Next Article in Journal
Effect of Oxide Supports on the Activity of Pd Based Catalysts for Furfural Hydrogenation
Next Article in Special Issue
Studies on the Binary MgO/SiO2 Mixed Oxide Catalysts for the Conversion of Ethanol to 1,3-Butadiene
Previous Article in Journal
Entrapping Immobilisation of Lipase on Biocomposite Hydrogels toward for Biodiesel Production from Waste Frying Acid Oil
Previous Article in Special Issue
Immobilization of Cellulolytic Enzymes in Mesostructured Silica Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Prediction of a Stable Organic Metal-Free Porous Material as a Catalyst for Water-Splitting

1
School of Mechanical & Automotive Engineering, Liaocheng University, Liaocheng 252000, China
2
School of Physics, Shandong University, Jinan 250100, China
3
Shandong Key Laboratory of Optical Communication Science and Technology, School of Physics Science and Information Technology, Liaocheng University, Liaocheng 252000, China
4
School of Chemistry and Chemical Engineering, Liaocheng University, Liaocheng 252000, China
*
Authors to whom correspondence should be addressed.
Catalysts 2020, 10(8), 836; https://doi.org/10.3390/catal10080836
Submission received: 6 May 2020 / Revised: 1 July 2020 / Accepted: 9 July 2020 / Published: 24 July 2020
(This article belongs to the Special Issue Porous Materials and Catalysts)

Abstract

:
It is of practical significance to find organic metal-free catalyst materials. We propose a new graphene-like carbon nitride structure, which was able to meet these requirements well. Its primitive cell consists of eight carbon atoms and six nitrogen atoms. Hence, we called this structure g–C8N6. The stability of the structure was verified by phonon spectroscopy and molecular dynamics simulations. Then its electronic structure was calculated, and its band edge position was compared with the redox potential of water. We analyzed its optical properties and electron–hole recombination rate. After the above analysis, it is predicted that it is a suitable photocatalyst material. To improve its photocatalytic performance, two methods were proposed: applied tensile force and multilayer stacking. Our research is instructive for the photocatalytic application of this kind of materials.

1. Introduction

Two problems have plagued people for a long time—the energy crisis and environmental pollution. Scientists have put much effort into solving these two problems. The development of hydrogen energy is promising [1]. If we can obtain much hydrogen, we can solve these two problems effectively.
Water-splitting to produce hydrogen using a photocatalyst is a feasible method [2]. In 1972, Honda and Fujishima first produced H2 by photochemical water-splitting using TiO2 [3]. With development of technology, there are increasing kinds of photocatalysts, including various oxides, sulfides, oxynitride semiconductors, etc. [4,5,6,7,8,9,10,11,12], which belong to inorganic materials. Inorganic photocatalysts materials have some drawbacks, for example limited concentration of active sites and heavy metal toxicity [13,14,15]. However, organic photocatalysts have some merits, for example low cost, ease of fabrication and mechanical flexibility [16,17]. Therefore, research of metal-free organic high-efficiency photocatalysts is very important.
Since the exfoliation of monolayer graphene, two-dimensional materials have attracted comprehensive research for excellent properties [18,19,20,21,22,23,24,25]. It is increasingly important to study and regulate the properties of two-dimensional materials [26,27,28]. Wang et al. first considered graphene-like carbon nitride (g–C3N4) as a photocatalyst. They predicted that the g–C3N4 photocatalyst exhibits interesting electronic properties that can decompose water to produce hydrogen [29]. Subsequently, many studies were carried out on the application of graphene-like carbon nitride in photocatalysis [30,31,32,33,34,35,36].
In this work, we propose a graphene-like carbon nitride structure with a primitive cell that consists of eight carbon atoms and six nitrogen atoms. Hence, we call this structure g–C8N6. First using basic calculations, we analyzed the electronic structure of g–C8N6 and compared its band edge position with the redox potential of water. At the same time, we calculated its optical properties and considered it as a suitable photocatalyst material.

