Kunyang Li,
Yidan Peng,
Shuangyu Li,
Fengying Luo,
Jing Li,
Haidong Ju,
Yepeng Yang and
Yizhou Li
*
Yunnan Key Laboratory of Metal–Organic Molecular Materials and Device, School of Chemistry and Chemical Engineering, Kunming University, Kunming 650214, China. E-mail: zh111111ou@kmu.edu.cn
First published on 1st August 2025
In this study, SiO2–TiO2/SnO2 composites were synthesized using a facile sol–gel method to facilitate the heterogeneous photocatalytic degradation of selected pharmaceuticals and personal care products (PPCPs), namely tetracycline hydrochloride (TC) and ciprofloxacin (CIP), under visible light. The effects of different tin precursors (SnCl4·5H2O, SnCl2 and Na2SnO3·3H2O) along with the addition of SiO2 on photocatalytic performance were systematically evaluated. The photocatalyst prepared by SnCl4·5H2O as the tin precursor showed the highest photodegradation performance for ciprofloxacin (CIP) and tetracycline (TC), and the photodegradation performance for tetracycline (TC) was 1.8, 4.33 and 43.33 times that of SnCl2 and Na2SnO3·3H2O as the tin precursor and pristine TiO2, respectively. Furthermore, the incorporation of SiO2 significantly improved the photocatalytic performance under visible light. The SiO2–TiO2/SnO2 (SnCl4) composites exhibited a 2.65-fold increase in tetracycline (TC) degradation compared to the SiO2-free TiO2–SnO2 (SnCl4) materials. The significant improvement of photocatalytic activity is attributed to the effects of tin precursors and SiO2 on the physicochemical properties, charge carrier dynamics and surface reactivity of the composites. This study presents a novel method for developing composite photocatalysts with exceptional activity by employing various tin precursors and incorporating SiO2 to enhance the removal of pharmaceutical and personal care products (PPCPs) under visible light. The findings have substantial implications for future research in the photocatalytic degradation of PPCPs, fostering environmentally friendly approaches in this field.
To enhance the visible light responsiveness and photocatalytic efficiency of TiO2, modifications such as doping, heterostructure construction and composite formation have been explored. Tin oxide (SnO2), characterized by suitable band alignment and high electron mobility, has been used as a coupling agent to improve charge separation in TiO2-based photocatalysts. Petronela Pascariu et al. investigated the effects of Sn loading and calcination temperature on the synthesis, structure, optical and catalytic properties of TiO2/SnO2. They found that the highest photocatalytic activity was obtained for the sample with a calcination temperature of 500 °C and a Sn loading of 1.5%.10 The highest photocatalytic efficiency of TiO2/SnO2 nanotube composite photocatalyst with 5 wt% SnO2 content was shown by Lin-Rui Hou et al.11 The effect of calcination temperature and tin dioxide content on TiO2/SnO2 have been extensively studied by the previous researchers.10–12 However, investigations into various tin precursors remain limited.
In general, tin(IV) chloride pentahydrate (SnCl4·5H2O), tin(II) choride (SnCl2), sodium stannate trihydrate (Na2SnO3·3H2O) and tin chloride (SnCl4) have been widely used as tin precursors for the preparation of TiO2/SnO2. Kateryna Bila et al. investigated the effect of SnCl2 and SnCl4 on TiO2/SnO2 composites and showed that the use of divalent tin led to the obtaining of nanocomposites, while the use of tetravalent tin led to the obtaining of solid solutions.12 Different tin precursors have significant effects on the properties and characteristics of the synthesized TiO2/SnO2. These precursors influence the material's specific surface area, the chemical state of tin, and the surface properties, which in turn impact the photocatalytic activity.
Silicon dioxide (SiO2) also plays an important factor in the preparation of composites. Previous studies have also shown that the specific surface area, surface morphology and optical properties of the materials are significantly influenced by SiO2.13 The quantitative deposition of titanium dioxide on silica was found to significantly increase the specific surface area of the materials and enhance the photocatalytic activity by M. Bellardita et al.14 M. r. S. Nivetha et al. prepared SiO2–TiO2/g-C3N4 materials, focusing on the role of SiO2. They found that SiO2 enhances the light reflectivity of the sample and reduces the complexation of photogenerated electrons and holes in addition to increasing the specific surface area and refining the grains.15
In this study, a simple sol–gel method was developed to synthesize SiO2–TiO2/SnO2 composites by varying the tin precursors and incorporated of SiO2. The effects of various tin precursors (SnCl4·5H2O, SnCl2 and Na2SnO3·3H2O) on the physicochemical properties and photocatalytic activities of the composites were systematically evaluated. Furthermore, the influence of SiO2 addition on the charge carrier dynamics, and surface reactivity was elucidated. The visible light photocatalytic activities of the samples were investigated through the degradation of selected pharmaceuticals and personal care products (PPCPs), specifically tetracycline hydrochloride (TC) and ciprofloxacin (CIP). These findings can be attributed to the differing capabilities of materials synthesized from various tin sources in transferring and delivering electrons during the photocatalytic process, as well as the physicochemical properties and surface reactivity enhanced by the incorporation of SiO2
For comparison, TiO2 (without SiO2 and tin precursors), SiO2–TiO2 (without tin precursors), SiO2–SnO2 (SnCl4) (without TiO2) and TiO2/SnO2 (SnCl4) (without SiO2) were prepared by similar methods.
The FTIR spectra of the materials between 4000 cm−1 and 500 cm−1 are shown in Fig. 2. The two absorption bands at 3450 cm−1 and 1630 cm−1 represent the stretching vibrations of the water and hydroxyl groups, respectively.21 The SiO2-added material shows three absorption bands at 1080 cm−1 and, 955 cm−1 and 792 cm−1. They represent the asymmetric telescopic vibration of Si–O–Si, the bending vibration of Si–OH and the symmetric telescopic vibration peak of Si–O, respectively.20 The absorption band of SiO2–SnO2 (SnCl4) at 1080 cm−1 is stronger than that of other materials at the same wavenumber. This enhancement can be attributed to the stretching vibration of the Sn–O bond, which may also contribute to the spectral feature near 1080 cm−1. Additionally, the vibrational modes of the Sn–O bond may overlap with those of the Si–O–Si bond, resulting in increased intensity of the absorption band for SiO2–SnO2 (SnCl4).22 The absorption bands at 1080 cm−1 for the composites SiO2–TiO2/SnO2 (SnCl4), SiO2–TiO2/SnO2 (SnCl2), and SiO2–TiO2/SnO2 (Na2SnO3) are weaker than those of SiO2–SnO2 (SnCl4). This attenuation is attributed to the interactions among the three oxides, which induce significant alterations in the electron cloud distribution.23 As a result, these alterations impact the vibrational properties of the Si–O–Si bond, leading to a reduced dipole moment and a correspondingly weaker absorption band.23 FT-IR spectra indicate the presence of Sn–O bonds, Si–O bonds, Si–O–Si bond energies in SiO2–TiO2/SnO2 (Na2SnO3), SiO2–TiO2/SnO2 (SnCl2) and SiO2–TiO2/SnO2 (SnCl4). Notably, the observed spectral changes show pronounced interactions among the three oxide components. These suggest that SiO2–TiO2/SnO2 composites have been successfully synthesised, while there may be synergistic and other effects among the three components.
![]() | ||
Fig. 2 FT-IR spectra of TiO2, TiO2/SnO2 (SnCl4), SiO2–TiO2, SiO2–SnO2 (SnCl4), SiO2–TiO2/SnO2 (SnCl4), SiO2–TiO2/SnO2 (SnCl2) and SiO2–TiO2/SnO2 (Na2SnO3). |
The SEM and EDS spectra of these materials are shown in Fig. 3. A comparison of Fig. 3a and b reveals that the TiO2/SnO2 (SnCl4) sample exhibits a smooth and compact surface with minimal porosity. In contrast, the SiO2–TiO2/SnO2 (SnCl4) sample, which incorporates additional SiO2, exhibits a rough and loose texture. The rough surface not only enhances the specific surface area of the catalyst, but also improves light utilization through multiple diffuse reflections.24 The incorporation of SiO2 into the samples leads to a looser surface texture, which can significantly enhance their specific surface area. This increase in surface area provides more active sites for both adsorption and photocatalytic processes.25 The comparison of Fig. 3b–d demonstrates that different tin precursors significantly influence the morphology of the composites. Among the three Sn-based materials analyzed, SiO2–TiO2/SnO2 (SnCl4), synthesized using SnCl4·5H2O as a precursor, exhibits the most sparse and porous structure, while SiO2–TiO2/SnO2 (Na2SnO3), derived from Na2SnO3·3H2O as a precursor, is characterized by the highest density. This suggests that different tin precursors may have an effect on the specific surface area of the composites, which affects the photocatalytic effect of the materials. The microscopic chemical ingredient analysis of the SiO2–TiO2/SnO2 (SnCl4) materials has been shown in Fig. 3e. The patterns show the presence of tin (Sn), oxygen (O), silicon (Si) and titanium (Ti) without any other element. Four elements, namely Sn, O, Si and Ti, were identified and their molar percentages were confirmed to be 0.84%, 50.32%, 25.41% and 23.43% respectively. The measured molar percentages of these four elements were consistent with expectations, indicating that SiO2–TiO2/SnO2 (SnCl4) was prepared accurately. Fig. 3f–i shows the elemental profiles of tin (Sn), oxygen (O), silicon (Si) and titanium (Ti) on SiO2–TiO2/SnO2 (SnCl4) with different colours. This homogeneity suggests a well-integrated structure on materials which is crucial for the photocatalysis.25
The morphology and microstructure of TiO2/SnO2 (SnCl4) and SiO2–TiO2/SnO2 (SnCl4) composites were analysed using TEM and HRTEM techniques. In the TEM analysis (Fig. 4a and b), a significant difference in the morphology of TiO2/SnO2 (SnCl4) and SiO2–TiO2/SnO2 (SnCl4) composites can be observed. The TiO2/SnO2 (SnCl4) composite is characterized by the presence of spherical nanoparticles that aggregate to form larger clusters. In contrast, the SiO2–TiO2/SnO2 (SnCl4) composite is characterized by more dispersed spherical nanoparticles, each enveloped by a thin film on the exterior, and this encapsulated film is reported by Yanyan He et al. to be possibly SiO2.26 The HRTEM image in Fig. 4c shows that the stripe spacing is approximately 0.352 nm, indicating that the sample growth direction of the anatase TiO2 (1 0 1) plane.25
![]() | ||
Fig. 4 TEM images of TiO2/SnO2 (SnCl4) (a) and SiO2–TiO2/SnO2 (SnCl4) (b), HRTEM images of SiO2–TiO2/SnO2 (SnCl4) (c). |
The N2 adsorption and desorption isotherm curves, as illustrated in Fig. S1 and S2, were recorded to investigate the specific surface area and corresponding pore size distribution of the samples.27 According to the IUPAC classification of nitrogen adsorption and desorption isotherms, all samples exhibited type IV isotherms, except for SiO2–SnO2 (SnCl4), which displayed type II isotherms. This indicates that the prepared materials possess macroporous and mesoporous properties. The pore size distribution of all the samples is primarily ranges from 0 to 80 nm, indicating the macroporous and mesoporous properties of the prepared materials. As shown in Table 1, the incorporation of SiO2 increased the specific surface area of the material and provided more active sites for the photocatalyst. The utilization of tin precursor significantly influences the specific surface area of the composites. Composites synthesized using SnCl4·5H2O exhibit the largest specific surface area, whereas those produced with Na2SnO3·3H2O have the smallest specific surface area. This may be due to the fact that Sn4+ has a higher charge than Sn2+. During the formation of the complex, Sn4+ may cause greater distortion in the lattice of SiO2–TiO2, thus creating more pores and defects within the material and increasing the specific surface area. In contrast, Sn2+ induces less lattice distortion.12 For Na2SnO3, the SnO32− anion exhibits distinct structural and property characteristics compared to Sn4+ and Sn2+. Consequently, it forms different chemical bonds and structures with SiO2–TiO2 in the system, which hinders the formation of structures with a high specific surface area.28
Samples | SBET (m2 g−1) | Band gap energy (eV) | Pore volume (cm3 g−1) | Pore size (nm) |
---|---|---|---|---|
TiO2 | 100.3152 | 3.32 | 0.1718 | 3.0606 |
TiO2/SnO2 (SnCl4) | 136.1575 | 2.92 | 0.1328 | 3.1390 |
SiO2–TiO2 | 442.9692 | 3.32 | 0.1914 | 2.7956 |
SiO2–SnO2 (SnCl4) | 484.9605 | 3.58 | 0.2632 | 4.6910 |
SiO2–TiO2/SnO2 (SnCl4) | 272.1116 | 2.81 | 0.1838 | 2.6483 |
SiO2–TiO2/SnO2 (SnCl2) | 180.7277 | 2.85 | 0.1200 | 2.8943 |
SiO2–TiO2/SnO2 (Na2SnO3) | 141.9377 | 2.87 | 0.1027 | 2.6569 |
To further investigate the chemical composition and binding states of the samples, the X-ray photoelectron spectroscopy of the samples was investigated. As shown in Fig. S4a, the binding energies of Ti 2p1/2 and Ti 2p3/2 in SiO2–TiO2/SnO2 (SnCl4) were located at 458.8 eV and 464.6 eV, respectively, attributed to Ti4+.29 The Sn 3d spectra of SiO2–TiO2/SnO2 (SnCl4) show spin–orbit double peaks at 495.6 eV and 486.8 eV, respectively, due to Sn4+ (Fig. S4b).10 The binding energy of Si 2p in SiO2–TiO2/SnO2 (SnCl4) is located at 103.2 eV, attributed to Si4+ (Fig. S4c).20 As shown in Fig. 5, the XPS curves of O 1s in SiO2–TiO2/SnO2 (SnCl4) can be identified as two peaks at 530.98 eV and 532.58 eV, which can be attributed to covalently bonded oxygen and surface oxygen species (Si–O, surface hydroxyl groups and adsorbed water) and metal oxide lattice oxygen (Sn–O and Ti–O).30 Compared to Fig. 5a, it can be seen that the amount of covalently bonded oxygen and surface oxygen species increases with the addition of SiO2. Firstly, the Si–O bond exhibits a peak at 532.58 eV, indicating that the introduction of SiO2 directly enhances the signal at this energy level.31 Secondly, the high specific surface area of SiO2 may result in a greater amount of adsorbed water (H2O) or hydroxyl residues, which in turn contributes to the increase of the 532.58 eV peak.32 Surface adsorbed water and –OH groups can trap photogenerated holes (h+) into reactive –OH radicals, thereby improving the photocatalytic performance of the materials.33
![]() | ||
Fig. 5 XPS spectra of O 1s region for samples, (a) the influence of SiO2, (b) the influence of Sn precursors . |
As can be seen from Fig. 4b, different tin precursors also influence the binding energy of O 1s. The O 1s binding energies of SiO2–TiO2/SnO2 (SnCl4) (532.58 eV and 530.98 eV) are higher than those of SiO2–TiO2/SnO2 (SnCl2) (532.28 eV and 530.68 eV) and SiO2–TiO2/SnO2 (Na2SnO3) (532.28 eV and 530.58 eV). In SiO2–TiO2/SnO2 (SnCl4), compared to SnCl2, Sn4+ carries a stronger positive charge which further polarises the Sn–O bond and reduces the electron density around oxygen, thereby increasing the O 1s binding energy. Sn2+ displays lower Lewis acidity and a diminished capacity to withdraw electron density from oxygen, resulting in a lower O 1s binding energy for SiO2–TiO2/SnO2 (SnCl2) O 1s.34 As for SiO2–TiO2/SnO2 (Na2SnO3), Na+ ions inhibit the formation of the SnO2 phase, maintaining the high purity Na2SnO3 structure. The O atoms in the SnO32− group have a higher electron density, resulting in the lowest O 1s binding energy of the three samples. A higher binding energy signifies a reduced electron cloud density of the lattice oxygen in metal oxides (Sn–O, Ti–O), which enhances the oxidative capacity of holes (h+) and facilitates the photocatalytic degradation of pollutants.35
The kinetic analysis of TC and CIP degradation was based on the fitting of pseudo-first-order equations, and the kinetic constants are shown in Fig. 7, with K values for the materials listed in Table 3. Among the catalysts tested, SiO2–TiO2/SnO2 (SnCl4) exhibited the highest K values, with values of 0.0130 min−1 for the photocatalytic degradation of TC and 0.0110 min−1 for the photocatalytic degradation of CIP, respectively. The higher K value showed that the material degraded faster. Notably, the K value for TC was 1.88 and 4.33 times higher than those of composites prepared with SnCl2 and Na2SnO3·3H2O as tin precursors, respectively. Furthermore, the incorporation of SiO2 resulted in a significant enhancement of the photocatalytic degradation rate, achieving a 2.65-fold increase in the K value for the degradation of TC compared to the TiO2/SnO2 (SnCl4). These results underscore the critical role of SiO2 incorporation and tin precursor selection in optimizing the degradation efficiency of TC and CIP.
![]() | ||
Fig. 7 Apparent first order rate constant k (min−1) for TC (20 mg L−1) (a) and CIP (20 mg L−1) (b) photocatalytic degradation over the as-prepared samples. |
Additionally, comparisons with previous studies have been conducted to evaluate the photocatalytic performance, as presented in Tables 2 and 4. Like Tables 2 and 4, TC and CIP were identified as the selected degradants in this study, and the prepared composites showed high removal of both compounds in visible light compared to previous studies.
Samples | K (TC, min−1) | K (CIP, min−1) |
---|---|---|
TiO2 | 0.0003 | 0.0002 |
SiO2–TiO2 | 0.0006 | 0.0005 |
SiO2–SnO2 (SnCl4) | 0.0013 | 0.0014 |
SiO2–TiO2/SnO2 (Na2SnO3) | 0.0030 | 0.0033 |
TiO2/SnO2 (SnCl4) | 0.0049 | 0.0046 |
SiO2–TiO2/SnO2 (SnCl2) | 0.0069 | 0.0061 |
SiO2–TiO2/SnO2 (SnCl4) | 0.0130 | 0.0100 |
![]() | ||
Fig. 8 (a) Transient photocurrent. (b) Nyquist plots of EIS. (c) PL spectra and. (d) UV-vis diffuse reflectance spectra. |
As shown in Fig. 8, the SiO2–TiO2/SnO2 (SnCl4) composites exhibit the highest transient photocurrent intensity (Fig. 8a) and the smallest electrochemical impedance spectroscopy arc radius (Fig. 8b), indicating excellent charge separation efficiency and minimum charge transfer resistance. The photoluminescence (PL) spectra (Fig. 8c) showed that the emission peak intensity of SiO2–TiO2 was significantly reduced compared to that of pure TiO2, whereas the weakest PL signal was obtained for SiO2–TiO2/SnO2 (SnCl4), confirming the effective suppression of charge recombination. The UV-vis absorption spectra (Fig. 8d) showed that the absorption edges of all the tin-containing samples were red-shifted, with SiO2–TiO2/SnO2 (SnCl4) having the widest absorption region with a band gap of 2.81 eV (the band gap of the prepared samples was estimated from the Tauc plot as shown in Fig. S5).
In summary, these findings indicated that the mechanism underlying the enhancement of the photocatalytic properties of the materials primarily involves two factors: the incorporation of SiO2 and the use of various tin precursors. The incorporation of SiO2 significantly improved the photocatalytic performance through two main mechanisms. Firstly, as shown in Fig. 8b and c, the incorporation of SiO2 effectively reduced the electrochemical impedance and increased the intensity of the photoluminescence(PL) of the materials. This improvement can be attributed to the role of SiO2 as a structural mediator, facilitating close contact between TiO2 and SnO2 to form a more efficient heterostructure. The resulting synergistic effect not only narrowed the band gap but also extended the visible light absorption while suppressing the recombination of photogenerated electron–hole pairs.44 Secondly, as shown in Table 1 and Fig. 3, the incorporation of SiO2 increased the specific surface area of the material and promoted a more looser structure. This structural modification provided additional active sites for photocatalytic reactions. Furthermore, the increased specific surface area improved light scattering efficiency, thereby increasing the material's overall light-harvesting capability. These combined effects synergistically contributed to the superior photocatalytic performance of the SiO2-modified composites.20
Fig. 8 demonstrates that various tin precursors significantly influence both the rate of photogenerated electron–hole separation and the lifetime of photogenerated carriers. The photogenerated carrier lifetime of SiO2–TiO2/SnO2 (SnCl4) is greater than that of SiO2–TiO2/SnO2 (SnCl2) and also exceeds that of SiO2–TiO2/SnO2 (Na2SnO3), indicating that the choice of tin precursor is crucial for the enhanced properties of the samples.
For the three tin precursors, Sn4+ in SnCl4·5H2O is more easily converted to SnO2 and forms a heterojunction structure with TiO2. This reduces the forbidden bandwidth, facilitates the enhancement of the separation rate of photogenerated carriers, reduces the complexation between photogenerated electrons and holes, and thus enhances the activity and strength of the photocatalyst.45 Sn is in the +2 valence state in SnCl2, likely due to incomplete oxidation. Although Sn in this state can influence the structure and properties of titanium dioxide, its ability to transfer and transmit electrons during the photocatalytic process is relatively weak because of its low valence.12 Tin exhibits a +4 oxidation state in Na2SnO3·3H2O; however, the presence of a substantial amount of sodium ions (Na+) in the system plays a crucial role. Na+ ions can inhibit the formation of the SnO2 phase, thereby preserving the high purity phase structure of Na2SnO3·3H2O, which is not favourable for the formation of TiO2/SnO2 heterojunctions.46
Free radical trapping tests were carried out on SiO2–TiO2/SnO2 (SnCl4) as shown in Fig. 9a and b to determine which are the active species in the photocatalytic process of SiO2–TiO2/SnO2 (SnCl4). The photodegradation efficiency decreased when Isopropanol (IPA), ethylenediamine tetraacetic acid disodium salt (EDTA-2Na), and p-benzoquinone (BQ) were added to the TC and CIP solutions. The degradation efficiencies decreased to 54.7% and 49.9% with the addition of IPA, 62.9% and 60.7% with the addition of EDTA-2Na, and 77.4% and 77.3% with the addition of BQ. This result indicates that ·OH and h+ are the main active species of SiO2–TiO2/SnO2 (SnCl4) photocatalysts during the photodegradation of CIP and TC. To verify this conclusion, electron paramagnetic resonance (EPR) spectroscopy is shown in Fig. 9c. In the absence of light, the EPR spectra in Fig. S6 do not show any distinct signal peaks. After 10 minutes of irradiation under a 500 W xenon lamp, the EPR spectrum of SiO2–TiO2/SnO2 (SnCl4) shows four distinct peaks in the range of 3460 to 3560 G, which corresponds to the characteristic signals of ·OH.20 The EPR results confirm that SiO2–TiO2/SnO2 is capable of generating ·OH in the presence of light.
To assess the reusability of the SiO2–TiO2/SnO2 (SnCl4) photocatalyst, a series of four reusability tests was conducted. The results are shown in Fig. 10a and b. After four cycles, the degradation rate of SiO2–TiO2/SnO2 (SnCl4) for TC and CIP changed from 84.9% and 81.3% to 81.0% and 77.5%, indicating that the photocatalyst has good stability in both TC and CIP degradation. The sample demonstrates excellent reusability as a photocatalyst.
In order to investigate the degradation process of CIP and TC, liquid chromatography-mass spectrometry (LC-MS) was used to analyse the major degradation products of CIP and TC. Three possible degradation pathways were detected for CIP, as shown in Fig. S6: (1) in pathway 1, the secondary amines on the piperazyl ring were first attacked. Subsequently, the carbon–nitrogen bonds broke down.47 (2) In pathway 2, the m/z = 348 organic matter was generated by CIP.48 (3) Pathway 3 was generated by CIP through the free radical elimination of CO, which then evolved through oxidation.16 As shown in Fig. S7, two possible degradation pathways for TC were identified. Pathway 1 could involve the dehydration of the TC molecule to generate an intermediate product with a mass-to-charge ratio (m/z) of 427, followed by a deamination reaction leading to the formation of an intermediate product with an m/z of 410. This intermediate product undergoes terminal oxidation induced by H+, ·O2− and ·OH, resulting in an intermediate product with an m/z of 302.49 (2) Pathway 2 may involve attack by SO4−· radicals, generating a product with m/z = 461, which is then further oxidised to produce a product with m/z = 477.50 The total organic carbon (TOC) removal experiments are shown in Table S1. The mineralisation of ciprofloxacin (CIP) and tetracycline hydrochloride (TC) was 29.46% and 33.11% respectively (see Table S1). This indicated that CIP and TC were partially mineralised to CO2 and H2O.
It is essential to assess the toxicity of the solution after photocatalytic treatment.51 As shown in Fig. 11, the toxicity of the treated solution was determined by comparing the effect of the solution on the viability of E. coli before and after the treatment by means of the E. coli viability assay. The plate smear count test was used to record the number of colonies. Fig. 11 shows that there were 101 colonies in the blank group, 84 in the pre-treatment solution group, and 97 in the post-treatment solution group. This suggests that the photocatalytically treated solution was essentially non-toxic to E. coli.
![]() | ||
Fig. 11 Effect of E. coli activity: (a) blank control, (b) ciprofloxacin solution before degradation, (c) degraded ciprofloxacin solution. |
S1: instruments and equipment; S2: measurement of photocatalytic activity; S3: electron paramagnetic resonance (EPR) measurements; Fig. S1: surface area analysis of composites. N2 adsorption/desorption isotherms of (a) TiO2, (b) SiO2–SnO2 (SnCl4); Fig. S2: surface area analysis of composites. N2 adsorption/desorption isotherms of (a) TiO2/SnO2 (SnCl4), (b) SiO2–TiO2/SnO2 (SnCl4), (c) SiO2–TiO2, (d) SiO2–TiO2/SnO2 (Na2SnO3), (e) SiO2–TiO2/SnO2 (SnCl2); Fig. S3: adsorption/desorption curves in the dark (a) TC, (b) CIP; Fig. S4: XPS spectra of Ti 2p, Sn 3d and Si 2p region for SiO2–TiO2/SnO2 (SnCl4 ); Fig. S5: plots of (αhν)2 − hν (a–f); Fig. S6: proposed degradation pathway of CIP over SiO2–TiO2/SnO2 (SnCl4); Fig. S7: proposed degradation pathway of TC over SiO2–TiO2/SnO2 (SnCl4); Table S1: the total organic carbon (TOC) values of TC and CIP. See DOI: https://doi.org/10.1039/d5ra03247d.
This journal is © The Royal Society of Chemistry 2025 |