Skip to main content

Hydrothermal derived nitrogen doped SrTiO3 for efficient visible light driven photocatalytic reduction of chromium(VI)

Abstract

In this work, we report on the synthesis of nitrogen doped SrTiO3 nanoparticles with efficient visible light driven photocatalytic activity toward Cr(VI) by the solvothermal method. The samples are carefully characterized by X-ray diffraction, scanning electron microscopy, transmission electron microscopy, UV–Vis diffuse reflectance spectroscopy and photocatalytic test. It is found that nitrogen doping in SrTiO3 lattice led to an apparent lattice expansion, particle size reduction as well as subsequent increase of Brunner–Emmet–Teller surface area. The visible light absorption edge and intensity can be modulated by nitrogen doping content, which absorption edge extends to about 600 nm. Moreover, nitrogen doping can not only modulate the visible light absorption feature, but also have consequence on the enhancement of charge separation efficiency, which can promote the photocatalytic activity. With well controlled particle size, Brunner–Emmet–Teller surface area, and electronic structure via nitrogen doping, the photocatalytic performance toward Cr(VI) reduction of nitrogen doped SrTiO3 was optimized at initial hexamethylenetetramine content of 2.

Background

Hexavalent chromium (Cr(VI)) is a common pollutant detected in groundwater originated from excessively released of electroplating, pigment production and tanning of leather, etc. (Wang et al. 2013a). Cr(VI) has raised considerable attention because of its high toxic, intense mobility and strong teratogenic activity to human organisms. The World Health Organization (WHO) has stipulated that Cr(VI) concentration in drinking water should be below 0.05 ppm (Chen et al. 2011). Precipitation (Gheju and Balcu 2011), adsorption (Sun et al. 2014a, b), ion exchange (Edebali and Pehlivan 2010) and membrane separation (Hsu et al. 2013) as conventional techniques are commonly used to eliminate Cr(VI) from wastewater. Precipitation and adsorption processes are economic and effective, but only efficient when the Cr(VI) concentration is relatively high (Abyaneh and Fazaelipoor 2016; Hokkanen et al. 2016). Ion exchange is high-efficiency in general, but it is rather expensive to maintain and operate (Ali et al. 2015). Some even cause secondary pollution. For example, solvent extraction method could bring in organic pollutants, sulfide precipitator as common precipitator may be residual and generate hydrogen sulfide (H2S). In general, conventional techniques are either low efficiency or cost too much when they are applied to low Cr(VI) concentration in wastewater (Wang et al. 2012a; Huang and Huang 1996).

Semiconductor photocatalytic reduction technology has attracted a lot of attention in recent years (Miseki et al. 2008; Kato and Kudo 2002; Wang et al. 2009; Mu et al. 2011; Nakhjavan et al. 2012; Duo et al. 2015; Zhang et al. 2009). Semiconductor photocatalytic technology has a promising prospect for wastewater Cr(VI) removing because it is efficiency and inexpensive to maintain and operate without secondary pollution (Hu et al. 2014; Meichtry et al. 2014; Gherbi et al. 2013; Alanis et al. 2013). Strontium titanate (SrTiO3) could be applied in Cr(VI) ion contaminant reduction with excellent photocatalyst performance, but it is only effective under ultraviolet irradiation which is about 4 % of the sunlight (Zheng et al. 2011). That is to say, strontium titanate is ineffective under visible light irradiation when applied to photocatalysis because it has a 3.2 eV band gap energy (Dong et al. 2012). Many leading groups also take advantage of the high bandgap. Li Ji group and Ib Chorkendorff group use SrTiO3 and TiO2 as protective window layers for Si photocathode during water splitting, and they achieve good result (Ji et al. 2015; Bae et al. 2016). Doping with nonmetal atoms to SrTiO3 material could hoist the valence band edge and extend its optical absorption edge towards the visible light range, resulting in visible light driven photocatalytic activity (Sulaeman et al. 2011; Zou et al. 2012). The perovskite phases materials characteristics depend on the anionic composition to a large extent. Therefore, replacing oxygen with other anions, take nitrogen for example, can greatly influence the physicochemical property of the material. There are many reports about doping action including anionic dopant species and metals ions, and anionic doping could narrow the desired semiconductor band gap better than cation ions doping (Khan et al. 2002; Chen and Burda 2008).

In our work, nitrogen-doped SrTiO3 powders are synthesized by hydrothermal method reaction. We take hexamethylenetetramine as doping sources and KOH as mineralizer to obtain the fine particles with excellent photocatalytic activity. The nitrogen doping effects on SrTiO3 nanoparticles are fully studied in an attempt to investigate the microstructure, optical properties and the relevance to the improved photocatalytic activity toward chromium(VI) reduction.