2. Results and Discussion

The structure of the g–C8N6 supercell (3 × 3 × 1) is shown as Figure 1a. The primitive cell is marked using the yellow shaded area. The lattice constant is 7.177 Å, and the lengths of two types C–N bonds are 1.35 Å and 1.34 Å, respectively. Unlike many carbon nitrides, the s-heptazines in g–C8N6 is connected by C–C bonds, rather than C–N bonds. The lengths of the C–C bonds in the s-heptazines are 1.426 Å, however the length of C–C bonds between two connected s-heptazines are 1.439 Å. It has a very uniform hexagonal symmetry structure; which angles are all 120°. Because it has many holes, it has enough contact with water when it is used as a photocatalyst.
Many similar carbon nitride structures have been synthesized [37,38,39], so perhaps a carbon-rich g–C8N6 layered structure may exhibit sufficient stability to be synthesized in future. We checked the stability of the structure theoretically through two methods. First, the force constant theory tested kinetic stability of the g–C8N6 through the phonon spectrum. A large (2 × 2) supercell was optimized in the calculation. The phonon spectral dispersion diagram along the highly symmetric direction in Brillouin zone is shown in Figure 1b. Obviously, the phonon spectral dispersion has no imaginary frequency, indicating that g–C8N6 framework is kinetically stable. This is similar to the phonon spectra of g–C13N13 [28], C2 N-h2D [40].
Second, the dynamic stability was checked with molecular dynamics simulations. The 126 atoms of the 3 × 3 supercell were simulated using a Nosé–Hoover thermostat at 300 K. The structure remained well after 5 ps, indicating its stability at room temperature. Temperature and total energy varied slightly with time, but the geometry of g–C8N6 remained stable.
Next, the electronic structures were analyzed. The energy-band structure distributions and electron density of states (DOS) by PBE method of g–C8N6 are shown in Figure 2a,b. It can be obtained that this material is a 2.13-eV band-gap semiconductor without magnetism. The conduction band minimum (CBM) is situated on K point and the valence band maximum (VBM) is situated on M point, which shows that g–C8N6 is an indirect band-gap semiconductor.
According to electronic structure of the material, we analyzed whether it is suitable for splitting water as a photocatalyst. The band of a photocatalyst should be greater than 1.23 eV, which is the difference between the water reduction potential and water oxidation potential. The positions of the conduction and valence band edges should straddle the redox potentials for water photolysis. The optical absorption properties should show that it can absorb partial visible light [41].
The band-edge potential affects whether it can be a photocatalyst. The CBM minus the Fermi energy and then minus the vacuum energy level is the conduction band edge. The VBM minus the Fermi energy and then minus the vacuum energy level is the valence band edge. The redox potentials of water are −4.5 eV and −5.73 eV, respectively. The conduction band edge of a suitable photocatalyst should be above the energy corresponding to the water reduction potential [ϕ(H+/H2)]. At the same time, the valence band edge of a suitable photocatalyst should be below the energy of the water oxidation potential [ϕ(O2 /H2O)]. As shown in Figure 3a, band gap of the g–C8N6 is 2.89 eV, which has ideal positions of the CBM and the VBM (ECBM = 1.22 eV and EVBM = 0.44 eV vs the normal hydrogen electrode, pH = 0). In other words, the electrons in the lowest unoccupied molecular orbital (LUMO) have sufficient reactivity to reduce water to hydrogen, whereas holes in the highest occupied molecular orbital (HOMO) have sufficient reduction potential to oxidize water to oxygen.
In order to analyze optical absorption properties of g–C8N6, we calculated the frequency-dependent dielectric matrix using hybrid HSE06 function and expanded a 10 × 10 × 1 k-point mesh. The random phase approximation was used during the calculation of the frequency dependent dielectric tensor. The complex dielectric constants at a given frequency can be defined as:
ε ( ω ) = ε 1 ( ω ) + i ε 2 ( ω )
The expression for the absorption coefficient I ( ω ) is given as [42]:
I ( ω ) = ω ε 2 / ( c n )
where c is the speed of light, and n is the index of refraction:
n = [ ε 1 2 ( ω ) + ε 2 2 ( ω ) + ε 1 ( ω ) ] 1 / 2 / 2
According to the above equation, the light absorption property can be obtained mainly from the value of the imaginary part, and slightly from the value of the real part. The absorption coefficient was above zero, when only if the imaginary part:
ε 2 ( ω ) > 0
The imaginary part is determined by the summation on the empty state. The following equation can be used:
ε α β ( 2 ) ( ω ) = 4 π 2 e 2 Ω lim q 0 1 q 2 c , v , k 2 w k δ ( ε c k ε v k ω ) × u c k + e α q | u v k u c k + e β q | u v k
