Mixed functionalization as a pathway to induce superconductivity in MXenes: vanadium and niobium carbide†
Received
27th February 2025
, Accepted 25th April 2025
First published on 25th April 2025
Abstract
The functionalization versatility of MXenes distinguishes them from other two-dimensional materials, enabling the design of numerous new materials with unique properties. By leveraging surface chemistry, functionalization allows for the manipulation of critical parameters such as the density of states and electron–phonon coupling, providing an excellent platform for exploring two-dimensional superconductivity. In this study, we investigate the impact of functionalization on vanadium carbide (V2C) MXene, which is intrinsically non-superconducting, by considering three different cases: (i) hydrogen adatoms, (ii) fluorine adatoms, and (iii) mixed functionalization with hydrogen and fluorine adatoms. We confirm the mechanical and dynamical stability of functionalized V2C using Born's stability criteria and phonon dispersion analyses. In all three cases, superconductivity emerges due to the presence of functional groups, which influence the electron–phonon interaction and electronic structure, leading to an enhanced electron–phonon coupling constant. The highest superconducting transition temperature is observed for mixed-functionalized V2C, attributed to the softening of the ZA phonon mode along with high-energy phonon modes induced by hydrogen. To further explore the potential of mixed functionalization in inducing superconductivity, we extend our approach to another non-superconducting MXene, Nb2C. The mixed-functionalized Nb2C exhibits a superconducting transition temperature of 9.2 K, which surpasses the reported values for Nb2CH2, Nb2CS2, and Nb2CBr2. These findings underscore the effectiveness of mixed functionalization in enabling superconductivity in MXenes, paving the way for future theoretical and experimental investigations.
1. Introduction
The dynasty of two-dimensional (2D) materials expanded by the inclusion of 2D layers of transition metal – carbides, nitrides, and carbonitrides, collectively known as MXenes.1,2 They exhibit variable surface chemistry and a wide range of electronic properties, making them promising candidates for ion batteries,3 supercapacitors,4 catalysis,5 sensing,6 thermoelectric materials,7 biomedicine,8 and energy storage devices.9 MXenes are typically synthesized from the corresponding MAX phases, where M stands for the transition metal and X stands for carbon or nitrogen, by selectively etching the element A (e.g., Al, Ga, Si, etc.) using hydrofluoric acid, molten inorganic salts, a mixture of fluoride salts and various acids.10–12 The etching results in 2D MX layers terminated with multiple functional groups, which reduce their chemical potential and increase their thermodynamic stability.13 The random distribution of terminal groups has been verified in various transition-metal carbides using electron-energy loss spectroscopy in transmission electron microscopy,14 nuclear magnetic resonance spectroscopy,15 and X-ray photoelectron spectroscopy.16 The presence of single or mixed terminations in MXenes depends on their synthesis or processing conditions.12,17 This capability to modify surface terminations enhances their extensive compositional diversity, setting them apart from other 2D materials.
Most of the MXenes exhibit metallic behavior, reflected from their electronic structure calculations, and tuning their electronic properties by surface functionalization makes them possible candidates for hosting superconductivity in the 2D limit.18 Some recent experimental and theoretical studies highlight the significance of surface terminations in inducing and enhancing superconductivity, leveraging the unsaturated surface chemistry of MXenes. Researchers are actively investigating the possible ways of surface functionalization of MXenes to explore superconductivity. For instance, pristine Nb2C has been reported as a non-superconducting material in experimental and theoretical studies.19,20 However, theoretical investigations have revealed the emergence of superconductivity in Nb2C through Cl adatoms, achieving a superconducting transition temperature (Tc) of approximately 6 K, which is further increased to 37.8 K with charge carrier doping.21 The experimental studies also report the induction of superconductivity in Nb2C under surface functionalization.19,22 Recent experimental reports have shown the feasibility of controlled surface functionalization of MXenes. For instance, Xia et al. synthesized fluorine-functionalized Mo2C by sintering a mixture of Mo2C and NH4F, observing an enhanced Tc of 5.7 K compared to the 3.2 K of pure Mo2C.17 Similarly, prior theoretical studies have demonstrated enhanced superconductivity in Mo2C through surface functionalization, with Tc values reaching up to 13 K for Br23 and H24 adatoms. In the nitride counterparts, such as Mo2N and W2N, hydrogenation significantly increases Tc from 16 K to 32.4 K and 10 K to 30.7 K, respectively.24 Further studies are needed to better understand the superconductivity induced by surface functionalization.25 Motivated by reports on superconductivity, we investigate the impact of functionalization on the superconducting properties of MXene, vanadium carbide (V2C), which is intrinsically non-superconducting.20
In this work, we investigate the structural, electronic, vibrational, and superconducting properties of V2C functionalized with (i) hydrogen, (ii) fluorine, and (iii) mixed hydrogen–fluorine, using density functional theory. The stability of these monolayers is confirmed by Born's mechanical stability criteria and the absence of imaginary frequencies in the phonon dispersion spectra. We also examine pristine V2C MXene, which is non-superconducting, consistent with previous reports.20 Upon functionalization, a significant enhancement in the electron–phonon coupling (EPC) constant is observed. Among the cases studied, mixed-functionalized V2C exhibits the highest EPC constant, corresponding to a superconducting transition temperature of 7 K. These findings highlight mixed functionalization as a promising strategy for inducing superconductivity in MXenes. To further explore this concept, we extend our study to Nb2C and observe an enhancement in its superconducting parameters.
