Pengfei
Li
a,
Rulong
Zhou
*b and
Xiao Cheng
Zeng
*ac
aDepartment of Chemical Physics, University of Science and Technology of China, Hefei, Anhui 230026, P. R. China. E-mail: xzeng1@unl.edu
bSchool of Science and Engineering of Materials, Hefei University of Technology, Hefei, Anhui 230009, P. R. China. E-mail: rlzhou@hfut.edu.cn
cDepartment of Chemistry and Nebraska Center for Materials and Nanoscience, University of Nebraska-Lincoln, Lincoln, Nebraska 68588, USA
First published on 5th August 2014
The most stable structures of two-dimensional (2D) silicon–carbon monolayer compounds with different stoichiometric compositions (i.e., Si:
C ratio = 2
:
3, 1
:
3 and 1
:
4) are predicted for the first time based on the particle-swarm optimization (PSO) technique combined with density functional theory optimization. Although the 2D Si–C monolayer compounds considered here are rich in carbon, many of the low-energy metastable and the lowest-energy silicon–carbon structures are not graphene (carbon monolayer) like. Phonon-spectrum calculations and ab initio molecular dynamics simulations were also performed to confirm the dynamical stability of the predicted most stable 2D silicon–carbon structures as well their thermal stability at elevated temperature. The computed electronic band structures show that all three predicted silicon–carbon compounds are semiconductors with direct or indirect bandgaps. Importantly, their bandgaps are predicted to be close to those of bulk silicon or bulk germanium. If confirmed in the laboratory, these 2D silicon–carbon compounds with different stoichiometric compositions may be exploited for future applications in nanoelectronic devices.
The desire for the continuous miniaturization of electronic devices calls for the development of new and novel low-dimensional materials. Besides graphene and silicene, a wide range of 2D materials, particularly monolayer sheets with atomic thickness, have been reported in the literature. Coleman et al.13 developed a liquid exfoliation technique that can efficiently produce monolayer 2D nanosheets from a variety of inorganic layered materials such as boron nitride (BN), molybdenum disulfide (MoS2), tungsten disulfide (WS2), molybdenum diselenide (MoSe2) and molybdenum telluride (MoTe2). Using a modified liquid exfoliation technique, Xie et al. successfully fabricated monolayer vanadium disulfide (VS2) and tin disulfide (SnS2) in the laboratory.14–16 On the theoretical side, increasing efforts have been devoted to predicting the structures and functional properties of novel 2D materials, such as monolayer boron sheets with low-buckled configurations,17,18 monolayer boron–carbon (BC) compounds,19,20 boron–silicon (BSi) compounds,21 aluminum carbon (AlC) compounds,22 carbon nitride (CN),23 germanene,24,25 tetragonal TiC,26 SnC27 and other group III–VI compounds.27,28
2D silicon–carbon (Si–C) monolayers can be viewed as composition-tunable materials between the pure 2D carbon monolayer – graphene – and the pure 2D silicon monolayer – silicene. Efforts have been made towards predicting the most stable structures of the SiC sheet. Recently, Li et al.29 and Zhou et al.30 reported a metallic pt-SiC2 2D sheet and semiconducting g-SiC2 siligraphene, respectively, based on density functional theory (DFT) calculations. Their studies indicate that the electronic properties of 2D silicon–carbon compounds can be strongly dependent on the structure and the stoichiometry. Moreover, few theoretical studies on 2D silicon–carbon compounds with different stoichiometries have been reported in the literature. 2D silicon–carbon sheets with different stoichiometric compositions are expected to possess different electronic properties from SiC and SiC2 sheets. Thus, it is timely to search for new 2D structures of silicon–carbon compounds with distinct stoichiometries and explore their structure–property relationships. In this study, we perform a comprehensive search for structures of 2D Si–C compounds with stoichiometric compositions (Si:
C ratios) of 2
:
3, 1
:
3 and 1
:
4, using particle-swarm optimization (PSO) techniques combined with density functional theory optimization. Our calculations suggest that the 2D Si–C compounds with higher carbon content over silicon are energetically more favored. The predicted lowest-energy structures of Si2C3-I, SiC3-I and SiC4-I exhibit semiconducting characteristics. Phonon-spectrum calculations and ab initio MD simulations further confirm the dynamic and thermal stability of the lowest-energy 2D structures. Finally, we show that the computed elastic constants of Si–C sheets are between those of graphene and silicene, suggesting that these newly predicted 2D Si–C compounds also possess good elastic properties.
