Next Article in Journal
Anisotropy of X-ray Absorption Cross Section in CeCoGe3 Single Crystal
Next Article in Special Issue
Fe2O3 Microcubes Derived from Metal–Organic Frameworks for Lithium-Ion Storage with Excellent Performance
Previous Article in Journal
Electron Polarons in Lithium Niobate: Charge Localization, Lattice Deformation, and Optical Response
Previous Article in Special Issue
Strain-Induced Tunable Band Offsets in Blue Phosphorus and WSe2 van der Waals Heterostructure
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

SiC3 as a Charge-Regulated Material for CO2 Capture

School of Metallurgy Engineering, Jiangxi University of Science and Technology, Ganzhou 341000, China
*
Authors to whom correspondence should be addressed.
Crystals 2021, 11(5), 543; https://doi.org/10.3390/cryst11050543
Submission received: 25 April 2021 / Revised: 11 May 2021 / Accepted: 11 May 2021 / Published: 13 May 2021
(This article belongs to the Special Issue 2D Crystalline Monolayer Nanosheets)

Abstract

:
The increasing CO2 emission rate is deteriorating the atmospheric environment, leading to global warming and climate change. The potential of the SiC3 nanosheet as a functioning material for the separation of CO2 from the mixture of CO2, H2, N2 and CH4 by injecting negative charges is studied by DFT calculations in this paper. The results show that in the absence of injecting negative charges, CO2 interacts weakly with the SiC3 nanosheet. While the interaction between CO2 and the SiC3 nanosheet can be strengthened by the injection of negative charges, the absorption mechanism of CO2 changes from physisorption to chemisorption when the injection of negative charges is switched on. H2/N2/CH4 are all physiosorbed on the SiC3 nanosheet with/without the injection of negative charges. The mechanism of CO2 adsorption/desorption on the SiC3 nanosheet could be tuned by switching on/off the injection of negative charges. Our results indicate that the SiC3 nanosheet can be regarded as a charge-regulated material for the separation of CO2 from the CO2/H2/N2/CH4 mixture.

1. Introduction

As the byproduct of human modernization activities, CO2 is regarded as one of the significant causes of global warming and climate change. One of the main sources of carbon emissions comes from the use of coal as fuel or coke to reduce ores during extraction metallurgy. Taking iron and steel making for example, about 1.3~1.7 tons of CO2 were exhausted per ton of steel produced in 2020, equating to about 8.0 percent of global carbon dioxide emissions. In the ferrous industry, most of the exhaust gases are reducing the atmosphere and contain the bulk of CO2 that is emitted at temperatures higher than the ambient temperature. The typical blast furnace gas composition is N2, CO, CO2 and H2 with the temperature of about 473~530 K, and the coke oven gas composition is N2, CH4, CO, CO2 and H2 with the temperature of about 1000 to 1300 K. CO2 in the hot exhaust gas is expected to be collected and then reduced to CO that can facilitate the reduction reaction of CO2 into synthetic fuels for a sustainable environment [1,2,3,4]. In the process of oxygen-converter steelmaking, CO2 is a weaker oxidizing agent compared with O2, and CO2 can be reduced to CO by the C in the liquid hot metal [5,6]. Therefore, it is expected to develop an adsorbent material with the highly selective, controllable, and reversible of CO2 capture that is high-temperature-resistant in the adsorption/desorption processes to fulfill the requirement of being in the metallurgy area.
Many 2D materials, due to their high surface area for CO2 capture, have been synthesized or theoretically predicted [7], such as BN2 [8], CN [9], C2N [10], C3N [11,12], PCn [13], and Me–N–C (Me = Fe, Cu, Co, etc.) nanosheets [14,15], borophene nanosheet [16,17], covalent triazine frameworks [18], nanoporous grapheme [19], and penta-graphene [20]. Adding/removing the electrons to/from the adsorbent materials allows alteration of the affinity between the adsorbent materials and gases that is in favor of the CO2 capture from the gas mixture [21]. Qin et al. [10] showed that the interaction between CO2 and the C2N nanosheet is enhanced by the negative charges or external electric field, and CO2 molecules are released from the C2N nanosheet once the charge state/electric field is switched off. Li et al. [11] found that CO2 molecules can be adsorbed and desorbed from the C3N nanosheet by adjustment of the applying charge density. Tao et al. [22] found that the charged calcite is highly selective for separating CO2 from the mixture of N2, H2 and CH4, and the optimal charge density range for CO2 capture and separation is 8.04~18.56 × 1013 e/cm2. Other previous studies [8,10,11,13,14,15,17,18,21,22,23,24,25] also showed that the negative charges could be used as a switch in CO2 capture and separation techniques by changing the affinity between CO2 and the substrates, such as penta-BN2, g-C3N4, nitrogen-doped porous carbons, etc.
Recently, works by Li and Shi et al. [26,27] showed that the silicon carbon monolayer possesses high thermal stabilities such that the Si2C3 and SiC3 sheets can retain their planar geometries below 3500 K and manifest excellent properties of semiconductivity and elasticity for applications in electronics and optoelectronics. Chabi et al. [28] found that 2D silicon carbide (SixCy) is a universal material, and exhibits the same properties as graphene, silicon or silicon carbide, depending on the composition (i.e., Si2C3, SiC3 and SiC4). For extensive exploration of novel nanostructures as potential materials for CO2 capture, a question raised in this study is whether SiC3 nanosheets are also a promising material for CO2 capture, and whether its capture/separation could be efficiently tuned by the charge/electric field.
Therefore, the adsorption behaviors of CO2, H2, N2 and CH4 on the SiC3 nanosheets with and without applying the injecting negative charges were studied by Density Functional Theory (DFT) calculations. Firstly, the stability of SiC3 nanosheet with different gas molecule (i.e., CO2, H2, N2 and CH4) adsorption was explored. Secondly, the effects of the injecting negative charges on the electronic structure, bonding features of adsorption systems and the mechanism of the CO2 adsorption/desorption on the SiC3 nanosheet were studied. Finally, the possibility of CO2 separation from the gas mixture with the assistance of negative charge, and the SiC3 nanosheet as a promising CO2 capture material were confirmed.