Experimental section

Synthesis of nitrogen doped SrTiO3 samples

Titanium tetraisopropoxide Ti(OC3H7)4 and strontium nitrate Sr(NO3)2·4H2O were used as starting materials, hexamethylenetetramine (HMT) as nitrogen source, and KOH as mineralizer. All of them were reagent grade and used without further purification. SrTiO3 was prepared by hydrothermal method. Ti(OC3H7)4 was dissolved in 10 mL 2-propanol firstly, Sr(NO3)2 aqueous solution was added to Ti(OC3H7)4 propanol solution dropwise with continuously stirring. Then, 0–8 g of HMT and 20 mL of 2 M KOH aqueous solution were added to the suspension in turn. The solution was placed into a Teflon container with a stainless steel autoclave outside and then the solution was heated at 200 °C for 3 h in an oven. After that, the autoclave was cooled to room temperature naturally, the obtained powder was washed with distilled water and alcohol three times and dried in vacuum at 60 °C overnight (Sulaeman et al. 2010). The final samples were labeled as pure SrTiO3, N-SrTiO3(0.5), N-SrTiO3(1), N-SrTiO3(2), N-SrTiO3(3), N-SrTiO3(4), N-SrTiO3(5), N-SrTiO3(6) and N-SrTiO3(8) with the increased HMT content.

Sample characterization

X-ray power diffraction (XRD) was applied to characterize the purity and crystallinity of all our samples (D8 Advance Bruker X-ray diffractometer, CuKα radiation, 2θ = 20–80o). Transmission electron microscopy (TEM) was used to determine the morphology of the as-prepared samples (JEM-2010 apparatus, 200 kVA acceleration voltage). Diffusive reflectance UV–Vis spectrophotometer (Perkin-Elmer Lambda35) was employed to measure the samples UV–Vis absorption, BaSO4 was taken as the reference sample. Barrett–Emmett–Teller (BET) technique was taken to determine the specific surface areas (Micromeritics ASAP 2000 Surface Area and Porosity Analyzer). X-ray photo spectrometer (XPS) analysis was employed for sample element state (ESCALab220i-XL). PGSTAT302 N potentiostat galvanostat Autolab electrochemical working station using a standard three-compartment cell was used for photoelectrochemical characteristics under 300 W Xe arc lamp (≥420 nm). The fluorine-doped tin oxide (FTO) glasses (0.6 cm2) were washed for 30 min using absolute ethanol with ultrasonication. 0.1 g sample mixed with 0.01 g Polyvinylidene fluoride (PVDF) and 0.5 mL N-methyl pyrrolidinone (NMP) were placed in an glass bottle under magnetic stirring for at least 8 h. Then the obtained mixtures were coated on the FTO glasses. Photocatalyst solution was coated onto the FTO glasses substrate by drop casting using 5 μL pipette tip, and 3 drops were enough. Then we use the pipette tip to smooth the film at room temperature in the air. Lastly, the coated FTO glasses were dried for 4 h at 60 °C in the air. Photocatalyst coated FTO glass, a piece of Pt sheet, an Ag/AgCl electrode and 0.5 M sodium sulfate were used as the working electrode, counter electrode, reference electrode and electrolyte, respectively.

Photocatalytic reactivity test

Disposal ability of Cr(VI) was evaluated through photo-reduction under 300 W mercury lamp with a filter (λ ≥ 400 nm) irradiation in photochemical reactor. 50 mL 5 mg L−1 Cr(VI) solution with pH adjusted to ≈1 containing 30 mg photocatalyst was magnetically stirred during photocatalytic test. 4 mL suspension was sampled at selected intervals and centrifuged (10,000 rpm, 5 min). The content of Cr(VI) was determined using diphenylcarbazide colorimetrical method (Ma et al. 2014). The reduction efficiency of Cr(VI) was calculated according to the formula:

$$\eta = ({\text{C}}_{0} - {\text{C}}_{t} )/{\text{C}}_{0} \times 100\;\%$$

where η is the photocatalytic efficiency, C0 is the initial concentration of Cr(VI), and Ct is the concentration of Cr(VI) after illumination (for time t).