The structure relaxation and total-energy calculations were performed using the VASP package38 within the generalized gradient approximation (GGA). An energy cutoff of 450 eV and an all-electron plane-wave basis set within the projector augmented wave (PAW) method were used. A dense k-point sampling with the grid spacing less than 2π × 0.04 Å−1 in the Brillouin zone was taken. To prevent interaction between the adjacent solid sheets, a 20 Å vacuum spacing was set along the direction (i.e., the direction normal to the monolayer). For the geometric optimization, both the lattice constants and atomic positions were relaxed until the forces on the atoms were less than 0.01 eV Å−1 and the total energy change was less than 1 × 10−5 eV. Phonon spectra of the low-energy crystalline structures were computed using the VASP package coupled with the PHONOPY program.39 The phonon spectrum calculation was to assure that the obtained 2D sheets entailed no negative phone modes.
To evaluate the relative stabilities among the predicted 2D C–Si compounds, we have computed their cohesive energy. The formula of cohesive energy for the 2D systems is defined as follows:
Ecoh = (xESi + yEC − ESixCy)/(x + y) |
2D Structure | Cohesive energy (eV per atom) |
---|---|
Si2C3-I | 7.2660 |
Si2C3-II | 7.1712 |
Si2C3-III | 7.1446 |
SiC3-I | 7.8561 |
SiC3-II | 7.8409 |
SiC3-III | 7.8365 |
SiC4-I | 8.0631 |
SiC4-II | 8.0434 |
Next, to ensure that the predicted lowest-energy structure for each Si:
C ratio is dynamically stable, phonon spectra of all of the three lowest-energy structures were computed using the supercell frozen phonon theory implemented in the PHONOPY program. The computed phonon spectra of the lowest-energy structures of Si2C3, SiC3 and SiC4 (Si2C3-I, SiC3-I and SiC4-I) are plotted in Fig. 2. Clearly, no negative phonon frequencies are present over the entire Brillouin zones for all three of the lowest-energy structures, indicating the inherent dynamical stability of these 2D Si–C sheets.
Moreover, the thermal stability of the Si2C3-I, SiC3-I and SiC4-I structures was examined using ab initio molecular dynamics (AIMD) simulations. In the AIMD simulation, the canonical ensemble (NVT ensemble) is adopted. The AIMD time step was 2 fs and the total simulation time was 15 ps for each given temperature. The structural features of each Si–C sheet prior to and after melting are shown in Fig. 3. It can be seen that the equilibrium structures of the Si2C3-I and SiC3-I sheets at the end of the 15 ps AIMD simulation show no sign of structural disruption at 3500 K, whereas both sheets exhibit disrupted structures at 4000 K. Thus we can conclude that the Si2C3-I and SiC3-I structures can maintain their structure integrity and planar geometry below 3500 K. The SiC4-I sheet appears to have the highest thermal stability among the three structures, as SiC4-I can still keep its geometric structure over the 15 ps AIMD simulation with the temperature controlled at 4000 K.
