Ayyakannu Sundaram Ganeshrajaab,
Subramani Thirumurugan‡
b,
Kanniah Rajkumar‡b,
Kaixin Zhuac,
Yanjie Wangac,
Krishnamoorthy Anbalagan*b and
Junhu Wang*a
aMössbauer Effect Data Center, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, Dalian 116023, China. E-mail: wangjh@dicp.ac.cn; Tel: +86 411 84379159
bDepartment of Chemistry, Pondicherry University, Pondicherry 605014, India. E-mail: kanuniv@gmail.com; Tel: +91 413 2654509
cUniversity of Chinese Academy of Sciences, Beijing 100049, China
First published on 16th December 2015
The structural, electronic, magnetic and photocatalytic properties of Sn doped TiO2 diluted magnetic semiconductor nanoparticles (NPs) prepared by a simple hydrothermal method were systematically investigated by various conventional techniques and 119Sn Mössbauer spectroscopy. Anatase, mixed (anatase–rutile) and rutile phases of Sn–TiO2 NPs were obtained by adding different amounts of SnCl4 into a titanium nitrate aqueous solution. Photocatalytic degradation of methyl orange and phenol derivatives (RPhOH) were studied under visible and UV light irradiation in water, respectively. The photocatalytic activities of prepared NPs were found to be drastically related to the structural, optical and ferromagnetic properties. A significant relationship was observed between the Hammett substitution constants of RPhOH and the photocatalytic activity. Among all the samples, the anatase phase with low Sn content performed with the best photocatalytic and ferromagnetic characteristics at room temperature.
Meanwhile, ferromagnetism in diluted magnetic semiconductors (DMS) has been another important subject of great scientific and technological interest for the past few years due to the possibility of manipulating charge and spin degrees of freedom.11 In particular, room-temperature ferromagnetic (RTFM) oxides have attracted great attention from magnetic fluids, biomedicine, magnetic resonance imaging, catalysis, and environmental remediation.12 Wang et al.13 developed a simple method to fabricate ZnO crystals with abundant Zn vacancies, and observed p-type conductivity, RTFM, and excellent photocatalytic activity. Anbalagan14 and Ganeshraja15 reported the ferromagnetic ordering in transition metal such as Co/Fe/Zn doped TiO2 NPs could be justified by the creation of some defect sites in the samples. The Sn–TiO2 NPs were also found to have RTFM property; however, the actual role of p-block element, in the present investigation, tin is still unclear.16
Very recently, we reported the ferromagnetism and photocatalytic activity in metal oxide coupled nanocomposites.17 The photocatalytic activity and magnetism studies of prepared metal oxide coupled nanocomposites might suggest a major role of surface oxygen vacancies and charge carriers.18 The RTFM nanocomposites were found to show better photocatalytic performance instead of diamagnetic commercially available photocatalysts under visible light irradiation.17,19 In order to take high activity of the coupled,19 doped or co-doped20 semiconductor nanocomposites, the concept of magnetic photocatalysts shows better charge carrier separation function. Hence, development of photocatalyst possessing ferromagnetic property in tune with visible-light activity has become an important topic in the photocatalysis research today.
In the present work, focus on the structural, electronic, magnetic and photocatalytic properties of Sn–TiO2 NPs with different Sn doping levels were studied. Using a simple hydrothermal method and various amounts of SnCl4 added into a titanium nitrate aqueous solution, anatase, mixture (anatase–rutile) and rutile phases of Sn–TiO2 NPs were obtained. In addition, methyl orange (MO) and RPhOH (where PhOH is phenol and R = 3-NH2, H and 4-Cl) were chosen as model pollutants to evaluate the photocatalytic activity of Sn–TiO2 samples under visible and UV light irradiation in water, respectively. Significant relation was observed between the Hammett substitution constant (σ) of RPhOH and the photocatalytic degradation efficiency (PDE) of Sn–TiO2 NPs. It was revealed that structural, electronic, magnetic and photocatalytic properties are closely related to the Sn doping level.