Results and discussion

Figure 1 showed the XRD patterns of the as-prepared samples. It was clearly seen that all diffraction peaks of pure SrTiO3 could be readily indexed to a well-crystallized perovskite phase of SrTiO3 (JCPDS no. 01-074-1296). There were no peaks belonging to other phases, indicating SrTiO3 was phase pure. As for the nitrogen doped SrTiO3 samples, it was observed that the diffraction peaks were identical to that of the pure SrTiO3 except for the weakened intensity and narrowed full width at half maximum (FWHM) of the corresponding peaks with the increasing initial HMT content, suggesting a particle size reduction may occurred by the nitrogen doping action. It was noted that further increase of HMT content to 3.0 g led to the appearance of additional diffraction peaks belonging to SrCO3(2θ = 25.15°, 25.79°) as shown in Fig. 1 and Additional file 1: Fig. S1. Structure refinement of the as prepared samples was carried out using general structure analysis system (GSAS) software (Larson and Von Dreele 1994). Rietveld refinement patterns of pristine SrTiO3 and N-SrTiO3 (2) samples were shown in Additional file 1: Fig. S2(a) and (b). The lattice parameters of pure SrTiO3 and N-SrTiO3(2) samples were listed in Additional file 1: Table S1. It was noted that the lattice volume of N-SrTiO3 sample expanded from 59.676 to 60.043 Å3. Several factors should be taken into consideration to clarify this phenomenon, including particle size, doping effect, surface feature, and so forth. Lattice expansion as a function of particle size reduction is not new and has been well documented in many semiconductors (Wang et al. 2013b; Mikhailovskaya et al. 2013). This can be ascribed to the surface structural relaxation due to the abundant surface exposed atoms. For the present SrTiO3 samples, the primary particle size was estimated to be ~48 and 39 nm for pure SrTiO3 and N-SrTiO3(2) samples, respectively, from the diffraction peak (110) using Scherrer formula, indicating much large particle size, which was also verified by TEM observations. Basically, large particle size implies less surface exposed atoms, which leads to negligible variation of lattice parameters. On the other hand, doping effects will not only alter the lattice parameters and electronic structure, but also have consequence on the physical properties (Wang et al. 2013b; Mikhailovskaya et al. 2013). Nitrogen tends to occupy at the oxygen site in SrTiO3 host matrix and results in lattice expansion because (1) the ionic radius of N3− is 0.146 nm, which is a bit larger than 0.138 nm of O2−, (2) Pauling’s electronegativity of nitrogen is 3.04, which is close to that of 3.44 for oxygen. This conception can be further verified by XPS observations.

Fig. 1
figure 1

XRD patterns of pure SrTiO3 and nitrogen doped SrTiO3 with different initial HMT content

Transmission electron microscopy (TEM) was employed for samples morphology and crystal structure. Figure 2a showed the TEM image of pure SrTiO3 nanoparticles. It can be seen that pure SrTiO3 exhibited irregular cubic-like morphology with an average diameter of about 200 nm, which is larger than that obtained from XRD results. High resolution TEM (HRTEM) image indicated that pure SrTiO3 exhibited high crystalline feature with the space between the adjacent of 0.2769 and 0.1931 nm, which was close to 0.276 and 0.194 nm for (110) and (200) plane of SrTiO3, respectively (Fig. 2b). As for nitrogen doped samples, the diameter of N-SrTiO3(2) showed a particle size reduction to ~100 nm with the same cubic-like morphology (Fig. 2c). HRTEM image for N-SrTiO3(2) was shown in Fig. 2d. As illustrated in Fig. 2d, the lattice plane space was determined to be 0.2790 and 0.1603 nm, which is compatible to (110) and (211) plane for SrTiO3.

Fig. 2
figure 2

TEM (a) and HRTEM (b) images of pure SrTiO3, TEM (c) and HRTEM (d) images of N-SrTiO3(2)

XPS technique was carried out to specify the chemical composition and charge state of the samples. Figure 3a shows the XPS global spectra of pure SrTiO3 and N-SrTiO3(2). From Fig. 3a, it was seen that all peaks can be ascribed to Sr, Ti, O, and C elements for pure SrTiO3 and Sr, Ti, O, N and C for nitrogen doped SrTiO3. Figure 3b showed the high resolution XPS spectra of Ti 2p for pure and nitrogen doped SrTiO3. The binding energies of Ti 2p3/2 and Ti 2p1/2 appeared at about 457.6 and 463.4 eV, respectively, which is nearly identical to the typical values of SrTiO3 (Sulaeman et al. 2011). Apparently, as shown in Fig. 3b, an about ~0.3 eV red shift was observed after the incorporation of nitrogen. This observation is related to the fact that an increase of the electron density on Ti occurred because the electronegativity of the N atom is smaller than the O atom (Wang et al. 2012b). The O 1 s peak in the XPS spectra of pure SrTiO3 and nitrogen doped SrTiO3 was shown in Fig. 3c. The peaks appeared at 529.0, 530.5 and 531.6 eV, which can be ascribed to lattice oxygen atoms, surface hydroxyl groups and surface chemisorbed oxygen, respectively (Su et al. 2015). The N-species in the N-SrTiO3 were displayed at 398.8 eV of N1 s XPS peak (Fig. 3d), which can be fitted into two peaks. The peak observed at binding energy of 399.2 eV was ascribed to sample surface adsorbed nitrogen while peak at 398.6 eV could be assigned to the shift of Ti–N bonds. The shift was very likely to be induced by the formation of Ovac and the Ti3+ cations adjacent to the vacancies (Zou et al. 2012).