2. Calculation Method and Details

The calculations about the adsorption behavior of gas (i.e., CO2, H2, N2 and CH4) on the SiC3 nanosheets were performed by using Density Functional Theory (DFT) methods, the implemented Dmol3 module [29] with Generalized Gradient Approximation (GGA) and the Perdew–Burke–Ernzerhof (PBE) functional exchange correction functional. The details about the PBE functional have been well documented in the previous paper [30]. The effects of the core electrons were treated as a single effective potential by using the DFT semi-core pseudopotentials (USPP). Double numerical plus polarization (DNP) with a cutoff radius of 4.7 Å, and the DFT + D method with the weak van der Waals (vdW) correction were adopted in our calculations. Additionally, 4 × 4 × 1 k-points sampling of the SiC3 nanosheet with 18 Si and 54 C atoms was adopted, and a 20 Å vacuum was set along the normal direction of the nanosheet for the elimination of the interactions between surface atoms. In geometric optimizations, the convergence tolerance was 1.0 × 10−5 Ha for the total energy, 0.002 Ha/Å for the maximum force, and 0.005 Å for the maximal displacement.
The adsorption energy (Eads) of gas i on the SiC3 nanosheet is described by the following equation [31,32]:
E a d s = E S i C 3 , i E S i C 3 E i
where E S i C 3 , i is the energy of the gas molecule i adsorbed on the SiC3 nanosheet, E S i C 3 is the energy of the pure SiC3 nanosheet, E i is the energy of the isolated gas molecule i, and i indicates the gas molecule (i.e., CO2, H2, N2 and CH4). A lower value of E a d s corresponds to the stronger interaction between the SiC3 nanosheet and gas molecule i, and thus, represents a more stable molecules adsorption structure.
The influence of the injecting negative charges on the stability of the SiC3 nanosheet was studied, where the negative charges of numbers 0 to 5 e were injected into the SiC3 nanosheet, equivalent to the negative charge densities of 0~2.06 × 1014 e/cm2, which were applied on the SiC3 nanosheet. For characterizing the adsorption of the gas molecule (i.e., CO2, H2, N2 and CH4) on the different, negatively charged SiC3 nanosheet, the negative-charge density of the SiC3 nanosheet (ρ) was investigated, and the negative charge density was defined as follows [10,17].
ρ = Q S
where Q is the total negative charges injected into (4 × 4 × 1) the supercell of the SiC3 nanosheet, and S is the corresponding surface area (2.43 × 10−14 cm2).
To study the structural stability of a SiC3 nanosheet under different negative-charge conditions, the phonon dispersion spectra and the cohesive energy are calculated. Generally, a larger cohesive energy corresponds to a more stable molecule-adsorption structure; also, no imaginary frequency in the phonon dispersion spectra indicates the stability of a material. The cohesive energy is defined as follows:
E c o h = n c E c + n S i E S i E S i C 3 n c + n S i
where E c , E S i and E S i C 3 are the energies of a single C atom, a single Si atom and the total energy of SiC3 nanosheet, respectively. n c and n S i are the number of Si and C atoms in the SiC3 nanosheet. Next, the effect of the injection of negative charges on the adsorption behavior of gas (i.e., CO2, H2, N2 and CH4) on the SiC3 nanosheets was studied.

