Pan Liab,
Xian Zhaoc,
Honggang Sun*a,
Li Wanga,
Bo Songd,
Baoyu Gaoe and
Weiliu Fan*c
aSchool of Mechanical, Electrical & Information Engineering, Shandong University, Weihai, 264209, China. E-mail: sunhg@sdu.edu.cn; Fax: +86-531-88364864; Tel: +86-531-88366330
bCollege of Physics and Information Engineering, Hebei Normal University, Shijiazhuang, 050024, China
cState Key Laboratory of Crystal Materials, Shandong University, Jinan, 250100, China
dMarine College, Shandong University, Weihai, 264209, China
eSchool of Environmental Science and Engineering, Shandong University, Jinan, 250100, China
First published on 1st August 2016
In the present work, hybrid density functional theory calculations were employed to analyze the electronic structures of ZnGa2O4 with different N impurity concentrations. The electronic transition energies in N-doped ZnGa2O4 with a molar ratio N/O of ∼3% were 3.35 and 1.43 eV for substituted and interstitial N-doping, respectively, which were much smaller than the value of 4.25 eV of pure ZnGa2O4. In 2N-doped ZnGa2O4 with a molar ratio N/O of ∼6%, three models with different N-doping sites (2Ns, 2Ni, and Ns + Ni) are analyzed. The calculated result indicated that the N impurity atoms preferred to form an N–N dimer rather than two apart N atoms in all three models. The electronic transition energies decreased by about 1.52, 0.91, and 2.51 eV for 2Ns-, 2Ni-, and Ns + Ni-doped ZnGa2O4, respectively, which mainly originated from the N–N π* states in the midgap. The half-filled impurity levels, which originated from the single electron of N-doping, were passivated due to the interaction of the two impurity atoms. The defect formation energies indicated that the oxygen vacancy would promote the introduction of a nitrogen impurity. Urea was recommended as a nitrogen source to easily achieve 2Ns-doping forms. Based on the present calculations, 2Ns-doped ZnGa2O4 was considered as the better photocatalyst due to a smaller band gap for visible light response and a suitable redox couple for overall water splitting. Our work provided a theoretical explanation for the origin of the enhanced visible light photocatalytic activity of N-doped ZnGa2O4.
To reduce the large band gap, some transition metals (TM) have been introduced in ZnGa2O4. Zhang's group synthesized Zn1−xCdxGa2O4 solid solution photocatalysts with better UV light activity for the methylene blue degradation, which exhibited the band gaps between 4.8 and 3.6 eV by different Cd components.26 Xu et al. narrowed the band gaps of ZnGa2O4 from 4.7 to 3.1 eV by introducing Cr to substitute Ga ions (ZnGa2−xCrxO4), presenting a better photocatalytic H2 production activity.27 Irie et al. demonstrated that Rh-doped ZnGa2O4 with a good visible light activity for water splitting had a 600 nm light adsorption onset due to the impurity states in the midgap.28 However, TM-doped materials have some disadvantages for efficient photocatalytic reactions, such as rapid electron–hole recombination and thermal instability. To solve these questions, nonmetal doping gives some promising advantages. Hereinto, the N-doped semiconductor photocatalysts, which displays favorable visible light activity, are the most widely investigated.29–34 N-Doped spinel ZnGa2O4 photocatalysts prepared by Parida et al. through solid–solid reaction method based on different N-precursors, which gave the band gap energy of minimal 2.6 eV.35 Lobo et al. also achieved N-doped ZnGa2O4 photocatalysts by combining sol–gel and nitridation steps, presenting the band gap energy of minimal 2.4 eV.9,15 Interestingly, in Parida's work, a significant band gap change from 3.0 to 2.6 eV was observed corresponding to the nitrogen content change from 0.04% to 0.15%. However, N-doped ZnGa2O4 prepared by nitridation steps gives a slight band gap variation from 2.47 to 2.40 eV when the nitrogen content increases from 0.58% to 2.85%. The result suggests that the different nitrogen sources and contents will influence the effect of N-doping on the band structure of ZnGa2O4. Although the effect of N-doping in the semiconductor photocatalyst was discussed by previous experimental and theoretical work, the form of N-doping and the origin of the improved visible light adsorption were still in debate. For example, the form of N-doping in TiO2 was speculated to be substituted, interstitial, NOx or NHx depending on experimental conditions.30–38 Experiments and theoretical calculations also noted that the forms of N-doping were connected with synthesis conditions.34,39 Lobo's work noted that the band gap narrowing of N-doped ZnGa2O4 should be contributed to the mixing states of N 2p and O 2p above the valence band, inducing by the substituted N-doping. However, Batzill's work and Li's work illustrated the p orbital energy of N was close to that of O, which induced less band gap narrowing.40,41 The fact suggests that the large band gap narrowing of about 2.0 eV of N-doped spinel ZnGa2O4 is still complicated and unclear. Therefore, further studies are necessary to reveal the origin of the experimentally observed large red shift.
