Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Enhanced photoelectric performance of (2Al, S) co-doped rutile SnO2

Minmin Guoa, Huimin Yanga, Mengting Gaoa, Erhui Zhanga, Zhenhai Liang*a and Peide Hanb
aCollege of Chemistry and Chemical Engineering, Taiyuan University of Technology, Taiyuan 030024, PR China. E-mail: liangzhenh@sina.com; liangzhenhai@tyut.edu.cn
bCollege of Materials Science and Engineering, Taiyuan University of Technology, Taiyuan 030024, PR China

Received 18th July 2017 , Accepted 30th August 2017

First published on 5th September 2017


Abstract

In this study, theoretical calculations and experiments have been carried out to investigate the photoelectric performance of (2Al, S) co-doped rutile SnO2. The electronic structures are studied by density functional theory (DFT). It is found that the metal Al can assist the bonding of the incorporated S with the neighboring O in SnO2, introducing new energy levels in the forbidden band of SnO2, which enhance the photoelectric performance. Meanwhile, the experiments are conducted to verify this. The (2Al, S) co-doped SnO2 with different doping ratios are prepared by a hydrothermal method. The samples are characterized by X-ray powder diffraction (XRD), scanning electron microscopy (SEM) and X-ray photoelectron spectroscopy (XPS). Results show that all the samples have rutile structure without any extra phase, and the dopant S2− ion was implanted into the crystalline lattice of (2Al, S) co-doped SnO2 and Al dopants replaced Sn atoms. The photoelectric performance tests show Al and S co-doping can improve the photoelectric performance, especially with a doping ratio of 5%, when the photocurrent reaches maximum of 3.0 μA cm−2 which is almost twice as much as pure SnO2, and the impedance is the smallest. The experiments results are consistent with our theoretical calculations. These findings are expected to be helpful for the design of highly active tin oxide-based photoelectric materials.


Introduction

SnO2 is a kind of rutile semiconductor oxide with wide band gap of 3.6 eV.1,2 It has advantages of good optical transmission to visible light, big ultraviolet absorption coefficient, low preparation temperature, stable chemical property and strong acid and alkali resistance at room temperature.3–6 Therefore, it can been widely used in organic light emitting diodes (OLED), liquid crystal displays (LCD), solar cells, gas sensors and field effect transistors.7–11 The preparation and performance of SnO2 materials have attracted wide attention. However, the photoelectric performance of pure SnO2 is poor. Doping can adjust the change of band gap and promote the separation of photo-induced electrons and holes, which improves the photoelectric performance. Therefore, research mainly focuses on the influence of doping on the photoelectric performance and stability of SnO2. Huang et al.12 prepared Zn-doped SnO2 nanorods by a simple hydrothermal method. The results show that photocatalytic activity of the synthesized Zn-doped SnO2 nanorods is much higher than pure SnO2 nanorods and bulk SnO2 powders. Lee et al.13 studied electrocatalytic activities and stabilities of Pt supported on Sb-doped SnO2 (ATO) for methanol (MOR) and ethanol (EOR) oxidation reactions. The results show that the Pt/ATO exhibited much higher electrochemical stabilities than Pt supported on carbon (Pt/C). Ahmed et al.14 synthesized Al doped SnO2 thin films by a sol–gel dip coating technique with different ratios of Al on glass and silicon substrates. The results show doping can enhance its photoelectric performance. With regard to SnO2 doping, there are many researches on single doping. The reports on co-doping is few, co-doping is firstly proposed during the study on semiconductor materials with single polarity and wide band gap.15 Two or more elements doping can further improve the photoelectric performance of materials by theory.16

In our paper, to gain detailed insight into the effect of Al and S dopants on SnO2 photoelectric performance, we used density function theory to investigate the electronic structures. Based on the theory calculation, we prepared (2Al, S) co-doping SnO2 photoelectrode. X-ray diffraction (XRD), scanning electron microscopy (SEM). X-ray photoelectron spectra (XPS) were taken to characterize pure SnO2 and different ratios of co-doping SnO2. Meanwhile, photoelectric performance of (2Al, S) co-doped rutile SnO2 are tested. As a result, the exceptionally high photoelectric performance observed for (2Al, S) co-doped SnO2 is ascribed to metal Al assisting S–O bonding.