where the indices c and v represent the conduction and valence band states and u c k is the cell periodic part of the orbitals at the k point. The summation of Equation (4) includes a large number of empty conduction band states. The number of empty conduction band states is almost twice the number of valence bands.
According to the theory described above, frequency-dependent dielectric function of g–C8N6 was calculated by the hybrid HSE06 function. Figure 3b shows the absorption coefficient and Figure 3c shows imaginary part of dielectric function of g–C8N6. From Figure 3b,c, the solar spectral range which can be absorbed is from 60 nm to 600 nm. Hence, it is obvious that g–C8N6 can be used as a photocatalyst for splitting water.
The recombination rate of electron–hole is an important aspect affecting the photocatalytic performance. In the process of photocatalytic water-splitting, after electronic transition, electrons and holes will decompose water into hydrogen and oxygen. However, there are many electrons that will go back to original orbitals and recombine with holes. If recombination rate of electrons and holes is relatively low, then there will be more electrons involved in the redox reaction process, and the catalytic performance will be improved.
Figure 4a is the charge density of the conduction band minimum (CBM), which is situated on K point of the 32nd energy band. Figure 4b is the charge density of the valence band maximum (VBM), which is located on M point of the 31st energy band. Comparing Figure 4a,b, we can find that there are few overlapping parts in the distribution of charge density, that is, the recombination rate of electrons and holes is relatively low. This result indicates that g–C8N6 may exhibit long-lived electrons and holes that can improve its photocatalytic performance.
The g–C8N6 is a semiconductor with a suitable band gap. Its band edge position relative to the redox potential of water is ideal and can absorb a part of visible light. More valuable, the recombination rate of its electron holes is relatively low. In summary, it has characteristics of photocatalyst and can be used for water-splitting.
Next, we analyze whether there are some ways to improve its performance. As we all know, tensile strain can regulate electronic structure of material and thus improves its properties [28,38,43]. Hence, we applied tensile strain to g–C8N6 to observe its band-gap change. Tensile strain is defined as:
ε = l l 0 l 0 × 100 %
where l 0 is length of original supercell structure, and l is length of supercell under tensile strain.
Figure 5 is the case of band-gap changes as the tensile force changes with PBE functional. The color of the squares in figure represents the color of the solar spectrum that can be absorbed. With the increase of tensile force, the length of C–C bonds and C–N bonds gradually increases, indicating that the orbitals binding ability of these bonds is weakened. At the same time, the splitting of bonding and antibonding decreases, resulting in narrow band gap, so absorption range of sunlight extends to green region. In contrast, when compressive stress is applied, the splitting of bonding and antibonding increases, thus widened the band gap, and the absorption of sunlight moves to red region. To sum up these two cases, tensile force can be used to capture a wider spectrum of the sun.
Multilayer stacking can also adjust electronic structure of material, thus improving its performance [44]. The multilayer structure is AA stacking. Work functions and energy-band gap of two-, three-, four-, five- and six-layer g–C8N6 calculated are listed in Table 1 with HSE06 functional. This is particularly important for layered structures as the interlayer interaction depends on the van der Waals interaction [45,46]. The results of Table 1 and Figure 6 were calculated using van der Waals force. Although the number of layers is increasing gradually, the work function maintain stability. In sharp contrast band gap decreases gradually with increase of the number of layers. Similar to the monolayer, all multilayer g–C8N6 remain an indirect gap semiconductor. When the number of layers is 2–6, the band gap is 2.50–2.35 eV. As number of layers increases, properties of the material gradually approach the bulk material with a gap of 2.35 eV. Compared with single-layer, multilayer stacking reduced the band gap by 0.54 eV.
To study the photocatalytic capacity of multilayer g–C8N6, the band alignment are shown in Figure 6. Compared with the monolayer, the VBM of the bilayer structure increased, while the CBM of the bilayer structure decreased. The other multilayer g–C8N6 systems were similar to the bilayer structure. The multilayered g–C8N6 system had ideal band edges that could decompose water to produce hydrogen and oxygen.