2. Computational details
Theoretical calculations have been performed using the density functional theory (DFT) framework as implemented in the QUANTUM ESPRESSO package.26 The norm-conserving Pseudo Dojo pseudopotentials of the Perdew–Burke–Ernzerhof (PBE) type have been employed.27 The generalized gradient approximation of the Perdew–Burke–Ernzerhof type28 is used for the exchange–correlation functional. For all the calculations, convergence criteria are set to 1.0 × 10−4 Ry Å−1 for forces and 1.0 × 10−5 Ry for energy. In order to avoid coupling between periodic images, we incorporated a vacuum space of around 20 Å along the z-axis in the structural models. The kinetic-energy cutoff for the wavefunctions is set to 100 Ry. A k-mesh of 36 × 36 × 1 and a Methfessel–Paxton smearing with a width of 0.01 Ry are adopted for the structures. The elastic constants of pristine and functionalized V2C are calculated with the help of Thermo pw code.29 The phonon dispersion and electron–phonon coupling (EPC) calculations are performed by density functional perturbation theory (DFPT)30 using a q-mesh of 6 × 6 × 1. Using the density of states at the Fermi level (NF), the matrix elements of the electron–phonon interaction (gνk,k+q), and the phonon (ωνq) and electronic (εk) band structures obtained from our ab initio calculations, we computed the Eliashberg spectral function,31 |  | (1) |
for the different MXenes. The corresponding electron–phonon coupling (EPC) constant, λ, is obtained from;32,33 |  | (2) |
The superconducting transition temperature (Tc) is then estimated using the Allen–Dynes modified McMillan equation:34
|  | (3) |
where
ωlog is the logarithmically averaged phonon frequency, and
μ* represents the Coulomb pseudopotential, set to 0.1 in this study.
3. Results and discussion
3.1. Structural properties and stability analysis
Pristine V2C crystallizes in the P3m1 space group, similar to the 1T-MoS2 structure.35 As shown in Fig. 1(a), V and C atoms in pristine V2C are arranged in a triple-layer configuration along the hexagonal c-axis, with a C layer sandwiched between two V layers. The optimized lattice constant, a, is 2.891 Å, and the layer thickness d (defined by the V–V distance) is 2.174 Å, which is consistent with previously reported values.36 Two types of hollow configurations are present on the surfaces: site A, located directly above or below the V atoms, and site B, located directly above or below the C atoms (Fig. 1(a)). Full surface adsorption is known to be energetically more favorable than partial adsorption.37 The AA – adsorption configuration is also known to be the most stable for most MXenes.21,24,36,38 We verified this by examining the relative stability of different adsorption configurations (see Table S1 in the ESI†), and found that the AA – adsorption configuration is indeed the most energetically favorable.