Results and discussion

Lattice parameters

Fig. 1 shows the optimized configurations of pure SnO2 and (2Al, S) co-doped SnO2. Compared with pure SnO2, the positions of Al atoms and S atom move to a certain degree accompanying the distortion of bonds. The distance of co-planar O and adjacent perpendicular O atom in pure SnO2 is 2.902 Å. However, for the (2Al, S) co-doped SnO2, the distance of S and O atoms is 1.641 Å, indicating the formation of S–O covalent bond. The definite lattice parameters are listed in Table 1. The lattice parameters are all bigger than that of pure one, this may because the radius of S2− is bigger than that of O2− radius (O2−: 1.22 Å, S2−: 1.70 Å).
image file: c7ra07891a-f1.tif
Fig. 1 2 × 2 × 2 SnO2 supercell (a) and (2Al, S) co-doped SnO2 supercell (b).
Table 1 The calculated lattice parameters of pure and (2Al, S) co-doped 2 × 2 × 2 SnO2
  a (Å) b (Å) c (Å)
Pure SnO2 9.475 9.475 6.373
(2Al, S) co-doped SnO2 10.065 9.761 6.519


Band structure and density of states

The calculated band gap of pure SnO2 is 1.07 eV, which is same with other theory calculations.17 However, the experiment value of SnO2 band gap is 3.6 eV. The difference may be caused by the underestimation of density functional theory to band gap.18 The defect of density functional theory leads to smaller forbidden band comparing to experiment value. In this paper, the calculated band gaps are used to compare the changes. Hence, the error does not affect the qualitative analysis on calculation results. As shown in Fig. 2, the band gap of (2Al, S) co-doped SnO2 is 0.92 eV, while the band gap of pure SnO2 is 1.07 eV. The difference suggests that (2Al, S) co-doping can narrow the band gap. Further analysis shows that the co-doping introduce fully occupied energy levels, which forms new top of valance band. At the same time, the bottom of conduct band move down.
image file: c7ra07891a-f2.tif
Fig. 2 Band structures of pure SnO2 (a) and (2Al, S) co-doped SnO2 (b).

The total density of states (TDOS) and partial density of states (PDOS) of pure and (2Al, S) co-doped SnO2 shown in Fig. 3 are discussed to explain the formation mechanism of fully occupied energy levels. The valence band of pure SnO2 is mainly composed of O 2p state mixing with small Sn 3d state, and the conduction band is constituted of O 2p and Sn 5p states. While for (2Al, S) co-doped SnO2, the fully occupied energy levels of valence band are the hybridization of O 2p and S 3p states, which indicate the formation of S–O bond. Similar S–O bond can be found for TiO2 doping.19 The bonding mechanism (Fig. 4) are as follows.


image file: c7ra07891a-f3.tif
Fig. 3 TDOS (a) and PDOS (b) of pure SnO2 and (2Al, S) co-doped SnO2.

image file: c7ra07891a-f4.tif
Fig. 4 Schematic plot of bonding mechanism for Al assisted S–O bonding in (2Al, S) co-doped SnO2.

Two metal acceptors Al substitute for two Sn atoms of SnO2 supercell, introducing two holes. Then, one S atom substitute for adjacent O atom, which make two electrons transfer from S atom to metal acceptors Al due to smaller electron negativity of S atom comparing with O atom. At this point, 3pz orbital of S atom is empty, which make the formation of covalent bond with adjacent O atom possible. The hybridization of pz orbitals of both S atom and O atom can form bonding orbital σ and antibonding orbital σ*, while their px and py orbitals can form bonding orbital π and antibonding orbital π*. The fully occupied energy levels are the antibonding orbital π*. The band gap is narrowed by the antibonding orbital π* and conductive band bottom. Therefore, the metal acceptors Al can assist S–O bonding by taking two electrons away from S atom.