Fig. 3
figure 3

XPS spectra of pure and nitrogen doped SrTiO3 samples: the whole scanning spectra (a) and the high resolution regional spectra of Ti 2p (b), O 1 s (c), and N 1 s (d)

The UV–Vis absorption performance of the samples was employed to investigate the electronic transitions and band gap energies of nitrogen doped SrTiO3. As shown in Fig. 4, absorption edge of pure SrTiO3 was about 386 nm, and the band gap energy was estimated to be 3.21 eV. Furthermore, a new absorption band at about 400–550 nm was observed in N-doped SrTiO3 samples which may be resulted from N 2p to Ti 3d orbital electron transition. Interestingly, the absorption intensity of N-doped SrTiO3 gradually increased with an increase of the initial HMT content. Moreover, the absorption edge also broadened to the higher wavelength more than 550 nm. It is presumed that the absorption above 550 nm is due to oxygen vacancy. When doping nitrogen into SrTiO3, there was possibility that O2− was replaced by N3− which would result in anion defects for the charge compensation of O2− and N3− (Zou et al. 2012). Therefore, the formation of the new band gaps in the visible light region could most possibly be ascribed to anion defects which were formed by nitrogen doping.

Fig. 4
figure 4

UV–visible diffuse reflection spectra of pure SrTiO3 and nitrogen doped SrTiO3 samples

BET measurement was adopted to detect the impacts of nitrogen doping on surface areas of SrTiO3. As shown in Fig. 5, the isotherm curves of SrTiO3 samples with different nitrogen content were characterized by a hybrid type between type I and IV in the Brunauer Deming-Dming-Teller (BDDT) classification. It is worth noting that the BET surface area of pure SrTiO3 nanoparticles is to be about 8.89 m2/g, which, however, increases to ~37.94 m2/g by nitrogen doping (inset of Fig. 5). The increase of surface area of the nitrogen doped SrTiO3 nanoparticles may predict greatly enhanced photocatalytic activities.

Fig. 5
figure 5

Nitrogen adsorption-desorption isotherms of pure SrTiO3 and nitrogen doped samples. The inset shows the BET surface area as a function of initial HMT content

Cr(VI) reduction was used as the model reaction to investigate the doping effects of nitrogen on the photocatalytic performance of SrTiO3. Cr(VI) suspensions were magnetically stirred to ensure the establishment of an adsorption/desorption equilibrium before illumination. Little Cr(VI) was adsorbed for all the as-prepared samples. Cr(VI) concentration was monitored using the standard diphenylcarbazide method with maximal adsorption at ~540 nm. The solution pH plays an important role on the reduction of Cr(VI). A series of experiments was conducted by varying solution pH from 0.5 to 4.3 to investigate pH influence. The pH influence of N-SrTiO3(2) sample was presented in Fig. 6a. It was clear that the photocatalytic efficiency increased by degrees when adjust the initial pH from 4.28 to 0.5. The lower of the pH led to the higher of the photocatalytic efficiency. On the basis of other group’s report, the relevant reactions during the reduction of Cr(VI) were as follows (Sun et al. 2014c; Mu et al. 2010):

Fig. 6
figure 6

Influence of initial pH on the Cr(VI) removal effect (a) and relationship between ln(C0/C) and the reaction time (b). Cr(VI) concentration as a function of visible light irradiation time in the presence of nitrogen doped SrTiO3 with various initial HMT content (c) and relationship between ln(C0/C) and the reaction time (d)

$$\begin{aligned} &{\text{N{-}SrTiO}}_{3} \mathop{\longrightarrow}\limits^{h\upsilon }e^{ - } + h^{ + } \hfill \\ &{\text{Cr}}_{2} {\text{O}}_{7}^{2 - } + 14{\text{H}}^{ + } + 6e^{ - } \to 2{\text{Cr}}^{3 + } + 7{\text{H}}_{ 2} {\text{O}} \hfill \\ & 2{\text{H}}_{2} {\text{O}} + 4h^{ + } \to {\text{O}}_{2} + 4{\text{H}}^{ + } \hfill \\ \end{aligned}$$