![]() | ||
Fig. 3 Snapshots of the three lowest-energy 2D Si–C compounds at the end of two independent 15 ps AIMD simulations: (a) Si2C3-I, (b) SiC3-I and (c) SiC4-I sheets. |
The Si2C3-II sheet is 94.8 meV per atom higher in energy than the Si2C3-I sheet, although all of the polygonal rings in the Si2C3-II sheet are hexagonal. Notably, this Si2C3-II sheet can be viewed as silicon-doped graphene. In contrast to the Si2C3-I sheet where all of the Si atoms are located separately (no Si–Si bonds were found in the sheet), there are both separately-distributed Si atoms and Si dimers in the Si2C3-II sheet. The percentage of Si atoms forming Si dimers is 50%. Since the sp2-hybridization is not favored by silicon Si–Si bonds, a planar structure should be energetically unfavorable, which is possibly a major reason why Si2C3-II is less stable than Si2C3-I.
The third lowest-energy structure of Si2C3, namely the Si2C3-III sheet, is 121.4 meV per atom higher in energy than the Si2C3-I sheet. Apparently, the Si2C3-III sheet is composed of pentagonal, hexagonal and octagonal rings, where each octagonal ring is surrounded by four pentagonal and four hexagonal rings. In this structure, all of the Si atoms form Si dimers so that its energy is much higher than that of the Si2C3-I or Si2C3-II sheet.
The structure of the SiC3-II sheet is also graphene like.41 The Si atoms in the SiC3-II sheet are also located separately, as in the SiC3-I sheet (see Fig. 1e). So, it is surprising that the SiC3-II sheet is 15.2 meV per atom higher in energy than the SiC3-I sheet. A closer examination of the structure indicates that the only difference between the structure of the SiC3-I and SiC3-II sheets is the location of the two Si atoms in every hexagonal ring. In the SiC3-I sheet, the two Si atoms are located at the 1 and 3 sites of every hexagonal ring (we denote the six sites of any hexagonal ring as site 1 to site 6), while in the SiC3-II sheet they are located at the 1 and 4 sites. From the viewpoint of doping, the SiC3-II sheet can be viewed as having 25% of the A-site and 25% of the B-site carbon atoms of the graphene substituted by Si atoms. The different location distribution of the Si atoms leads to more of the C atoms being connected with one another in the SiC3-I sheet compared to SiC3-II, which should be the main reason why the SiC3-I sheet has a lower energy than the SiC3-II sheet. Due to the different Si distributions, the structure of the SiC3-II sheet cannot be decomposed into C chains and Si–C chains, while that of the SiC3-I sheet can.
The structure of the SiC3-III sheet is very different from those of the SiC3-I and SiC3-II sheets (see Fig. 1f). The SiC3-III sheet is composed of octagonal, hexagonal, and pentagonal rings, and possesses a much higher symmetry compared to the SiC3-I and SiC3-II sheets. The Si atoms in the SiC3-III sheet form dimers while the C atoms form complete hexagonal rings. It is known that Si dimers in a planar structure are energetically not favored whereas C hexagonal rings are favored. So, even though Si–Si bonds exist, the total energy of the SiC3-III sheet is only a little higher than that of SiC3-II sheet.
Finally, we make a comparison of the C–C/C–Si bond length in graphene/SiC and the SiC compounds reported here. The C–C bond lengths in Si2C3-I, SiC3-I, and SiC4-I are 1.438 Å, 1.455 Å, and 1.432 Å respectively, slightly longer than that in graphene (1.42 Å). The C–Si bonds are slightly longer in Si2C3-I (1.792 Å) than those either in the SiC sheet (1.786 Å), or in SiC3-I (1.781 Å) and SiC4-I (1.770 Å). We have also computed the Si–Si distance between two parallel stacked (in registry) Si2C3-I monolayers. As shown in ESI Fig. S1,† the minimum Si–Si distance is about 3.4 Å. Hence, new Si–Si bonds are not expected to form when two Si2C3-I monolayers are stacked on top of one another.
In summary, although the 2D Si–C compounds considered here are all C-rich, most of the lowest-energy structures and low-energy metastable structures are not akin to Si-doped graphene. Pentagonal and heptagonal rings are occasionally formed in the lowest-energy structures, and octagonal rings normally appear in the metastable structures. The Si atoms tend to be located separated from one another, i.e. the Si atoms prefer to be bonded with C atoms but not Si atoms, which is a main factor that influences the relative stability of the 2D Si–C structures.
