Next Article in Journal
Preparation of DNC Solid Dispersion by a Mechanochemical Method with Glycyrrhizic Acid and Polyvinylpyrrolidone to Enhance Bioavailability and Activity
Previous Article in Journal
Effect of Textile Wastewater Secondary Effluent on UF Membrane Characteristics
Previous Article in Special Issue
Reactive Oxygen Species-Responsive Miktoarm Amphiphile for Triggered Intracellular Release of Anti-Cancer Therapeutics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

One-Pot Synthesis of SnO2-rGO Nanocomposite for Enhanced Photocatalytic and Anticancer Activity

Department of Physics and Astronomy, College of Sciences, King Saud University, Riyadh 11451, Saudi Arabia
*
Author to whom correspondence should be addressed.
Polymers 2022, 14(10), 2036; https://doi.org/10.3390/polym14102036
Submission received: 3 April 2022 / Revised: 13 May 2022 / Accepted: 13 May 2022 / Published: 16 May 2022
(This article belongs to the Special Issue Intelligent Self-Assembled Polymer for Targeted Cancer Treatment)

Abstract

:
Metal oxide and graphene derivative-based nanocomposites (NCs) are attractive to the fields of environmental remediation, optics, and cancer therapy owing to their remarkable physicochemical characteristics. There is limited information on the environmental and biomedical applications of tin oxide-reduced graphene oxide nanocomposites (SnO2-rGO NCs). The goal of this work was to explore the photocatalytic activity and anticancer efficacy of SnO2-rGO NCs. Pure SnO2 NPs and SnO2-rGO NCs were prepared using the one-pot hydrothermal method. X-ray diffraction (XRD), transmission electron microscopy (TEM), scanning electron microscopy (SEM), X-ray photoelectron spectroscopy (XPS), Fourier transform infrared (FTIR), UV–Vis spectrometry, photoluminescence (PL), and Raman scattering microscopy were applied to characterize the synthesized samples. The crystallite size of the SnO2 NPs slightly increased after rGO doping. TEM and SEM images show that the SnO2 NPs were tightly anchored onto the rGO sheets. The XPS and EDX data confirmed the chemical state and elemental composition of the SnO2-rGO NCs. Optical data suggest that the bandgap energy of the SnO2-rGO NCs was slightly lower than for the pure SnO2 NPs. In comparison to pure SnO2 NPs, the intensity of the PL spectra of the SnO2-rGO NCs was lower, indicating the decrement of the recombination rate of the surfaces charges (e/h+) after rGO doping. Hence, the degradation efficiency of methylene blue (MB) dye by SnO2-rGO NCs (93%) was almost 2-fold higher than for pure SnO2 NPs (54%). The anticancer efficacy of SnO2-rGO NCs was also almost 1.5-fold higher against human liver cancer (HepG2) and human lung cancer (A549) cells compared to the SnO2 NPs. This study suggests a unique method to improve the photocatalytic activity and anticancer efficacy of SnO2 NPs by fusion with graphene derivatives.

1. Introduction

Hepatocellular carcinoma is the second leading cause of death after lung cancer, globally [1]. The World Health Organization (WHO) reported that there were approximately 19 million new cancer cases and approximately 10 million deaths from cancer worldwide in 2020 [2]. There are several ways, such as surgery, radiation, and chemotherapy, to treat cancer. However, drug resistance and selectivity are still two major hurdles in cancer therapy. On the other hand, the release of environmental pollutants (e.g., dyes and drugs) from various factories challenge the intactness of the ecosystem. Hence, it is important to develop a nanostructure that can be used in biomedical as well as environmental applications [3].
Metal nanoparticles, such as silver nanoparticles (AgNPs), have recently become attractive to different areas such as catalysis, bio-sensing, antimicrobial activity, and bio-medicine due to the fact of their potential applications [4,5]. Metal oxide nanoparticles (NPs) have shown potential for their application in cancer therapy and environmental remediation [6,7]. For example, NPs of ZnO, TiO2, WO3, and In2O3 have been commonly studied for catalytic, anticancer, antibacterial, optoelectronics, and photocatalytic degradation [8,9,10] due to the fact of their unique physicochemical properties. Particularly, tin oxide (SnO2) NPs, an n-type semiconductor with a wide bandgap of ~3.6 eV, was studied for these purposes because of their several advantages, e.g., low cost, stability, facile synthesis, and low toxicity [11]. The hydrothermal method has been commonly used to synthesize various nanoforms of SnO2 such as NPs, nanorods, and nanowires [12,13]. Earlier studies reported the photocatalytic activity of SnO2 NPs under UV or visible light illumination [14,15]. However, there is a large scope for further improvement of the photocatalytic performance of such nanostructures.
Graphene is a 2D nanocrystal consisting of single layer of carbon atoms arranged in a honeycomb lattice. Graphene oxide (GO) and reduced graphene oxide (rGO) are the two most important derivatives of graphene [16,17]. GO and rGO are promising materials for various applications due to the fact of their interesting properties such as high surface area, high mechanical strength, excellent optical property, and presence of functional groups on their surfaces. Due to the presence of functional groups, GO and rGO have attracted the attention of material scientists for the synthesis of nanocomposites with metal oxide NPs [18,19]. Earlier studies focused on improving the physicochemical properties of metal oxide NPs by integration of GO or rGO [20,21]. Moreover, GO has different advantages that enhance the properties natural polymer as a nanocomposite. For example, natural polymer/GO composites (KGM/GO) are strong and biocompatible [22]. Different synthesis procedures, such as laser irradiation, gas–liquid interface interaction, co-precipitation, and hydrothermal, were employed to prepare metal oxide NPs anchored on GO/rGO nanosheets [23]. Such nanocomposites have superior characteristics than their individual one. For instance, ZnO-rGO NCs have better anticancer efficiency than pure ZnO NPs [20]. Au-rGO NCs were prepared by green synthesis for PTT of MCF7 cancer cell treatment with promising application in the field of nanobiomedicine [24]. Another study demonstrated that the photocatalytic activity of Fe3O4 NPs improved after doping with rGO [25]. Chen et al. [26] reported that the prepared Pd nanocatalyst exhibited an outstanding catalytic applicability for reducing extremely poisonous Cr(VI) and a broad variety of azo dyes.
Keeping the above in mind, we aimed to improve the anticancer and photocatalytic activity of SnO2 NPs by incorporating rGO. Pure SnO2 NPs and SnO2-rGO NCs were produced using the one-pot hydrothermal method. X-ray diffraction (XRD), field emission scanning electron microscopy (FE-SEM), field emission transmission electron microscopy (FE-TEM), X-ray photoelectron microscopy (XPS), Raman scattering microscopy, Fourier transmission infrared (FTIR) microscopy, UV–Vis spectrometry, and photoluminescence (PL) spectrometry were applied to characterize the physicochemical properties of the prepared samples. The photocatalytic activity of SnO2 NPs and SnO2-rGO NCs was examined against methylene blue (MB) dye under UV illumination. The anticancer potential of these samples were explored in two different cancer cells: human liver cancer (HepG2) and human lung cancer (A549).

2. Experimental Section

2.1. Materials and Reagents

Tine chloride (SnCl4·4H2O), graphite powder, sodium nitrate (NaNO3), sulphuric acid (H2SO4), potassium permanganate (KMnO4), hydrogen peroxide (H2O2), hydrazine monohydrate (NH2NH2·H2O), and methyl blue (MB) dye were obtained from Sigma Aldrich (Millipore-Sigma, St. Louis, MO, USA). Distilled water (DW) was employed as a medium for preparation. The chemicals were utilized as received without any further purification.