It was seen that the higher H+ concentration predicted higher reduction activity of Cr(VI) over nitrogen doped SrTiO3. Thus, the photocatalytic reduction efficiency of Cr(VI) can be promoted at a lower pH condition. Considering the harm of low pH value to the facilities in application, we adopted pH≈1 in the later Cr(VI) reduction test. It is noted that there is a linear relationship between ln(C0/C) and reaction time as seen in Fig. 6b, indicating the reduction of Cr(VI) proceeds through a pseudo-first-order kinetic reaction, ln(C0/C) = kappt, where kapp is the apparent rate constant. The optimal apparent rate constant was determined to be 1.37 h−1 at pH value of 1. As seen in Fig. 6c, the photocatalytic performance of N-SrTiO3 can be flexibly tuned by the initial HMT content. With the increase of the initial nitrogen content from 0 to 2 g, N-SrTiO3 showed increasing photocatalytic activity and N-SrTiO3(2) revealed the optimized photocatalytic activity. The reduction efficiency of 5 mg/L−1 Cr(VI) achieved to about 100 % after about 2 h under visible light irradiation. However, as for N-SrTiO3(3) sample, there was a decrease in the photocatalytic performance. Moreover, the reduction of Cr(VI) proceeds through a pseudo-first-order kinetic reaction over all nitrogen doped SrTiO3 samples as depicted in Fig. 6d. Two synergistic effects may be the reasons for the photocatalytic performance of the N-SrTiO3 samples. One was that nitrogen doping action affected the surface feature of SrTiO3 nanoparticles strongly. Doping action resulted in an increase of the specific surface area of the nitrogen doped SrTiO3 samples which provided a better access for the Cr(VI) molecules to approach the as-prepared samples and then the Cr(VI) reduction reaction can easily proceed. Another factor is related to the band gap narrowing and efficient charge separation which was resulted from both N-dopants and the Ovac. Because of the Ovac, conduction band minimum (CBM) of nitrogen doped SrTiO3 shifted more positive. At the same time, valence band maximum of nitrogen doped SrTiO3 moved more negative for the reason of N-dopants (Zou et al. 2012). The efficient charge separation in nitrogen doped SrTiO3 can be verified by Nyquist plot and photocurrent measurements. As shown in Fig. 7a, the observed arcs for both pure and nitrogen doped SrTiO3 samples were depressed which was often observed in solid electrolytes impedance spectra. The arc radius on the Nyquist plot of N-SrTiO3 electrode was almost the same with that of SrTiO3 electrode. Generally speaking, photocurrent value could give an index to the sample’s ability to generate and transfer the photogenerated charge carriers indirectly, which impact samples photocatalytic performance to a big degree (Dong et al. 2014). The photocurrent plots of the SrTiO3 and N-SrTiO3 electrodes under visible light was shown in Fig. 7b which was obtained at a bias voltage of 0.5 V. SrTiO3 sample showed a photocurrent of 8.85 × 10−8 A while an enhanced photocurrent of 3.98 × 10−7 A for N-SrTiO3 sample was observed at the same condition. The photocurrent for N-SrTiO3 was nearly 4.5 times as much as the SrTiO3 sample which indicated that the faster charge transfer and more effective separation of electron-hole pairs through the N-SrTiO3 electrode interface, which can drastically enhance the photocatalytic activity toward Cr(VI) reduction.

Fig. 7
figure 7

Electrochemical impedance spectroscopy (EIS) plots (a) and photocurrent responses of the SrTiO3 and N-SrTiO3(2) electrodes (b)

To check the stability of the N-SrTiO3(2) photocatalyst, recycle test was carried out. The XRD pattern and UV–visible diffuse reflection spectra of N-SrTiO3(2) after the reduction of Cr(VI) was measured, and was shown in Fig. 8b, c. The XRD patterns peak of the N-SrTiO3(2) sample did not change much before and after the Cr(VI) reduction test, suggesting that our photocatalyst was stable in structure. But as seen in Fig. 8c, visible light absorption decreased much after 5rd recycle test which may result from nitrogen loss in the nitrogen doped SrTiO3, being often observed in several nitrogen doped oxides (Kumar et al. 2014; Ruzimuradov et al. 2015). This may also be the reason for decrease of photocatalytic performance in Fig. 8a as 55 % of the photocatalytic activity was preserved after 5 cycles. In summary, nitrogen doped SrTiO3 possess excellent photocatalytic performance towards Cr(VI) reduction under visible light, but the stability of the nitrogen doped SrTiO3 sample still needs to be further improved.