In addition, we emphasize a combined study of photocatalytic and ferromagnetic characters at room temperature on Sn–TiO2 samples, which is the most emerging field in environmental remediation and in the development of magnetic material. Even after a decade-long research, the actual mechanism of ferromagnetism combined with photocatalytic behavior in these materials is still not understood, although hints about some of the key factors that contribute to the magnetism have been pointed out.21 It is believed that oxygen vacancy, phase change and doping level play a major role in the RTFM semiconductor oxides.22 However, demonstration of a direct correlation between the magnetism, dopant concentration, oxygen vacancy and photocatalytic activity has been difficult.23 Because of these reasons, in this work, we made an effort to investigate the true role of tin ions on the above properties of prepared Sn–TiO2 NPs.
![]() | ||
Fig. 1 XRD patterns (a) 10–80°, (b) 23–29° and (c) 31–37° of Sn–Ti-x samples, where anatase (A), rutile (R) and casseterite (C). |
Fig. 1b and c clearly show that the peak intensities of anatase Sn–Ti-x samples decreased and rutile peak intensities increased with the increase of tin content. The rutile (110) reflection appears at ∼27°, which is intermediate to that of the pure TiO2 (27.4°, 2θ) and pure SnO2 (26.6°, 2θ) positions.25 It can also be ascribed that some of the lattice constants in rutile, SnO2 and TiO2 were observed, in addition, lattice distortion was induced by the substitution of Sn for Ti.26 Average particle sizes were estimated from the XRD patterns using the Scherrer equation from 25 and 27° reflections (Table 1).27 In the case of anatase phases, the crystalline size of Sn–Ti-x (x = 0, 0.05 and 0.12) samples are bigger than that of rutile Sn–Ti-1.37 samples. According to the previous report, a critical particle size should be required for the crystalline phase transformation from anatase to rutile, thus the rutile size should be smaller than anatase one in the coexistence of anatase and rutile phases of TiO2.28,29
Samples | D (nm) | Eg (eV) | Life time (ns) | Amplitude (×10−3) | χ2 | ||||
---|---|---|---|---|---|---|---|---|---|
τ1 | τ2 | τ3 | a1 | a2 | a3 | ||||
Sn–Ti-0 | 27.91 | 3.30 | 1.48 | 10.07 | 0.22 | 22.05 | 1.22 | 1259.21 | 1.16 |
Sn–Ti-0.05 | 14.48 | 3.18 | 2.80 | 38.29 | 0.17 | 12.74 | 9.70 | 791.41 | 1.28 |
Sn–Ti-0.12 | 9.54 | 3.19 | 2.88 | 37.37 | 0.34 | 17.47 | 13.32 | 354.40 | 1.11 |
Sn–Ti-0.17 | 10.60 | 3.20 | 2.59 | 37.74 | 0.09 | 7.02 | 5.54 | 2056.93 | 1.21 |
Sn–Ti-0.30 | 8.21 | 3.20 | 2.71 | 37.84 | 0.02 | 11.17 | 9.44 | 924.56 | 1.32 |
Sn–Ti-1.14 | 6.66 | 3.21 | 2.39 | 33.91 | 0.15 | 10.58 | 8.26 | 983.07 | 1.12 |
Sn–Ti-1.37 | 6.57 | 3.24 | 2.53 | 34.82 | 0.11 | 6.99 | 5.41 | 1630.39 | 1.71 |
The crystalline phase analysis of the Sn–Ti-x samples obtained in this work was further investigated by Raman spectroscopy as shown in Fig. 2. Predominant anatase peaks were observed at 144, 399, 521 and 638 cm−1 in Sn–Ti-x (x = 0, 0.05 and 0.12) samples. The Raman peak intensities of anatase Sn–Ti-x samples decreased with the increase of tin content. With further increase in Sn content, four strong peaks were observed at 149, 402, 522 and 638 cm−1, which are similar to that of anatase TiO2,30 moreover, one additional weak rutile peak is observed at 433 cm−1 in Sn–Ti-0.17 and Sn–Ti-0.30. In the high Sn content samples, three strong peaks were observed at 262, 436, and 618 cm−1, which are quite different to those of pure rutile TiO2 (235, 446 and 611 cm−1) and can be explained due to SnO2 peaks in coexistence with rutile TiO2 phase. Additional formation of SnO2 crystalline phase was observed at ∼776 and ∼448 cm−1, which indicates that some amount of tin exists as tin oxide crystalline phase coupled with rutile TiO2 in Sn–Ti-1.14 and Sn–Ti-1.37 samples.31 Therefore, the results of Raman spectra of the samples confirm the results of XRD patterns.