SEM analysis

Fig. 5 shows SEM images of different ratios of (2Al, S) co-doped SnO2. The pure SnO2 is spherical particle and the surface is nonuniform. For (2Al, S) co-doped SnO2, the particle sizes decrease. Small particle size can short spreading time of charge carrier from the body to the surface and reduce electron–hole recombination rate, which promote the photocatalytic activity. Different levels of reunion appear in all samples. When the doping ratio is 5%, the particles are more dispersed and the granularity is more uniform comparing with others. Hence, the photoelectric performance is the best.
image file: c7ra07891a-f5.tif
Fig. 5 SEM images of different ratios of (2Al, S) co-doped SnO2. (a) 0%, (b) 5%, (c) 10% and (d) 15%.

XRD analysis

Fig. 6 show the XRD patterns of different co-doping ratios of SnO2 crystal. The peaks appeared at 2θ = 26.506°, 33.945° and 51.94°, which are the characteristic peaks of rutile SnO2.20,21 The corresponding crystal faces are (110), (101) and (211). No characteristic diffraction peaks of other SnO2 crystal were found, indicating that they still keep the structure of rutile SnO2 after co-doping. In addition, comparing with pure SnO2, no significant shift in the peak positions was found.
image file: c7ra07891a-f6.tif
Fig. 6 XRD patterns of pure SnO2 and different ratios of (2Al, S) co-doped SnO2.

XPS analysis

The XPS spectrum of 5% (2Al, S) co-doping SnO2 are displayed in Fig. 7. Fig. 7a shows O 1s spectrum, it has two asymmetry peaks, which indicate the presence of two different types of oxygen states.22 The peaks at 530.88 eV and 531.84 eV belong to lattice oxygen (O2−) and adsorbed oxygen23,24 respectively. The peak of O 1s at 530.88 eV belongs to SnO2−x structural O atoms or oxygen inside non-stoichiometric oxides within the surface region. The peak at 531.84 eV suggests that oxygen from O2 molecules of the ambient atmosphere adsorbed on the grains or surface of SnO2. Compared to O 1s spectrum of pure SnO2 (Fig. 7e), the peaks move to lower binding energy, which indicates that (2Al, S) co-doped SnO2 contain more oxygen deficiency. The spectrum of Sn 3d in Fig. 7b have two binding energy peaks. The peaks of Sn 3d5/2 and Sn 3d3/2 respectively locate at 486.98 eV and 495.43 eV, suggesting the presence of Sn4+. Also, the different value between Sn 3d5/2 and Sn 3d3/2 is 8.45 eV, which is consistent with the standard spectrum of Sn as reported in the Handbook of X-ray photoelectron spectroscopy.26 However, compared to Sn 3d spectrum of pure SnO2 (Fig. 7f), the peaks moves towards high binding energy. The generation of chemical displacement is not only the potential energy change caused by valence electron transfer, but also the contribution of lattice field. For (2Al, S) co-doped SnO2, due to the small difference between the electronegativity of Al and Sn, there is no chemical displacement caused by valence electron transfer. The doped Al dispersed in the lattice, causing SnO2 lattice defects, which can lead to the change of the lattice field, resulting in the increase of binding energy. S 2p XPS spectrum is shown in Fig. 7c, the peaks at 160.95 eV and 162.81 eV indicate the presence of S2−. Therefore, we can conclude that S2− ion replace the O2− ion in the lattice of SnO2. However, it also has a peak at 167.49 eV except the above two peaks, which suggests the existence of S(+6). This may be caused by SO42− ions adsorbed on the surface.25 As shown in Fig. 7d, the binding energy of Al 2p peak at 74.55 eV shows that the Al3+ ions replace the Sn4+ ions of SnO2.
image file: c7ra07891a-f7.tif
Fig. 7 XPS spectrum of 5% (2Al, S) co-doped SnO2 (a) O 1s, (b) Sn 3d, (c) S 2p, (d) Al 2p and pure SnO2 (e) O 1s, (f) Sn 3d.