2.2. Synthesis of Graphene Oxide and Reduced Graphene Oxide

Graphene oxide (GO) was synthesized via Hummer’s method [27]. A reduction method was used to synthesize rGO to form GO. The 200 mg of GO was dispersed in 30 mL of DW. Then, 0.5 mL of hydrazine monohydrate (NH2NH2·H2O) was added to the above solution. The mixture solution was further warmed to 80 °C for 3 h to reduce GO into rGO. Subsequently, the suspension was centrifuged and rinsed several times with ethanol and distilled water and finally dried at 70 °C for 18 h to obtain the rGO sheets.

2.3. Synthesis of SnO2-rGO Nanocomposites

A one-pot hydrothermal approach was applied to prepare the SnO2-rGO NCs. Briefly, 1 g of tine chloride (SnCl4·4H2O) and 0.2 g of GO were dispersed in 30 mL of DW and sonicated for 30 min. Then, 1.5 mL of hydrazine (NH2NH2·H2O) was added to this solution. The NaOH solution was further slowly added to the mixture to reach pH 13. Next, the mixture was stirred for 1 h to obtain a homogenous solution. Fifty milliliters was transferred to a stainless autoclave and heated at 140 °C for 24 h. After, the product was centrifuged, washed several times with ethanol and DW, and dried at 70 °C for 18 h to obtain the SnO2-rGO NCs. A similar procedure was applied for the synthesis of pure SnO2 NPs without mixing of GO. The synthesis protocol of the SnO2-rGO NCs is depicted in Scheme 1.

2.4. Characterization Techniques

X-ray diffraction (PanAnalytic XPert Pro, Malvern Instruments, Malvern, Worcestershire, UK), with Cu-Kα radiation (λ = 0.15405 nm, at 45 kV and 40 mA) and an angle ranging from 30° to 80°, was used to investigate the crystal structure and phase purity of the prepared samples. The morphology was studied via field emission transmission electron microscopy (FETEM) (200 kV, 2100F, JEOL, Inc., Tokyo, Japan). Mapping of the elemental distribution of the SnO2-rGO NCs was carried out by field emission scanning electron microscopy (FE-SEM) (JSM-7600F, JEOL, Inc.). Moreover, the chemical state and elemental composition of SnO2-rGO NCs were determined by X-ray photoelectron spectroscopy (XPS) (PHI-5300 ESCA PerkinElmer, Boston, MA, USA) and energy-dispersive X-ray spectroscopy (EDX) [28]. Moreover, the typical working conditions of XPS were 10−9 Torr. This is necessary because the released photoelectrons have a low energy and are easily absorbed by the surrounding environment. Al Kα with Mgα X-rays were used to excite the samples. A micro-Raman spectroscopic analysis was also performed using a Raman microscope (Thermo Scientific, Waltham, MA, USA) with a 532 nm (6 mW) laser and 32 two-second scans (IY-Horiba-T64000). The absorption spectra of the SnO2 NPs and SnO2-rGO NCs were recorded using a UV–Visible spectrophotometer (Hitachi U-2600). The PL peak intensity of SnO2 NPs and SnO2-rGO NCs were assessed by a fluorescent spectrometer (Hitachi F-4600, Hitachi, Tokyo, Japan) with an excitation wavelength of 300 nm. The effect of the rGO sheets on the microstructure of the SnO2 NPs was studied by Fourier transform infrared (FTIR) (PerkinElmer Paragon 500, Waltham, MA, USA).

2.5. Cell Culture

Human liver cancer (HepG2) and human lung cancer (A549) cell lines were used to assess the anticancer activity of NPs and NCs. Cell lines were purchased from American Type Culture Collection (ATTC, Manassas, WV, USA). Cells were cultured in DMEM (Invitrogen, Carlsbad, CA, USA) with 10% fetal bovine serum (FBS) and antibiotics (100 µg/mL streptomycin + 100 U/mL penicillin). Cells were maintained in a humidified incubator at 37 °C with a 5% CO2 supply.

2.6. Exposure Protocol

An amount of 1000 µg/mL of both of SnO2 NPs and SnO2-rGO NCs were dissolved in DMEM as stock suspension. The suspension was further diluted to different concentrations (i.e., 0, 5, 10, 25, 50, 100, and 200 µg/mL). The different dilutions NPs and NCs were sonicated for 20 min at 80 W at room temperature in an ultrasonic water bath sonicator before being exposed to cells.

2.7. Cell Viability

To evaluate the effect of the synthesized samples against two types of cancer (i.e., HepG2 and A549), MTT assays were performed. At a density of 1 × 104 cells/well, 100 µL of cells were placed onto 96-well plates and incubated for 24 h. Then, 100 µL of each concentration (i.e., 0, 5, 10, 25, 50, 100, and 200 µg/mL) of the SnO2 NPs and the SnO2-rGO NCs were added into each well and incubated for 24 h. After, 100 µL of MTT solution was added to each well and incubated for 3 h. Next, 100 µL of CHAPS detergent (3-(3-cholamidopropyl) (dimethylammonio)-1-propanesulfonate) solution was added into each well to solubilize the MTT dye. Cell viability was then measured using a microplate reader (BioTek ELx800 Universal) at a wavelength of 570 nm.

2.8. Photocatalytic Evaluation

The photocatalytic experiments were conducted through degradation of MB dye under UV illumination (λ ≥ 420 nm), which was provided using a 400 W Xe lamp (CEL-HXF300, Beijing China Education Au-light Co., Ltd., Beijing, China). The 10 ppm of MB dye was dispersed in 50 mL of DI water under continuous stirring, and 2 mL of MB solution was taken as dye without catalysts. Then, 20 mg of both of SnO2 NPs and SnO2-rGO NCs was added to the above solution as a suspension solution. Before exposure to UV irradiation, an adsorption–desorption equilibrium was achieved after 30 min of stirring the suspension solution in the dark. At regular intervals, 2 mL of suspension under light was taken and centrifuged to remove the catalyst powders. The degradation efficiency of MB dye was estimated by following Equation (1).
Degradation efficiency % = C 0 C t C 0 × 100
where C0 represents the absorbance before irradiation, and Ct indicates the absorbance throughout the irradiation. To examine cycle stability, the catalysts were separated from suspension by centrifugation and subsequent washing with ethanol and deionized water. The catalysts were further dried at 60 °C for 12 h and utilized several times.

3. Results and Discussion

3.1. Crystallographic Study

The crystalline structure and phase formation of GO, rGO, SnO2 NPs, and SnO2-rGO NCs were examined by X-ray diffraction (XRD). The XRD spectra of GO and rGO nanosheets showed 2θ values of peaks at 10.57°, 27.94°, and 24.86° corresponding to the (001), (100), and (002) planes, respectively. In Figure 1, various 2θ (hkl) values of peaks are shown in the SnO2-rGO NCs at 26.24° (110), 33.92° (101), 37.94° (200), 51.60° (211), 58.30° (220), 61.8° (002), 78.48° (202), and 83.31° (321) planes. The results reveal that the peak of the (002) plane of graphitic carbon in the SnO2-rGO nanocomposites (NCs) overlapped with the (110) plane of the SnO2 NPs [29]. All of the SnO2 peaks in the XRD spectra associated to the standard SnO2 (JCPDS card No: 01-0803912). The average crystalline sizes of the SnO2 NPs and SnO2-rGO NCs were calculated using the Sherrier equation applying a (110) peak.
D = k λ β cos θ
where k = 0.90 is the shape factor of the crystallite, λ is the wavelength, β is the full width at half maximum, and θ is the reflection angle.
According to Scherrer’s equation, the average crystallite size of the SnO2 NPs and SnO2-rGO NCs were 7.64 and 7.98 nm, respectively (Table 1). The crystallite size of the SnO2 NPs decreased after loading on rGO as shown in other reports [30,31].