3. Method and Computational Details

The calculations were carried out within the framework of density functional theory (DFT), which was implemented in the Vienna ab initio simulation package known as VASP [47,48,49]. The electron–electron interactions were handled by exchange correlation functions with Perde–Burke–Ernzerhof (PBE) in the form of generalized gradient approximation (GGA) [50]. Energy cutoff value for plane wave expansion of electron wave function is set to 600 eV. Projector-augmented wave (PAW) potentials were used to describe electron–ion interactions [51,52]. Valence electrons include four electrons of carbon (2s22p2) and five electrons of nitrogen (2s22p3). A vacuum region was set to 15 Å. The Brillouin zone was divided into 9 × 9 × 1 k points. Structural optimization was performed by conjugate gradient (CG) method until the residual force of each atom was less than 0.001 eV/Å. A more accurate Heyd–Scuseria–Ernzerhof (HSE) screening potential method [53,54] was used to calculate band gap. Optical absorption spectra of the two dimension covalent triazine frameworks (CTFs) were computed from the dielectric function by Kramers–Kronig dispersion relation [55]. Work function means the minimum energy that must be provided for an electron to escape immediately from the surface of the material. The work function can be calculated by: φ = V ( ) E F , where V ( ) and EF are the electrostatic potential in a vacuum region far from the neutral surface and the Fermi energy of the neutral surface system, respectively.

4. Conclusions

We propose a new graphene-like carbon nitride structure g–C8N6, the stability of the structure of which was verified by phonon spectroscopy and molecular dynamics simulations. Then its electronic structure was calculated, and its band edge position was compared with the redox potential of water. We analyzed its optical properties and electron–hole recombination rate. After the above analysis, it is theoretically a suitable water-splitting photocatalyst material. To improve its photocatalytic performance, two methods were proposed: applied tensile force and multilayer stacking. Our research is instructive for the photocatalytic application of this kind of materials.

Author Contributions

Conceptualization, H.L.; data curation, Q.W. and Z.F.; formal analysis, M.C.; funding acquisition, C.L.; resources, C.B. (Chunjiang Bao) and K.Q.; software, C.B. (Chenglin Bai) and F.G.; writing—original draft, H.Y.; writing—review & editing, H.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Introduction and Cultivation Plan of Youth Innovation Talents for Universities of Shandong Province in 2019 (Research and Innovation Team on Materials Modification and Optoelectronic Devices at extreme conditions), Shandong Provincial Key Research and Development Program (Public Welfare Science and Technology Research) (No. 2019GGX103010), Science and Technology Plan of Youth Innovation Team for Universities of Shandong Province (No. 2019KJJ019), Shandong Natural Science Foundation (No. ZR2019PEE006), National Natural Science Foundation of China (No. 11504153, 11504386 ), Science and Technology Planning Project of Higher School in Shandong Province (No. J18KA243) and Liaocheng University High-level Talents & PhD Research Startup Foundation (No. 318051619). The APC was funded by Introduction and Cultivation Plan of Youth Innovation Talents for Universities of Shandong Province in 2019 (Research and Innovation Team on Materials Modification and Optoelectronic Devices at extreme conditions).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kroke, E.; Schwarz, M. Novel group 14 nitrides. Coord. Chem. Rev. 2004, 248, 493–532. [Google Scholar] [CrossRef]
  2. Chen, X.; Shen, S.; Guo, L.; Mao, S.S. Semiconductor-based photocatalytic hydrogen generation. Chem. Rev. 2010, 110, 6503–6570. [Google Scholar] [CrossRef] [PubMed]
  3. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  4. Kumaravel, V.; Kang, M. Photocatalytic Hydrogen Evolution. Catalysts 2020, 10, 492. [Google Scholar] [CrossRef]
  5. Ming, H.; Ming, J.; Oh, S.-M.; Lee, E.-J.; Huang, H.; Zhou, Q.; Zheng, J.; Sun, Y.-K. High dispersion of TiO2 nanocrystals within porous carbon improves lithium storage capacity and can be applied batteries to LiNi0.5Mn1.5O4. J. Mater. Chem. A 2014, 2, 18938–18945. [Google Scholar] [CrossRef]
  6. Xu, L.; Cui, Q.; Zhang, H.; Jiao, A.; Tian, Y.; Li, S.; Li, H.; Chen, M.; Chen, F. Ultra-clean PtPd nanoflowers loaded on GO supports with enhanced low-temperature electrocatalytic activity for fuel cells in harsh environment. Appl. Surf. Sci. 2020, 511, 145603. [Google Scholar] [CrossRef]
  7. Park, J.H.; Kim, S.; Bard, A.J. Novel carbon-doped TiO2 nanotube arrays with high aspect ratios for efficient solar water splitting. Nano Lett. 2006, 6, 24–28. [Google Scholar] [CrossRef]
  8. Hua, J.; Liu, Y.L.; Li, H.S.; Zhao, M.W.; Liu, X.D. Effect of the alloying element titanium on the stability and trapping of hydrogen in pure vanadium: A first-principles study. Int. J. Mod. Phys. B 2014, 28, 1450207. [Google Scholar] [CrossRef]
  9. Ksibi, M.; Rossignol, S.; Tatibouët, J.-M.; Trapalis, C. Synthesis and solid characterization of nitrogen and sulfur-doped TiO2 photocatalysts active under near visible light. Mater. Lett. 2008, 62, 4204–4206. [Google Scholar] [CrossRef]
  10. Cheng, Y.W.; Chan, R.C.; Wong, P. Disinfection of Legionella pneumophila by photocatalytic oxidation. Water Res. 2007, 41, 842–852. [Google Scholar] [CrossRef]
  11. Yu, J.; Xiang, Q.; Zhou, M. Preparation, characterization and visible-light-driven photocatalytic activity of Fe-doped titania nanorods and first-principles study for electronic structures. Appl. Catal. B: Environ. 2009, 90, 595–602. [Google Scholar] [CrossRef]
  12. Liu, S.; Yu, J.; Jaroniec, M. Tunable photocatalytic selectivity of hollow TiO2 microspheres composed of anatase polyhedra with exposed {001} facets. J. Am. Chem. Soc. 2010, 132, 11914–11916. [Google Scholar] [CrossRef] [PubMed]
  13. Zhang, X.; Xie, X.; Wang, H.; Zhang, J.; Pan, B.; Xie, Y. Enhanced photoresponsive ultrathin graphitic-phase C3N4 nanosheets for bioimaging. J. Am. Chem. Soc. 2013, 135, 18–21. [Google Scholar] [CrossRef] [PubMed]
  14. Su, D.S.; Zhang, J.; Frank, B.; Thomas, A.; Wang, X.; Paraknowitsch, J.; Schlögl, R. Metal-free heterogeneous catalysis for sustainable chemistry. Chemsuschem: Chem. Sustain. Energy Mater. 2010, 3, 169–180. [Google Scholar] [CrossRef]
  15. Hua, J.; Liu, Y.L.; Li, H.S.; Zhao, M.W.; Liu, X.D. Energetics of carbon and nitrogen impurities and their interactions with vacancy in vanadium. Chin. Phys. B 2016, 25, 036104. [Google Scholar] [CrossRef]
  16. Xiao, W.; Jin, X.; Chen, G.Z. Up-scalable and controllable electrolytic production of photo-responsive nanostructured silicon. J. Mater. Chem. A 2013, 1, 10243–10250. [Google Scholar] [CrossRef]
  17. Ng, C.H.; Winther-Jensen, O.; Ohlin, C.A.; Winther-Jensen, B. Exploration and optimisation of poly (2, 2′-bithiophene) as a stable photo-electrocatalyst for hydrogen production. J. Mater. Chem. A 2015, 3, 11358–11366. [Google Scholar] [CrossRef]
  18. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A. Electric field effect in atomically thin carbon films. Science 2004, 306, 666–669. [Google Scholar] [CrossRef] [Green Version]
  19. Bunch, J.S.; Van Der Zande, A.M.; Verbridge, S.S.; Frank, I.W.; Tanenbaum, D.M.; Parpia, J.M.; Craighead, H.G.; McEuen, P.L. Electromechanical resonators from graphene sheets. Science 2007, 315, 490–493. [Google Scholar] [CrossRef] [Green Version]
  20. Lee, C.; Wei, X.W.; Kysar, J.W.; Hone, J. Measurement of the elastic properties and intrinsic strength of monolayer grapheme. Science 2008, 321, 385–388. [Google Scholar] [CrossRef]
  21. Meyer, J.C.; Geim, A.K.; Katsnelson, M.I.; Novoselov, K.S.; Booth, T.J.; Roth, S. The structure of suspended graphene sheets. Nature 2007, 446, 60. [Google Scholar] [CrossRef] [PubMed]
  22. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Katsnelson, M.I.; Grigorieva, I.V.; Dubonos, S.V.; Firsov, A.A. Two-dimensional gas of massless Dirac fermions in graphene. Nature 2005, 438, 197–200. [Google Scholar] [CrossRef] [PubMed]
  23. Geim, A.K.; Novoselov, K.S. The rise of graphene. Nat. Mater. 2007, 6, 183. [Google Scholar] [CrossRef]
  24. Li, H.; Hu, H.; Bao, C.; Zhang, X.; Wang, A.; Zhou, H.; Zhao, M. The strability and electronic structure of Fe atoms embedded in zigzag graphene nanoribbons. Phys. B: Condens. Matter 2014, 441, 28–32. [Google Scholar] [CrossRef]
  25. Xu, M.; Liang, T.; Shi, M.; Chen, H. Graphene-like two-dimensional materials. Chem. Rev. 2013, 113, 3766–3798. [Google Scholar] [CrossRef]
  26. Zhang, H.; Cui, Q.; Xu, L.; Jiao, A.; Tian, Y.; Liu, X.; Li, S.; Li, H.; Chen, M.; Chen, F. Blue laser-induced photochemical synthesis of CuAg nanoalloys on h-BN supports with enhanced SERS activity for trace-detection of residual pesticides on tomatoes. J. Alloy. Compd. 2020, 825, 153996. [Google Scholar] [CrossRef]
  27. Saleem, Z.; Pervaiz, E.; Yousaf, M.U.; Niazi, M.B.K. Two-dimensional materials and composites as potential water splitting photocatalysts: A review. Catalysts 2020, 10, 464. [Google Scholar] [CrossRef]
  28. Li, H.; Hu, H.; Bai, C.; Bao, C.; Feng, Z.; Guo, F. The metal-free magnetism and ferromagnetic narrow gap semiconductor properties in graphene-like carbon nitride. Phys. B: Condens. Matter 2019, 555, 91–95. [Google Scholar] [CrossRef]
  29. Wang, X.; Maeda, K.; Thomas, A.; Takanabe, K.; Xin, G.; Carlsson, J.M.; Domen, K.; Antonietti, M. A metal-free polymeric photocatalyst for hydrogen production from water under visible light. Nat. Mater. 2009, 8, 76–80. [Google Scholar] [CrossRef]
  30. Liu, G.; Niu, P.; Sun, C.; Smith, S.C.; Chen, Z.; Lu, G.Q.; Cheng, H.-M. Unique electronic structure induced high photoreactivity of sulfur-doped graphitic C3N4. J. Am. Chem. Soc. 2010, 132, 11642–11648. [Google Scholar] [CrossRef]
  31. Zhang, J.; Sun, J.; Maeda, K.; Domen, K.; Liu, P.; Antonietti, M.; Fu, X.; Wang, X. Sulfur-mediated synthesis of carbon nitride: Band-gap engineering and improved functions for photocatalysis. Energy Environ. Sci. 2011, 4, 675–678. [Google Scholar] [CrossRef]
  32. Wu, F.; Liu, Y.; Yu, G.; Shen, D.; Wang, Y.; Kan, E. Visible-light-absorption in graphitic C3N4 bilayer: Enhanced by interlayer coupling. J. Phys. Chem. Lett. 2012, 3, 3330–3334. [Google Scholar] [CrossRef]
  33. Maeda, K.; Wang, X.; Nishihara, Y.; Lu, D.; Antonietti, M.; Domen, K. Photocatalytic activities of graphitic carbon nitride powder for water reduction and oxidation under visible light. J. Phys. Chem. C 2009, 113, 4940–4947. [Google Scholar] [CrossRef]