 |
| Fig. 1 Top and side view of (a) V2C, (b) V2CH2, (c) V2CF2, and (d) V2CFH. Here grey, red, orange, and yellow balls represent vanadium (V), carbon (C), hydrogen (H), and fluorine (F) atoms, respectively. In (a), A and B indicate the two types of hollow sites and d is the layer thickness. | |
In this study, we consider V2CX2 structures (X = H, F, or mixed H and F) functionalized at the AA – adsorption configurations of V2C. The optimized crystal structures of V2CH2, V2CF2, and V2CFH are shown in Fig. 1(b)–(d), respectively, with structural parameters summarized in Table 1. Relative to pristine V2C, hydrogen functionalization leads to an in-plane contraction, reducing the lattice constant from 2.891 Å to 2.874 Å due to the shorter V–H bond (1.93 Å). In contrast, fluorine functionalization induces in-plane expansion, increasing the lattice constant to 2.986 Å because of the longer V–F bond (2.14 Å). The structural parameters for V2CF2 agree well with previous reports.36 For the mixed-functionalized V2C, the lattice constant increases to 2.956 Å, the V–V distance decreases to 2.067 Å, and the layer thickness expands to 4.371 Å, reflecting the combined influence of H and F. The V–C bond remains at 1.99 Å, indicating a balance between the two functional groups (Table 1).
Table 1 The calculated structural parameters and elastic constants for pristine and functionalized V2C, wherein a and d denote the lattice constant and thickness of the compounds, and X is the functional group
|
a (Å) |
V–C (Å) |
V–X (Å) |
d (Å) |
C
11 (N m−1) |
C
12 (N m−1) |
C
22 (N m−1) |
C
66 (N m−1) |
V2C |
2.891 |
1.99 |
— |
2.174 |
153.376 |
69.620 |
153.376 |
41.878 |
V2CH2 |
2.874 |
2.00 |
1.93 |
4.213 |
191.554 |
13.917 |
191.554 |
88.818 |
V2CF2 |
2.986 |
2.00 |
2.14 |
4.595 |
173.812 |
72.167 |
173.812 |
50.822 |
V2CFH |
2.956 |
1.99 |
1.98, 2.14 |
4.371 |
190.221 |
56.902 |
190.221 |
66.659 |
To evaluate the mechanical stability of these compounds under external load, we calculated their elastic constants. The Born mechanical stability criteria39 for 2D hexagonal crystals require C11 > 0, C66 > 0 (with 2C66 = C11 − C12), and C11 > C12. All the functionalized compounds V2CX2 (X = H, F, or mixed H and F) meet these criteria (see Table 1), suggesting that they may be synthesized under suitable experimental conditions. Notably, the elastic constants of these materials fall within the range observed for other 2D MXenes40 but are lower than those of graphene,41 indicating that they are less stiff and more flexible compared to graphene.
3.2. Electronic and vibrational properties
We calculated the band structure of pristine and functionalized V2C using DFT to investigate their electronic properties. The projected band structures are plotted along the high-symmetry path Γ–M–K–Γ in the left panel of Fig. 2 and 3, with the corresponding projected density of states (PDOS) shown in the rightmost panels. The band crossings at the Fermi level indicate the metallic behavior of these materials, though variations in the electronic band structures are observed depending on the functional groups.
 |
| Fig. 2 Projected electronic band structure and projected density of states of (a) V2C, (b) V2CH2, and (c) V2CF2. | |
 |
| Fig. 3 Projected electronic band structure and projected density of states of V2CFH. | |
In pristine V2C, the V-dz2, V-dzx, and V-dxy orbitals predominantly contribute to the bands near the Fermi level (Fig. 2(a)). In contrast, for V2CH2, the dominant contributions arise from V-dz2, V-dx2–y2, and V-dzy orbitals (Fig. 2(b)), while in V2CF2, the V-dz2 and V-dx2–y2 orbitals contribute primarily (Fig. 2(c)). Notably, a Dirac cone is observed above the Fermi level at approximately 0.8 eV at the K-point in V2CH2 (Fig. 2(b)), whereas in V2CF2, a Dirac cone appears around 0.3 eV below the Fermi level at the K-point (Fig. 2(c)), highlighting the potential of these materials for Dirac-physics-based applications. For the mixed-functionalized V2C (V2CFH), the V-dz2 and V-dx2–y2 orbitals (Fig. 3) predominantly contribute at the Fermi level. The dominance of the V-d orbitals in all compounds is further confirmed by the PDOS plots, where only minor contributions from carbon and the functional groups are observed (rightmost panels of Fig. 2 and 3).
The phonon dispersion of pristine and functionalized V2C is calculated using density functional perturbation theory (DFPT). The absence of imaginary frequencies in the phonon spectra indicates the dynamical stability of all compounds (Fig. 4–6).
 |
| Fig. 4 (a) and (b) Phonon dispersion color-coded with atoms' contributions, (c) projected phonon density of states, and (d) Eliashberg spectral function (solid blue region) and EPC constant (dotted grey line) of V2C. | |
 |
| Fig. 5 Phonon dispersion color-coded with atoms' contributions and projected phonon density of states, and Eliashberg spectral function (solid blue region) and EPC constant (dotted grey line) of (a)–(e) V2CH2, and (f)–(j) V2CF2. | |
 |
| Fig. 6 (a)–(e) Phonon dispersion color-coded with atoms' contributions, (f) projected phonon density of states, and (g) Eliashberg spectral function (solid blue region) and EPC constant (dotted grey line) of V2CFH. | |
 |
| Fig. 7 (a) Phonon dispersion and (b) Eliashberg spectral function with EPC constant of Nb2CFH. | |
The atom-resolved phonon dispersion is shown in the left panels, while the phonon density of states is presented in the rightmost panels. For V2C, the phononic bands corresponding to constituent atoms are well-separated: the low-frequency phonon modes (up to 400 cm−1) are primarily associated with V atom vibrations, while the high-frequency modes originate from C atom vibrations. These two sets of modes are separated by a phonon band gap of 254 cm−1 (31.491 meV) (Fig. 4(a)–(c)). Upon functionalization, this phononic band gap undergoes modifications.
In the case of hydrogen functionalization, additional high-frequency phonon modes emerge, leading to three distinct frequency regions in the phonon spectra: (i) low-frequency modes (up to 350 cm−1), primarily dominated by V atom vibrations. (ii) Intermediate-frequency modes (650 cm−1 to 730 cm−1), dominated by C atom vibrations. (iii) High-frequency modes (above 900 cm−1), originating from H atom vibrations. This results in two phononic gaps of 315 cm−1 and 198 cm−1, respectively (Fig. 5(a)–(d)).
Similarly, fluorine functionalization predominately leads to three frequency regions: (i) low-frequency phononic bands (up to 500 cm−1), arising from both V and F contributions. (ii) Intermediate-frequency modes dominated by C vibrations. (iii) High-frequency modes, also dominated by C vibrations. The functionalization introduces a phononic gap of 134 cm−1 between the first and second regions and another gap of 99 cm−1 between the second and third regions (Fig. 5(f)–(i)).
For mixed functionalization with hydrogen and fluorine (V2CFH), the phonon spectrum exhibits four distinct regions: (i) a low-frequency region (up to 500 cm−1) with contributions from both V and F atoms. (ii) Two intermediate-frequency regions: 619 cm−1 to 654 cm−1 and 724 cm−1 to 768 cm−1, dominated by C vibrations. (iii) Two high-frequency regions: 850 cm−1 to 914 cm−1 and 996 cm−1 to 1058 cm−1, associated with H vibrations (Fig. 6(a)–(f)). V2CFH features four phononic gaps: one large gap of 120 cm−1 and three smaller gaps of 74 cm−1, 85 cm−1, and 82 cm−1. Notably, in both fluorine-functionalized (V2CF2) and mixed-functionalized (V2CFH) cases, phonon softening occurs in the V-based ZA mode, particularly near the M point and along the K–Γ direction (Fig. 5(f) and 6(a)). In summary, functionalization modifies the phonon spectrum by introducing additional modes at higher frequencies and inducing phonon softening in the lower-frequency region. These changes may influence electron–phonon interactions and potentially affect superconducting properties.
3.3. Induced superconductivity in V2C via functionalization
To investigate the superconducting properties of pristine and functionalized V2C, we calculated their Eliashberg spectral function and electron–phonon coupling (EPC) constant from our DFPT results. We obtained a weak EPC constant of 0.28 for pristine V2C (Fig. 4(d)), corresponding to a superconducting transition temperature (Tc) of 0.08 K, which is in agreement with earlier theoretical reports.20 The EPC constant increases from 0.28 to 0.33 upon hydrogen functionalization of V2C, corresponding to an increased superconducting transition temperature of Tc = 1.53 K (Fig. 5(e) and Table 2).
Table 2 The average electron–phonon coupling constant λ and the estimated superconducting transition temperature Tc (K) for V2C, V2CH2, V2CF2, and V2CFH
|
λ
|
T
c (K) |
V2C |
0.28 |
0.08 |
V2CH2 |
0.33 |
1.53 |
V2CF2 |
0.55 |
4.5 |
V2CFH |
0.6 |
7.0 |
Nb2CFH |
0.74 |
9.2 |
Up to the intermediate region, the EPC constant of V2CH2 remains comparable to that of pristine V2C (0.28), as the phonon modes are predominantly associated with vanadium and carbon vibrations. However, due to the lighter mass of hydrogen, high-frequency phonon modes are introduced, providing additional EPC channels and enhancing the EPC constant to 0.33. It is worth noting that the Tc enhancement in V2C is lower than that reported for other hydrogenated MXenes (Mo2CH2, Mo2NH2), likely due to the presence of phonon softening in those cases.24 Upon fluorine functionalization, the EPC constant for V2CF2 increases to 0.55 (Fig. 5(j)), leading to a higher superconducting transition temperature (Tc) of 4.5 K. The enhancement in λ arises from two key factors: (i) the softening of the ZA phonon mode, which increases λ due to its frequency dependence (see Section 2), and (ii) the increased spectral weight in the low-frequency region (up to 500 cm−1). The Eliashberg spectral function shows that this low-frequency region, dominated by vanadium and fluorine vibrations, contributes 94% (0.52 out of 0.55) to the total EPC constant. Compared to pristine V2C and V2CH2, fluorine functionalization introduces additional phonon modes in the low-frequency region that couple with vanadium vibrations, playing a dominant role in enhancing the EPC constant and increasing Tc (Fig. 5(j) and Table 2). Upon mixed functionalization with hydrogen and fluorine, the EPC constant further increases to ∼ 0.6 (Fig. 6(g)), corresponding to a Tc of 7 K. The enhancement in superconductivity arises from the synergistic effect of fluorine-induced softening of the ZA phonon mode in the low-frequency region and the introduction of high-frequency phonon modes by hydrogen, which creates additional EPC channels. This study suggests that multi-element functionalization of MXenes could be a promising strategy for enhancing superconductivity. Further experimental and theoretical investigations are required to explore this effect in other MXene materials.
3.4. Inducing superconductivity in Nb2C via mixed functionalization
We further extended our investigation of mixed functionalization to Nb2C, a material previously reported to be non-superconducting both theoretically and experimentally.19,20 As discussed above, AA – adsorption configurations are the most favorable for MXenes; hence, we consider the AA – adsorption configuration for mixed functionalization of Nb2C with fluorine and hydrogen. The calculated lattice constant of Nb2CFH is 3.193 Å, and the electronic band structure indicates metallic behavior (ESI† Fig. S1). The density of states at the Fermi level is 1.07 states per eV (Fig. S1, ESI†). We then calculated the phonon dispersion of Nb2CFH to assess its stability. The absence of imaginary frequencies confirms its dynamical stability (Fig. 7(a)). Finally, the Eliashberg spectral function was computed to evaluate the electron–phonon coupling, revealing an enhanced EPC constant of 0.74 (Fig. 7(b)), corresponding to a Tc of 9.2 K (Table 2). The observed enhancement in EPC constant and Tc suggests a synergistic effect of fluorine and hydrogen vibrations, surpassing the EPC constant achieved in other single-atom functionalized Nb2C systems such as Nb2CH2, Nb2CBr2, Nb2CS2, and Nb2CCl2, and approaches the highest reported Tc in Nb2CO2 (12 K).21,38 These findings highlight the potential of mixed functionalization as an effective strategy for inducing and enhancing superconductivity in MXenes, offering a promising approach for tuning their superconducting properties.
4. Conclusions
We investigate the electronic structure, vibrational properties, and electron–phonon coupling (EPC) in V2C-based MXenes, including pristine V2C and functionalized V2CX2 (X = H, F, or mixed H–F functionalization). While pristine V2C exhibits negligible EPC, rendering it unsuitable as a phonon-mediated superconductor, functionalization significantly enhances the EPC constant and induces superconductivity. In the case of mixed functionalization, the EPC constant reaches λ = 0.6, corresponding to a superconducting transition temperature (Tc) of 7 K, attributed to the synergistic vibrations of fluorine and hydrogen and the softening of the ZA phonon mode. Extending our study to another non-superconducting MXene, Nb2C, we find that mixed functionalization further enhances the EPC constant to 0.74, resulting in a Tc of 9.2 K—higher than previously reported values for functionalized Nb2C, including Nb2CH2, Nb2CCl2, Nb2CS2, and Nb2CBr2. These findings highlight the potential of mixed functionalization in MXenes for inducing superconductivity, offering promising avenues for designing high-performance two-dimensional superconductors.
Author contributions
Prarena Jamwal: conceptualization, methodology, data curation, investigation, formal analysis, visualization, validation, writing – original draft, writing – reviewing & editing. Rajeev Ahuja: writing – reviewing. Rakesh Kumar: supervision, writing – reviewing.
Data availability
The data supporting this study were obtained from first-principles density functional theory (DFT) calculations. All relevant computational details and results are provided in the manuscript and ESI.† Additional data can be made available upon request.
Conflicts of interest
There are no conflicts to declare.
Acknowledgements
P. J. acknowledges the National Supercomputing Mission (NSM) for providing computing resources of ‘PARAM SMRITI’ at NABI, Mohali, which is implemented by C-DAC and supported by the Ministry of Electronics and Information Technology (MeitY) and Department of Science and Technology (DST), Government of India. R. K. would like to acknowledge MeiTY (approval no. Y-20/8/2024 R&D-E) for the support.
Notes and references
- M. Naguib, V. N. Mochalin, M. W. Barsoum and Y. Gogotsi, Adv. Mater., 2014, 26, 992–1005 CrossRef CAS PubMed.
- A. VahidMohammadi, J. Rosen and Y. Gogotsi, Science, 2021, 372, eabf1581 CrossRef CAS PubMed.
- M. Narayanasamy, S. Zaman and C. M. Koo, Mater. Today Energy, 2023, 37, 101405 CrossRef CAS.
- M. Hu, H. Zhang, T. Hu, B. Fan, X. Wang and Z. Li, Chem. Soc. Rev., 2020, 49, 6666–6693 RSC.
- Á. Morales-García, F. Calle-Vallejo and F. Illas, ACS Catal., 2020, 10, 13487–13503 CrossRef.
- D. H. Ho, Y. Y. Choi, S. B. Jo, J.-M. Myoung and J. H. Cho, Adv. Mater., 2021, 33, 2005846 CrossRef CAS PubMed.
- M. Khazaei, M. Arai, T. Sasaki, M. Estili and Y. Sakka, Phys. Chem. Chem. Phys., 2014, 16, 7841–7849 RSC.
- I.-C. Lee, Y.-C. E. Li, J. L. Thomas, M.-H. Lee and H.-Y. Lin, Mater. Horiz., 2024, 11, 876–902 RSC.
- S. Sun, C. Liao, A. M. Hafez, H. Zhu and S. Wu, Chem. Eng. J., 2018, 338, 27–45 CrossRef CAS.
- M. Naguib, M. W. Barsoum and Y. Gogotsi, Adv. Mater., 2021, 33, 2103393 CrossRef CAS PubMed.
- Y. Wei, P. Zhang, R. A. Soomro, Q. Zhu and B. Xu, Adv. Mater., 2021, 33, 2103148 CrossRef CAS PubMed.
- 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.
- B. Anasori, M. R. Lukatskaya and Y. Gogotsi, Nat. Rev. Mater., 2017, 2, 16098 CrossRef CAS.
- L. H. Karlsson, J. Birch, J. Halim, M. W. Barsoum and P. O. Å. Persson, Nano Lett., 2015, 15, 4955–4960 CrossRef CAS PubMed.
- K. J. Harris, M. Bugnet, M. Naguib, M. W. Barsoum and G. R. Goward, J. Phys. Chem. C, 2015, 119, 13713–13720 CrossRef CAS.
- J. Halim, K. M. Cook, M. Naguib, P. Eklund, Y. Gogotsi, J. Rosen and M. W. Barsoum, Appl. Surf. Sci., 2016, 362, 406–417 CrossRef CAS.
- H. Xia, L. Zhi, G. Feng, H. Cao, X. Xu, S. Zhang, J. Feng, C. Li, J. Li and P. Zhang, J. Phys. Chem. Solids, 2024, 193, 112163 CrossRef CAS.
- M. Khazaei, A. Ranjbar, M. Arai, T. Sasaki and S. Yunoki, J. Mater. Chem. C, 2017, 5, 2488–2503 RSC.
- 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.
- J. Bekaert, C. Sevik and M. V. Milošević, Nanoscale, 2020, 12, 17354–17361 RSC.
- C. Sevik, J. Bekaert and M. V. Milošević, Nanoscale, 2023, 15, 8792–8799 RSC.
- K. Wang, H. Jin, H. Li, Z. Mao, L. Tang, D. Huang, J.-H. Liao and J. Zhang, Surf. Interfaces, 2022, 29, 101711 CrossRef CAS.
- J.-J. Zhang and S. Dong, J. Chem. Phys., 2017, 146, 034705 CrossRef PubMed.
- J. Bekaert, C. Sevik and M. V. Milošević, Nanoscale, 2022, 14, 9918–9924 RSC.
- M. Mozafari and M. Soroush, Mater. Adv., 2021, 2, 7277–7307 RSC.
- P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, G. L. Chiarotti, M. Cococcioni, I. Dabo, A. D. Corso, S. de Gironcoli, S. Fabris, G. Fratesi, R. Gebauer, U. Gerstmann, C. Gougoussis, A. Kokalj, M. Lazzeri, L. Martin-Samos, N. Marzari, F. Mauri, R. Mazzarello, S. Paolini, A. Pasquarello, L. Paulatto, C. Sbraccia, S. Scandolo, G. Sclauzero, A. P. Seitsonen, A. Smogunov, P. Umari and R. M. Wentzcovitch, J. Phys.: Condens. Matter, 2009, 21, 395502 CrossRef PubMed.
- M. van Setten, M. Giantomassi, E. Bousquet, M. Verstraete, D. Hamann, X. Gonze and G.-M. Rignanese, Comput. Phys. Commun., 2018, 226, 39–54 CrossRef CAS.
- J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef CAS PubMed.
- A. Dal Corso, J. Phys.: Condens. Matter, 2016, 28, 075401 CrossRef PubMed.
- S. Baroni, S. de Gironcoli, A. Dal Corso and P. Giannozzi, Rev. Mod. Phys., 2001, 73, 515–562 CrossRef CAS.
- W. L. McMillan, Phys. Rev., 1968, 167, 331–344 CrossRef CAS.
- P. B. Allen and R. C. Dynes, Phys. Rev. B, 1975, 12, 905–922 CrossRef CAS.
- P. B. Allen, Phys. Rev. B, 1972, 6, 2577–2579 CrossRef CAS.
- R. Dynes, Solid State Commun., 1972, 10, 615–618 CrossRef CAS.
- Y. Linghu, N. Li, Y. Du and C. Wu, Phys. Chem. Chem. Phys., 2019, 21, 9391–9398 RSC.
- A. Champagne, L. Shi, T. Ouisse, B. Hackens and J.-C. Charlier, Phys. Rev. B, 2018, 97, 115439 CrossRef CAS.
- 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.
- Y. Yang and C. S. Ting, J. Phys. D: Appl. Phys., 2020, 53, 485301 CrossRef CAS.
- F. Mouhat and F. M. C.-X. Coudert, Phys. Rev. B: Condens. Matter Mater. Phys., 2014, 90, 224104 CrossRef.
- U. Yorulmaz, A. Özden, N. K. Perkgöz, F. Ay and C. Sevik, Nanotechnology, 2016, 27, 335702 CrossRef PubMed.
- D. Çakr, F. M. Peeters and C. Sevik, Appl. Phys. Lett., 2014, 104, 203110 CrossRef.
|
This journal is © the Owner Societies 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.