3. Results and Discussion

3.1. Electronic Properties and Stability of SiC3 Nanosheet

The SiC3 nanosheet is a graphene-like structure conformation with a P6/mmm space group. The calculated lattice constants of the primitive SiC3 nanosheet are a = b = 5.584 Å. Figure 1a shows the deformation electron density of the 4 × 4 × 1 fully relaxed supercell structure of the SiC3 nanosheet. One can see that electrons are accumulated around the six-membered ring of carbon due to the larger, stronger electronegativity of carbon than that of silicon. The SiC3 nanosheet has two different bonds, which are Si–C and C–C. The bonds lengths of Si–C and C–C are 1.796 and 1.428 Å, respectively, and the angle of C–C–C, C–C–Si and C–Si–C are all 120°. Those calculated results are in good agreement with the reported value [33]. For a pure SiC3 nanosheet, the obtained Local Density of States (LDOS) of C and Si atoms are shown in Figure 1b. LDOS results show that the SiC3 nanosheet exhibits a metallic characteristic due to several bands across the Fermi level that might be primarily contributed by the Si-p orbital [26]. The finding matches with the obtained results by Chabi et al. [28] that show that the SiC3 nanosheet is a potential material for use as a semiconductor. Figure 1c shows the cohesive energy of the SiC3 nanosheet under different charged conditions. The cohesive energy of a neutral SiC3 nanosheet is 9.86 eV/atom, which is slightly larger than the reported value of neutral SiC3 (7.84 eV/atom) [26]. The cohesive energy of the SiC3 nanosheet is 9.86 eV/atom without the applied negative charge, and it decreases with increasing the applied negative charge, finally achieving a stable value (8.37 eV/atom) when the applied negative charge exceeds 3 e−1. It is indicated that the SiC3 nanosheet is a strongly bonded network and is a stable structure under the negative charge conditions. There is no imaginary frequency in the phonon dispersion spectra (Figure 1d). Therefore, all these results reflect the dynamic stability of the SiC3 nanosheet under the condition of injecting negative charges. The SiC3 nanosheet can withstand a rather high temperature under atmospheric pressure [26]. Therefore, the SiC3 nanosheet is a stable material for CO2 separation application after the injection of negative charges.

3.2. CO2/H2/N2/CH4 Adsorption on Uncharged SiC3 Nanosheet

The gas adsorption behaviors of a neutral SiC3 nanosheet is firstly investigated by analyzing the most stable relaxation configurations of the SiC3 nanosheet with gas adsorption. Various initial adsorption sites are considered, including the top sites of C and Si atoms and bridge sites of C–C and C–Si bonds, as well as the center of hexagon holes. Stable relaxation configurations of the SiC3 nanosheet with different gas molecule (i.e., CO2, H2, N2 and CH4) adsorptions are obtained and shown in Figure 2. For the most stable relaxation configurations, the CO2, H2, N2 and CH4 molecule distances from the SiC3 nanosheet are 3.005, 2.6850, 3.235 and 3. 145 Å, respectively.
For the CO2 adsorbed on the SiC3 nanosheet, the O–C–O angle and the C–O bond length exhibit a minor change in comparison to the isolated carbon dioxide molecule, indicating that the absorption of CO2 is classical physisorption. Similar results are observed for the H2, N2 and CH4 molecule adsorbed on the SiC3 nanosheet. Thus, it could be concluded that the gas molecules (i.e., CO2, H2, N2 and CH4) adsorbed on the SiC3 nanosheet surface are caused by physisorption.
The adsorption energies (Eads) of CO2, H2, N2 and CH4 adsorbed on the SiC3 nanosheet are −0.324, −0.227, −0.299 and −0.281 eV, respectively. Among those four gases, the most stable relaxation configurations of gas molecules adsorbed on SiC3 nanosheet are CO2, followed by N2, CH4 and H2, which indicates that the SiC3 material is a potential material for use as a CO2 adsorbent.