Transient photocurrent response test

Fig. 8 display the transient photocurrent response of pure SnO2 and different doping ratios of (2Al, S) co-doped SnO2 photoelectrode under 0.5 V bias. The photocurrent value of pure SnO2, 5%, 10%, and 15% (2Al, S) co-doped SnO2 reached 1.7, 3.0, 2.4 and 1.2 μA cm−2. When the co-doping amount is 5%, the photocurrent value reaches maximum (the photocurrent values of 0–10% are shown in Fig. S1 and S2). The common photoelectric metal oxides materials are TiO2 and ZnO, which show excellent photoelectric properties among metal oxides. Our photocurrent value of 5% (2Al, S) co-doped SnO2 is almost equal to TiO2 with etching depth of 387 nm[thin space (1/6-em)]27 and ZnO under UV light irradiation.28 As the amount of co-doping continuously increase, the photocurrent value decreases. The results show that suitable amount of co-doping can improve the sunlight utilization rate of SnO2. 5% (2Al, S) co-doped SnO2 has small particles and large surface area, which can increase contact area with sunlight. This is consistent with above SEM analysis. Meanwhile, a proper amount of co-doping can inhibit the recombination of photoelectric charge by promoting the separation of photoelectron and hole pairs. In addition, the photocurrent decreases with illumination time, which may due to the slight corrosion of electrodes.
image file: c7ra07891a-f8.tif
Fig. 8 Transient photocurrent response curve of pure SnO2 and different ratios of (2Al, S) co-doped SnO2.

Electrochemical impedance spectroscopy

Electrochemical impedance curve of pure SnO2 and different doping amount of (2Al, S) co-doped SnO2 electrodes are shown in Fig. 9. The curve is composed of teratogenic arc of high frequency area and straight line of low-frequency area. The arc diameter of 5% (2Al, S) co-doped SnO2 is the smallest, suggesting minimal electrode reaction resistance and more active points of electrochemical reaction. Meanwhile, the slope is relatively higher than others, which shows faster diffusion of ions. In addition, 5% (2Al, S) co-doped SnO2 has smaller solution resistance, indicating that moderate amount of co-doping can increase the effective contact area of the material and shorten the transmission path of electron and ion.
image file: c7ra07891a-f9.tif
Fig. 9 Electrochemical impedance plots of pure SnO2 and different ratios of (2Al, S) co-doped SnO2.

Conclusions

In summary, DFT calculations have been carried out to detailedly study the effect of photoelectric performance of (2Al, S) co-doped SnO2. Compared to pure SnO2, the (2Al, S) co-doped SnO2 has smaller band gap of 0.92 eV, introducing new energy levels in the forbidden band, which enhance the photoelectric performance in a certain degree. Furthermore, the S–O bonding mechanism assisting by metal Al was explored. The theory calculation can provide evidence supporting experiment conduct. Based on this, different ratios of (2Al, S) co-doped SnO2 were synthesized by hydrothermal method. In order to study the photoelectric performance, we prepared (2Al, S) co-doped SnO2 electrodes. The photocurrent of 5% (2Al, S) co-doped SnO2 reached maximum value of 3.0 μA cm−2. At this point, the impedance is the smallest. These findings can pave the way for developing electrodes with high photoelectric performance in photoelectric catalysis field.

Method

Calculation methods and calculation model

In this study, first-principles calculation was carried out using the Cambridge Serial Total Energy Package (CASTEP) code29,30 based on density functional theory (DFT).31 The exchange correlation potential was described with generalised gradient approximation (GGA) in the scheme of Perdew–Burke–Ernzerhof function (PBE).32,33 The Brillouin zone sampling point K values are 4 × 4 × 3. The cutoff energy for plane waves is set as 340 eV. All geometry structures are fully relaxed until the convergence criteria of energy and force are less than 2.0 × 10−5 eV per atom and 0.03 eV Å−1 respectively. The maximum stress is 0.1 GPa.

In calculation, we choose a 2 × 2 × 2 SnO2 supercell with 48 atoms as calculation model. It has two types of Sn atoms and oxygen atoms respectively. The Sn atoms respectively locate in the vertex and body position. The oxygen atoms can be divided into two classes: one is coplanar with Sn atom, the other is perpendicular to the surface. For (2Al, S) co-doped rutile SnO2 supercell, two metal Al replace the body-centered Sn atoms, and S atom replace O atom perpendicular to the surface. For an investigation on the electronic structure and ground-state property, the valence electronic configurations are O-2s22p4, S-3s23p4, Sn-5s25p2and Al-3s23p1 states. To obtain reliable results, structural optimization for bulk SnO2 obtained the following parameters: a = 4.737 Å, c = 3.186 Å, in good agreement with experiment (a = 4.734 Å, c = 3.187 Å).34

Synthesis of (2Al, S) co-doped SnO2 electrodes

The synthesis of (2Al, S) co-doped SnO2 used a one-step hydrothermal method.35,36 The process is as follows: SnCl4·5H2O (10.058 g) was firstly dissolved in a beaker with 60 mL deionized water and stirred about 1 h. The sulfourea and AlCl3·6H2O, as S and Al sources, were added into the solution. After stirring for 4 h, the solution was transferred to a 100 mL stainless-steel autoclave with a Teflon liner. The autoclave was sealed and heated at 433 K for 12 h without shaking or stirring during this period. Black precipitate was collected after the autoclave cooled down to room temperature and washed with distilled water several times in order to remove any impurities. After that, the samples were obtained by high temperature calcinations at 573 K for 2 h. Three (2Al, S) co-doped SnO2 samples with initial doping molar ratios of 5%, 10% and 15% were synthesized. In the preparation, the molar ratio of Al to S is 2[thin space (1/6-em)]:[thin space (1/6-em)]1. Pure SnO2 was synthesized by the same procedure but without any additives.

The above samples were grinded and dispersed in glycerine forming 4 suspensions: 0%, 5%, 10% and 15%. The suspensions were ultrasonicated for 5 min. Then electrodes were prepared by dispensing the suspensions on glass substrates. Finally, the electrodes were burned for 2 h to reserve.

Characterization

The crystal structures of the samples were identified by X-ray diffraction (XRD, X'pert PRO). The morphologies of the samples were characterized by scanning electron microscopy (SEM) with a JSM-7001F (JEOL, Japan). Elemental compositions of the samples were detected by X-ray photoelectron spectroscopy (XPS, Thermo, ESCALAB 250Xi, USA).

Photoelectric performance tests

A trielectrode system was carried out to test the photoelectric performance of SnO2. The pure SnO2 and different ratios of (2Al, S) co-doped electrodes, Pt electrode and saturated calomel electrode (SCE) were respectively used as tested electrodes, auxiliary electrode and reference electrode. Na2SO4 solution with concentration of 1.0 mol L−1 was used as electrolyte solution. The bias voltage is set as 0.5 V.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work is jointly funded by the National Natural Science Foundation of China and Shenhua Group Corp. (Grant No. U1261103), the Natural Science Foundation of Shanxi Province of China. (Grant No. 201601D011023) and Graduate Science and Technology Innovation Fund Project of Shanxi (No. 2017BY053).

References

  1. C. Drake and S. Seal, Appl. Phys. Lett., 2007, 90, 104 CrossRef.
  2. S. H. Nam and J. H. Boo, J. Nanosci. Nanotechnol., 2012, 12, 1559–1562 CrossRef CAS PubMed.
  3. N. Chiodini, A. Paleari, G. Spinolo and P. Crespi, J. Non-Cryst. Solids, 2003, 322, 266–271 CrossRef CAS.
  4. T. R. Giraldi, M. T. Escote, A. P. Maciel, E. Longo and E. R. Leite, Thin Solid Films, 2006, 515, 2678–2685 CrossRef CAS.
  5. M. V. Reddy, L. Y. Tse, K. Z. B. Wen and B. V. R. Chowdari, Mater. Lett., 2015, 138, 231–234 CrossRef CAS.
  6. S. Yu, W. Yang, L. Li and W. Zhang, Sol. Energy Mater. Sol. Cells, 2016, 144, 652–656 CrossRef CAS.
  7. Y. F. Li, W. J. Yin, R. Deng, R. Chen, J. Chen, Q. Y. Yan, B. Yao, H. D. Sun, S. H. Wei and T. Wu, NPG Asia Mater., 2012, 4, e30 CrossRef.
  8. S. Mihaiu, A. Braileanu, D. Crisan and M. Zaharescu, Rev. Roum. Chim., 2003, 48, 939–946 CAS.
  9. A. N. M. Green, E. Palomares, S. A. Haque, J. M. Kroon and J. R. Durrant, J. Phys. Chem. B, 2005, 109, 12525–12533 CrossRef CAS PubMed.
  10. N. Barsan, M. Schweizer-Berberich and W. Göpel, Fresenius' J. Anal. Chem., 1999, 365, 287–304 CrossRef CAS.
  11. Y. Cheng, R. Yang, J. P. Zheng, Z. L. Wang and P. Xiong, Mater. Chem. Phys., 2012, 137, 372–380 CrossRef CAS.
  12. H. Huang, S. Tian, J. Xu, Z. Xie, D. Zeng, D. Chen and G. Shen, Nanotechnology, 2012, 23, 105502 CrossRef PubMed.
  13. K. S. Lee, I. S. Park, Y. H. Cho, D. S. Jung, N. Jung, H. Y. Park and Y. E. Sung, J. Catal., 2008, 258, 143–152 CrossRef CAS.
  14. S. F. Ahmed, S. Khan, P. K. Ghosh, M. K. Mitra and K. K. Chattopadhyay, J. Sol–Gel Sci. Technol., 2006, 39, 241–247 CrossRef CAS.
  15. T. Yamamoto and H. Katayama-Yoshida, Jpn. J. Appl. Phys., Part 2, 1999, 38, L166–L169 CAS.
  16. T. Yamamoto and H. Katayama-Yoshida, Phys. B, 2001, 302–303, 155–162 CrossRef.
  17. F. E. H. Hassan, A. Alaeddine, M. Zoaeter and I. Rachidi, Int. J. Mod. Phys. B, 2005, 19, 4081–4092 CrossRef.
  18. C. Stampf and C. G. Van de Walle, Phys. Rev. B, 1999, 59, 5521–5535 CrossRef.
  19. M. Niu, D. Cheng and D. Cao, Int. J. Hydrogen Energy, 2013, 38, 1251–1257 CrossRef CAS.
  20. R. K. Mishra and P. P. Sahay, Ceram. Int., 2012, 38, 2295–2304 CrossRef CAS.
  21. A. K. Singh and U. T. Nakate, Adv. Nanopart., 2013, 02, 66–70 CrossRef.
  22. B. Thomas and B. Skariah, J. Alloys Compd., 2015, 625, 231–240 CrossRef CAS.
  23. P. Y. Liu, J. F. Chen and W. D. Sun, Vacuum, 2004, 76, 7–11 CrossRef CAS.
  24. M. Kennedy, Ph.D. Dissertation, University of DuisburgEssen, 2004.
  25. T. Ohno, Appl. Catal., A, 2004, 265, 115–121 CrossRef CAS.
  26. C. D. Wagner, W. M. Riggs, L. E. Davis and J. F. Moulder, Handbook of X-Ray Photoelectron Spectroscopy, Perkin Elmer, Eden Prairie, 1979 Search PubMed.
  27. S. Y. Luo, B. X. Yan and J. Shen, ACS Appl. Mater. Interfaces, 2014, 6, 8942–8946 CAS.
  28. Y. J. Wang, R. Shi, J. Lin and Y. F. Zhu, Energy Environ. Sci., 2011, 4, 2922–2929 CAS.
  29. B. Andriyevsky, M. Romanyuk and V. Stadnyk, J. Phys. Chem. Solids, 2009, 70, 1109–1112 CrossRef CAS.
  30. S. J. Clark, M. D. Segall, C. J. Pickard, P. J. Hasnip, M. J. Probert, K. Refson and M. C. Payne, Z. Kristallogr., 2005, 220, 567–570 CAS.
  31. F. M. Bickelhaupt and E. J. Baerends, Rev. Comput. Chem., 2000, 15, 1–86 CAS.
  32. E. Fabiano, L. A. Constantin and F. D. Sala, Int. J. Quantum Chem., 2013, 113, 673–682 CrossRef CAS.
  33. J. Nisar, C. Århammar, E. Jämstorp and R. Ahuja, Phys. Rev. B, 2011, 84, 2250–2262 CrossRef.
  34. B. Thangaraju, Thin Solid Films, 2002, 402, 71–78 CrossRef CAS.
  35. L. M. Fang and X. T. Zu, Adv. Mater. Res., 2007, 26–28, 683–686 CrossRef CAS.
  36. S. Blessi, M. M. L. Sonia, S. Vijayalakshmi and S. Pauline, Int. J. ChemTech Res., 2014, 6, 2153–2155 CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c7ra07891a

This journal is © The Royal Society of Chemistry 2017