Previously, many photocatalysts were prepared and studied with considerable photocatalytic activity. However, the number of visible-light-sensitive photocatalysts with overall water splitting ability is limited due to the strict demand of the band relative positions. That is the conduction band minimum (CBM) must be higher than the hydrogen producing potential and the valence band maximum (VBM) must be lower than the water oxidation potential. Therefore, a great deal of effort has been spent on finding semiconductors with suitable band structures. The potentials of VBM and CBM of ZnGa2O4 are ca. 3.0 and −1.2 V (vs. SHE), respectively, which indicates the ZnGa2O4 has a suitable redox couple for overall water splitting.42 Although some transition metals (TM) and nonmetals have been introduced into ZnGa2O4 to reduce the large band gap of 4.2 eV to response visible light, the influence of impurities on the VBM and CBM relative positions is unclear. Generally, the TM-doping introduces some d impurity states in the gap of semiconductor photocatalysts. Although the d states narrowed the original bandgap and simultaneously retains the potential, others depressed the CBM to suppress the reduction potential for H2 production. The redox ability of photogenerated electron–hole pairs decreases as Cd concentration increases in Zn1−xCdxGa2O4 due to the CB and VB shift to Fermi level.26 Irie's work demonstrates that the introduction of Rh-doping in a regular octahedral coordination (Ga sites) creates fully occupied t62g and empty e0g midgap states due to ligand-field splitting of the octahedrally coordinated Rh3+, and these midgap states had suitable potentials to produce hydrogen/oxygen under visible-light irradiation.28 The nonmetal-doping, such as N, and C introduces p impurity states in the gap of semiconductor photocatalysts. N-Doping induces the VB contraction and stronger oxidation potential in rutile TiO2 and a slight VB extension in anatase TiO2.43 The carbon self-doping in g-C3N4 could shift the potentials of both valence and conduction bands positively, which would increase the photooxidation ability and reduce the photoreduction ability.44 These previous studies indicate that the doping in ZnGa2O4 requires careful design for the visible-light sensitive photocatalyst toward overall water splitting.
Theoretical calculations based on local density functional theory (DFT) are considered as a benefit tool to analyze the properties of semiconductors. However, although the structural parameters are agreement with experiments, the band gaps are severely underestimated by the DFT calculations.45 The band gaps of ZnGa2O4 calculated by local density approximation (LDA) and generalized gradient approximation (GGA) are about 2.6–2.8 eV,46,47 which are much less than the experimental value. Some methods, such as GW, LDA + U, hybrid and screened functionals, were developed and used to correct the gap underestimated problem. The GW approximation48 based on the many-body perturbation theory is to date the top method for calculating quasiparticle band structures. Dixit's work gives the band gap of about 4.6 eV of ZnGa2O4 using the GW approximation, which is in good agreement with experiments.49 However, although the GW approximation shows methodological advances, DFT is still widely used for investigating electronic structure features due to much more computationally demanding by the GW calculations. The LDA + U method,50–53 which includes the on-site Coulomb interaction in the Hamiltonian, has been used to descript the effect of the d-electrons. The band gap of ZnGa2O4 calculated by the LDA + U method was about 3.1 eV,46 which slightly suppressed the error of the LDA calculation. However, the underestimate of band gap is still severe in the LDA + U calculations. Hybrid and screened functionals, which mix a fraction of nonlocal potentials such as Hartree–Fock (HF) exchange potentials or screened nonlocal exchange potentials within Kohn–Sham type approaches, go beyond the local and semi-local approximations of DFT.54 A new class of calculations based on hybrid and screened functionals have been reported to give good band gaps in semiconductors.55,56 These researches suggest that the combination of local DFT and nonlocal functionals will be favorable for accurately calculating the band structure of ZnGa2O4 and its changes induced by N-doping.
In the present paper, theoretical calculations with different local, nonlocal, screened exchange and hybrid functionals are implemented to achieve the accurate band structure of cubic spinel ZnGa2O4. Based on the calculated results, the Perdew–Burke–Ernzerhof (PBE0) hybrid functional is employed to obtain microscopic insight into the effect of N-doping on the electronic properties of ZnGa2O4 correlated with N impurity forms and content. The results provide a solid basis for the rationalization of the experimentally observed red shift and the changes of the VB and CB potentials as a result of N-doping in ZnGa2O4.
The N-doped models were constructed based on the fully relaxed pure ZnGa2O4. For the substitutional N-doped model, one oxygen atom in the ZnGa2O4 crystal structure was replaced by one nitrogen atom labeled as Ns. For the interstitial N-doped model, one nitrogen atom was set at the interstitial sites of the ZnGa2O4 crystal lattice labeled as Ni. For the higher N impurity concentration, two nitrogen atoms were introduced in the lattice labeled as 2Ns, 2Ni, and Ns + Ni. Different doping positions of two nitrogen atoms were tested, and the relative stability of the doped models was studied to find the most stable structure according to N–N distance. All the electronic structures were calculated on the corresponding optimized models. The local, nonlocal, screened and hybrid functionals are implemented to calculate the band structures of the pure ZnGa2O4 as shown in Fig. S2.† The calculated results suggest that the hybrid PBE0 functional63 will be favorable for analyzing the changes of the electronic structures of ZnGa2O4 induced by N-doping.
To examine the thermodynamic stability of the different N-doped models, the formation energies (Ef) were calculated according to the following equation,64–67
Ef = Et(Doping) + nOμO − Et(Pure) − nNμN, |
![]() | ||
Fig. 1 The geometrical structures and electron density of (a) pure, (b) Ns-doped, and (c) Ni-doped ZnGa2O4. The unit of the electron density labeled in the table is e Å−3. |
To study the variation of chemical bonding induced by the N-doping for ZnGa2O4, we calculate the total electron density and Mulliken charge population (as shown in Table 1) for the Ns and Ni models. For the pure ZnGa2O4, Fig. 1a shows that the O ions connect with adjacent Ga ions through a common electron cloud, presenting covalent O–Ga bonds. For the Ns-doped ZnGa2O4, Fig. 1b shows more electron cloud between Ns ion and Ga cations (0.672 e Å−3) than that between O ions and Ga cations (0.609 e Å−3), which is consistent with the shorter Ns–Ga and Ns–Zn bonds after geometrical optimization. The charges of −0.87 |e| on the Ns ion are close to −0.84 |e| on the original O ions. The result suggests that N2− ion is localized in the lattice, which will introduce one acceptor level in the ZnGa2O4. For the Ni-doped ZnGa2O4, Fig. 1c shows a covalent interaction between the Ni ion and the adjacent O ion. The charges on the Ni ion and the O ion are −0.48 |e| and −0.50 |e|, respectively. The total charges of −0.98e on the NO species suggests that the NO2− species exist in the lattice.
Models | Bond lengths (Å) | Charges (|e|) | ||||||
---|---|---|---|---|---|---|---|---|
O/N–Ga | O/N–Zn | N–O | N–N | Ga | Zn | O | N | |
Pure | 2.030 | 2.060 | 1.20 | 0.97 | −0.84 | |||
Ns | 2.019 | 1.956 | 1.21 | 0.96 | −0.84 | −0.91 | ||
Ni | 1.960 | 1.951 | 1.368 | 1.22 | 0.96 | −0.50 | −0.48 | |
2Ns | 1.994 | 1.950 | 1.696 | 1.20 | 0.93 | −0.84 | −0.77 | |
1.987 | 1.945 | −0.77 | ||||||
2Ni | 2.084 | 1.797 | 1.176 | 1.22 | 0.93 | −0.75 | −0.21 | |
2.014 | −0.13 | |||||||
Ns + Ni | 1.971 | 2.160 | 1.284 | 1.20 | 0.97 | −0.84 | −0.46 | |
1.975 | 2.131 | −0.46 | ||||||
2Ns + VO | 1.918/1.945/1.988 | 1.929 | 1.21 | 0.95 | −0.84 | −1.00 | ||
1.932/1.962/1.978 | 1.996 |
![]() | ||
Fig. 2 The band structures, and density of states of (a) pure, (b) Ns-doped, and (c) Ni-doped ZnGa2O4. The dash line presents the Fermi level. |
For the Ns-doped ZnGa2O4, as shown in Fig. 2b, the electronic transition energy decreases to about 3.35 eV due to three impurity levels localized at the VBM. The host band gap of 4.35 eV from the host VBM to CBM is slightly larger than 4.25 eV of the pure ZnGa2O4. We contribute the larger band gap to a reduction of the Coulomb repulsion and a contraction of the band, resulting from the removal of one electron when one N atom replaces one O atom in the cell. The PDOS shows that the impurity levels are mainly contributed by the mixing states of N 2p states with little O 2p states. It is clear that the substituted N-doping improve the solar energy utilization of ZnGa2O4 through some continuous states above the VBM, and Ns-doped ZnGa2O4 still shows overall water splitting ability. However, the narrowing of the electronic transition energy is still not enough to response the visible light, which demands the energy of less than 2.8 eV.
For the Ni-doped ZnGa2O4, Fig. 2c shows two isolated impurity levels in the gap. The theoretical gap from the host VBM to the CBM has a slight increasing of about 0.18 eV, which is contributed to the p–p repulsion of the Ni-doping and the VBM. However, the electron excited energy from the top impurity level to the CBM is about 1.43 eV, which is much smaller than 4.25 eV of pure phase. The PDOS indicates that the two midgap levels are mainly originated from the mixing states of N 2p states and O 2p states. The further highest occupied molecular orbital (HOMO) analysis as shown in Fig. 3c shows a remarkable NO π-antibonding character. Therefore, we conclude that the electronic excitations from the NO π* states are one of the reasons of the visible light response in Ni-doped ZnGa2O4. Importantly, the calculated band structure indicates that the overall water splitting ability of ZnGa2O4 has been break due to the deep impurity levels. In addition, the calculated band structures of Ns- and Ni-doped ZnGa2O4 show the top impurity level crossing the Fermi level presents half occupied characteristic, which is agreement with the charge population analysis. However, the partially occupied impurity states go against the photocatalytic activity of semiconductors because these states can act as recombination centers to suppress the photogenerated current and reduce the UV light activity. Thus, the passivated approach is desired to compensate the half occupied impurity states.
![]() | ||
Fig. 4 The geometrical structures and electron density of (a) 2Ns-doped, (b) 2Ni-doped, and (c) Ns + Ni-doped ZnGa2O4. The unit of the electron density labeled in the table is e Å−3. |
Fig. 4 shows that the two N-doping ions are bounded together by a common electron cloud, presenting a strong covalent-like characteristic. For 2Ns-doped ZnGa2O4, more electrons are shared by Ns-doping and Ga ions (more than 0.672 e Å−3), which give stronger Ns–Ga covalent bonds. The interaction of the two Ns ions reduces the charges on each Ns ion to −0.77 |e|. For 2Ni-doped ZnGa2O4, the total electron density map indicates that the Ni–O bonds are much weaker than that in Ni-doped ZnGa2O4. This result is consistent with the smaller total charges of −0.34 |e| on the two interstitial N ions, which is contributed to the interaction of the two Ni ions. For Ns + Ni-doped ZnGa2O4, the effect of the N2 dimer is similar to the original O ions. However, the two N–Ga bonds present stronger covalent-like characteristic by capturing more electrons (−0.92 |e|). The present results give strong interaction between the two N-doping for the three 2N-doped models, which will significantly affect the electronic structures of ZnGa2O4.
![]() | ||
Fig. 5 The band structures and density of states of (a) 2Ns-doped, (b) 2Ni-doped, and (c) Ns + Ni-doped ZnGa2O4. The dash line presents the Fermi level. |
![]() | ||
Fig. 6 The highest occupied molecular orbital (HOMO) of (a) 2Ns-doped, (b) 2Ni-doped, and (c) Ns + Ni-doped ZnGa2O4. |
Fig. 5b gives the band structure of 2Ni-doped ZnGa2O4. Only one isolated level localizes in the gap, which induces the electronic transition energy decrease to 3.34 eV. The host band gap has a slight change of about 0.08 eV, and there are no half-filled levels in the band. The calculated band structure shows a significant difference compared with that of Ni-doped ZnGa2O4, in which two isolated levels, smaller electronic transition energy of 1.43 eV and larger host band gap increasing of 0.18 eV are observed. The DOS and PDOS of 2Ni-doped ZnGa2O4 show that the midgap level is mainly contributed by O 2p states with little N 2p states. The HOMO analysis as shown in Fig. 6b gives an –Ni–Ni–O– mixing characteristic, which consists of Ni–Ni π* and Ni–O σ orbitals. Importantly, some N 2p states localize above the Ga 4s states of the CBM mixing with empty states. Based on the calculated results and combined with charge analysis, we conclude that the unpaired electrons of the Ni-doping are paired, which induce one empty anti-bond molecular orbital moving up and the other occupied anti-bond molecular orbital moving down. The up orbital results in the weakening of the p–p repulsion between the midgap states and the VBM. Thus the broadening of the host band gap is smaller than that in Ni-doped ZnGa2O4. The down orbital increases the electronic transition energy of about 1.91 eV from the impurity states to CBM compared with that in Ni-doped ZnGa2O4. Therefore, although the Ni–Ni impurity induces the electronic transition energy decreasing than pure phase, the interaction of two Ni-doping atoms goes against the further improving of the visible light response of N-doped ZnGa2O4.
The band structure of Ns + Ni-doped ZnGa2O4 as shown in Fig. 5c gives the host band gap of 4.46 eV and no half-filled level localizes in the band structure. One isolated level localizes in the gap, and the electronic transition energy from the level to CBM is 1.74 eV, which is much smaller than 4.25, 3.35, 3.34, and 2.73 eV of pure, Ns-, 2Ni-, and 2Ns-doped ZnGa2O4, and just larger than 1.43 eV of Ni-doped ZnGa2O4. The DOS and PDOS show the N 2p states are responsible for the midgap level. The HOMO analysis presents a remarkable Ns–Ni π anti-bonding character. Therefore, the electronic transition of π*–s is one of reasons that inducing visible light absorption in experiments. The character is agreement with that in 2Ns-doped ZnGa2O4, while is different to that in 2Ni-doped ZnGa2O4. In addition, there are some N 2p impurity states localizing above the Ga 4s states of the CBM. Obviously, this character is similar to that in 2Ni-doped ZnGa2O4, which originates from the charge compensation of the two N-doping. Although the charge compensation may lower the impurity levels to broaden the electronic transition energy as discussed above, two effects play a positive role to narrow the energy in Ns + Ni-doped ZnGa2O4. One is that the Ni-doping introduces more electrons resulting in a strong p–p repulsion of the impurity states and valence bands, the other is that the mixing of the N 2p states and conduction bands as shown in Fig. 5c lowers the CBM. Therefore, based on the calculated results, we conclude that the Ns + Ni-doping units the advantages of substituted and interstitial N-doping, which is favorable for improving the solar energy utilization of ZnGa2O4. However, the redox potentials of the impurity levels indicate that Ns + Ni-doped ZnGa2O4 has no O2 producing ability.
Models | Etr (eV) | Ef (eV) | ||||
---|---|---|---|---|---|---|
N2 | NH3 | N2H4 | Urea | Hexamine | ||
Pure | 4.25 | |||||
Ns | 3.35 | 5.28 | 3.83 | 4.84 | 2.47 | 4.43 |
Ni | 1.43 | 5.12 | 3.67 | 4.68 | 2.41 | 4.27 |
2Ni | 2.73 | 6.80 | 3.90 | 5.93 | 1.39 | 5.10 |
2Ns | 3.34 | 9.62 | 6.62 | 8.75 | 4.21 | 7.92 |
Ns + Ni | 1.74 | 7.17 | 4.27 | 6.30 | 1.76 | 5.47 |
2Ns + VO | 3.38 | 3.25 | 0.35 | 2.38 | −2.16 | 1.55 |
The oxygen vacancy has a significant effect on the band structure of N-doped ZnGa2O4. As shown in Fig. 7c, some impurity levels form a continuous impurity band and extend the width of about 0.8 eV of the VB. The band gap decreases to 3.38 eV and the half-filled level induced by Ns-doping disappears. Obviously, the electron transferring between oxygen vacancy and Ns-doping will be favor of improving the photocatalytic activity. Fig. 7d shows the impurity levels are mainly contributed by the mixing states of N 2p states and O 2p states. The energy of the mixing states is lower, which suggests stronger oxidizing ability of than that in Ns-doped model VB. In addition, the calculated band structure indicates that 2Ns + VO-doped ZnGa2O4 still keep the overall water splitting ability.
Footnote |
† Electronic supplementary information (ESI) available: The convergence tests, the band structures of the pure ZnGa2O4 calculated by different functionals, the relative energies of 2N-doped ZnGa2O4 with different doping sites. See DOI: 10.1039/c6ra09655g |
This journal is © The Royal Society of Chemistry 2016 |