3.3. Adsorption of CO2/H2/N2/CH4 by Charged SiC3 Nanosheet

With the injection of the negative charges with density of 1.23 × 1014 e/cm2, the stable relaxation configurations of the SiC3 nanosheet with different gas molecule (i.e., CO2, H2, N2 and CH4) adsorption are obtained and shown in Figure 3.
After the injection of negative charges into the SiC3 nanosheet with CO2 absorption, the CO2 molecule strongly interacts with the SiC3 nanosheet with the adsorption energy Eads of −1.628 eV; the adsorption energy is significantly smaller than that of the uncharged SiC3 nanosheet (−0.324 eV). Furthermore, the O–C–O angle changes from 180° to 131°, the C–O bond length increases from 1.117 Å to 1.259 Å, and the distance between CO2 and the SiC3 nanosheet decreases from 3.005 Å to 2.002 Å. Obvious electron density distribution overlap is observed between CO2 and the SiC3 nanosheet (Figure 4e), indicating that the strong interaction between CO2 and the SiC3 nanosheet, and the absorption of the CO2 molecule on the SiC3 nanosheet is due to chemisorption.
The optimized adsorption configurations of H2, N2 and CH4 after the injection of negative charges are shown in Figure 3b–d. After the injection of negative charges into the SiC3 nanosheet with H2 absorption, for the H2 molecule, the H–H bond length increases slightly from 0.751 to 0.769 Å, the distance between H2 and the SiC3 nanosheet increases slightly from 2.685 to 2.964 Å, and the adsorption energy Eads decreases from −0.227 to -0.499 eV. For the N2 molecule, the N–N bond length increases slightly from 1.111 to 1.123 Å, the distance between N2 and the SiC3 nanosheet decreases slightly from 3.235 to 3.151 Å, and the adsorption energy Eads decreases from −0.299 to −0.578 eV. For the CH4 molecule, the C–H bond length increases slightly from 1.100 to 1.102 Å, and the distance between CH4 and the SiC3 nanosheet decreases from 3.514 to 2.700 Å, and the adsorption energy Eads decreases from −0.281 to −0.525 eV. Furthermore, no evident electron density distribution overlaps are observed between gas molecules (H2, N2 and CH4) and the SiC3 nanosheet after the injection of negative charges (Figure 4e). Thus, it is suggested that the adsorptions of H2, N2 and CH4 on the SiC3 nanosheet are due to physisorption with/without the injecting negative charges, which differs from the chemisorption of CO2 on the SiC3 nanosheet with the injection of negative charges.
Figure 4 shows no obvious electron density distribution overlaps observed between CO2/H2/N2/CH4 molecules and the SiC3 nanosheet without the injecting negative charges. It indicates that CO2/H2/N2/CH4 interact weakly with the SiC3 nanosheet without the injecting negative charges. However, after the injection of the negative charge with density of 1.23 × 1014 e/cm2, obvious electron density distribution overlap between CO2 and the SiC3 nanosheet is observed. In other words, after the injection of the negative charges, the adsorptions of H2, N2 and CH4 on the SiC3 nanosheet are also due to physisorption with weak interactions between these molecules; when the adsorption mechanism of CO2 on the SiC3 nanosheet changes from physisorption to chemisorption, the interaction between CO2 and the SiC3 nanosheet gets stronger. These results suggest that CO2 could be separated from the mixture of CO2, H2, N2 and CH4 that might be tuned by switching on/off the injection of negative charges.

3.4. Mechanism of CO2 Adsorption/Desorption

The mechanism of CO2 adsorption/desorption on the SiC3 nanosheet are investigated by switching on/off the injection of negative charges. Figure 5 shows that the adsorption of CO2 on the SiC3 nanosheet is due to physisorption without the injection of negative charges. With the injection of negative charges with a density of 1.23 × 1014 e/cm2, for the CO2 molecule, the O–C–O angle changes from 180° to 131°, the C–O bond length increases from 1.117 Å to 1.259 Å, and the distance between CO2 and the SiC3 nanosheet decreases from 3.005 Å to 2.002 Å. This result implies that the adsorption mechanism changes from physisorption to chemisorption; the analysis is consistent with the obtained results from Figure 4, and the process is exothermic with a reaction energy of −1.172 eV. After switching off the injection of negative charges, the adsorption mechanism of CO2 changes from chemisorption to physisorption. The adsorbed CO2 molecule is rebound to its physisorption state, and the distance between CO2 and the SiC3 nanosheet increases from 2.002 to 3.005 Å. This transition is also exothermic, 0.776 eV, without any energy barrier. In conclusion, the processes of CO2 adsorption/desorption on the SiC3 nanosheet is reversible, and CO2 adsorption and separation would be tuned by switching on/off the injection of negative charges.

3.5. Separation of CO2 from H2, N2 and CH4 on Negatively Charged SiC3

The effect of the injection of negative charges on the interactions between gases and the SiC3 nanosheet is shown in Figure 6. With the increase in negative charge density, the adsorption energy of CO2 is smaller than those values of N2, CH4 and H2. When the negative charge intensity exceeds 3.75 × 1013 e/cm2, the adsorption energy of CO2 adsorption decreases with the negative charge intensity observably, while the decrement of the adsorption energies for H2, N2 and CH4 are relatively small. It is suggested that the CO2–SiC3 adsorption structure is more stable than the adsorption structures of N2–SiC3, CH4–SiC3, and H2–SiC3 under different charged conditions.
To further explore the CO2 separation ability of the SiC3 nanosheet by switching on/off the injection of negative charges, the co-adsorption behaviors of CO2, H2, N2 and CH4 on the SiC3 nanosheet are investigated. As Figure 7 shows, after the injecting of the negative charges with a density of 1.23 × 1014 e/cm2 into the SiC3 nanosheet with gases co-absorption, the co-adsorption energy decreases from −0.255 to −0.976 eV. It implies that with the injection of negative charges, the SiC3 nanosheet with gases (i.e., CO2, H2, N2 and CH4) co-absorption gets more stable in comparison with the one without the injection of negative charges.
Switching on the injection of negative charges, the modules of H2, N2 and CH4 move away from the SiC3 nanosheet. The distance between the molecule and the SiC3 nanosheet increases from 2.817 to 3.506 Å for H2–SiC3, from 3.243 to 3.170 Å for N2–SiC3, and from 3.491 to 3.506 Å for CH4–SiC3. No obvious electron density distribution overlaps are observed between H2/N2/CH4 molecules and the SiC3 nanosheet with/without the injection of negative charges (Figure 8). It indicates that the effect of the injection of negative charges on the adsorptions of H2/N2/CH4 is not obvious, and H2/N2/CH4 are all physiosorbed on the SiC3 nanosheet with/without the injection of negative charges.
With the injection of negative charges into the SiC3 nanosheet with gases co-absorption, the distance between the CO2 molecule and the SiC3 nanosheet gets shorter, decreasing from 3.271 to 1.995 Å, and the O–C–O angle of CO2 changes from 178.9° to 119.4°. No obvious electron density distribution overlap is observed between the CO2 molecule and the SiC3 nanosheet without the injection of negative charges, while an obvious electron density distribution overlap is observed with the injection of negative charges (Figure 8), implying that the absorption mechanism of CO2 changes from physisorption to chemisorption after the injection of negative charges, which is also consistent with the obtained results from Figure 5.
In summary, H2/N2/CH4 all interact weakly with the SiC3 nanosheet with/without the injection of negative charges. While the interaction between CO2 and the SiC3 nanosheet can be strengthened by the injection of negative charges, the absorption mechanism of CO2 changes from physisorption to chemisorption after the injection of negative charges. Therefore, it could be concluded that the separation CO2 from the mixture of CO2, H2, N2 and CH4 can be achieved by switching on/off the injection of negative charges. All of the above results demonstrate that the SiC3 nanosheet is a promising material for the separation CO2 from the CO2/H2/N2/CH4 mixture by using the injection of negative charges.

4. Conclusions

The potential of SiC3 nanosheets as functional materials for the separation of CO2 from the mixture of CO2, H2, N2 and CH4 by injecting negative charges is studied in this paper. The results show that the SiC3 nanosheets are a promising material for the separation of CO2 from the CO2/H2/N2/CH4 mixture. The main results are summarized as follows.
(1)
In the absence of injecting negative charges, CO2 interacts weakly with the SiC3 nanosheet. While the interaction between CO2 and the SiC3 nanosheet can be strengthened by the injection of negative charges, the absorption mechanism of CO2 changes from physisorption to chemisorption when the injection of negative charges is switched on.
(2)
The effect of injecting negative charges on the SiC3 nanosheet with H2/N2/CH4 adsorption is not obvious, and H2/N2/CH4 are all physiosorbed on the SiC3 nanosheet with/without the injection of negative charges.
(3)
The mechanism of CO2 adsorption/desorption on the SiC3 nanosheet could be tuned by switching on/off the injection of negative charges. The separation of CO2 from the mixture of CO2, H2, N2 and CH4 can be achieved by switching on/off the injection of negative charges.

Author Contributions

Conceptualization, H.Z. and H.X.; methodology and software, H.X. and W.L.; validation and investigation, W.L. and H.X.; formal analysis, H.Z. and H.X.; writing—original draft preparation, W.L. and H.Z.; writing—review and editing, H.Z. and H.X.; supervision, H.X.; project administration, H.Z.; funding acquisition, H.Z. and H.X. All authors have read and agreed to the published version of the manuscript.

Funding

The authors gratefully acknowledge financial supports by the grants from the National Natural Science Foundation of China (No. 52074135), the Natural Science Foundation of Jiangxi Province, China (No. 20202BAB214016) and the Program of Qingjiang Excellent Young Talents, Jiangxi University of Science and Technology (No. JXUSTQJYX2020013).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Endrődi, B.; Samu, A.; Kecsenovity, E.; Halmágyi, T.; Sebők, D.; Janáky, C. Operando cathode activation with alkali metal cations for high current density operation of water-fed zero-gap carbon dioxide electrolysers. Nat. Energy 2021, 6, 439–448. [Google Scholar] [CrossRef]
  2. Anwar, M.; Fayyaz, A.; Sohail, N.F.; Khokhar, M.; Baqar, M.; Khan, W.; Rasool, K.; Rehan, M.; Nizami, A. CO2 capture and storage: A way forward for sustainable environment. J. Environ. Manag. 2018, 226, 131–144. [Google Scholar] [CrossRef] [PubMed]
  3. Hu, Z.; Wang, Y.; Shah, B.; Zhao, D. CO2 capture in metal–organic framework adsorbents: An engineering perspective. Adv. Sustain. Syst. 2019, 3, 1800080. [Google Scholar] [CrossRef] [Green Version]
  4. Senocrate, A.; Battaglia, C. Electrochemical CO2 reduction at room temperature: Status and perspectives. J. Energy Storage 2021, 36, 102373. [Google Scholar] [CrossRef]
  5. Lv, M.; Zhu, R.; Bi, X.; Lin, T. Application research of carbon dioxide in BOF steelmaking process. J. Univ. Sci. Technol. Beijing 2011, 33, 126–130. [Google Scholar]
  6. Dong, K.; Wang, X. CO2 utilization in the ironmaking and steelmaking process. Metals 2019, 9, 273. [Google Scholar] [CrossRef] [Green Version]
  7. Pardakhti, M.; Jafari, T.; Tobin, Z.; Dutta, B.; Moharreri, E.; Shemshaki, N.; Suib, S.; Srivastava, R. Trends in solid adsorbent materials development for CO2 capture. ACS Appl. Mater. Int. 2019, 11, 34533–34559. [Google Scholar] [CrossRef] [PubMed]
  8. Xiong, H.; Zhang, H.; Gan, L. CO2 capture and separation on the penta-BN2 monolayer with the assistance of charge/electric field. J. Mater. Sci. 2021, 56, 4341–4355. [Google Scholar] [CrossRef]
  9. Talapaneni, S.; Singh, G.; Kim, I.; AlBahily, K.; Al-Muhtaseb, A.; Karakoti, A.; Tavakkoli, E.; Vinu, A. Nanostructured carbon nitrides for CO2 capture and conversion. Adv. Mater. 2020, 32, 1904635. [Google Scholar] [CrossRef]
  10. Qin, G.; Du, A.; Sun, Q. Charge-and Electric-Field-Controlled Switchable Carbon Dioxide Capture and Gas Separation on a C2N Monolayer. Energy Technol. 2018, 6, 205–212. [Google Scholar] [CrossRef]
  11. He, C.; Zhang, M.; Li, T.; Zhang, W. A novel C6N2 monolayer as a potential material for charge-controlled CO2 capture. J. Mater. Chem. C 2020, 8, 6542–6551. [Google Scholar] [CrossRef]
  12. Li, X.; Guo, T.; Zhu, L.; Ling, C.; Xue, Q.; Xing, W. Charge-modulated CO2 capture of C3N nanosheet: Insights from DFT calculations. Chem. Eng. J. 2018, 338, 92–98. [Google Scholar] [CrossRef]
  13. Zhou, S.; Wang, M.; Wang, J.; Xin, H.; Liu, S.; Wang, Z.; Wei, S.; Lu, X. Carbon phosphides: Promising electric field controllable nanoporous materials for CO2 capture and separation. J. Mater. Chem. A 2020, 8, 9970–9980. [Google Scholar] [CrossRef]
  14. Li, X.; Zhu, L.; Chang, X.; He, D.; Xue, Q.; Xing, W. Me–N–C (Me = Fe, Cu, and Co) nanosheet as a promising charge-controlled CO2 capture material. J. Mater. Chem. A 2018, 6, 12404–12410. [Google Scholar] [CrossRef]
  15. He, C.; Wang, R.; Xiang, D.; Li, X.; Fu, L.; Jian, Z.; Huo, J.; Li, S. Charge-regulated CO2 capture capacity of metal atom embedded graphyne: A first-principles study. Appl. Surf. Sci. 2020, 509, 145392. [Google Scholar] [CrossRef]
  16. Wang, Z.; Lü, T.; Wang, H.; Feng, Y.; Zheng, J. Review of borophene and its potential applications. Front. Phys. 2019, 14, 1–20. [Google Scholar] [CrossRef] [Green Version]
  17. Tan, X.; Tahini, H.; Smith, S. Borophene as a promising material for charge-modulated switchable CO2 capture. ACS Appl. Mater. Int. 2017, 9, 19825–19830. [Google Scholar] [CrossRef]