  34. Zhu, Y.-P.; Li, M.; Liu, Y.-L.; Ren, T.-Z.; Yuan, Z.-Y. Carbon-doped ZnO hybridized homogeneously with graphitic carbon nitride nanocomposites for photocatalysis. J. Phys. Chem. C 2014, 118, 10963–10971. [Google Scholar] [CrossRef]
  35. Du, A.; Sanvito, S.; Li, Z.; Wang, D.; Jiao, Y.; Liao, T.; Sun, Q.; Ng, Y.H.; Zhu, Z.; Amal, R. Hybrid graphene and graphitic carbon nitride nanocomposite: Gap opening, electron–hole puddle, interfacial charge transfer, and enhanced visible light response. J. Am. Chem. Soc. 2012, 134, 4393–4397. [Google Scholar] [CrossRef]
  36. Takanabe, K.; Kamata, K.; Wang, X.; Antonietti, M.; Kubota, J.; Domen, K. Photocatalytic hydrogen evolution on dye-sensitized mesoporous carbon nitride photocatalyst with magnesium phthalocyanine. Phys. Chem. Chem. Phys. 2010, 12, 13020–13025. [Google Scholar] [CrossRef]
  37. Suter, T.M.; Brazdova, V.; Mccoll, K.; Miller, T.S.; Nagashima, H.; Salvadori, E.; Sella, A.; Howard, C.H.; Kay, C.W.M.; Cora, F.; et al. Synthesis, structure and electronic properties of graphitic carbon nitride films. J. Phys. Chem. C 2018, 122, 25183–25194. [Google Scholar] [CrossRef] [Green Version]
  38. Li, H.; Hu, H.; Bao, C.; Hua, J.; Zhou, H.; Liu, X.; Liu, X.; Zhao, M. Tensile strain induced half-metallicity in graphene-like carbon nitride. Phys. Chem. Chem. Phys. 2015, 17, 6028–6035. [Google Scholar] [CrossRef]
  39. Mahmood, J.; Lee, E.K.; Jung, M.; Shin, D.; Jeon, I.-Y.; Jung, S.-M.; Choi, H.-J.; Seo, J.-M.; Bae, S.-Y.; Sohn, S.-D. Nitrogenated holey two-dimensional structures. Nat. Commun. 2015, 6, 6486. [Google Scholar] [CrossRef] [Green Version]
  40. Li, H.; Hu, H.; Bao, C.; Feng, Z.; Guo, F.; Tian, G.; Liu, Y. Potential application of a porous graphitic carbon nitride as an organic metal-free photocatalyst for water splitting. Diam. Relat. Mater. 2018, 87, 50–55. [Google Scholar] [CrossRef]
  41. Li, H.; Hu, H.; Bao, C.; Guo, F.; Zhang, X.; Liu, X.; Hua, J.; Tan, J.; Wang, A.; Zhou, H. Forming heterojunction: An effective strategy to enhance the photocatalytic efficiency of a new metal-free organic photocatalyst for water splitting. Sci. Rep. 2016, 6, 29327. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Guan, S.; Cheng, Y.; Liu, C.; Han, J.; Lu, Y.; Yang, S.A.; Yao, Y. Effects of strain on electronic and optic properties of holey two-dimensional C2N crystals. Appl. Phys. Lett. 2015, 107, 231904. [Google Scholar] [CrossRef] [Green Version]
  43. Li, H.; Hu, H.; Bai, C.; Bao, C.; Guo, F.; Feng, Z.; Liu, Y. The charge regulation of electronic structure and optical properties of graphitic carbon nitride under strain. Rsc. Adv. 2019, 9, 7464–7468. [Google Scholar] [CrossRef] [Green Version]
  44. Yang, G.; Li, L.; Lee, W.B.; Ng, M.C. Structure of graphene and its disorders: A review. Sci. Technol. Adv. Mater. 2018, 19, 613–648. [Google Scholar] [CrossRef] [Green Version]
  45. Van, T.B.; Torrent, M.; Gonze, X. Interatomic force constants including the DFT-D dispersion contribution. Phys. Rev. B 2016, 93, 144304. [Google Scholar]
  46. Pike, N.; Dewandre, A.; Van Troeye, B.; Gonze, X.; Verstraete, M. Vibrational and dielectric properties of the bulk transition metal dichalcogenides. Phys. Rev. Mater. 2018, 2, 063608. [Google Scholar] [CrossRef] [Green Version]
  47. Kresse, G.; Hafner, J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B 1993, 47, 558. [Google Scholar] [CrossRef]
  48. Kresse, G.; Hafner, J. Ab initio molecular-dynamics simulation of the liquid-metal–amorphous-semiconductor transition in germanium. Phys. Rev. B 1994, 49, 14251. [Google Scholar] [CrossRef]
  49. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169. [Google Scholar] [CrossRef]
  50. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [Green Version]
  51. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953. [Google Scholar]
  52. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758. [Google Scholar] [CrossRef]
  53. Heyd, J.; Scuseria, G.E.; Ernzerhof, M. Hybrid functionals based on a screened Coulomb potential. J. Chem. Phys. 2003, 118, 8207–8215. [Google Scholar] [CrossRef] [Green Version]
  54. Heyd, J.; Scuseria, G.E. Efficient hybrid density functional calculations in solids: Assessment of the Heyd–Scuseria–Ernzerhof screened Coulomb hybrid functional. J. Chem. Phys. 2004, 121, 1187–1192. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Gajdoš, M.; Hummer, K.; Kresse, G.; Furthmüller, J.; Bechstedt, F. Linear optical properties in the projector-augmented wave methodology. Phys. Rev. B 2006, 73, 045112. [Google Scholar]
Figure 1. (a) Atomic structure of g–C8N6 supercell (3 × 3 × 1), the primitive cell is marked using the yellow shaded area, brown atoms are C and gray atoms are N; (b) phonon spectral dispersion diagram of g–C8N6 supercell (2 × 2 × 1).
Figure 1. (a) Atomic structure of g–C8N6 supercell (3 × 3 × 1), the primitive cell is marked using the yellow shaded area, brown atoms are C and gray atoms are N; (b) phonon spectral dispersion diagram of g–C8N6 supercell (2 × 2 × 1).
Catalysts 10 00836 g001
Figure 2. (a) Energy-band structure distributions and (b) electron density of states (DOS) of g–C8N6 by PBE method; (c) energy-band structure distributions and (d) electron density of states (DOS) of g–C8N6 by the HSE06 hybrid function method.
Figure 2. (a) Energy-band structure distributions and (b) electron density of states (DOS) of g–C8N6 by PBE method; (c) energy-band structure distributions and (d) electron density of states (DOS) of g–C8N6 by the HSE06 hybrid function method.
Catalysts 10 00836 g002
Figure 3. (a) Band alignments of g–C8N6 with respect to the standard water redox potentials; (b) absorption coefficient and (c) imaginary part of the dielectric function (in-plane polarization) for g–C8N6 using HSE06 method.
Figure 3. (a) Band alignments of g–C8N6 with respect to the standard water redox potentials; (b) absorption coefficient and (c) imaginary part of the dielectric function (in-plane polarization) for g–C8N6 using HSE06 method.
Catalysts 10 00836 g003
Figure 4. (a) Charge density of the conduction band minimum (CBM), which is located on K point of the 32nd energy band; (b) charge density of the valence band maximum (VBM), which is located on M point of the 31st energy band. The isosurface value is set to 0.003 Å−3.
Figure 4. (a) Charge density of the conduction band minimum (CBM), which is located on K point of the 32nd energy band; (b) charge density of the valence band maximum (VBM), which is located on M point of the 31st energy band. The isosurface value is set to 0.003 Å−3.
Catalysts 10 00836 g004
Figure 5. Case of band-gap changes as the tensile force changes. The color of the squares in figure represents the color of the solar spectrum that should be absorbed.
Figure 5. Case of band-gap changes as the tensile force changes. The color of the squares in figure represents the color of the solar spectrum that should be absorbed.
Catalysts 10 00836 g005
Figure 6. Band-edge potentials of multilayer g–C8N6 compare to standard water redox potentials.
Figure 6. Band-edge potentials of multilayer g–C8N6 compare to standard water redox potentials.
Catalysts 10 00836 g006
Table 1. Work function (WF) and band gap (Eg) of two-, three-, four-, five- and six-layer 2D g–C8N6 computed with HSE06 functional. The band-gap differences (ΔEg) between the multilayers and mono-layer g–C8N6 are shown.
Table 1. Work function (WF) and band gap (Eg) of two-, three-, four-, five- and six-layer 2D g–C8N6 computed with HSE06 functional. The band-gap differences (ΔEg) between the multilayers and mono-layer g–C8N6 are shown.
N123456N
WF (eV)5.305.285.305.265.295.305.30
Eg (eV)2.892.502.402.382.362.352.35
ΔEg (eV)0.000.390.490.510.530.540.54

Share and Cite

MDPI and ACS Style

Li, H.; Hu, H.; Bai, C.; Bao, C.; Liu, C.; Wang, Q.; Guo, F.; Feng, Z.; Yu, H.; Chen, M.; et al. Prediction of a Stable Organic Metal-Free Porous Material as a Catalyst for Water-Splitting. Catalysts 2020, 10, 836. https://doi.org/10.3390/catal10080836

AMA Style

Li H, Hu H, Bai C, Bao C, Liu C, Wang Q, Guo F, Feng Z, Yu H, Chen M, et al. Prediction of a Stable Organic Metal-Free Porous Material as a Catalyst for Water-Splitting. Catalysts. 2020; 10(8):836. https://doi.org/10.3390/catal10080836

Chicago/Turabian Style

Li, Hengshuai, Haiquan Hu, Chenglin Bai, Chunjiang Bao, Cailong Liu, Qinglin Wang, Feng Guo, Zhenbao Feng, Hanwen Yu, Ming Chen, and et al. 2020. "Prediction of a Stable Organic Metal-Free Porous Material as a Catalyst for Water-Splitting" Catalysts 10, no. 8: 836. https://doi.org/10.3390/catal10080836

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