Fig. 3 presents the 119Sn Mössbauer spectra of the Sn–Ti-x samples at room temperature. A summary of the obtained 119Sn Mössbauer parameters are given in Table 2. The isomer shift (IS) values of all components indicated that tin atoms in all prepared samples were tetravalent (Sn4+). The absence of absorption peaks in the Doppler velocity range of 2–4 mm s−1 indicates Sn2+ ion free in the Sn–Ti-x samples.32 The difference in the quadrupole splitting (QS) values of the component peaks implied that Sn atoms existed under different asymmetry of coordination polyhedron around Sn atoms in the TiO2 lattice. The spectra of low level Sn doped samples of Sn–Ti-x (x = 0.05, 0.12 and 0.17) were fitted by one doublet, respectively. The obtained IS values are very close to that of typical Sn4+ bonded to oxygen in octahedral environments.33 More precisely, IS are found to be positive (IS = 0.10–0.14 mm s−1 and QS = 1.02–1.30 mm s−1), which are consistent with the values reported for the Sn4+ dopant in transition metal oxides.34,35 This fact thus suggests the occurrence of Ti–O–Sn–O–Ti chain fragments in the lattice and consequently insertion of dopant into the matrix.36 The spectra of high level Sn doped samples of Sn–Ti-x (x = 0.30, 1.14 and 1.37) were decomposed into two doublets, both of these with unusually high QS were assigned to the defects playing very important role in the realization of ferromagnetism in these materials. The parameters of doublet D1 correspond well to those characteristics of Sn doped TiO2 with lattice crystal defects (IS = 0.11–0.14 mm s−1 and QS = 1.34–1.76 mm s−1), while doublet D2 (IS = 0.09–0.16 mm s−1 and QS = 4.28–5.86 mm s−1) can be associated with a much distorted Sn microenvironment due to a neighboring crystal defect like a vacancy.37 The spectra show that defects occur in a very high amount in these simple hydrothermally prepared tin doped titania samples. The doublets having very large QS are associated with defects, vacancies, stabilized by doping of Sn4+ in TiO2 system. Considering the unusual large QS and line width (LW) values, it is also possible to result in magnetic interaction. From XRD results small amount of the SnO2 coupled with TiO2, this may also possible to exist large QS and LW values. Anyway, the 119Sn Mössbauer spectral results clearly showed the existence of structural defects, which have an essential role in the appearance of ferromagnetism.