3.2. Morphological Study

High-resolution TEM (HRTEM) images were selected to investigate the morphology, chemical compositions, and particle size of the synthesized samples (Figure 2). TEM and HRTEM images show that GO exhibited thick, flattened nanosheet-like surfaces. Navazani et al. [32] reported that the increased thickness of the graphene sheets was attributed to organic functional groups and the electrostatic interaction of oxides on the surface. TEM images (Figure 2G,H) show that the distribution of the SnO2 NPs was randomly agglomerated with a uniform particles size owing to smaller particle size and higher surface energy of SnO2 NPs as reported earlier [33]. TEM images (Figure 2L,M) of the SnO2-rGO NCs show that the SnO2 NPs were loaded onto the rGO nanosheet. These results suggest the successful formation of SnO2-rGO NCs. Figure 2I,N show that the lattice spacing (d) of the SnO2 NPs and SnO2-rGO NCs were found to be 0.220 and 0.330 nm, respectively, which matched to the (210) and (110) planes of the SnO2 structure [34,35]. These values of interplanar spacing (d) were supported by the XRD data.

3.3. SEM Study

FE-SEM measurements were used to study the surface morphology of the prepared samples. Figure 3 displays the FE-SEM images of GO, rGO, SnO2 NPs, and SnO2-rGO NCs. In Figure 3A,B, the structure of GO and rGO nanosheets can be observed as a carbon layer. Figure 3C,D shows that the SnO2 NPs were spherical shaped with a uniform distribution and an anchored rGO sheet, which was also confirmed by TEM analysis. Ahamed et al. [36] observed that rGO improved the anticancer therapy of metal oxide NPs by tuning its physicochemical properties. Such a phenomenon is crucial and useful for photocatalytic degradation and cancer therapy. EDX spectra (Figure 4A) confirmed the presence of Sn, C, and O elements in the SnO2-rGO NCs. SEM mapping (Figure 4C,E) of SnO2-rGO NCs also exhibited a homogeneous distribution of Sn, O, and C elements in the SnO2-rGO NCs.

3.4. XPS Study

The electronic structure and the chemical composition of the SnO2-rGO NCs samples were also investigated by X-ray photoelectron spectroscopy (XPS), thus providing information on the electronic communication between SnO2 and rGO [37]. The binding energy peaks of Sn, O, and C elements were shown in the XPS wide spectra (Figure 5A) of the SnO2-rGO NCs. It can be seen that peaks of the Sn element were matched to the spin-orbit peaks of Sn 3p, Sn 3d, and Sn 4p [38]. The high-resolution spectra of the Sn, O, and C are presented in Figure 5B–D, respectively. The results suggest that the presence of Sn, O, and C peaks confirm the fabrication of SnO2 on rGO as supported by the EDX data (Figure 4A). The binding energy of the Sn3d5/2 and Sn3d3/2 peaks were assigned at 486.55 and 495.3 eV, respectively, which could be attributed to the SnO2 NPs [39]. According to the curve fit of the O1s spectra (Figure 5C), the binding energy peaks of O1s at 532 and 533 eV corresponded to Sn-O and Sn-O-C, respectively, which agrees with previous works [40,41]. The C1s peak (Figure 5D) was subdivided into three individual peaks after the fitting of the main peak. The peaks at 284, 285, and 288 eV were observed to correspond to C=C, C-OH, and O-C=O, respectively, as shown in a previous study [42]. These results reveal that the rGO had a residual oxygen group [40].

3.5. Raman Analysis

The electronic and vibration bands of prepared samples were analyzed via Raman spectroscopy. The Ramen spectra of the SnO2 NPs and SnO2-rGO NCs are shown in Figure 6A. It can be seen that Eu, As, A1g, and B2g vibration bands of the SnO2 NPs’ peaks were located at 352.33, 579, 620, and 773.32 cm−1, respectively [43]. The peak positions at 579 and 620 cm−1 of the SnO2-rGO NCs were similar to the SnO2 NPs as shown in Figure 6A. These results indicate that the intensity of the Raman peaks of the SnO2-rGO NCs was lower than the SnO2 NPs due to the interaction between SnO2 NPs and rGO [39]. Figure 6B shows that two new scattering bands (i.e., band D and band G) appeared in the SnO2-rGO NCs compared to the SnO2 NPs. Moreover, the scattering bands of the SnO2-rGO NCs were observed at 1349.47 cm−1 (band D) and 1590.50 cm−1 (band G), which were attributed to instability in the hexagonal graphitic layers and vibration bonds of the carbon atom, respectively [44]. Our results from the Raman analysis confirm that the SnO2 NPs attached onto the surface of the rGO sheets [45].

3.6. FTIR Study

The functional groups of the GO, rGO, SnO2 NPs, and SnO2-rGO NCs were determined by FTIR spectrometry. Figure 7 shows the FTIR spectra of the prepared samples in a range of 450–4000 cm−1. The results reveal the changes in the microstructural characteristics in the first region (500–650 cm−1) of the SnO2 NPs during loading on rGO. However, three stretching vibrations of C-O, C=C, and C=O assigned at 1048.73, 1620, and 1734.92 cm−1, respectively, can be seen in the GO. The presence of C=O and C-O at the 1713.96 and 1048.37 cm−1 bands indicate the reduction of GO into rGO [46]. Moreover, O-H stretching vibrations were observed at 3433.69, 1646.43, and 1353 cm−1 for both SnO2 NPs and SnO2-rGO NCs due to the existence water molecules during the hydrothermal process [47]. The band observed at 593 cm−1 was assigned to the O-Sn-O stretching mode in the SnO2 NPs [48], while two strong peaks at 530.20 and 608.6 cm−1 in the SnO2-rGO NCs were assigned to Sn-O stretching vibration owing to the strong interaction between SnO2 and rGO [49]. These results indicate the formation of single-structure SnO2-rGO NCs as supported by the XRD data.

3.7. Optical Study

The optical properties of the SnO2 NPs and SnO2-rGO NCs were studied by UV–Vis spectroscopy. The UV–Vis spectra of the SnO2 NPs and SnO2-rGO NCs are presented in Figure 8A. The absorption edge was shifted toward a higher wavelength for the SnO2-rGO NCs in comparison to pure SnO2 NPs. The difference in absorption peaks tuned the bandgap energy of the SnO2-rGO NCs in comparison to the SnO2 NPs. Taucs formula was used to assess the bandgap energy [50]. The bandgap energy (Eg) (Figure 8B) of the SnO2 NPs and SnO2-rGO NCs were 3.79 and 3.45 eV, respectively. We observed a slight decrease in the bandgap energy of the SnO2 NPs after rGO integration due to the increase in particle size as shown Table 1. Reduction of the bandgap energy improves UV light absorption, which can be applied in the photocatalytic degradation of pollutants [51].

3.8. Photoluminescence (PL)

Photoluminescence (PL) was used to measure the migration rate of electrons and holes in the SnO2NPs and SnO2-rGO NCs NPs. The PL spectra of the SnO2NPs and SnO2-rGO NCs NPs are presented in Figure 9. Moreover, the PL spectra of the SnO2 NPs and SnO2-rGO NCs at room temperature were obtained using an excitation wavelength of 355 nm. After adding rGO, an emission peak at 402 nm was greatly reduced. Our results observed that the PL intensity of the SnO2-GO NCs system (black line) was lower than that of pure SnO2 NPs, which may be attributed to greater charge separation, a longer life of the electron–hole pair, and a higher efficiency of charge [52]. This process has the potential to be used in a large variety of applications including cancer treatment and the photocatalytic degradation of environmental contaminants.

3.9. Photocatalytic Study

The degradation of MB dye under UV irradiation was used to investigate the photocatalytic activity of the SnO2 NPs and SnO2-rGO NCs. Figure 10A,B shows the absorption spectra of the MB solution of the SnO2 NPs and SnO2-rGO NCs under UV irradiation within 50 min. The absorption wavelength of the MB solution was at 664 nm as shown in Figure 10. We observed that the intensity of the absorption at 664 nm in the MB spectra decreased with increasing exposure time. Additionally, the color of the MB solution faded gradually, indicating the breaking of the dye’s structure by oxidation [53]. Ali et al. [54] also observed that MB dye was degraded by nanocomposites (NCs) under UV irradiation over a short exposure time.
Figure 11A depicts the variations in (Ct/C0) as a function of exposure time for pure SnO2 NPs and SnO2-rGO NCs at different time intervals (0–50 min) under UV irradiation. The photocatalytic activity of the nanocomposites was enhanced by rGO integration [55]. The results show that the degradation efficiency of the SnO2-rGO NCs (93%) under UV light after 50 min was higher than for the SnO2 NPs (54%). The reaction kinetics of the current catalysis can be described as pseudo-first-order kinetics, i.e., ln(Ct/C0) = −kt (k = kinetic rate constant, t = reaction time, C0 = initial absorbance, and Ct = absorbance at time t). The rate constant (K) values of the SnO2 NPs and SnO2-rGO NCs were 0.0231 and 0.0102 min−1, respectively (Figure 11B). Figure 11D displays the recyclability of the SnO2-rGO NCs after a four-time run with the same experimental conditions. The catalysts were washed in ethanol and deionized water and centrifuged to test for cycle stability. Then, catalysts were dried at 60 °C for 12 h and used multiple times. Even after four runs, the degradation efficiency of the MB dye was still approximately 93%. This indicates that the photocatalytic performance of the SnO2-rGO NCs can be continuously used without damage during the oxidation of contaminants. Our results indicate that the SnO2-rGO NCs had outstanding stability and potential for application in environmental remediation [56]. A comparison of the present samples’ degradation efficiency of MB dye with previous studies is shown in Table 2.

3.10. Mechanism of Photocatalysis

The photocatalytic mechanism for the degradation of MB dye by SnO2-rGO NCs is shown in Figure 12. The MB dye degradation process occurs via a series of chemical reactions, which generate superoxide and hydroxyl radicals. Prepared samples (SnO2 NPs and SnO2-rGO NCs) and MB dye were exposed under UV radiation ( λ < 400 nm ) . Upon irradiating the SnO2, excited electrons in the valance band (VB) transferred to the conduction band (CB), producing holes in the VB (Equation (3)). The electrons in the CB of the SnO2 NPs could transfer to the rGO due to the SnO2 NPs attached to the rGO surfaces (Equation (4)). The generated electrons (e) and holes (h+) reacted with water (H2O) to produce new free oxygen radicals (Equation (5)). These radicals can be transferred to the hydroxyl radicals (Equations (6) and (7)). Hydroxyl radicals were produced by water molecules and holes (Equation (8)). These free oxygen radicals can decompose MB dye into smaller molecules (Equation (9)). The rGO reduced the rate of electron–hole pair recombination in SnO2-rGO NCs, which further increased the degradation efficiency of the SnO2-rGO NCs.
SnO 2 rGO SnO 2 rGO e + h +
SnO 2 e e trap ( rGO )
e trap rGO + O 2 O 2
O 2 + 2 HO + H + H 2 O 2 + O 2
H 2 O 2 2 HO
h + VB + OH / H 2 O HO + H +
HO + MB dye degraded product

3.11. Anticancer Study

Several recent studies showed that the anticancer potential of metal oxide NPs can be enhanced by integration with graphene derivatives [62,63]. For instance, Ahamed et al. [36] observed that SnO2-ZnO-rGO NCs displayed higher cytotoxicity toward human breast cancer cells (MCF-7) in comparison to pure ZnO NPs to improve their physiochemical properties such as the bandgap energy, which was lower compared to SnO2-ZnO NPs and pure ZnO NPs. In this work, we focused on improving the cytotoxicity of SnO2 NPs against cancer cells by rGO incorporation. The cytotoxicity at different concentrations (5-200 µg/mL) of SnO2 NPs and SnO2-rGO NCs was assessed in human liver cancer (HepG2) and lung cancer (A549) cells (Figure 13A,B). The results indicate that both SnO2 NPs and SnO2-rGO NCs induced dose-dependent cytotoxicity in the two types of cancer cells. Moreover, the SnO2-rGO NCs exerted a higher cytotoxicity than the pure SnO2 NPs, and the anticancer potential of the SnO2 NPs increased with rGO doping. The IC50 values of the SnO2-rGO NCs were almost 1.5-fold higher in both cancer cells than for pure SnO2 NPs (Table 3). These results indicate that rGO integration effectively enhanced the anticancer efficacy of the SnO2 NPs. In addition, these data warrant further research on the anticancer potential of SnO2-rGO NCs in suitable in vivo models.

4. Conclusions

In summary, SnO2 NPs and SnO2-rGO NCs were successfully prepared using a novel one-pot hydrothermal approach. The XRD data showed that the average crystallite size of the SnO2 NPs decreased after rGO doping. HRTEM and SEM results revealed that SnO2 NPs tightly anchored the rGO sheets, and SnO2 and rGO were uniformly distributed in the SnO2-rGO NCs with high-quality lattice fringes without distortion. The XPS and EDX data confirmed the chemical state and elemental composition of the SnO2-rGO NCs. Optical data suggested that the bandgap energy of the SnO2-rGO NCs was slightly lower than that of the pure SnO2 NPs. The intensity of the PL spectra was lower in the SnO2-rGO NCs in comparison to the pure SnO2 NPs. In the photocatalytic test, the MB degradation efficiency of the SnO2-rGO NCs (93%) was higher than for the SnO2 NPs (54%). This can be explained by a lower bandgap energy and lower PL intensity, which reduced the separation of charge carries (i.e., electrons and holes). Moreover, the anticancer efficacy of SnO2-rGO NCs was almost 1.5-fold higher against human liver cancer HepG2 and lung cancer A549 cells in comparison to pure SnO2 NPs. These data suggest a novel approach to enhance the photocatalytic activity and anticancer performance of SnO2 NPs via rGO fusion. This study warrants further research on SnO2-rGO NCs for the photocatalytic degradation of different pollutants and anticancer efficacy in various cancer and normal cells.

