect on the mechanical properties of MXenes was further explored via introducing a carbon vacancy. The results indicated that the vacancy significantly weakened the tensile properties of MXene systems with their ideal strength, critical strain, in-plane stiffness constant, and Young’s modulus in ranges of 40–50%, 33–50%, 17–21%, and 20–27%, respectively. Furthermore, the defect also affected the bonding, resulting in a certain anisotropy of stress along armchair and zigzag directions even under the biaxial tension condition. The aforementioned anisotropy can be mainly attributed to the slight difference in the linear concentration of defects along the two orthogonal directions.
25th Anniversary Article: MXenes: A New Family of Two-Dimensional Materials
. Adv. Mater. 26, 992-1005 (2014). doi: 10.1002/adma.201304138.Two-Dimensional Transition Metal Carbides
. ACS Nano. 6, 1322-1331 (2012). doi: 10.1021/nn204153h.Synthesis of Two-Dimensional Materials by Selective Extraction
. Acc. Chem. Res. 48, 128-135 (2015). doi: 10.1021/ar500346b.MXene: A Promising Transition Metal Carbide Anode for Lithium-Ion Batteries
. Electrochem. Commun. 16, 61-64 (2012). doi: 10.1016/j.elecom.2012.01.002.Flexible and Conductive MXene Films and Nanocomposites with High Capacitance
. Proc. Natl. Acad. Sci. 111, 16676-16681 (2014). doi: 10.1073/pnas.1414215111.Efficient U(VI) Reduction and Sequestration by Ti2CTx MXene
. Environ. Sci. Technol. 52, 10748-10756 (2018). doi: 10.1021/acs.est.8b03711.Cation Intercalation and High Volumetric Capacitance of Two-Dimensional Titanium Carbide
. Science. 341, 1502-1505 (2013). doi: 10.1126/science.1241488.Are MXenes Promising Anode Materials for Li Ion Batteries? Computational Studies on Electronic Properties and Li Storage Capability of Ti3C2 and Ti3C2X2 (X = F, OH) Monolayer. 40)
, J. Am. Chem. Soc. 134, 16909-16916 (2012). doi: 10.1021/ja308463r.Computational Studies on the Structural, Electronic and Optical Properties of Graphene-like MXenes (M2CT2, M = Ti, Zr, Hf; T = O, F, OH) and Their Potential Applications as Visible-Light Driven Photocatalysts
. J. Mater. Chem. A. 4, 12913-12920 (2016). doi: 10.1039/C6TA04628B.Searching for Highly Active Catalysts for Hydrogen Evolution Reaction Based on O-Terminated MXenes through a Simple Descriptor
. Chem. Mater. 28, 9026-9032 (2016). doi: 10.1021/acs.chemmater.6b03972.Effective Removal of Anionic Re(VII) by Surface-Modified Ti2CTx MXene Nanocomposites: Implications for Tc(VII) Sequestration
. Environ. Sci. Technol. 53, 3739-3747 (2019). doi: 10.1021/acs.est.8b07083.Research progress of MXene materials in radioactive element and heavy metal ion sequestration
. Sci. Sin. Chim. 49, 27-38 (2019). doi: 10.1360/N032018-00155.Synthesis and Characterization of 2D Molybdenum Carbide (MXene)
. Adv. Funct. Mater. 26, 3118-3127 (2016). doi: 10.1002/adfm.201505328.Two-Dimensional Nanocrystals Produced by Exfoliation of Ti3AlC2
. Adv. Mater. 23, 4248-4253 (2011). doi: 10.1002/adma.201102306.Synthesis of a New Graphene-like Transition Metal Carbide by de-Intercalating Ti3AlC2
. Mater. Lett. 109, 295-298 (2013). doi: 10.1016/j.matlet.2013.07.102.New Two-Dimensional Niobium and Vanadium Carbides as Promising Materials for Li-Ion Batteries
. J. Am. Chem. Soc. 135, 15966-15969 (2013). doi: 10.1021/ja405735d.Metal–Semiconductor Transition of Two-Dimensional Mg2C Monolayer Induced by Biaxial Tensile Strain
. Phys. Chem. Chem. Phys. 19, 32086-32090 (2017). doi: 10.1039/C7CP06150A.Element Replacement Approach by Reaction with Lewis Acidic Molten Salts to Synthesize Nanolaminated MAX Phases and MXenes
. J. Am. Chem. Soc. 141, 4730-4737 (2019). doi: 10.1021/jacs.9b00574.Novel Electronic and Magnetic Properties of Two-Dimensional Transition Metal Carbides and Nitrides
. Adv. Funct. Mater. 23, 2185-2192 (20130). doi: 10.1002/adfm.201202502.Vibrational and Mechanical Properties of Single Layer MXene Structures: A First-Principles Investigation
. Nanotechnology. 27, 335702 (2016). doi: 10.1088/0957-4484/27/33/335702.Stabilities and Electronic Properties of Vacancy-Doped Ti2CO2
. Comp. Mater. Sci. 159, 127-135 (2019). doi: 10.1016/j.commatsci.2018.12.007.First Principles Study of Two-Dimensional Early Transition Metal Carbides
. MRS Commun. 2, 133-137 (2012). doi: 10.1557/mrc.2012.25.Flexible Two-Dimensional Tin+1Cn (n = 1, 2 and 3) and Their Functionalized MXenes Predicted by Density Functional Theories
. Phys. Chem. Chem. Phys. 17, 15348-15354 (2015). doi: 10.1039/C5CP00775E.Observations of Intergranular Stress Corrosion Cracking in a Grain-Mapped Polycrystal
. Science. 321, 382-385 (2008). doi: 10.1126/science.1156211.Intrinsic Piezoelectricity in Two-Dimensional Materials
. J. Phys. Chem. Lett. 3, 2871-2876 (2012). doi: 10.1021/jz3012436.Outstanding Mechanical Properties of Monolayer MoS2 and Its Application in Elastic Energy Storage
. Phys. Chem. Chem. Phys. 15, 19427 (2013). doi: 10.1039/c3cp52879k.Manipulating the Mechanical Properties of Ti2C MXene: Effect of Substitutional Doping
. Phys. Rev. B. 95, 184106 (2017). doi: 10.1103/PhysRevB.95.184106.Monolayer Mg2C: Negative Poisson's Ratio and Unconventional Two-Dimensional Emergent Fermions
. Phys. Rev. Mater. 2, 104003 (2018). doi: 10.1103/PhysRevMaterials.2.104003.Superior Structural, Elastic and Electronic Properties of 2D Titanium Nitride MXenes over Carbide MXenes: A Comprehensive First Principles Study
. 2D Mater. 5, 045004 (2018). doi: 10.1088/2053-1583/aacfb3.Intrinsic Structural, Electrical, Thermal, and Mechanical Properties of the Promising Conductor Mo2C MXene
. J. Phys. Chem. C. 120, 15082-15088 (2016). doi: 10.1021/acs.jpcc.6b04192.Ultra-Strength Materials
. Prog. Mater. Sci. 55, 710-757 (2010). doi: 10.1016/j.pmatsci.2010.04.001.The Ideal Strength of Iron in Tension and Shear
. Acta Mater. 51, 2271-2283 (2003). doi: 10.1016/S1359-6454(03)00033-8.Ab Initio Calculation of Ideal Strength and Phonon Instability of Graphene under Tension
. Phys. Rev. B. 76, 064120 (2007). doi: 10.1103/PhysRevB.76.064120.Phonon-Mediated Stabilization and Softening of 2D Transition Metal Carbides: Case Studies of Ti2CO2 and Mo2CO2