Samples | MB | Relative area (%) | VSM | PDE (%) | ||||
---|---|---|---|---|---|---|---|---|
IS (mm s−1) | QS (mm s−1) | LW (mm s−1) | Ms (memu g−1) | Hc (G) | Mr (memu g−1) | |||
a PDE were calculated from photodegradation of phenol in water with Sn–Ti-x samples under ultraviolet light irradiation (368 nm) at 240 min. | ||||||||
Sn–Ti-0 | — | — | — | — | 10.30 | 109.33 | 1.08 | 42.59 |
Sn–Ti-0.05 | 0.14 | 1.02 | 2.53 | 100 | 16.66 | 105.33 | 1.30 | 70.89 |
Sn–Ti-0.12 | 0.14 | 1.30 | 2.53 | 100 | 129.56 | 57.12 | 6.23 | 68.45 |
Sn–Ti-0.17 | 0.10 | 1.13 | 2.54 | 100 | 0.16 | 277.13 | 0.02 | 57.16 |
Sn–Ti-0.30 | 0.13 | 1.34 | 2.53 | 88 | 0.12 | 296.71 | 0.02 | 7.17 |
0.09 | 4.28 | 2.53 | 12 | — | — | — | — | |
Sn–Ti-1.14 | 0.14 | 1.65 | 2.53 | 73 | 0.13 | 821.14 | 0.04 | 25.73 |
0.12 | 5.64 | 2.53 | 27 | — | — | — | — | |
Sn–Ti-1.37 | 0.11 | 1.76 | 2.53 | 74 | 0.10 | 318.63 | 0.02 | 23.39 |
0.16 | 5.86 | 2.53 | 26 | — | — | — | — |
XPS analysis was performed to further study the chemical state of tin in TiO2 matrix. Fig. 4 shows the survey, Ti 2p, Sn 3d and O 1s XPS spectra of a typical Sn–Ti-0.17 sample. It can be seen that XPS peak positions of Ti 2p3/2 and Ti 2p1/2 located at 458.4 and 464.3 eV, indicate that Ti element mainly exist in the chemical state of Ti4+. The peak was observed with the B.E. of 529.6 eV corresponds to oxygen in TiO2 lattice (OL). The doublet peaks observed at 486.1 and 494.6 eV in the Sn 3d XPS spectrum are ascribed to Sn 3d5/2 and Sn 3d3/2 of the substituted Sn4+ ion as dopants in the TiO2 lattice, since the peak position of Sn 3d5/2 (486.1 eV) was located between that of SnO2 (486.6 eV) and metallic Sn (484.4 eV).38 The XPS analysis infers that the Sn4+ ion can conveniently replace Ti4+ ion in the lattice and enter the lattice of TiO2 as explored by 119Sn Mössbauer spectroscopy.
SEM analysis was employed to understand the surface morphology of Sn–Ti-x crystals. Fig. S1† depicts SEM images, which illustrate that the particles mainly belong to anatase and anatase–rutile mixed phases, loosely agglomerated, spherical with nanosized particles, and significantly good in crystalline quality. The compositions of the Sn–Ti-x samples were analyzed by EDX as shown in Fig. S2–S5.† The atomic percentages of Ti, Sn and O elements obtained by EDX analysis are presented in Table 3. The distribution of Ti, O and Sn elements in a typical Sn–Ti-0.05 sample could be seen on the elemental mapping and line spectrum, where the relative location of each constituent can be spotted in different colors as shown in Fig. 5. Although this percentage of Sn/Ti has approached upto the detection limit of EDX and some background noise have kicked in, we can find that Sn atoms have been incorporated over the entire TiO2 NPs obviously as in Fig. 5d. The EDX elemental maps and line scan spectra confirmed the uniform distribution of O, Ti, and Sn species in lattice of nanosized Sn–Ti-x samples.