Fig. 8
figure 8

Cycling performance of N-SrTiO3(2) toward Cr(VI) reduction (a), XRD patterns of N-SrTiO3(2) after 5 cycles recycle test toward Cr(VI) reduction (b), UV–visible diffuse reflection spectra of of N-SrTiO3(2) after 5 cycles recycle test toward Cr(VI) reduction (c)

Conclusions

In summary, nitrogen doped SrTiO3 nanoparticles with controlled particle size, electronic structure and efficient visible light driven photocatalytic activity toward Cr(VI) were successfully prepared by a solvothermal method. XRD, BET and TEM analyses indicated that nitrogen doped SrTiO3 nanoparticles with cube-like morphology exhibited an apparent lattice expansion, particle size reduction as well as subsequent increase of BET surface area via nitrogen doping. The visible light absorption edge and intensity can be modulated by nitrogen doping content, which absorption edge extends to about 600 nm. Moreover, nitrogen doping can not only modulate the visible light absorption feature, but also have consequence on the enhancement of charge separation efficiency, which can promote the photocatalytic activity. With well controlled particle size, BET surface area, and electronic structure via nitrogen doping, the visible light driven photocatalytic performance toward Cr(VI) reduction of nitrogen doped SrTiO3 was optimized at initial HMT content of 2. Such a finding may help to provide hints for developing and designing new photocatalytic semiconductors.

References

  • Abyaneh AS, Fazaelipoor MH (2016) Evaluation of rhamnolipid (RL) as a biosurfactant for the removal of chromium from aqueous solutions by precipitate flotation. J Environ Manage 165:184–187

    Article  Google Scholar 

  • Alanis C, Natividad R, Barrera-Diaz C, Martínez-Miranda V, Prince J, Valente JS (2013) Photocatalytically enhanced Cr(VI) removal by mixed oxides derived from MeAl (Me: Mg and/or Zn) layered double hydroxides. Appl Catal B Environ 140–141:546–551

    Article  Google Scholar 

  • Ali SW, Mirza ML, Bhatti TM (2015) Removal of Cr(VI) using iron nanoparticles supported on porous cation-exchange resin. Hydrometallurgy 157:82–89

    Article  Google Scholar 

  • Bae D, Shayestehaminzadeh S, Thorsteinsson EB, Pedersen T, Hansen O, Seger B, Vesborg PCK, Ólafsson S, Chorkendorff I (2016) Protection of Si photocathode using TiO2 deposited by high power impulse magnetron sputtering for H2 evolution in alkaline media. Sol Energy Mater Sol Cells 144:758–765

    Article  Google Scholar 

  • Chen XB, Burda C (2008) The electronic origin of the visible-light absorption properties of C-, N-and S-Doped TiO2 nanomaterials. J Am Chem Soc 130:5018–5019

    Article  Google Scholar 

  • Chen H, Shao Y, Xu ZY, Wan HQ, Wan YQ, Zheng SR, Zhu DQ (2011) Effective catalytic reduction of Cr(VI) over TiO2 nanotube supported Pd catalysts. Appl Catal B Envrion 105:255–262

    Article  Google Scholar 

  • Dong WJ, Li XY, Yu J, Guo WC, Li BJ, Tan L, Li CR, Shi JJ, Wang G (2012) Porous SrTiO3 spheres with enhanced photocatalytic performance. Mater Lett 67:131–134

    Article  Google Scholar 

  • Dong HJ, Sun JX, Chen G, Li CM, Hua YD, Lv CD (2014) An advanced Ag-based photocatalyst Ag2Ta4O11 with outstanding activity, durability and universality for removing organic dyes. Phys Chem Chem Phys 16:23915–23921

    Article  Google Scholar 

  • Duo F, Wang Y, Mao X, Zhang X, Wang Y, Fan C (2015) A BiPO4/BiOCl heterojunction photocatalyst with enhancedelectron-hole separation and excellent photocatalytic performance. Appl Surf Sci 340:35–42

    Article  Google Scholar 

  • Edebali S, Pehlivan E (2010) Evaluation of Amberlite IRA96 and Dowex 1 × 8 ion-exchange resins for the removal of Cr(VI) from aqueous solution. Chem Eng J 161:161–166

    Article  Google Scholar 

  • Gheju M, Balcu I (2011) Removal of chromium from Cr(VI) polluted wastewaters by reduction with scrap iron and subsequent precipitation of resulted cations. J Hazard Mater 196:131–138

    Article  Google Scholar 

  • Gherbi R, Trari M, Nasrallah N (2013) Influence of light flux and hydrodynamic flow regime on the photoreduction of Cr(VI) on the CuAl2O4/TiO2 hetero-junction. Chem Eng 1:1275–1282

    Google Scholar 

  • Hokkanen S, Bhatnagar A, Repo E, Lou S, Sillanpää M (2016) Calcium hydroxyapatite microfibrillated cellulose composite as a potential adsorbent for the removal of Cr(VI) from aqueous solution. Chem Eng J 283:445–452