. Phys. Chem. Chem. Phys. 20, 14608-14618 (2018). doi: 10.1039/C8CP00752G.Measurement of the Elastic Properties and Intrinsic Strength of Monolayer Graphene
. Science. 321, 385-388 (2008). doi: 10.1126/science.1157996.Stretching and Breaking of Ultrathin MoS2
. ACS Nano. 5, 9703-9709 (2011). doi: 10.1021/nn203879f.Ideal Strength and Phonon Instability in Single-Layer MoS2
. Phys. Rev. B. 85, 235407 (2012). doi: 10.1103/PhysRevB.85.235407.Ab Initio Molecular Dynamics for Liquid Metals
. Phys. Rev. B. 47, 558-561 (1993). doi: 10.1103/PhysRevB.47.558.Efficient Iterative Schemes for Ab Initio Total-Energy Calculations Using a Plane-Wave Basis Set
. Phys. Rev. B. 54, 11169-11186 (1996). doi: 10.1103/PhysRevB.54.11169.Projector Augmented-Wave Method
. Phys. Rev. B. 50, 17953-17979 (1994). doi: 10.1103/PhysRevB.50.17953.Generalized Gradient Approximation Made Simple
. Phys. Rev. Lett. 77, 3865-3868 (1996). doi: 10.1103/PhysRevLett.77.3865.First-Principles Study on Structural, Electronic and Elastic Properties of Graphene-like Hexagonal Ti2C Monolayer
. Comput. Mater. Sci. 83, 290-293 (2014). doi: 10.1016/j.commatsci.2013.11.025.Electronic Strengthening of Graphene by Charge Doping
. Phys. Rev. Lett. 109, 226802 (2012). doi: 10.1103/PhysRevLett.109.226802.Failure Mechanisms of Graphene under Tension
. Phys. Rev. Lett. 105, 245502 (2010). doi: 10.1103/PhysRevLett.105.245502.Superior Mechanical Flexibility of Phosphorene and Few-Layer Black Phosphorus
. Appl. Phys. Lett. 104, 251915 (2014). doi: 10.1063/1.4885215.The Response of Mechanical and Electronic Properties of Graphane to the Elastic Strain
. Appl. Phys. Lett. 96, 091912 (2010). doi: 10.1063/1.3353968.Auxetic Materials: Functional Materials and Structures from Lateral Thinking! Adv
. Mater. 12, 617-628 (2000). doi: 10.1002/(SICI)1521-4095(200005)12:9<617::AID-ADMA617>3.0.CO;2-3.Advances in Negative Poisson's Ratio Materials
. Adv. Mater. 5, 293-296 (1993). doi: 10.1002/adma.19930050416.Thermodynamic Properties and Structural Stability of Thorium Dioxide
. J. Phys. Condens. Matter. 24, 225801 (2012). doi: 10.1088/0953-8984/24/22/225801.First-Principles Study of Noble Gas Stability in ThO2
. J. Nucl. Mater. 490, 181-187 (2017). doi: 10.1016/j.jnucmat.2017.04.031.Charged Point Defects in the Flatland: Accurate Formation Energy Calculations in Two-Dimensional Materials
. Phys. Rev. X. 4, 031004 (2014). doi: 10.1103/PhysRevX.4.031044.Atomic Defects in Monolayer Titanium Carbide (Ti3C2Tx) MXene
. ACS Nano. 10, 9193-9200 (2016). doi: 10.1021/acsnano.6b05240.Discovery of Carbon-Vacancy Ordering in Nb4AlC3–x under the Guidance of First-Principles Calculations
. Sci. Rep. 5, 14192 (2015). doi: 10.1038/srep14192.Carbon Vacancies in Ti2CT2 MXenes: Defects or a New Opportunity?
Phys. Chem. Chem. Phys. 19, 31773-31780 (2017). doi: 10.1039/C7CP06593K.The online version of this article (https://doi.org/10.1007/s41365-019-0688-x) contains supplementary material, which is available to authorized users.