Samples | EDX, atomic (%) | BET | ||||
---|---|---|---|---|---|---|
Ti | Sn | O | Surface area (m2 g−1) | Langmuir surface area (m2 g−1) | Pore size (nm) | |
Sn–Ti-0 | 33.33 | — | 66.67 | 45.80 | 73.48 | 25.49 |
Sn–Ti-0.05 | 31.49 | 1.59 | 66.93 | 75.42 | 688.68 | 8.48 |
Sn–Ti-0.12 | 28.49 | 3.44 | 68.07 | 107.43 | 593.05 | 5.34 |
Sn–Ti-0.17 | 32.09 | 5.49 | 62.42 | 70.98 | 648.47 | 7.99 |
Sn–Ti-0.30 | 40.04 | 12.18 | 47.78 | 70.92 | 112.65 | 13.21 |
Sn–Ti-1.14 | 15.49 | 17.71 | 66.80 | 60.82 | 368.69 | 5.86 |
Sn–Ti-1.37 | 14.15 | 19.33 | 66.52 | 65.89 | 105.21 | 6.17 |
Fig. S6† shows the N2 adsorption–desorption isotherms and Barret–Joyner–Halenda (BJH) pore size distribution curves of Sn–Ti-x samples with a significant hysteresis loop observed in the relative pressure (p/p0) range of 0.95–0.99. This property implies the presence of mesopores. This result is further confirmed by the pore size distribution curve which indicates the predominance of pores 5.34–13.21 nm in diameter. The BET and Langmuir surface area results are presented in Table 3, respectively.
Fig. 6 shows the TEM images of typical Sn–Ti-0.05, Sn–Ti-0.30 and Sn–Ti-1.37 samples. NPs with a diameter of 5–10 nm were observed in all as prepared samples. The electron diffraction patterns of the selected area on anatase Sn–Ti-0.05 sample (inset Fig. 6a) showed the strong Debye–Scherrer rings and additionally complicated bright spots were observed for Sn–Ti-0.30 and Sn–Ti-1.37 samples (inset Fig. 6b and c), indicating the coexistence of anatase–rutile or rutile–casseterite crystalline phases.39
![]() | ||
Fig. 6 TEM images and selected area electron diffraction patterns of (a) Sn–Ti-0.05, (b) Sn–Ti-0.30 and (c) Sn–Ti-1.37 samples. |
Fig. 7 displays the HRTEM images of Sn–Ti-0.05, Sn–Ti-0.30 and Sn–Ti-1.37 samples. For the pure anatase structure, the fringe spacing (d) of (101) crystallographic plane was determined to be 3.52 Å,9 while Sn4+ doped TiO2 samples, it was determined to fall in the range 3.52 to 3.55 Å. This implies that Sn4+ ions were doped in the TiO2 lattice in substitutional mode, since the ionic radius of Sn4+ (0.69 Å) is larger than that of the lattice Ti4+ (0.53 Å).40 Furthermore, a fringe spacing of ∼3.22 Å corresponds to the (110) plane of rutile TiO2 phase was observed for Sn–Ti-0.30 crystals. The crystallographic planes of SnO2 cassiterite (101) fringe spacing observed at 2.64 Å, further confirmed that SnO2 was coupled with TiO2 in Sn–Ti-1.37 sample. The structural, chemical and morphological characterizations performed using XRD, XPS, Raman, 119Sn Mössbauer spectroscopy, SEM-EDX and HRTEM clearly explored the octahedral substitution of Sn4+ for Ti4+ in the Ti–Sn-x (x = 0.5, 0.12 and 0.17) crystal lattice, in addition SnO2 was coupled with TiO2 in remaining Ti–Sn-x (x = 0.30, 1.14 and 1.37) samples.
Fig. 8a shows the repetitive scan spectra of photodegradation of MO with the Sn–Ti-x samples suspended in water. It is observed that a gradual decrease in the absorption bands at 268 and 464 nm of MO with the increase of the irradiation time.41 In Fig. 8b, the PDE of MO is plotted as a function of reaction time which indicates that the PDE of Sn–Ti-x samples can be improved by doping an appropriate density of Sn4+ dopant. The Sn–Ti-x samples with x = 0.05 and 0.12 exhibited higher photocatalytic activity. However, if x increases steadily, the photocatalytic activity begins to fall down appropriately. This behavior is due to the phase transformation of the crystals as the dopant content is enhanced. In general, the photocatalytic performance of anatase is considered superior to that of the more stable rutile phase of titania. Some of the earlier reports42,43 attributed that anatase is very effective in photocatalytic activity due to higher density of localized states, the existence of surface-adsorbed hydroxyl species and slower charge carrier recombination in the crystals.
![]() | ||
Fig. 8 (a) Repetitive scan spectra of photodegradation of MO in water over Sn–Ti-0.05 sample and (b) PDE of SnO2, TiO2 (rutile) and Sn–Ti-x samples vs. time under visible light irradiation. |
The UV-light-induced photocatalysis was confirmed by the degradation of RPhOH as shown in Fig. 9. In Fig. 10a, the PDE of phenol is plotted as a function of reaction time. There was low observable degradation of phenol over the undoped anatase, rutile TiO2 and SnO2 samples. However, in the presence of Sn–Ti-x (x = 0.05, 0.12 and 0.17) a rapid degradation of phenol occurred by UV light irradiation. Phenol was removed upto ∼70% on 240 min irradiation using Sn–Ti-x catalyst. It can be concluded that the tin doping into the anatase phase of TiO2 is efficient for the photocatalytic degradation of the sample under study.
![]() | ||
Fig. 9 Repetitive scan spectra of photodegradation of RPhOH in water over Sn–Ti-0.05 sample under UV light irradiation (368 nm). |
It is known that the properties of phenolic compounds can be drastically affected by the electronic nature of R group substituents. In the present work, substituent effect of R group on the photocatalytic activity over Sn–Ti-0.05 sample was studied using RPhOH (R = H, 3-NH2 and 4-Cl). The Hammett's substitution constant (σ) was used to establish a relation between the PDE with the substituted phenols as shown in Fig. 10b. These were defined from the ionization constants of the appropriately substituted benzene derivatives such as benzoic acid, phenol, aniline etc., as σ(X) = logKX − log
KH, where KH is the ionization constant for benzene derivative in water at 25 °C and KX is the corresponding constant for a meta- (σm) or a para substituted (σp) benzene derivatives. The PDE was found to be dependent on the electron donating/withdrawing effect of the substituent R group on the aromatic ring of RPhOH. The results indicate that the highest PDE was observed for 3-NH2PhOH (σ = −0.16) and lowest for 4-ClPhOH (σ = 0.23) due to electrophilic nature of organic substrates of phenolic compounds.
Information regarding the presence of surface states, formation of photoinduced charge carriers, and their recombination kinetics can be drawn from the photoluminescence (PL) spectra of semiconductor materials.47 PL spectra of Sn–Ti-x samples with excitations at 330 nm are presented in Fig. 11. Broad emission in the spectral range from 350 to 600 nm was observed as well as well-resolved peaks/shoulders at 410, 417, 438, 449, 466, 480, 490 and 560 nm corresponding energy of 3.03, 2.98, 2.83, 2.76, 2.66, 2.59, 2.53 and 2.22 eV, respectively were seen. A change in Sn4+ dopant density in TiO2 altered the intensity and or change the shape and peak position of the PL spectra compared to spectrum of predominantly anatase Sn–Ti-x (0 and 0.05) samples. It was found that the steady state emission spectrum contains a narrower UV emission located near 390 nm (3.18 eV) and a broader emission range from 450 to 491 nm.48 A strong green emission was observed at 560 nm (2.22 eV) as shown Fig. 11. Therefore, emissions likely originated from surface defects, such as ionizable oxygen vacancies and the recombination of self-trapped excitons (STEs) which are localized within TiO6 octahedra.49 The formation of oxygen vacancies were confirmed from EPR measurement for Sn–Ti-0.12 sample in our previous study.17 The STEs in the present samples were originated from band-to-band excitation where the excited electron and hole created a local deformation of TiO6 octahedra and thus localized themselves into surface defect states of TiO2. Two peaks around 480 and 560 nm were observed for Sn–Ti-x samples (Fig. 11), which are attributed to the transition from oxygen vacancies with two-trapped electrons and one-trapped electron in the valence band of Sn–Ti-x samples, respectively. The energy levels relate to these two types of oxygen vacancies were located at 0.51 eV and 0.87 eV below the conduction band of Sn–Ti-x NPs, respectively,50 which implies that the recombination of charge carriers were effectively suppressed upon Sn–Ti-x samples.
Similar PL spectra were reported by Serpone et al.51 and Chetri et al.52 PL intensities resulting from efficient charge separation was already reported for TiO2 and ZnO doped with Sn, W, Ag, and Au.53,54 These semiconductors showing lower PL intensities were also reported to exhibit higher photocatalytic activity.55 The defect concentrations and life times of photoexcited species plays an important role in photocatalysis, investigation of photocatalysts through PL spectroscopy is important to obtain critical reasons behind the enhanced photocatalytic activity. A strong correlation between PL intensities and photocatalytic activities were established by previous researches.47,53 PL emission results from the recombination of photoinduced charge carriers, the stronger the PL signal, the higher the recombination rate of the photoinduced charge carriers.56 When the recombination rate decreases, more photoinduced charge carriers can participate in dye photodegradation, resulting in the enhancement of photocatalytic activity.57 Anatase Sn–Ti-0.05 sample, which are less luminescent with decreased intensity than rutile Sn–Ti-1.37, has a wider band gap resulting in efficient electron–hole separation. The reduction in PL intensity can also be caused by an increase in non-radiative or radiative recombination at longer wavelengths outside the recorded PL range. These effects are occurring in nanocomposite system as a consequence of charge trapping on surfaces or the inter phases between the two oxide phases.52
The defects in Sn–Ti-x samples were further characterized by lifetime measurements. The PL lifetime decay profiles of Sn–Ti-x samples with different dopant concentrations are shown in Fig. S10† and PL lifetime parameters are presented in Table 1. The decay follows tri-exponential kinetics. The photoinduced charge carriers relax to the shallow-trap levels, radiantly recombine with the lifetime of τ1. The component of shorter lifetime (τ1) was caused by the free annihilation of positrons in defect-free crystal. In a disordered system, small vacancies or shallow positron traps can reduce the surrounding electron density and thus increase the lifetime of τ1.58 The τ1 values for Sn–Ti-x are from 2.39 to 2.88 ns, longer than that of undoped Sn–Ti-0 (1.48 ns), indicating the existence of monovacancies or point defects in TiO2.59,60 The photoinduced charge carriers could be relaxed to the deep-trap levels related to the oxygen vacancies of the nanostructure and consequently recombine radiantly with a much longer lifetime of τ2, which was also increased from undoped TiO2 to Sn doped TiO2. The fast component with lifetime τ3 may be due to this near band edge relaxation of TiO2. The lifetime becomes longer due to a lack of significant overlap between electron and hole wave functions.
From Table 1, the values of τ3 decrease from undoped to Sn doped TiO2 samples, whereas the values of τ1 and τ2 corresponding to the defect-related emissions decrease in the opposite order. Thus, the lattice defects and oxygen vacancies are more pronounced in the Sn doped TiO2 samples than that of undoped sample. Consequently, it may be concluded that the Sn doping might increase the concentrations of photoinduced charge carriers in the different trap levels and oxygen vacancies, which implies the longer lifetime of the doped nanostructure than the undoped one.61 Among all samples, longer PL lifetime observed for Sn–Ti-0.05 and Sn–Ti-0.12 samples indicates lower recombination rate of the electron–hole pairs and more efficient photocatalytic performance.62
Considering the relation between magnetic and photocatalytic activities of Sn–Ti-x NPs, the anatase RTFM Sn–Ti-x (0.05 and 0.12) NPs showed high photocatalytic activity (Table 3). RTFM observed in these undoped and Sn doped TiO2 NPs can be attributed to oxygen vacancies and/or defects. Doping these NPs with low concentrations of Sn4+ ions increased both photocatalytic activity and ferromagnetism, presumably due to the creation of additional oxygen or titanium vacancies. However, when Sn doping concentration x increased above 0.12, while both photocatalytic activity and magnetic moment of Sn–Ti-x NPs rapidly decreased. Since the decrease in PDE could be attributed to the high rate of e−–h+ recombination facilitated by the doped Sn4+ ions and due to formation of diamagnetic rutile phases, we tentatively assign the reduction in Ms also to the reduction of charge carriers. High electron–hole concentration is necessary for both interfacial electron transfer models in photocatalytic process as well as the charge-transfer ferromagnetism to work. The experimental results show that the magnetic property of Sn–Ti-x were caused by the double exchange between Sn4+ and Ti4+ as well as the defective state in them, the Sn–Ti-0.05 and Sn–Ti-0.12 samples show higher ferromagnetic and photocatalytic properties than that of other Sn–Ti-x (x = 0.17, 0.30, 1.14 and 1.37) samples because of its higher defective states or oxygen vacancies and stronger visible light absorption.
To make sure that the RTFM was related to the presence of oxygen vacancies in the Sn–Ti-x samples, PL measurements are correlated with magnetic and photocatalytic results for the samples. To correlate the presence of RTFM, PL spectra and photocatalytic activity studies of the Sn–Ti-x (x = 0.05, 0.17, 0.30, 1.14 and 1.37) samples, the observed Mr, PDE values and intensity of the emission band have been plotted with Sn doping level and shown in Fig. 13. The result suggests that the doping level and oxygen vacancies in the Sn–Ti-x nanocrystals are mainly responsible for RTFM, photocatalytic activity and intensity variation of PL emission band of Sn–Ti-x samples. It has been shown that low levels of Sn doping in TiO2 can strongly reduce the energy for oxygen vacancy formation, resulting in increasing number of oxygen vacancies.68 This will reduce e−–h+ recombination rate and result in the higher PDE and alternatively high level of oxygen vacancies69 makes strong ferromagnetic behavior in metal oxide semiconductor NPs. The decrease in the value of Mr can be attributed to the decrease in the density of oxygen vacancies at higher Sn doping level. The decrease in the density of oxygen vacancies at higher Sn doping level was also confirmed by PL spectroscopic results. Oxygen vacancies are found to be the main cause for enhanced photocatalytic activity of metal oxide nanostructures.23 Contrast behavior was observed for Sn–Ti-0.17 sample, low Mr value and high PDE observed was due to predominant synergetic effect between anatase–rutile mixed phases70 than oxygen vacancies or magnetic nature. These new weak ferromagnetic Sn–Ti-0.05 and Sn–Ti-0.12 samples were extremely effective for the MO and RPhOH degradation, and maintained relatively high activity. In addition, we investigate the combined photocatalyst and ferromagnetic character at room temperature in the Sn–Ti-x NPs, which has the potential to collect the powders during waste water treatment. The room temperature hysteresis loops of Sn–Ti-0.05 sample before and after isolated from waste water treatment were measured to clarify the role of photocatalyst to make strong ferromagnetic character (Fig. S11†). The ferromagnetic character is somewhat higher for typical waste water treated Sn–Ti-0.05 photocatalyst under UV light irradiation (Ms = 21.66 memu g−1, Hc = 117.04 G and Mr = 2.07 memu g−1) instead of untreated samples (Ms = 16.66 memu g−1, Hc = 105.33 G and Mr = 1.30 memu g−1). This is due to creation of more oxygen vacancies on the UV illuminated Sn–Ti-x samples,71 but further investigations are needed to determine creation of ferromagnetism due to waste water treated photocatalysts. Hence the RTFM should be extended to various potential applications, such as spintronics, photodegradation, catalysis, separation, and purification processes.
![]() | ||
Fig. 13 Variation of intensity of emission band, Mr and PDE (phenol degradation) values of Sn–Ti-x with different Sn doping level. |
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra25609g |
‡ These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2016 |