    Article  Google Scholar 

  • Hsu HT, Chen SS, Tang YF, Hsi HC (2013) Enhanced photocatalytic activity of chromium(VI) reduction and EDTA oxidization by photoelectrocatalysis combining cationic exchange membrane processes. J Hazard Mater 248–249:97–106

    Article  Google Scholar 

  • Hu XF, Ji HH, Chang F, Luo YM (2014) Simultaneous photocatalytic Cr(VI) reduction and 2,4,6-TCP oxidation over g-C3N4 under visible light irradiation. Catal Today 224:34–40

    Article  Google Scholar 

  • Huang C, Huang CP (1996) Application of aspergillus oryzae and rhzopus oryzae for Cu(II) removal. Water Res 30:1985–1990

    Article  Google Scholar 

  • Ji L, McDaniel MD, Wang SJ, Posadas AB, Li XH, Huang HY, Lee JC, Demkov AA, Bard AJ, Ekerdt JG, Yu ET (2015) A silicon-based photocathode for water reduction with an epitaxial SrTiO3 protection layer and a nanostructured catalyst. Nat Nanotechnol 10:84–90

    Article  Google Scholar 

  • Kato H, Kudo A (2002) Visible-light-response and photocatalytic activities of TiO2 and SrTiO3 photocatalysts codoped with antimony and chromium. J Phys Chem B 106:5029–5034

    Article  Google Scholar 

  • Khan SUM, Al-Shahry M, Ingler WB Jr (2002) Efficient photochemical water splitting by a chemically modified n-TiO2. Science 297:2243–2246

    Article  Google Scholar 

  • Kumar S, Tonda S, Baruah A, Kumar B, Shanker V (2014) Synthesis of novel and stable g-C3N4/N-doped SrTiO3 hybrid nanocomposites with improved photocurrent and photocatalytic activity under visible light irradiation. Dalton Trans 43:16105–16114

    Article  Google Scholar 

  • Larson AC, Von Dreele RB (1994) General structure analysis system (GSAS). Los Alamos National Laboratory Report LAUR, pp 86–748

  • Ma YL, Liu XQ, Li Y, Su YG, Chai ZL, Wang XJ (2014) K4Nb6O17·4.5H2O: a novel dual functional material with quick photoreduction of Cr(VI) and high adsorptive capacity of Cr(III). J Hazard Mater 279:537–545

    Article  Google Scholar 

  • Meichtry JM, Colbeau-Justin C, Custo G, Litter MI (2014) Preservation of the photocatalytic activity of TiO2 by EDTA in thereductive transformation of Cr(VI). Studies by time resolved microwave conductivity. Catal Today 224:236–243

    Article  Google Scholar 

  • Mikhailovskaya ZA, Buyanova ES, Petrova SA, Morozova MV, Zhukovskiy VM, Zakharov RG, Tarakina NV, Berger IF (2013) Cobalt-doped Bi26Mo10O69: crystal structure and conductivity. J Solid State Chem 204:9–15

    Article  Google Scholar 

  • Miseki Y, Kato H, Kudo A (2008) Water splitting into H2 and O2 over niobate and titanate photocatalysts with (111) plane-type layered perovskite structure. Energy Environ Sci 2:306–314

    Article  Google Scholar 

  • Mu RX, Xu ZY, Li LY, Shao Y, Wan HQ, Zheng SR (2010) On the photocatalytic properties of elongated TiO2 nanoparticles for phenol degradation and Cr(VI) reduction. J Hazard Mater 176:495–502

    Article  Google Scholar 

  • Mu S, Long Y, Kang SZ, Mu J (2011) Surface modification of TiO2 nanoparticles with a C60 derivative and enhanced photocatalytic activity for the reduction of aqueous Cr(VI) ions. Catal Commun 11:741–744

    Article  Google Scholar 

  • Nakhjavan B, Tahir MN, Natalio F, Panthofer M, Gao H (2012) Ni@Fe2O3 heterodimers: controlled synthesis and magnetically recyclable catalytic application for dehalogenation reactions. Nanoscale 4:4571–4577

    Article  Google Scholar 

  • Ruzimuradov O, Sharipov K, Yarbekov A, Saidov K, Hojamberdiev M, Prasad RM, Cherkashinin G, Riedel R (2015) A facile preparation of dual-phase nitrogen-doped TiO2-SrTiO3 macroporous monolithic photocatalyst for organic dye photo degradation under visible light. J Eur Ceram Soc 35(6):1815–1821

    Article  Google Scholar 

  • Su YG, Huang SS, Wang TT, Peng LM, Wang XJ (2015) Defect-mediated efficient catalytic activity toward p-nitrophenol reduction: a case study of nitrogen doped calcium niobate system. J Hazard Mater 295:119–126

    Article  Google Scholar 

  • Sulaeman U, Yin S, Sato T (2010) Solvothermal synthesis and photocatalytic properties of nitrogen-doped SrTiO3 Nanoparticles. J Nanomater

  • Sulaeman U, Yin S, Sato T (2011) Visible light photocatalytic activity induced by the carboxyl group chemically bonded on the surface of SrTiO3. Appl Catal B Environ 102:286–290

    Article  Google Scholar 

  • Sun XT, Yang LR, Li Q, Zhao JM, Li XP, Wang XQ, Liu HZ (2014a) Amino-functionalized magnetic cellulose nanocomposite as adsorbent for removal of Cr(VI): synthesis and adsorption studies. Chem Eng J 241:175–183

    Article  Google Scholar 

  • Sun YY, Yue QY, Mao YP, Gao BY, Gao Y, Huang LH (2014b) Enhanced adsorption of chromium onto activated carbon by microwave-assisted H3PO4 mixed with Fe/Al/Mn activation. J Hazard Mater 265:191–200

    Article  Google Scholar 

  • Sun Q, Li H, Zheng SL, Sun ZM (2014c) Characterizations of nano-TiO2/diatomite composites and theirphotocatalytic reduction of aqueous Cr(VI). Appl Surf Sci 311:369–376

    Article  Google Scholar 

  • Wang DF, Kako T, Ye JH (2009) New Series of Solid-Solution Semiconductors (AgNbO3)1-x(SrTiO3)x with Modulated Band Structure and Enhanced Visible-Light Photocatalytic Activity. J Phys Chem C 113:3785–3792

    Article  Google Scholar 

  • Wang Q, Guan YP, Ren XF, Yang MZ, Liu X (2012a) Removal of low concentration Cr(VI) from aqueous solution by magnetic-fluids fixed bed using the high gradient magnetic separation. J Colloid Interf Sci 374:325–330

    Article  Google Scholar 

  • Wang DH, Jia L, Wu XL, Lu LQ, Xu AW (2012b) One-step hydrothermal synthesis of N-doped TiO2/C nanocomposites with high visible light photocatalytic activity. Nanoscale 4:576–584

    Article  Google Scholar 

  • Wang L, Li XY, Teng W, Zhao QD, Shi Y, Yue RL, Chen YF (2013a) Efficient photocatalytic reduction of aqueous Cr(VI) over flower-like SnIn4S8 microspheres under visible light illumination. J Hazard Mater 244–245:681–688

    Article  Google Scholar 

  • Wang M, Liu Q, Che YS, Zhang LF, Zhang D (2013b) Characterization and photocatalytic properties of N-doped BiVO4 synthesized via a sol-gel method. J Alloys Compd 548:70–76

    Article  Google Scholar 

  • Zhang L, Wong KH, Chen Z, Yu JC, Zhao J, Hu C, Chan CY, Wong PK (2009) AgBr-Ag-Bi2WO6 nanojunction system: a novel and efficient photocatalyst with double visible-light active components. Appl Catal A Gen 363:221–229

    Article  Google Scholar 

  • Zheng ZK, Huang BB, Qin XY, Zhang XY, Dai Y (2011) Facile synthesis of SrTiO3 hollow microspheres built as assembly of nanocubes and their associated photocatalytic activity. J Colloid Interf Sci 358:68–72

    Article  Google Scholar 

  • Zou F, Jiang Z, Qin XQ, Zhao YX, Jiang LY, Zhi JF, Xiao TC, Edwards PP (2012) Template-free synthesis of mesoporous N-doped SrTiO3 perovskite with high visible-light-driven photocatalytic activity. Chem Commun 48:8514–8516

    Article  Google Scholar 

Download references

Authors’ contributions

The manuscript was conceived and designed by SYG and XGJ. ZLX and ST performed acquisition of data. WXJ made some revisions of the manuscript. All authors read and approved the final manuscript.

Acknowledgements

This work is financially supported by the National Natural Science Foundation of China (Grants 21267041, 21367018, 21563021), the Project of Scientific and Technological Innovation Team of Inner Mongolia University (12110614).

Competing interests

The authors declare that they have no competing interests.

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Yiguo Su or Xiaojing Wang.

Additional file

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Xing, G., Zhao, L., Sun, T. et al. Hydrothermal derived nitrogen doped SrTiO3 for efficient visible light driven photocatalytic reduction of chromium(VI). SpringerPlus 5, 1132 (2016). https://doi.org/10.1186/s40064-016-2804-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40064-016-2804-2

Keywords