![]() | ||
Fig. 4 Computed electronic band structures of (a) Si2C3-I, (b) SiC3-I and (c) SiC4-I monolayer sheets. The Fermi level is set to 0 eV. |
2D Structure | Bandgap | |
---|---|---|
GGA | HSE06 | |
Si2C3-I | 0.83 eV (D) | 1.37 eV (D) |
SiC3-I | 0.86 eV (D) | 1.40 eV (D) |
SiC4-I | 0.14 eV (I) | 0.51 eV (I) |
The computed partial density of states (PDOSs) of the predicted 2D Si–C compounds was also analyzed. The representative PDOSs for Si2C3-I, SiC3-I and SiC4-I sheets are plotted in Fig. 5. It is clear that in all three cases the higher valence bands and lower conduction bands (about −2.0 to 2.5 eV of the energy windows) are contributed by the sp2 orbitals of Si and C, while the pz orbitals of Si and C only have contributions to the lower valence bands (below −2.0 eV) and higher conduction bands. So, the electronic properties of these sheets are only determined by the in-plane σ and σ* bonds rather than the π and π* states, in contrast to graphene and graphite where the conjugate π states have a major influence on the electronic properties such as excellent conductivity.
![]() | ||
Fig. 5 Computed PDOSs for (a) Si2C3-I, (b) SiC3-I and (c) SiC4-I monolayer sheets. The Fermi level is set at 0 eV. |
For the Si2C3-I sheet, it is clearly shown that the valence band maximum (VBM) is mainly contributed by the s, px and py orbitals of the C atoms, while the contribution of Si is about half that of C. On the other hand, the conduction band minimum (CBM) is mainly contributed by Si atoms and the contribution of C atoms is about half that of Si. For the SiC3-I sheet, both the VBM and CBM are contributed mainly by the sp2 hybridization states of C. And for the SiC4-I sheet, it is obvious that the C and Si atoms contribute to the VBM and CBM nearly equally. It is worth noting that there are such large differences in the contribution to the VBM and CBM states for different 2D Si–C compounds.
To gain a deeper understanding of the nature of the bonding for the predicted 2D Si–C compounds, we applied the electron localization function (ELF) analysis, which can be used to classify chemical bonds rigorously. Due to the more localized characteristic of σ states than π states, a relatively large value of ELF distribution (e.g., 0.725) for the Si–C compounds can mostly characterize the in-plane σ states. The plotted iso-surfaces of ELFs for the lowest-energy Si–C sheets are shown in Fig. 6. It can be seen that in all the cases the ELFs of the C–C bonds are localized just at the center of the bonds and those of the Si–C bonds are localized closer to the C atoms. This is due to the fact that the electronegativity of the C atoms is stronger than that of the Si atoms. For the Si2C3-I sheet, apparently there are more Si–C bonds. The incline of ELFs to the C atoms suggests that both the VBM and CBM states mainly originate from the in-plane sp2 hybridization states of C and Si, respectively. For the other two cases, most of the C atoms connect with one another, forming C chains. The composition of Si–C bonds in SiC3-I and SiC4-I sheets are lower than that in the Si2C3-I sheet, hence the charge transfer from Si to C is weaker as reflected by the fact that the contribution of the Si sp2 states to the VBM is minor in both cases.
![]() | ||
Fig. 6 Iso-surfaces of the ELF with the value of 0.725 for (a) Si2C3-I, (b) SiC3-I and (c) SiC4-I sheets. |
2D Structure | Elastic constants (GPa) | |||
---|---|---|---|---|
c 11 | c 12 | c 22 | c 66 | |
Si2C3-I | 453.8 | 171.0 | 473.6 | 160.2 |
SiC3-I | 594.5 | 180.8 | 640.2 | 231.5 |
SiC4-I | 629.4 | 192.9 | 584.6 | 197.4 |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c4nr03247k |
This journal is © The Royal Society of Chemistry 2014 |