  18. Buyukcakir, O.; Je, S.H.; Talapaneni, S.N.; Kim, D.; Coskun, A. Charged covalent triazine frameworks for CO2 capture and conversion. ACS Appl. Mater. Int. 2017, 9, 7209–7216. [Google Scholar] [CrossRef] [PubMed]
  19. Sun, C.; Zhu, S.; Liu, M.; Shen, S.; Bai, B. Selective molecular sieving through a large graphene nanopore with surface charges. J. Phys. Chem. Lett. 2019, 10, 7188–7194. [Google Scholar] [CrossRef] [PubMed]
  20. Wang, M.; Zhang, Z.; Gong, Y.; Zhou, S.; Wang, J.; Wang, Z.; Wei, S.; Guo, W.; Lu, X. Penta-graphene as a promising controllable CO2 capture and separation material in an electric field. Appl. Surf. Sci. 2020, 502, 144067. [Google Scholar] [CrossRef]
  21. Tan, X.; Kou, L.; Smith, S.C. Layered graphene-hexagonal BN nanocomposites: Experimentally feasible approach to charge-induced switchable CO2 capture. ChemSusChem 2015, 8, 2987–2993. [Google Scholar] [CrossRef]
  22. Tao, L.; Huang, J.; Dastan, D.; Wang, T.; Li, J.; Yin, X.; Wang, Q. CO2 capture and separation on charge-modulated calcite. Appl. Surf. Sci. 2020, 530, 147265. [Google Scholar] [CrossRef]
  23. Sathishkumar, N.; Wu, S.; Chen, H. Charge-regulated, electric-field and combined effect controlled switchable CO2 capture and separation on penta-C2N nanosheet: A computational study. Chem. Eng. J. 2021, 407, 127194. [Google Scholar] [CrossRef]
  24. Tan, X.; Tahini, H.; Smith, S. Hexagonal boron nitride and graphene in-plane heterostructures: An experimentally feasible approach to charge-induced switchable CO2 capture. Chem. Phys. 2016, 478, 139–144. [Google Scholar] [CrossRef]
  25. Darvishnejad, M.; Reisi-Vanani, A. DFT-D3 calculations of the charge-modulated CO2 capture of N/Sc-embedded graphyne: Compilation of some factors. J. CO2 Util. 2021, 46, 101469. [Google Scholar] [CrossRef]
  26. Li, P.; Zhou, R.; Zeng, X. The search for the most stable structures of silicon–carbon monolayer compounds. Nanoscale 2014, 6, 11685–11691. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Shi, Z.; Zhang, Z.; Kutana, A.; Yakobson, B.I. Predicting two-dimensional silicon carbide monolayers. ACS Nano 2015, 9, 9802–9809. [Google Scholar] [CrossRef] [PubMed]
  28. Chabi, S.; Kadel, K. Two-Dimensional Silicon Carbide: Emerging Direct Band Gap Semiconductor. Nanomaterials 2020, 10, 2226. [Google Scholar] [CrossRef] [PubMed]
  29. Delley, B. From molecules to solids with the dmol3 approach. J. Chem. Phys. 2000, 113, 7756–7764. [Google Scholar] [CrossRef]
  30. Qin, X.; Wu, Y.; Liu, Y.; Chi, B.; Li, X.; Wang, Y.; Zhao, X. Origins of dirac cone formation in ab3 and a3b (a, b = C, Si, and Ge) binary monolayers. Sci. Rep-UK 2017, 7, 1–13. [Google Scholar] [CrossRef] [Green Version]
  31. Sharma, A.; Khan, M.; Husain, M. Adsorption of phosgene on Si-embedded MoS2 sheet and electric field assisted desorption: Insights from DFT calculations. J. Mater. Sci. 2019, 54, 11497–11508. [Google Scholar] [CrossRef]
  32. Jiao, Y.; Du, A.; Zhu, Z.; Rudolph, V.; Lu, G.; Smith, S. A density functional theory study on CO2 capture and activation by graphene-like boron nitride with boron vacancy. Catal. Today 2011, 175, 271–275. [Google Scholar] [CrossRef]
  33. Ding, Y.; Wang, Y. Geometric and electronic structures of two-dimensional SiC3 compound. J. Phys. Chem. C 2014, 118, 4509–4515. [Google Scholar] [CrossRef]
Figure 1. (a) Deformation electron density of SiC3 nanosheet structure (4 × 4), (b) LDOS of the Si and C atoms of SiC3 nanosheet with Fermi level set to 0 eV, (c) the cohesive energy of the SiC3 nanosheet under different negative charge density conditions, and (d) the phonon dispersion spectra of SiC3 nanosheet.
Figure 1. (a) Deformation electron density of SiC3 nanosheet structure (4 × 4), (b) LDOS of the Si and C atoms of SiC3 nanosheet with Fermi level set to 0 eV, (c) the cohesive energy of the SiC3 nanosheet under different negative charge density conditions, and (d) the phonon dispersion spectra of SiC3 nanosheet.
Crystals 11 00543 g001
Figure 2. Top and side views of the stable adsorption configurations of different gas molecules adsorbed on the SiC3 nanosheet without the injecting negative charges. (a) CO2, (b) H2, (c) N2, and (d) CH4 adsorbed on the SiC3.
Figure 2. Top and side views of the stable adsorption configurations of different gas molecules adsorbed on the SiC3 nanosheet without the injecting negative charges. (a) CO2, (b) H2, (c) N2, and (d) CH4 adsorbed on the SiC3.
Crystals 11 00543 g002
Figure 3. Top and side views of the stable adsorption configurations of different gas molecules adsorbed on SiC3 nanosheet after the injection of the negative charges with density of 1.23 × 1014 e/cm2. (a) CO2, (b) H2, (c) N2, and (d) CH4 adsorbed on the SiC3.
Figure 3. Top and side views of the stable adsorption configurations of different gas molecules adsorbed on SiC3 nanosheet after the injection of the negative charges with density of 1.23 × 1014 e/cm2. (a) CO2, (b) H2, (c) N2, and (d) CH4 adsorbed on the SiC3.
Crystals 11 00543 g003
Figure 4. The electron density distribution of the most stable adsorption configurations of gas molecule (i.e., CO2, H2, N2 and CH4) adsorbed on the SiC3 nanosheet. (ad) Without the injecting negative charges, and (eh) after the injection of negative charges with density of 1.23 × 1014 e/cm2.
Figure 4. The electron density distribution of the most stable adsorption configurations of gas molecule (i.e., CO2, H2, N2 and CH4) adsorbed on the SiC3 nanosheet. (ad) Without the injecting negative charges, and (eh) after the injection of negative charges with density of 1.23 × 1014 e/cm2.
Crystals 11 00543 g004
Figure 5. The process of (a) CO2 adsorption on the SiC3 nanosheet after injecting negative charges with density of 1.23 × 1014 e/cm2 for the SiC3 nanosheet with CO2 adsorption, and (b) CO2 desorption from the SiC3 nanosheet after switching off the injection of negative charges.
Figure 5. The process of (a) CO2 adsorption on the SiC3 nanosheet after injecting negative charges with density of 1.23 × 1014 e/cm2 for the SiC3 nanosheet with CO2 adsorption, and (b) CO2 desorption from the SiC3 nanosheet after switching off the injection of negative charges.
Crystals 11 00543 g005
Figure 6. Effect of the injection of negative charges on the adsorption energy of the most stable configurations of SiC3 nanosheet with CO2, H2, N2 and CH4 adsorption.
Figure 6. Effect of the injection of negative charges on the adsorption energy of the most stable configurations of SiC3 nanosheet with CO2, H2, N2 and CH4 adsorption.
Crystals 11 00543 g006
Figure 7. Top and side views of the optimized configurations of the SiC3 nanosheet with gases molecules (i.e., CO2, H2, N2 and CH4) co-adsorption (a) without the injection of negative charges, and (b) with the injection of negative charges (1.23 × 1014 e/cm2).
Figure 7. Top and side views of the optimized configurations of the SiC3 nanosheet with gases molecules (i.e., CO2, H2, N2 and CH4) co-adsorption (a) without the injection of negative charges, and (b) with the injection of negative charges (1.23 × 1014 e/cm2).
Crystals 11 00543 g007
Figure 8. Top and side views of the charge density distribution of the SiC3 nanosheet with gases molecules (i.e., CO2, H2, N2 and CH4) co-adsorption (a) without the injection of negative charges, and (b) with the injection of negative charges (1.23 × 1014 e/cm2).
Figure 8. Top and side views of the charge density distribution of the SiC3 nanosheet with gases molecules (i.e., CO2, H2, N2 and CH4) co-adsorption (a) without the injection of negative charges, and (b) with the injection of negative charges (1.23 × 1014 e/cm2).
Crystals 11 00543 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, H.; Xiong, H.; Liu, W. SiC3 as a Charge-Regulated Material for CO2 Capture. Crystals 2021, 11, 543. https://doi.org/10.3390/cryst11050543

AMA Style

Zhang H, Xiong H, Liu W. SiC3 as a Charge-Regulated Material for CO2 Capture. Crystals. 2021; 11(5):543. https://doi.org/10.3390/cryst11050543

Chicago/Turabian Style

Zhang, Haihui, Huihui Xiong, and Wei Liu. 2021. "SiC3 as a Charge-Regulated Material for CO2 Capture" Crystals 11, no. 5: 543. https://doi.org/10.3390/cryst11050543

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop