Brought to you by:
Paper

SnO2 nanorods/graphene nanoplatelets nanocomposites: towards fast removal of malachite green and pathogen control

, , , and

Published 22 December 2021 © 2021 IOP Publishing Ltd
, , Focus on nanomedicine Citation Aqsa Arshad et al 2022 Nanotechnology 33 115101 DOI 10.1088/1361-6528/abfdef

0957-4484/33/11/115101

Abstract

The world is facing alarming challenges of environmental pollution due to uncontrolled water contamination and multiple drug resistance of pathogens. However, these challenges can be addressed by using novel nanocomposites materials such as, SnO2/graphene nanopaletelets (GNPs) nanocomposites remarkably. In this work, we have prepared SnO2 nanorods and SnO2/GNPs nanocomposites (GS-I and GS-II) with size of 25 ± 6 nm in length and 4 ± 2 nm in diameter. The optical bandgap energies change from 3.14 eV to 2.80 eV in SnO2 and SnO2/GNPs nanocomposite. We found that SnO2/GNPs nanocomposite (GS-II) completely removes (99.11%) malachite green in 12 min, under UV light exposure, while under same conditions, SnO2 nanorods removes only 37% dye. Moreover, visible light exposure resulted in 99.01% removal of malachite green in 15 min by GSII as compared to 24.7% removal by SnO2. In addition, GS-II nanocomposite inhibits 79.57% and 78.51% growth of P. aeruginosa and S. aureus respectively. A synchronized contribution of SnO2 and GNPs makes SnO2/GNPs nanocomposites (GS-II) an innovative multifunctional material for simultaneous fast and complete removal of malachite green and inhibition of drug resistant pathogens.

Export citation and abstract BibTeX RIS

1. Introduction

The negative biological impact of environmental pollution, particularly contaminated water is a great challnge. Chemical industries release dangerous dyes into the water effluents which contributes towards water contamination, thus badly affecting the life in every form [1]. It is hard to eliminate the dyes, as they don't degrade naturally. Photocatalysis is an effective method of wastewater remediation among the existing water treatment methods [2]. Carbon based biocompatible materials [3] have been excessively used for decontamination of water, nevertheless, the need of more efficient water cleaning solutions urges the researchers to find novel and cost-effective materials with unique features like re-usability and high efficiency.

Disease is another threat to humanity. Covid-19 pandemic has made it clear that the health sector needs investment in terms of time, money and research [4, 5]. Multi-drug resistance of microbes demands the researchers to encounter them by developing novel antibiotics. These drug-resistant pathogens can cause multiple nosocomial infections, thus affecting public health. The seriousness of this issue can be assessed by an alarming statistics of total affected number of patients. CDC reported that 2 million people worldwide are affected by the pathogens. Consequently, the death rate is about 23,000 persons per annuum. Also, another notable issue is the treatment of diseases associated with pathogens, which is beyond the reach of a common man [6]. Therefore, synthesis and testing of novel nanomedicine are the areas that demand urgent attention of research community.

Combining metal oxides and graphene can serve as an effective strategy for a photocatalytic reaction and pathogen control. Based on their potential to be modified by doping and compositing, metal oxides have numerous applications in all the branches of sciences and technology, e.g. in sensors, fuel cells, piezoelectric devices, photovoltaic cells, nanoelectronics and photocatalysis [7, 8]. Among metal oxides, tin (IV) oxide (SnO2), an n-type inorganic semiconductor, has gained great attention due to its remarkable properties. SnO2 has admirable chemical and optical properties based on its wide band gap (ca. 3.6 eV). It finds many interesting applications which include transistors, solar cells, organic light emitting diodes, display screens, rechargeable batteries, optical coating and gas sensing devices etc [9]. SnO2 can be used as a photocatalyst for the degradation of toxic chemical dyes, phenols, and pesticides. It's remarkable chemical stability, high photosensitivity, low cost, and wide energy bandgap makes it suitable for photocatalysis. Moreover, it is not harmful to human health. It can also promote light induced generation of electron-hole pairs, which is vital for a photocatalytic reaction [10, 11]. Making its nanocomposites with graphene family materials can possibly provide the solution to address the issues that are faced during oxidative degradation of chemical dyes and growth inhibition pathogens like P. aeruginosa and S. aureus [6].

There are some basic issues that need to be addressed before the use of a material in applications e.g. catalysis and nanomedicine. A high electron mobility and oxidation capability of a nanomaterial is crucial to determine its efficiency as photocatalyst and nanomedicine. Electrostatically coupled semiconductors (that don't detach easily from each other) allow photogenerated charge carriers to transport between the constituent species. Consequently, the photogenerated electrons (traveling across the interfaces) and valence band (VB) holes, both can generate reactive oxygen species (ROS). These ROS can be detrimental for bacteria and are also responsible to degrade the dye. Based on this idea, ultra-small SnO2 nanostructures anchored on graphene nanoplatelets (GNPs) can give a high surface area material, with provision of plenty of nano-channels that stop or delay the electron-hole recombination process. GNPs being a member of graphene wonder family can introduce various desirable features in nanocomposites, when combined with different metals/non-metal oxides [1216]. Electron acceptor GNPs provide electronic pathways, in this case. Graphene family materials have also been used effectively in nanomedicine, as reported previously. Keeping these factors in view, herein, we report the photocatalytic and bio-medical applications of SnO2/GNPs nanocomposites.

2. Experimental

2.1. Synthesis of SnO2/GNPs nanocomposites

Graphene nanoplatelets (Knano, 99.9%), tin (IV) chloride pentahydrate (sigma-Aldrich, 98%), urea (sigma-Aldrich, 99%), hydrochloric acid (HCl) and distilled water were used as received. Solvothermal method was used to prepare SnO2 nanorods. First, 1.2 g metal precursor and 12 g urea was dissolved in distilled water. Few drops of HCl were added to 400 ml precursor solution to obtain pH 4, under constant stirring for 15 min. This solution was transferred to a Teflon lined autoclave, which was then placed in an electric oven for heat treatment at 90 °C for 24 h. Two nanocomposites were prepared by selecting two different quantities of GNPs (12 wt% and 25 wt%). Two graphene solutions were prepared in mixed solvents. Their sonicated dispersions were introduced to SnO2 solution. The process was carried out under continuous stirring. This solution was transferred to autoclave and was given a heat treatment at 90 °C for 24 h. The precipitates were obtained via washing, several times in ethanol and water, followed by overnight drying. A schematic presentation of the process is depicted in ESI figure 4(a) (available online at stacks.iop.org/NANO/33/115101/mmedia).

2.2. Photocatalytic activities

A photocatalytic chamber equipped with UV light source (90 W, Type C and emission peak = 254 nm) was used to investigate the effect of UV light irradiation. Malachite green (C23H25N2) dye was used to investigate the photocatalytic activity of prepared nanomaterials. A 20 ppm solution of malachite green was prepared from 500 ppm stock solution. 0.028 g sample was added to the solution, with a drop of HCl. The beaker was well covered with aluminum foil and was stirred for 45 min in the dark to achieve adsorption-desorption equilibrium, and 4 ml of solution was collected at t = 0 min. This solution was again stirred in the UV light equipped photocatalytic chamber. The test-solutions were collected in a regular manner. Absorbance profiles were then obtained to monitor the changes. To study the impact of visible light, a same process was repeated using same experimental conditions except a visible light source (200 W) was used for irradiation.

2.3. Antibacterial activity of nanocomposites

P. aeruginosa and S. aureus were used as model bacteria to evaluate the antibacterial activity of prepared samples. The protocol is briefly described here. The isotonic solution was prepared without sample which was considered as control. Under incubator conditions, SnO2 and SnO2/GNPs nanocomposites (GS-I and GS-II) were used in the same concentration (10 mg ml−1) for 24 h, to test the antibacterial activity. 200 μl of prepared samples was added to the 5 ml of Luria Bertani (LB). The 100 μl bacterium culture with medium and samples was incubated for 24 h at 37 °C. The optical density at 600 nm was recorded for 24 h. The protocol is described in detail in one of our previous work [17].

3. Characterizations

X-ray diffraction spectra were obtained by a PANalytical X'Pert PRO diffractometer, using Cu Kα radiation (wavelength λ = 0.15406 nm) via powder diffraction method. The step size was 0.03°. The size, morphology, microstructure, and crystallinity of SnO2 and SnO2/GNPs nanocomposites (GS-I and GS-II) were investigated using a transmission electron microscope (TEM), high resolution transmission electron microscope (HRTEM) and selected area electron diffraction (SAED) (JOEL 2100 F). The powdered samples were dispersed in ethanol by ultrasonication for 10 min. As a result, nearly transparent solutions were obtained. KBr pellets method was employed using FTIR spectrometer (IR Tracer-100 Shimadzu). Raman spectra were obtained using Raman spectrometer (Ramboss). Photocatalytic and optical properties were studied using a PerkinElmer spectrometer (Lambda 25 UV). To study the optical bandgap energy a transparent aqueous solution of each powder sample was prepared. The powder was dispersed in the distilled water by sonication for 15 min. The step size was 2 nm.

4. Results and discussions

4.1. Phase analysis

The crystal behavior of the prepared samples is shown in figure 1. The data is obtained from 20° to 90°. The SnO2 shows peaks located at 26.6° (110), 32.6° (020), 38.8° (101), 51.5° (211), 58.2° (002) and 64.7° (112). The result agrees well with a previous work on SnO2 [18], and also matches with the standard diffraction peaks data (JCPDS No. 00-001-0625). The findings show that SnO2 possesses tetragonal crystalline structure with lattice parameters, a = b = 4.7 Å and c = 3.17 Å. The diffraction peaks of SnO2/GNPs nanocomposites found at 26.6°, 32.6°, 38.8°, 51.5°, 54.6° 58.2°, 61.8°, 64.7° and 73.1° are assigned to (110), (020), (101), (211), (220), (002), (310), (112) and (202) planes, respectively [19, 20]. Carbon registers its presence by showing a peak at 26.6° (002), which confirms the presence of graphene in nanocomposites. The crystallite size of SnO2 in GS-I and GS-II nanocomposites is 3.98 nm and 5.78 nm respectively (obtained via Scherrer's formula). It is observed that the presence of GNPs has affected the growth of SnO2 in certain crystallographic directions (figure 1) in GS-I and GS-II nanocomposites, e.g. growth of (101) plane has been supressed in presence of greater content of GNPs in GS(II) nanocomposite.

Figure 1.

Figure 1. XRD analysis of SnO2 and SnO2/GNPs nanocomposites.

Standard image High-resolution image

4.2. Morphology

The size, morphology, and microstructure of SnO2 nanorods and SnO2/GNPs nanocomposites have been investigated using TEM. The results are shown in figures 2(a)–(c). Figure 2(a) illustrates that SnO2 possesses nanorods like morphology. The length and diameter of nanorods are ca. 25 ± 6 nm and 4 ± 2 nm, respectively. Particle size tunes the electronic and optical properties. Particle size also determines the surface area which strongly influences photocatalytic properties. Figure 2(b) demonstrates that SnO2 nanorods are attached on graphene sheets. Figure 2(c) depicts the detailed view of GS-II nanocomposite. The TEM images show that graphene sheets have well dispersed nanorods. This combined assembly offers high surface area and interfacial contacts, which could possibly play a significant role in photocatalysis and ROS generation.

Figure 2.

Figure 2. (a) The TEM image of SnO2 nanorods, (b) SnO2/GNPs nanocomposite (GS-II) and (c) a detailed image of SnO2/GNPs nanocomposite (GS-II), the scale is 20 nm.

Standard image High-resolution image

4.3. HR-TEM and SAED

The HRTEM and SAED images are shown in figure 3. These images verify the XRD findings, i.e. SnO2 and SnO2/GNPs nanocomposites have crystalline nature. Figure 3(a) indicates the spacing between the atomic planes of SnO2 nanorods. It is 0.288 nm associated with (101) plane of SnO2. Moreover, figure 3(b) illustrates the spacing between the lattice plane of SnO2/GNPs nanocomposites. It is found that (101), (110) planes of SnO2 and (002) plane of carbon are observed with lattice spacing 0.269 nm, 0.326 nm and 0.346, respectively. We performed the selected area electron diffraction (SAED) experiments to further probe the crystallinity of constituent materials and hybrid nature of SnO2/GNPs (GS-II) nanocomposite. SAED pattern of SnO2 is presented in figure 3(c). The figure 3(c) shows that the SnO2 shows (110), (101), (211) and (002) planes which verify XRD results given in figure 1.

Figure 3.

Figure 3. (a) The HRTEM image of SnO2, (b) the HRTEM image of SnO2/GNPs nanocomposites, (c) the SAED image of SnO2 and (d) indicates the SAED pattern of SnO2/GNPs (GS-II) nanocomposite.

Standard image High-resolution image

The figure 3(d) also confirms the presence of SnO2 and carbon in GS-II sample. The presence of carbon (002) plane establishes the inclusion of graphene in the nanocomposite. These findings confirm the biphasic nature of SnO2/GNPs nanocomposites.

4.4. FTIR spectroscopy and Raman spectroscopy

The prepared samples were further analyzed using FTIR spectroscopy to investigate the chemical bonds formation. The successful formation of SnO2 is presented by a band at ca. 667 cm−1, which is assigned to vibrations associated with O–Sn–O bond [20, 21]. The bands around 1500 and 1620 cm−1 indicate the presence of adsorbed water molecules in SnO2 and SnO2/GNPs nanocomposites. The atmospheric CO2 adsorbed on the sample's surface show a band around 2375 cm−1. Another band centered around 3400 cm−1 is due to –OH bond. No signatures of C–N and H–N bonds are observed in all the samples, which eliminate the possibility of carbon nitride (gC3N4) formation [22].

The Raman spectra shown in figure 4(b) indicates the D-band and G-band of neat GNPs that appear at 1350 cm−1 and 1587 cm−1 respectively. These two bands are fingerprint signatures of graphene-based systems. The D-band shows increased intensity for SnO2/GNPs nanocomposites as compared to neat GNPs, which is due to incorporation of SnO2 on the surface of GNPs. The G-band shows a shift to higher values i.e., 1601 cm−1 and 1607 cm−1 in GS-I and GS-II nanocomposites, respectively which is due to charge transfer between SnO2 and GNPs. It indicates the successful formation of nanocomposites.

Figure 4.

Figure 4. (a) FTIR spectra of SnO2 and SnO2/GNPs nanocomposites, (b) Raman spectra of GNPs and SnO2/GNPs nanocomposites.

Standard image High-resolution image

4.5. Optical properties analysis

The optical properties of SnO2 and SnO2/GNPs nanocomposites are investigated to find their suitability for photocatalytic applications. In figure 5(a), the absorbance of prepared nanomaterial has been obtained for wavelengths from 300 to 800 nm.

Figure 5.

Figure 5. (a) Absorbance profiles of SnO2 and SnO2/GNPs nanocomposites, and (b) the corresponding energy band gap diagram (Tauc's plot).

Standard image High-resolution image

The band gap energies are found using Tauc's relation, which is, ${(\alpha h\nu )}^{n}=A(h\nu -{E}_{g}).$ Here, hν, A and α represent the photon energy, constant and absorbance coefficient, respectively. The value of n is determined by indirect bandgap transition and (or) direct bandgap transition.

The optical bandgap energies, calculated using Tauc's equation, are shown in figure 5(b). The bandgap energies change from 3.14 to 2.80 eV in SnO2 and SnO2/GNPs nanocomposite (GS-I and GS-II). The band gap energies are reduced with inclusion of graphene. Since, the defects are present in the localized regions and new energy levels may be created between the valance band and conduction band due to the chemical reaction of SnO2 and GNPs in nanocomposites samples, therefore, a reduction in bandgap energy is observed. This reduction in bandgap energy indicates the potential of the prepared materials towards photocatalysis [23].

4.6. Photocatalysis

The photocatalytic activity of SnO2 and SnO2/GNPs nanocomposites has been investigated by degradation of malachite green under UV light and visible light. The variant concentration of malachite green under UV light and visible light has been monitored by obtaining absorbance spectra, with maximum absorption observed at 618 nm. The absorbance profiles of aqueous malachite green solution in presence of SnO2 and SnO2/GNPs nanocomposites, under UV light irradiation, is depicted in ESI figures 2(a)–(c). Only a slight fraction of dye is decomposed by SnO2 nanorods, after a 12 min exposure to UV-light. The absorbance profiles in ESI figures 2(b), and (c) show that GS-I degrades the dye to some extent, in same time, whereas, it has been removed completely, using GS-II nanocomposite. The reaction kinetics of the process are explained in figures 6(a), and (b). Langmuir-Hinshelwood model is used to obtain pseudo-first order rate kinetics. It has been obtained that GS-II has the highest rate constant, k = 0.385 min−1, which is an order higher than that of neat SnO2 nanorods, i.e. k = 0.038 min−1. GS-I possesses an intermediate rate constant k = 0.117 min−1. The photodegradation efficiency of malachite green has been improved from 37.2% (for SnO2) to 99.11% (for GS-II), whereas GS-I shows only 78.26% photodegradation of malachite green. This is attributed to optimized GNPs content in the nanocomposite.

Figure 6.

Figure 6. (a) Concentration variation of malachite green upon exposure to UV light, (b) reaction kinetics of SnO2 and SnO2/GNPs nanocomposites (in UV light), (c) concentration variation of malachite green during exposure to visible light for neat SnO2 and SnO2/GNPs nanocomposites, (d) the corresponding reaction kinetics of SnO2 and SnO2/GNPs nanocomposites (in visible light) and (e) photoluminescence spectra of SnO2 nanorods and GS-II nanocomposite.

Standard image High-resolution image

To monitor the photodegradation of SnO2, GS-I and GS-II nanocomposites under visible light irradiation, a similar process was followed. The absorbance profiles are shown in ESI figures 3(a)–(c). The rate kinetics of the photocatalytic reaction taking place under visible light are presented in the figures 6(c), and (d). The reaction rate kinetics have been studied using Langmuir-Hinshelwood model. The highest rate constant has been observed for GS-II nanocomposite i.e. k = 0.332 min−1, whereas k = 0.116 and 0.019 min−1 is obtained for GS-I and SnO2 nanorods, respectively. Malachite green has been degraded upto 24.7%, 85.52% and 99.1% in 15 min by SnO2, GS-I and GS-II, respectively. The photocatalysis results are further enriched by photoluminescence results. Figure 6(e) shows that SnO2 shows a broad band from 625 to 675 nm, however the intensity of the same band is greatly quenched after the inclusion of graphene in GS-II nanocomposite. This can be attributed to the fact that the electronic properties of semiconducting SnO2 nanorods are altered by graphene. The photoinduced charge carriers behave differently in SnO2 and GS-II nanocomposite. The PL results agree well with the photocatalytic results. The photoexcited electrons get trapped by the GNPs that are adjacent to SnO2 nanorods. GNPs being electron acceptor provide conducting channels for the interfacial charge transfer, which delays the recombination of photoinduced charge carriers. Therefore, the PL intensity is greatly reduced in GS-II nanocomposite. This also explains the excellent photocatalytic ability of GS-II nanocomposite. The reduced recombination of charge carriers in GS-II promotes reactive oxygen species to interact with organic dye for longer time, which in turn results in higher photodegradation of malachite green. The PL supported catalytic activity has also been explained previously [2426]. Surface area is another factor that contributes towards the efficient adsorption of contaminant on the surface of photocatalyst. Smaller the particle size, greater will be the available surface area for the adsorption of MG on the nanocomposite. SnO2 nanorods have 25 ± 6 nm length and 4 ± 2 nm diameter. These small sized nanorods greatly enhance the surface area, which promotes efficient photocatalytic activity.

A schematic process explaining the reaction mechanism is shown in figure 7. SnO2 nanorods show a very low photocatalytic activity. The higher band gap and comparatively low surface area is responsible for low photocatalytic activity. When SnO2 nanorods get exposed to the UV light in malachite green solution, the VB electrons excite to the conduction band (CB), but these excited electrons de-excite from CB to VB quickly. A few conduction electrons of SnO2 nanorods can be trapped by dissolved oxygen. Therefore, the reactive oxygen species, so generated, are responsible for slight degradation of dye. SnO2/GNPs nanocomposites exhibit greatly enhanced photocatalytic activity, due to inclusion of graphene. Its good electrical conductivity is responsible for provision of effective charge separation. The low band gap energy assists in the fast photoexcitation of electrons. These electrons are readily accepted by adjacent GNPs, thus enhancing the separation lifetime of the charge carriers. The photogenerated holes and electrons are then used for the generation of ROS. The ROS attack the organic structure of malachite green, and completely decompose it to CO2 and H2O [17, 23]. In addition, the increased surface area of nanocomposites also serves for better adsorption of dye on the catalyst surface thus enhancing the photocatalytic activity. The lower bandgap, increased surface area and GNPs availability, synergistically promote higher photodegradation of dye. It is also observed that GS-II degrades the dye with almost same efficiency in 15 min.

Figure 7.

Figure 7. Schematic description of photocatalytic reaction taking place at the interface of SnO2 and GNPs in malachite green solution under UV light irradiation. A same process takes place under visible light irradiation with different reaction kinetics.

Standard image High-resolution image

The recyclability performance of a nanocatalyst is an essential criterion that determines its practical use, particularly in countries with economical constraints. The recyclability test of GS-II nanocomposite was conducted by washing and drying the sample and re-using it four times (figures 8(a), and (b)) both under UV light and visible light irradiation. GS-II shows sustainable photodegradation efficiency for 4 cycles, with only a very slight reduction in efficiency in the 4th cycle, in both cases.

Figure 8.

Figure 8. Recyclability performance of GS-II nanocomposite under UV light irradiation (a), and visible light irradiation (b), microstructure of GS-II nanocomposite after 4 cycles of photocatalysis (c), and crystallinity of GS-II nanocomposite after 4th cycle (d).

Standard image High-resolution image

The GS-II nanocomposite was also tested, after 4 recycles, for microstructure, and crystal structure. The results are presented in figures 8(c), and (d) respectively. Figure 8(c) depicts that SnO2 nanorods are anchored on graphene nanosheet even after 4 cycles of use as photocatalyst, which shows that the microstructure of the catalyst remains intact. Figure 8(d) presents the crystallinity of GS-II after 4 cycles of photocatalytic activity.

The interplanar spacing in SnO2 nanorods and graphene is also presented. The result shows that the crystal structure is not deteriorated after photocatalysis.

Graphene and SnO2 based nanocomposites have rarely been investigated for the removal of malachite green (MG), however, some relevant literature can be found for the removal of other organic contaminants [27, 28]. A systematic comparison of this study with the previous work is cataloged in the table 1. First, a comparison has been established for the degradation of malachite green dye using different materials [2938], where adsorption and photocatalysis both have been covered. It can be readily inferred from this comparison that SnO2/GNPs nanocomposite (GS-II) has shown a superior performance for the fast and complete removal of MG (99.11% in 12 min only). In a previous work, a competitive material, Ag3PO4@MWCNTs@Cr:SrTiO3 shows comparable performance of MG removal, but involves expensive material (Ag3PO4) and tedious preparation method. Whereas SnO2/GNPs nanocomposite could be prepared using few and cheap precursors via a simple synthesis method (see ESI figure 3), which makes it superior. In the second part, the table 1 anthologizes the use of graphene-based materials for the elimination of malachite green [3943], which also establishes the merits of prepared nano-photocatalyst. In addition, the literature shows that graphene-based nanocomposites have been used for decontamination of variety of organic dyes, but few reports are available for the removal of malachite green [44, 45]. In the end, a comparison of SnO2 based materials, to degrade malachite green dye, has also been presented [4649]. It not only depicts the outstanding ability of SnO2/GNPs nanocomposite (GS-II) to remove selected dye (MG) but also shows that SnO2/GNPs nanocomposite has not been investigated so far, for the removal of malachite green. Therefore, this work adds significantly to the existing knowledge, by making SnO2/GNPs nanocomposite superior to the contemporary materials in terms of affordability and re-usability.

Table 1. A comparison for the removal of malachite green (MG), based on different wastewater remediation strategies.

MaterialsDyeMethodTimeDegradation%References
CdO-AgMGUV light PC175 min78.02%[29]
Chitosan/ZnOMGVisible light PC300 min54%[30]
Chitosan/Ce-ZnOMGVisible light PC300 min87%[30]
Organobentonite/Co3O4 MGVisible light PC180 min100%[31]
MgFe2O4/BiMoO6 MGVisible light PC120 min97%[32]
Ag3PO4@MWCNTs@Cr:SrTiO3 MGVisible light PC10 min100%[33]
Starch nanocomposite hydrogelMGSolar light PC300 min91%[34]
Magnetic BCN/GOPAMGAdsorption35 min91%[35]
Graphite/NaA-cl-AACMGAdsorption120 min95%[36]
Magnetic cobalt oxide NPsMGAdsorption120 min95%–97%[37]
C3N4/Fe3O4/ZIF-8MGAdsorption225 min100%[38]
rGO/TiO2/polyMGAdsorption60 min91.4%[39]
rGO/ZnOMGUV light PC90 min78%[40]
MRGO/ZIFMGUV light PC120 min100%[50]
ZnO/GOMGSolar light PC150 min60%[41]
CuWO4-RGOMGUV light PC60 min93%[42]
RGO/TiO2/PANCMA NFsMGUV light PC62 min90.6%[51]
SnO2 nanoparticlesMGUV light PC150 min[46]
rGO/CdO/SnO2 MGVisible Light PC120 min94%[47]
SnO2/TiO2 MethyleneVisible light PC75 min96%[48]
 Green    
Ag/AgO2/SnO2 MGVisible light PC120 min99%[49]
SnO2 nanorods, SnO2/GNPs GS-I and GS-IIMGUV light PC12 min37%, 78.26%This study
    99.11% 
SnO2 nanorods, SnO2/GNPs GS-I and GS-IIMGVisible light PC15 min24.7%, 85.52%This study
    99.10% 

4.7. Antibacterial properties

The testing of novel nanomedicine is inevitable to diffuse the long lasting and harmful effects of drug resistant pathogens. To determine the synergic antibacterial effects of SnO2 and GNPs, we have tested all the prepared materials against two model Gram-positive and Gram-negative bacterial strains (P. aeruginosa and S. aureus). Figures 9(a) and (c) represent the OD profiles and cell viabilities of the selected bacterial strains in the presence of SnO2, and SnO2/GNPs nanocomposites (GS-I and GS-II). The results show that SnO2 inhibits 26.03% and 13.46% growth of P. aeruginosa and S. aureus, respectively. SnO2/GNPs nanocomposite (GS-I) inhibits 42.41% and 44% growth of P. aeruginosa and S. aureus, whereas GS-II nanocomposite has resulted in 79.57% and 78.51% inhibition of P. aeruginosa and S. aureus, respectively. Hence, GS-II has exhibited a significant antimicrobial effect.

Figure 9.

Figure 9. Antibacterial activity of SnO2 and SnO2/GNPs nanocomposites (GS-I, GS-II), (a) and (b) show the results obtained for P. aeruginosa, (c) and (d) show the results for S. aureus.

Standard image High-resolution image

The overall performance of GS-I and GS-II is considerably better than that of SnO2, which is due to an increased amount of graphene. GS-II has significantly controlled the life threatening pathogens from 2 to 24 h [17, 52].

The same experiment was conducted in dark as well. The results are shown graphically in figures 10(a), and (b). The OD profiles show that the growth inhibition of both bacterial strains in dark is significantly less than the one carried out in ordinary setup. Figure 10(a) shows that SnO2, GS-I, and GS-II inhibit 6.36%, 14.05% and 24.39% growth of P. aeruginosa in 24 h. Figure 10(b) shows that SnO2, GS-I, and GS-II inhibit 7.46%, 14.26% and 26.36% growth of S. aureus in 24 h.

Figure 10.

Figure 10. Cell viabilities of P. aeruginosa (a), and S. aureus (b) in dark.

Standard image High-resolution image

There have been several explanations about pathogen control by graphene-based nanocomposites. Figure 11 details a schematic presentation of the possible pathways that lead to the bacterial death. Graphene has wrinkled surface (being flexible) and sharp edges (like cutters). It has been proposed that the edges play a role of cutter by directly disrupting bacterial membranes. The damaged bacterial membrane can ultimately lead to the leakage of cytoplasm [53]. This membrane disruption may result immediate cell death. The TEM image (figure 2(b)) shows that the graphene has a sheet like structure with a rough surface due to presence of SnO2 nanorods. The thin and sharp edges of graphene are also visible. This morphology has been experimentally reported to impair cell membranes in Gram-negative bacteria like E. coli. It has been explained by direct experimental evidence, using TEM images, that edges of graphene act like blade that cut the bacterial membranes. It has been clearly demonstrated that graphene may get inserted into E. coli to damage it completely. The loss of membrane integrity has been found to be associated with the lipid extraction within few nanoseconds, mainly caused by graphene wrapping around the cell and its cutter like activity [5457]. The membrane disintegration is schematically described in figure 11.

Figure 11.

Figure 11. A possible mechanisms of bacterial death induced by the SnO2/GNPs nanocomposite.

Standard image High-resolution image

It has been experimentally proven that the bacterial membranes are electron rich surfaces. The graphene sheets being an outstanding electron acceptor material, are readily available for membrane-electrons, thus creating an electron deficit at the membrane surface. The unbalanced charges appear as another cause to damage the toxic bacterial cells [53].

The antibacterial performance of a nanomaterial depends strongly on its particle size. The dimensions of a nanomaterial determine its penetrability into the bacterial cell. Nanomaterials with small particle sizes can easily penetrate in the bacterial cell membranes by passing through the porous cell membranes. SnO2 nanorods with 25 ± 6 nm length and 4 ± 2 nm diameter can easily penetrate through the permeable membranes pores. The easy inflow of nanostructures is only possible if their size is smaller than that of membrane pores [58]. When the SnO2 nanorods enter the cytoplasm, they may interrupt normal function of cell along with many other destructive mechanisms. This also explains the slight antibacterial activity exhibited by neat SnO2 nanorods.

The formation of ROS has been a common theme for toxicity of nanomaterials. ROS initiated oxidation-reduction reactions cause cell death with onset of DNA destruction and protein denaturation. It is an established fact that the positive metal ions, free electrons could also lead towards ROS initiated oxidative stress, which is primarily responsible to damage the pathogenic bacteria cells [6]. All the above-mentioned activity, in a coherent manner, ultimately results the significant inhibition of bacterial growth. However, it is important to address the effect of light in ROS generation. In normal day light experiment, the generation of ROS is considerably enhanced due to creation of photoinduced charge carriers. These electrons, holes, and ROS are ultimately used to generate oxidative stress in bacterial cells. The basic conclusion derived from the antibacterial assessments is that ROS generation has significant effect in bacterial growth inhibition. The ROS generation varies in ambient light and in dark. In the absence of light, the charge carriers (e−h pairs) can't be generated which reduces the growth inhibition of bacteria. The large cell viabilities have been observed in dark which are mainly due to the absence or a smaller number of ROS. The slight bacterial death that has been found in the dark is largely due to sheet like cutting effect of GNPs in the nanocomposite.

There are other factors that play an important role for the observed antibacterial activity of SnO2 and SnO2/GNPs nanocomposites in the dark. First, the electrostatic interaction between SnO2 nanorods and bacterial cell membrane leads to transport of Sn2+ ions into the cell cytoplasm. The interaction between Sn2+ and cytoplasmic organelles may cause DNA damage which may lead to bacterial cell death. Also, the membrane malfunctioning due to attached SnO2 nanorods is one of the reasons of bacterial growth inhibition. These mechanisms have been proposed previously for experiments that are conducted in dark [59, 60].

Table 2 depicts the performance of various graphene/metal oxide nanocomposites. It can be readily concluded from the comparison that SnO2/GNPs nanocomposite (GS-II) possesses significant control over bacterial growth, both for Gram-positive and Gram-negative bacteria

Table 2. A comparative description of different graphene-based nanocomposites for bacterial growth inhibition.

MaterialBacteriaMorphologyBacteria typeZOI/% cell inactivation for bacteriaReferences
rGO-CuO nanocomposites (rGO-CuO1, rGO-CuO2) P. aeruginosa (PAO1)FilmsGram-negative71.5%, 72%[61]
Graphene-tin oxide nanocomposites (WC, CSC) P. aeruginosa NPsGram-negative27 ± 1 mm, 38 ± 0.7 mm[62]
Graphene-SnO2 nanocomposite P. aeruginosa NPsGram-negative160.5 mm, ∼79%[21]
GO-Ag nanocomposite P. aeruginosa, S. aureus NPsGram-negative and Gram-positive24 h 68% and 42%[63]
Cu2O-GO S. aureus, E. coli NanosheetsGram-positive and Gram- negative32% and 68%[64]
SnO2/GNPs nanocomposites (GS-II) P. aeruginosa, S. aureus Nanorods on nanosheetsGram-negative and Gram-positive79.57% and 78.51%This study

5. Conclusions

SnO2 nanorods and SnO2/GNPs nanocomposites have been prepared using a chemical route. SnO2 exists in tetragonal crystalline phase. The optical bandgap energies have been tuned from 3.14 (for SnO2), to 2.85 eV and 2.80 eV in GS-I and GS-II, respectively. SnO2/GNPs nanocomposite, (GS- II) has shown complete dye's degradation (99.11% in 12 min) under UV light irradiation. In visible light, GS-II removes 99.01% malachite green in 15 min, while SnO2 removes the same only upto 24.7% in the same time. GS-II has also exhibited significant antimicrobial effect by inhibiting 79.57% growth of P. aeruginosa, and 78.51% of S. aureus. This outstanding performance of GS-II nanocomposite is attributed to the synergistic effects of SnO2 and GNPs, that give rise to increased adsorption, delayed carrier's recombination, and ROS generation thus proving it an excellent multi-functional nanocomposite, for the elimination of malachite green from aqueous solution and pathogen control. The findings will pave the way for new prospects for designing wastewater treatment strategies and to encounter antibiotic drug resistance of bacteria.

Acknowledgments

A A pays gratitude to Dr Mahavir Sharma for his unconditional support during this work. A A acknowledges Dr B G Mendis and Dr Ian Terry (Durham University, UK) for extending assistance to carry out TEM measurements. Dr J I Saggu, and Dr K Nadeem are acknowledged for facilitating experiments. M Z acknowledges EPSRC for funding. The similarity index checked via Turnitin (ID: 1299992513) is well below the allowed limit (9%).

Data availability statement

The data that support the findings of this study are available upon reasonable request from the authors.

Please wait… references are loading.