Author Contributions

Conceptualization, M.A.; investigation and methodology, Z.M.A., H.A.A., S.A., M.J.A. and M.A.; writing—original draft preparation, Z.M.A. and M.A.; writing—review and editing, Z.M.A., H.A.A., S.A. and M.A. All authors have read and agreed to the published version of the manuscript.

Funding

The authors extend their sincere appreciation to the researchers who supported project number: RSP-2021/129, King Saud University, Riyadh, Saudi Arabia, for funding this research.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bray, F.; Ferlay, J.; Soerjomataram, I.; Siegel, R.L.; Torre, L.A.; Jemal, A. Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2018, 68, 394–424. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Ahamed, M.; Akhtar, M.J.; Khan, M.A.M.; Alhadlaq, H.A. Enhanced anticancer performance of eco-friendly-prepared Mo-ZnO/RGO nanocomposites: Role of oxidative stress and apoptosis. ACS Omega 2022, 7, 7103–7115. [Google Scholar] [CrossRef] [PubMed]
  3. Giofrè, S.V.; Tiecco, M.; Celesti, C.; Patanè, S.; Triolo, C.; Gulino, A.; Spitaleri, L.; Scalese, S.; Scuderi, M.; Iannazzo, D. Eco-friendly 1,3-dipolar cycloaddition reactions on graphene quantum dots in natural deep eutectic solvent. Nanomaterials 2020, 10, 2549. [Google Scholar] [CrossRef] [PubMed]
  4. Huang, L.; Sun, Y.; Mahmud, S.; Liu, H. Biological and environmental applications of silver nanoparticles synthesized using the aqueous extract of ginkgo biloba leaf. J. Inorg. Organomet. Polym. Mater. 2020, 30, 1653–1668. [Google Scholar] [CrossRef]
  5. Liu, Y.; Huang, L.; Mahmud, S.; Liu, H. Gold nanoparticles biosynthesized using ginkgo biloba leaf aqueous extract for the decolorization of azo-dyes and fluorescent detection of Cr(VI). J. Clust. Sci. 2020, 31, 549–560. [Google Scholar] [CrossRef]
  6. Singh, M.P.; Sirohi, P.; Afzal, S. Biosynthesis zinc oxide nanoparticles using senna occidentalis leaf extract and evaluation of their cytotoxic effect on SW480 colon cancer cell line. Res. Sq. 2022, 1–20. [Google Scholar] [CrossRef]
  7. Spitaleri, L.; Nicotra, G.; Zimbone, M.; Contino, A.; Maccarrone, G.; Alberti, A.; Gulino, A. Fast and efficient sun light photocatalytic activity of Au_ZnO core-shell nanoparticles prepared by a one-pot synthesis. ACS Omega 2019, 4, 15061–15066. [Google Scholar] [CrossRef] [Green Version]
  8. Sharma, D.K.; Shukla, S.; Sharma, K.K.; Kumar, V. A Review on ZnO: Fundamental properties and applications. Mater. Today Proc. 2020, 49, 3028–3035. [Google Scholar] [CrossRef]
  9. Liu, X.; Zhai, H.; Wang, P.; Zhang, Q.; Wang, Z.; Liu, Y.; Dai, Y.; Huang, B.; Qin, X.; Zhang, X. Synthesis of a WO 3 photocatalyst with high photocatalytic activity and stability using synergetic internal Fe3+ doping and superficial Pt loading for ethylene degradation under visible-light irradiation. Catal. Sci. Technol. 2019, 9, 652–658. [Google Scholar] [CrossRef]
  10. Zeng, J.; Wu, M.; Lan, S.; Li, J.; Zhang, X.; Liu, J.; Liu, X.; Wei, Z.; Zeng, Y. Facile preparation of biocompatible Ti2O3 nanoparticles for second near-infrared window photothermal therapy. J. Mater. Chem. B 2018, 6, 7889–7897. [Google Scholar] [CrossRef]
  11. Kar, A.; Olszówka, J.; Sain, S.; Sloman, S.R.I.; Montes, O.; Fernández, A.; Pradhan, S.K.; Wheatley, A.E.H. Morphological effects on the photocatalytic properties of SnO2 nanostructures. J. Alloys Compd. 2019, 810, 151718. [Google Scholar] [CrossRef]
  12. Firooz, A.A.; Mahjoub, A.R.; Khodadadi, A.A. Preparation of SnO2 nanoparticles and nanorods by using a hydrothermal method at low temperature. Mater. Lett. 2008, 62, 1789–1792. [Google Scholar] [CrossRef]
  13. Li, Y.; Guo, Y.; Tan, R.; Cui, P.; Li, Y.; Song, W. Synthesis of SnO2 nano-sheets by a template-free hydrothermal method. Mater. Lett. 2009, 63, 2085–2088. [Google Scholar] [CrossRef]
  14. Viet, P.V.; Thi, C.M.; Hieu, L. Van the high photocatalytic activity of SnO2 nanoparticles synthesized by hydrothermal method. J. Nanomater. 2016, 2016, 4231046. [Google Scholar] [CrossRef] [Green Version]
  15. Kim, S.P.; Choi, M.Y.; Choi, H.C. Photocatalytic activity of SnO2 nanoparticles in methylene blue degradation. Mater. Res. Bull. 2016, 74, 85–89. [Google Scholar] [CrossRef]
  16. Shams, M.; Guiney, L.M.; Huang, L.; Ramesh, M.; Yang, X.; Hersam, M.C.; Chowdhury, I. Environmental science nano paper influence of functional groups on the degradation of graphene oxide nanomaterials. Environ. Sci. Nano 2019, 6, 2203. [Google Scholar] [CrossRef]
  17. Shyamala, R.; Gomathi Devi, L. Reduced graphene oxide/SnO2 nanocomposites for the photocatalytic degradation of rhodamine B: Preparation, characterization, photosensitization, vectorial charge transfer mechanism and identification of reaction intermediates. Chem. Phys. Lett. 2020, 748, 137385. [Google Scholar] [CrossRef]
  18. Balsamo, S.A.; Fiorenza, R.; Condorelli, M.; Pecoraro, R.; Brundo, M.V.; Lo Presti, F.; Sciré, S. One-pot synthesis of TiO2-rGO photocatalysts for the degradation of groundwater pollutants. Materials 2021, 14, 5938. [Google Scholar] [CrossRef]
  19. Raslan, A.; Saenz del Burgo, L.; Ciriza, J.; Luis Pedraz, J. Graphene oxide and reduced graphene oxide-based scaffolds in regenerative medicine. Int. J. Pharm. 2020, 580, 119226. [Google Scholar] [CrossRef]
  20. Ahamed, M.; Javed Akhtar, M.; Majeed Khan, M.A.; Alhadlaq, H.A. Facile green synthesis of ZnO-RGO nanocomposites with enhanced anticancer efficacy. Methods 2022, 199, 28–36. [Google Scholar] [CrossRef]
  21. Zhu, J.; Zhu, T.; Zhou, X.; Zhang, Y.; Lou, X.W.; Chen, X.; Zhang, H.; Hng, H.H.; Yan, Q. Facile Synthesis of Metal Oxide/Reduced Graphene Oxide Hybrids with High Lithium Storage Capacity and Stable Cyclability. Nanoscale 2011, 3, 1084–1089. [Google Scholar] [CrossRef] [PubMed]
  22. Zhu, W.K.; Cong, H.P.; Yao, H.B.; Mao, L.B.; Asiri, A.M.; Alamry, K.A.; Marwani, H.M.; Yu, S.H. Bioinspired, ultrastrong, highly biocompatible, and bioactive natural polymer/graphene oxide nanocomposite films. Small 2015, 11, 4298–4302. [Google Scholar] [CrossRef] [PubMed]
  23. Hoon Suh, D.; Park, S.K.; Nakhanivej, P.; Kang, S.W.; Park, H.S. Microwave synthesis of SnO2 nanocrystals decorated on the layer-by-layer reduced graphene oxide for an application into lithium ion battery anode. J. Alloys Compd. 2017, 702, 636–643. [Google Scholar] [CrossRef]
  24. Yousefimehr, F.; Jafarirad, S.; Salehi, R.; Zakerhamidi, M.S. Facile fabricating of RGO and Au/RGO nanocomposites using Brassica oleracea var. gongylodes biomass for non-invasive approach in cancer therapy. Sci. Rep. 2021, 11, 11900. [Google Scholar] [CrossRef] [PubMed]
  25. Khan, M.A.M.; Khan, W.; Ahamed, M.; Alhazaa, A.N. Investigation on the structure and physical properties of Fe3O4/RGO nanocomposites and their photocatalytic application. Mater. Sci. Semicond. Process. 2019, 99, 44–53. [Google Scholar] [CrossRef]
  26. Chen, J.; Wei, D.; Liu, L.; Nai, J.; Liu, Y.; Xiong, Y.; Peng, J.; Mahmud, S.; Liu, H. Green synthesis of konjac glucomannan templated palladium nanoparticles for catalytic reduction of Azo compounds and hexavalent chromium. Mater. Chem. Phys. 2021, 267, 124651. [Google Scholar] [CrossRef]
  27. Bansode, S.R.; Khare, R.T.; Jagtap, K.K.; More, M.A.; Koinkar, P. One step hydrothermal synthesis of SnO2-RGO nanocomposite and its field emission studies. Mater. Sci. Semicond. Process. 2017, 63, 90–96. [Google Scholar] [CrossRef]
  28. Matthew, J. Surface Analysis by Auger and X-Ray Photoelectron Spectroscopy; Briggs, D., Grant, J.T., Eds.; IMPublications: Chichester, UK; SurfaceSpectra: Manchester, UK, 2003; ISBN 1-901019-04-7. [Google Scholar]
  29. Nurzulaikha, R.; Lim, H.N.; Harrison, I.; Lim, S.S.; Pandikumar, A.; Huang, N.M.; Lim, S.P.; Thien, G.S.H.; Yusoff, N.; Ibrahim, I. Graphene/SnO2 nanocomposite-modified electrode for electrochemical detection of dopamine. Sens. Bio-Sens. Res. 2015, 5, 42–49. [Google Scholar] [CrossRef] [Green Version]
  30. Pham, V.T.; le Trung, H.; Tran, N.K.; Chu Manh, H.; Nguyen Duc, H.; Tran Thi Quynh, H.; Pham, T.H. Hydrothermal synthesis, structure, and photocatalytic properties of SnO2/RGO nanocomposites with different GO concentrations. Mater. Res. Express 2018, 5, 095506. [Google Scholar] [CrossRef]
  31. Priyadharsan, A.; Vasanthakumar, V.; Karthikeyan, S.; Raj, V.; Shanavas, S.; Anbarasan, P.M. Multi-functional properties of ternary CeO2/SnO2/RGO nanocomposites: Visible light driven photocatalyst and heavy metal removal. J. Photochem. Photobiol. A Chem. 2017, 346, 32–45. [Google Scholar] [CrossRef]
  32. Navazani, S.; Shokuhfar, A.; Hassanisadi, M.; Askarieh, M.; Di Carlo, A.; Agresti, A. Facile synthesis of a SnO2@rGO nanohybrid and optimization of its methane-sensing parameters. Talanta 2018, 181, 422–430. [Google Scholar] [CrossRef] [PubMed]
  33. Drzymała, E.; Gruzeł, G.; Depciuch, J.; Budziak, A.; Kowal, A.; Parlinska-Wojtan, M. Structural, chemical and optical properties of SnO2 NPs obtained by three different synthesis routes. J. Phys. Chem. Solids 2017, 107, 100–107. [Google Scholar] [CrossRef]
  34. Ahmad, S.; Ahmad, A.; Khan, S.; Ahmad, S.; Khan, I.; Zada, S.; Fu, P. Algal extracts based biogenic synthesis of reduced graphene oxides (RGO) with enhanced heavy metals adsorption capability. J. Ind. Eng. Chem. 2019, 72, 117–124. [Google Scholar] [CrossRef]
  35. Zhang, B.; Zhang, C.; Liu, B.; Zhou, X.; Huang, G. Anchoring SbxOy/SnO2 nano-heterojunction on reduced graphene oxide as lithium ion batteries anodes with remarkable rate performance and excellent cycle stability. Ionics 2021, 27, 4205–4216. [Google Scholar] [CrossRef]
  36. Ahamed, M.; Akhtar, M.J.; Khan, M.A.M.; Alhadlaq, H.A. SnO2-doped ZnO/reduced graphene oxide nanocomposites: Synthesis, characterization, and improved anticancer activity via oxidative stress pathway. Int. J. Nanomed. 2021, 16, 89–104. [Google Scholar] [CrossRef]
  37. Gulino, A. Structural and electronic characterization of self-assembled molecular nanoarchitectures by X-ray photoelectron spectroscopy. Anal. Bioanal. Chem. 2013, 405, 1479–1495. [Google Scholar] [CrossRef]
  38. Tian, R.; Zhang, Y.; Chen, Z.; Duan, H.; Xu, B.; Guo, Y.; Kang, H.; Li, H.; Liu, H. The effect of annealing on a 3D SnO2/graphene foam as an advanced lithium-ion battery anode. Sci. Rep. 2016, 6, 19195. [Google Scholar] [CrossRef] [Green Version]
  39. Pi, S.; Zhang, X.; Cui, H.; Chen, D.; Zhang, G.; Xiao, S.; Tang, J. Facile fabrication of Au nanoparticles/Tin oxide/reduced graphene oxide ternary nanocomposite and its high-performance SF6 decomposition components sensing. Front. Chem. 2019, 7, 476. [Google Scholar] [CrossRef] [Green Version]
  40. Al-Gaashani, R.; Najjar, A.; Zakaria, Y.; Mansour, S.; Atieh, M.A. XPS and structural studies of high quality graphene oxide and reduced graphene oxide prepared by different chemical oxidation methods. Ceram. Int. 2019, 45, 14439–14448. [Google Scholar] [CrossRef]
  41. Jiang, S.; Huang, R.; Zhu, W.; Li, X.; Zhao, Y.; Gao, Z.; Gao, L.; Zhao, J. Free-standing SnO2@rGO anode via the anti-solvent-assisted precipitation for superior lithium storage performance. Front. Chem. 2019, 7, 878. [Google Scholar] [CrossRef]
  42. Bhangare, B.; Ramgir, N.S.; Jagtap, S.; Debnath, A.K.; Muthe, K.P.; Terashima, C.; Aswal, D.K.; Gosavi, S.W.; Fujishima, A. XPS and kelvin probe studies of SnO2/RGO nanohybrids based NO2 sensors. Appl. Surf. Sci. 2019, 487, 918–929. [Google Scholar] [CrossRef]
  43. Wang, X.; Wang, X.; Di, Q.; Zhao, H.; Liang, B.; Yang, J. Mutual effects of fluorine dopant and oxygen vacancies on structural and luminescence characteristics of F doped SnO2 nanoparticles. Materials 2017, 10, 1398. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Rahimi, A.; Kazeminezhad, I.; Mousavi Ghahfarokhi, S.E. Synthesis and investigation of SnS2/RGO nanocomposites with different GO concentrations: Structure and optical properties, photocatalytic performance. J. Mater. Sci. Mater. Electron. 2018, 29, 4449–4456. [Google Scholar] [CrossRef]
  45. Du, F.; Yang, B.; Zuo, X.; Li, G. Dye-sensitized solar cells based on low-cost nanoscale SnO2@RGO composite counter electrode. Mater. Lett. 2015, 158, 424–427. [Google Scholar] [CrossRef]
  46. Akada, K.; Terasawa, T.O.; Imamura, G.; Obata, S.; Saiki, K. Control of work function of graphene by plasma assisted nitrogen doping. Appl. Phys. Lett. 2014, 104, 131602. [Google Scholar] [CrossRef]
  47. Khan, M.A.M.; Khan, W.; Ahamed, M.; Ahmed, J.; Al-Gawati, M.A.; Alhazaa, A.N. Silver-decorated cobalt ferrite nanoparticles anchored onto the graphene sheets as electrode materials for electrochemical and photocatalytic applications. ACS Omega 2020, 5, 31076–31084. [Google Scholar] [CrossRef] [PubMed]
  48. Wang, Y.; Ding, J.; Liu, Y.; Liu, Y.; Cai, Q.; Zhang, J. SnO2@reduced graphene oxide composite for high performance lithium-ion battery. Ceram. Int. 2015, 41, 15145–15152. [Google Scholar] [CrossRef]
  49. Guo, W.; Zhou, Q.; Zhang, J.; Fu, M.; Radacsi, N.; Li, Y. Hydrothermal synthesis of Bi-Doped SnO2/RGO nanocomposites and the enhanced gas sensing performance to benzene. Sens. Actuators B Chem. 2019, 299, 126959. [Google Scholar] [CrossRef]
  50. Orek, C.; Gündüz, B.; Kaygili, O.; Bulut, N. Electronic, optical, and spectroscopic analysis of TBADN organic semiconductor: Experiment and theory. Chem. Phys. Lett. 2017, 678, 130–138. [Google Scholar] [CrossRef]
  51. Mani, R.; Vivekanandan, K.; Subiramaniyam, N.P. Photocatalytic activity of different organic dyes by using pure and fe doped SnO2 nanopowders catalyst under UV light irradiation. J. Mater. Sci. Mater. Electron. 2017, 28, 13846–13852. [Google Scholar] [CrossRef]
  52. Van Tuan, P.; Tuong, H.B.; Tan, V.T.; Thu, L.H.; Khoang, N.D.; Khiem, T.N. SnO2/reduced graphene oxide nanocomposites for highly efficient photocatalytic degradation of methylene blue. Opt. Mater. 2022, 123, 111916. [Google Scholar] [CrossRef]
  53. Sujatmiko, F.; Sahroni, I.; Fadillah, G.; Fatimah, I. Visible light-responsive photocatalyst of SnO2/RGO prepared using pometia pinnata leaf extract. Open Chem. 2021, 19, 174–183. [Google Scholar] [CrossRef]
  54. Ali, G.; Zaidi, S.J.A.; Abdul Basit, M.; Park, T.J. Synergetic performance of systematically designed G-C3N4/RGO/SnO2 nanocomposite for photodegradation of rhodamine-B dye. Appl. Surf. Sci. 2021, 570, 151140. [Google Scholar] [CrossRef]
  55. Gomari, N.; Kazeminezhad, I.; Ghahfarokhi, S.E.M. Impact of morphology evolution of ZnSn(OH)6 microcubes on photocatalytic activity of ZnSn(OH)6/SnO2/RGO ternary nanocomposites for efficient degradation of organic pollutants. Opt. Mater. 2021, 113, 110878. [Google Scholar] [CrossRef]
  56. Nenavathu, B.P.; Kandula, S.; Verma, S. Visible-light-driven photocatalytic degradation of safranin-T dye using functionalized graphene oxide nanosheet (FGS)/ZnO nanocomposites. RSC Adv. 2018, 8, 19659–19667. [Google Scholar] [CrossRef] [Green Version]
  57. Al-Rawashdeh, N.A.F.; Allabadi, O.; Aljarrah, M.T. Photocatalytic activity of graphene oxide/zinc oxide nanocomposites with embedded metal nanoparticles for the degradation of organic dyes. ACS Omega 2020, 5, 28046–28055. [Google Scholar] [CrossRef]
  58. Tian, H.; Wan, C.; Xue, X.; Hu, X.; Wang, X. Effective electron transfer pathway of the ternary TiO2/RGO/Ag nanocomposite with enhanced photocatalytic activity under visible light. Catalysts 2017, 7. [Google Scholar] [CrossRef]
  59. Asgharian, M.; Mehdipourghazi, M.; Khoshandam, B.; Keramati, N. Photocatalytic degradation of methylene blue with synthesized RGO/ZnO/Cu. Chem. Phys. Lett. 2019, 719, 1–7. [Google Scholar] [CrossRef]
  60. Shih, K.-Y.; Kuan, Y.-L.; Wang, E.-R. One-step microwave-assisted synthesis and visible-light photocatalytic activity enhancement of BiOBr/RGO nanocomposites for degradation of methylene blue. Materials 2021, 14, 4577. [Google Scholar] [CrossRef]
  61. Li, B.; Shao, X.; Liu, T.; Shao, L.; Zhang, B. Construction of metal/WO2.72/RGO ternary nanocomposites with optimized adsorption, photocatalytic and photoelectrochemical properties. Appl. Catal. B Environ. 2016, 198, 325–333. [Google Scholar] [CrossRef]
  62. Tammina, S.K.; Mandal, B.K.; Ranjan, S.; Dasgupta, N. Cytotoxicity study of piper nigrum seed mediated synthesized SnO2 nanoparticles towards colorectal (HCT116) and lung cancer (A549) cell lines. J. Photochem. Photobiol. B Biol. 2017, 166, 158–168. [Google Scholar] [CrossRef] [PubMed]
  63. Ahamed, M.; Akhtar, M.J.; Khan, M.A.M.; Alaizeri, Z.A.M.; Alhadlaq, H.A. Evaluation of the cytotoxicity and oxidative stress response of CeO2-RGO nanocomposites in human lung epithelial A549 cells. Nanomaterials 2019, 9, 1709. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Scheme 1. Synthesis protocol of the SnO2-rGO NCs.
Scheme 1. Synthesis protocol of the SnO2-rGO NCs.
Polymers 14 02036 sch001
Figure 1. XRD spectra of GO, rGO, SnO2 NPs, and SnO2-rGO NCs.
Figure 1. XRD spectra of GO, rGO, SnO2 NPs, and SnO2-rGO NCs.
Polymers 14 02036 g001
Figure 2. TEM characterization: TEM and HRTEM images of GO (AC); rGO (DF); SnO2 NPs (GI); SnO2-rGO NCs (LN).
Figure 2. TEM characterization: TEM and HRTEM images of GO (AC); rGO (DF); SnO2 NPs (GI); SnO2-rGO NCs (LN).
Polymers 14 02036 g002
Figure 3. FE-SEM images of GO (A); rGO (B); SnO2 NPs (C); SnO2-rGO NCs (D).
Figure 3. FE-SEM images of GO (A); rGO (B); SnO2 NPs (C); SnO2-rGO NCs (D).
Polymers 14 02036 g003
Figure 4. The elemental composition of the SnO2-rGO NCs by EDX (A). SEM elemental mapping of the SnO2-rGO NCs: (B) tin (Sn); (C) oxygen (O); (D); carbon (C) (E).
Figure 4. The elemental composition of the SnO2-rGO NCs by EDX (A). SEM elemental mapping of the SnO2-rGO NCs: (B) tin (Sn); (C) oxygen (O); (D); carbon (C) (E).
Polymers 14 02036 g004
Figure 5. XPS wide spectra of SnO2-rGO NCs (A); Sn3d high-resolution XPS spectra (B); O1s high-resolution XPS spectra (C); C1s high-resolution XPS spectra (D).
Figure 5. XPS wide spectra of SnO2-rGO NCs (A); Sn3d high-resolution XPS spectra (B); O1s high-resolution XPS spectra (C); C1s high-resolution XPS spectra (D).
Polymers 14 02036 g005
Figure 6. Raman spectra of the SnO2 NPs and SnO2-rGO NCs (A); magnified Raman spectra of the SnO2 NPs and SnO2-rGO NCs (B).
Figure 6. Raman spectra of the SnO2 NPs and SnO2-rGO NCs (A); magnified Raman spectra of the SnO2 NPs and SnO2-rGO NCs (B).
Polymers 14 02036 g006
Figure 7. FTIR spectra of the prepared GO, rGO, SnO2 NPs, and SnO2-rGO NCs.
Figure 7. FTIR spectra of the prepared GO, rGO, SnO2 NPs, and SnO2-rGO NCs.
Polymers 14 02036 g007
Figure 8. UV–Vis spectra of the SnO2NPs and SnO2-rGO NCs (A); the bandgap energy of the SnO2 NPs and SnO2-rGO NCs (B).
Figure 8. UV–Vis spectra of the SnO2NPs and SnO2-rGO NCs (A); the bandgap energy of the SnO2 NPs and SnO2-rGO NCs (B).
Polymers 14 02036 g008
Figure 9. PL spectra of the SnO2NPs and SnO2-rGO NCs.
Figure 9. PL spectra of the SnO2NPs and SnO2-rGO NCs.
Polymers 14 02036 g009
Figure 10. UV–Vis absorbance spectra for the degradation of the MB dye at different time intervals under UV irradiation: SnO2 NPs (A); SnO2-rGO NCs (B).
Figure 10. UV–Vis absorbance spectra for the degradation of the MB dye at different time intervals under UV irradiation: SnO2 NPs (A); SnO2-rGO NCs (B).
Polymers 14 02036 g010
Figure 11. Variations in (Ct/C0) as a function of irradiation time for SnO2 NPs and SnO2-rGO NCs (A), The plot of ln(C0/Ct) as a function of light exposure time (B), the degradation efficiency of the SnO2 NPs and SnO2-rGO NCs (C), and four cycling runs of the SnO2-rGO NCs for the degradation of the MB dye under UV irradiation (D).
Figure 11. Variations in (Ct/C0) as a function of irradiation time for SnO2 NPs and SnO2-rGO NCs (A), The plot of ln(C0/Ct) as a function of light exposure time (B), the degradation efficiency of the SnO2 NPs and SnO2-rGO NCs (C), and four cycling runs of the SnO2-rGO NCs for the degradation of the MB dye under UV irradiation (D).
Polymers 14 02036 g011
Figure 12. Photocatalytic mechanism for the degradation of MB dye by the SnO2-rGO NCs.
Figure 12. Photocatalytic mechanism for the degradation of MB dye by the SnO2-rGO NCs.
Polymers 14 02036 g012
Figure 13. Cytotoxicity of rGO, SnO2 NPs, and SnO2-rGO NCs in human liver cancer HepG2 cells (A) and lung cancer A549 cells (B); one-way ANOVA followed by the post hoc test was applied for statistical analysis. * p < 0.05, statistically significant difference when compared to the control group.
Figure 13. Cytotoxicity of rGO, SnO2 NPs, and SnO2-rGO NCs in human liver cancer HepG2 cells (A) and lung cancer A549 cells (B); one-way ANOVA followed by the post hoc test was applied for statistical analysis. * p < 0.05, statistically significant difference when compared to the control group.
Polymers 14 02036 g013
Table 1. Structural and Optical Properties of the SnO2 NPs and SnO2-rGO NCs.
Table 1. Structural and Optical Properties of the SnO2 NPs and SnO2-rGO NCs.
PropertiesSnO2 NPsSnO2-rGO NCs
XRD size (nm)7.647.98
TEM size (nm)9.429.84
Optical bandgap (eV)3.793.45
Table 2. Comparison of the MB dye degradation efficiency of the present samples with previous studies.
Table 2. Comparison of the MB dye degradation efficiency of the present samples with previous studies.
SampleModel Dye PollutantReaction TimeDegradation Efficiency (%)Concentration
of MB
Source of LightReference
SnO2-rGO NCsMethylene blue (MB)50 min93%10 ppmUV irradiationThis work
ZnO/GONCsMethylene blue (MB)40 min100%10 ppmSunlight irradiation[57]
TiO2/rGO/Ag NCsMethylene blue (MB)120 min91.2%10 ppmUV irradiation[58]
rGO/ZnO/CuMethylene blue (MB)60 min95.14%5 ppmSunlight irradiation[59]
BiOBr/rGO NCsMethylene blue (MB)75 min96%7 ppmSunlight irradiation[60]
Au/WO2.72/
rGONCs
Methylene blue (MB)60 min91.2%10 ppmSunlight irradiation[61]
Table 3. IC50 values of SnO2 NPs and SnO2-rGO NCs for two types of human cancer cells.
Table 3. IC50 values of SnO2 NPs and SnO2-rGO NCs for two types of human cancer cells.
SamplesHuman Liver Cancer (HepG2) CellsHuman Lung Cancer (A549) Cells
SnO2 NPs160.97 μg/mL153.13 μg/mL
SnO2-rGO NCs100 μg/mL95.72 μg/mL
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Alaizeri, Z.M.; Alhadlaq, H.A.; Aldawood, S.; Akhtar, M.J.; Ahamed, M. One-Pot Synthesis of SnO2-rGO Nanocomposite for Enhanced Photocatalytic and Anticancer Activity. Polymers 2022, 14, 2036. https://doi.org/10.3390/polym14102036

AMA Style

Alaizeri ZM, Alhadlaq HA, Aldawood S, Akhtar MJ, Ahamed M. One-Pot Synthesis of SnO2-rGO Nanocomposite for Enhanced Photocatalytic and Anticancer Activity. Polymers. 2022; 14(10):2036. https://doi.org/10.3390/polym14102036

Chicago/Turabian Style

Alaizeri, ZabnAllah M., Hisham A. Alhadlaq, Saad Aldawood, Mohd Javed Akhtar, and Maqusood Ahamed. 2022. "One-Pot Synthesis of SnO2-rGO Nanocomposite for Enhanced Photocatalytic and Anticancer Activity" Polymers 14, no. 10: 2036. https://doi.org/10.3390/polym14102036

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop