You are on page 1of 37

Journal Pre-proof

An environmental approach for the


photodegradation of toxic pollutants from
wastewater using silver nanoparticles decorated
titania-reduced graphene oxide

Ankita Ojha, Pardeep Singh, Ramesh Oran,


Dhanesh Tiwary, Ajay K. Mishra, Ayman A.
Ghfar, Mu. Naushad, Tansir Ahamad, Binota
Thokchom, K. Vijayaraghavan, S Rangabhashiyam

PII: S2213-3437(21)00599-6
DOI: https://doi.org/10.1016/j.jece.2021.105622
Reference: JECE105622

To appear in: Journal of Environmental Chemical Engineering


Received date: 11 March 2021
Revised date: 22 April 2021
Accepted date: 1 May 2021
Please cite this article as: Ankita Ojha, Pardeep Singh, Ramesh Oran, Dhanesh
Tiwary, Ajay K. Mishra, Ayman A. Ghfar, Mu. Naushad, Tansir Ahamad,
Binota Thokchom, K. Vijayaraghavan and S Rangabhashiyam, An environmental
approach for the photodegradation of toxic pollutants from wastewater using
silver nanoparticles decorated titania-reduced graphene oxide, Journal of
Environmental Chemical Engineering, (2021)
doi:https://doi.org/10.1016/j.jece.2021.105622
This is a PDF file of an article that has undergone enhancements after acceptance,
such as the addition of a cover page and metadata, and formatting for readability,
but it is not yet the definitive version of record. This version will undergo
additional copyediting, typesetting and review before it is published in its final
form, but we are providing this version to give early visibility of the article.
Please note that, during the production process, errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
© 2021 Published by Elsevier.
An environmental approach for the photodegradation of toxic pollutants from wastewater

using silver nanoparticles decorated titania-reduced graphene oxide

Ankita Ojha1, Pardeep Singh2*, Ramesh Oran3, Dhanesh Tiwary1, Ajay K. Mishra4, Ayman A.

Ghfar5, Mu. Naushad5,6, Tansir Ahamad5,7, Binota Thokchom8, K. Vijayaraghavan9,

Rangabhashiyam S10*
1
Department of Chemistry, Indian Institute of Technology (BHU), Varanasi, India -221005.
2
Department of Environmental Studies, PGDAV College, University of Delhi, New Delhi India -110067

f
3
Dept. of Nanoscience and Technology (DNST), Central University of Jharkhand (CUJ) Brambe, Ranchi Jharkhand

oo
(835205)
4
Institute for Nanotechnology and Water Sustainability, College of Science, Engineering and Technology,
University of South Africa, Science Campus, Florida, 1710, Johannesburg, South Africa
5

pr
Department of Chemistry, College of Science, King Saud University, Riyadh-11451, Saudi Arabia
6
Yonsei Frontier Lab, Yonsei University, Seoul, Korea
7
School of Science & Technology, Glocal University, Saharanpur, India
8

9
e-
DST-Inspire Faculty, Department of Forestry and Environmental Science., Manipur University, Indo-Myanmar
Road, Canchipur – 795003, Imphal, Manipur, India
Department of Civil and Environmental Engineering, 1Engineering Drive 2, National University of Singapore,
Pr
Singapore 117576, Singapore
10
Department of Biotechnology, School of Chemical and Biotechnology, SASTRA Deemed University, Thanjavur -
613401, Tamilnadu, India.
al

Corresponding authors’ email addresses: psingh.rs.apc@itbhu.ac.in (Pardeep Singh); rambhashiyam@gmail.com


(Rangabhashiyam S)
Abstract
n
ur

Light-induced demineralization of organic pollutants is a recent development in the field of


waste treatment. The present study focuses on TiO 2-rGO-Ag hybrid nanocomposite through a
Jo

two-step simultaneous synthesis pathway. The methods primarily involving facile solvothermal
treatment (TiO2-rGO nanocomposite) in a water-ethanol mixture, followed by microwave
irradiation method for deposition of silver (Ag) nanoparticles over the Binary nanocomposite
synthesized. The resulting hybrid photocatalyst was well, and in-detail characterized using
analytical techniques such as High-Resolution X-ray Diffractometer (HR-XRD), Transmission
Electron Microscopy and High-Resolution Transmission Electron Microscopy (TEM/HR-TEM),
High-Resolution Scanning Electron Microscopy (HR-SEM), Fourier-Transform Infra-Red
Spectroscopy (FT-IR), Energy Dispersive X-Ray Spectroscopy (EDX), Raman Spectroscopy,
Solid-State UV–Vis Spectroscopy and X-Ray Photoelectron spectroscopy (XPES). The
nanomaterial was examined to determine photocatalytic property by conducting a

1
photodegradation experiment of an aqueous solution of Methylene Blue (MB) dye and
monitoring the changes. The photocatalytic efficiency of the nano-hybrid synthesized was also
analyzed in detail for the degradation of compounds that active components of petrochemical
pollutants such as Benzene, Toluene, and Phenol the Visible region of radiation in a
photochemical reactor under ambient reaction conditions. The set of experiments suggested that
the photocatalytic efficiency of ternary nanocomposite synthesized was quite noticeable. The
variation in the catalyst's photoactivity with the change in the functional group was monitored,
and the rate of reaction has been correlated with attached substituents.

f
Keywords: Environmental remediation; Surface Plasmon Resonance; Ternary; Microwave;

oo
Solvothermal

pr
1. Introduction
e-
Heterogeneous catalysis (a catalysis process where catalysts and reactants are in different phases)

has undoubtedly emerged as a breakthrough in the field of environmental protection.


Pr
Photocatalysis, one such kind of heterogeneous catalysis method, occurs between solid, i.e.,

catalyst (nanocomposite as for the system we investigated) primarily made up of semiconductor


al

containing TiO2 and gas or liquid phase of the pollutant [1]. The photocatalyst activation is
n

stimulated by irradiation that generates electron-hole pairs in the semiconductor region [2,3,4].
ur

These techniques have been implied for the degradation of varying and multiple classes of
Jo

Volatile Organic Compounds (VOCs), such as benzene, toluene, methyl/ethylbenzene,

dichloromethane, etc. [5]. There has been high contamination of natural resources, such as soil,

air, and water, due to the bulk generation of waste. Their undesirable and assorted effects are

much visible not only on human beings but also on the ecosystem. Such a threat comes from a

high level of contaminated water due to petrochemical wastes [6].

The mining and petrochemical industries play a leading role in the fiscal growth of such waste

materials. Many industrial goods cover an enormous range of applications worldwide [7,8,9].

2
Wastewater released by petrochemical industries poses a severe threat to human beings as well

as flora fauna. Such petrochemical wastes contain an extensive range of chemicals comprising

both aromatic and polyaromatic hydrocarbons such as phenolic compounds, their derivatives,

and metal salts, few surface-active substances (alkyl trimethyl or dimethyl ammonium

compounds; CTAB; Mono- and dicarboxylic sulfophenyl acids), sulfide compounds, naphthenic

acids and many other chemicals [10,11]. Petroleum compounds mostly contain complex

f
mixtures of hydrocarbons, organometallics, kinds of paraffin, cycloalkanes, aromatics (benzene

oo
and related compounds), and polyaromatic hydrocarbons (naphthalene, anthracene, pyrene, etc.)

pr
too. Benzene and its derivatives compounds, along with other polyaromatic hydrocarbons, were

stated among the most carcinogenic to human beings and are responsible for causing a variety of
e-
disorders like lymphoma, chromosomal breakage, leukaemia, and interference with DNA
Pr
segregation as well as act as endocrine disruptors and accountable for reproductive problems,

hypospadias, miscarriages, and infertility, etc. [12,13,14,15]. With the increasing industrial
al

activities and exploitation of petrochemicals, these compounds are certainly a part of the system.
n

Not only surface water but groundwater is getting affected by these compounds through
ur

underground pipelines and leachates. The entrance of benzene and its derivatives has increased
Jo

in water bodies in the past few years with uncontrolled and extensive human activities when

using personal products or agricultural sources [16]. Several techniques, such as physical,

chemical, and even sometimes biological, have been employed to degrade these pollutants from

wastewater. Several methods, such as advanced oxidation, biofiltration, separation by the

membrane, and sorption, have been evolved to treat wastewater containing such contaminants

[5,17]. However, these have limited applications due to either these being expensive or being

associated with environmental consequences. The best method among these technologies was

3
selected based on wastewater and waste chemistry, the cost-effectiveness of the technique,

proper availability of space for carrying out the process, plans for reprocessing and discharging,

robust operational hands, and by-products generated during the process. Of all the various

approaches available, the physical method for the abatement of organic pollutants, adsorption,

and coagulation is the most commonly used process [18]. Advanced Oxidation Processes (AOPs)

are new and rapidly surfacing techniques for ecological confiscation of chemically stable / less

f
reactive or organic pollutants less susceptible to biodegradation. They aim at generating highly

oo
reactive hydroxyl radicals (.OH). These radicals tend to attack large molecules and break them

pr
into smaller fragments. These smaller fragments are mineralized to CO2, H2O, and other small

molecules [20]. This technology aims to apply laboratory scale techniques to running at pilot-
e-
scale and develop environmentally friendly catalysts that support the sustainable treatment
Pr
technology with the 'zero’ concepts of industrial wastewater [21].

A recent breakthrough in this advanced method is applying Titania (Titanium dioxide, TiO2)
al

based nanocomposite. Such Titania-based nanocomposites own high reactivity, low toxicity,
n

resistance to photo-induced reactions, high photocatalytic activity, Photo- and chemical stability
ur

even at very low pH and low production cost. It has the potential to degrade a wide range of
Jo

petrochemical wastes. Due to the relatively high band energy gap of 3.2 eV, Titania shows the

UV region's radiation absorbance. Titania absorbs the UV radiations generates an electron-hole

(e--h+) pair, which reacts with water molecules that are adsorbed on the surface of the catalyst to

produce hydroxyl radicals (.OH) on reacting with h+ [13,22,23]. The only drawback affecting

the application of Titania as a semiconductor is the recombination of the electron-hole pair.

Since the bandgap for Titania is very high most of the research activities have stressed the

shifting of Titania bandgap to lower values and overcome electron-hole recombination issue. The

4
mechanism involving TiO2–UV photocatalysis can be categorized into four sequential steps. In

the first step, there is an excitation of TiO2 by UV irradiation that follows the second step, i.e.,

the generation of excitons (electron-hole pairs) for initiation. The third step is the generation of

several reactive oxidative species and, finally, the oxidation of the organic compounds by these

oxidative species [24,25]. A variety of dopants such as CdS, Pt, Au, and plasmonic nanoparticles

have been introduced to reduce the semiconductor device's bandgap. However, the interaction

f
between nanoparticles and pollutants may reduce the nanoparticles' photocatalytic activity [26].

oo
Hence a combination of Titania and graphene nanosheets has been a better approach to overcome

pr
the problem.

Carbon Nanomaterials have revolutionarily introduced a new thrust area of research, especially
e-
in environmental science, such as Graphene, fullerene, and quantum dots leading to recent and
Pr
much sustainable advancements with time [27]. Graphene is acclaimed to have high mechanical

strength, larger surface area (theoretical value lying around 2600 m2 g-1), relatively high electron
al

mobility (approx. 15 X 103 m2 V-1 s-1 even at room temperature) Fermi velocity (106 m/s)
n

[28,29] and effective thermal conductivity. The extensive p-electron conjugation in Graphene
ur

(which levels the superfluous electrons turns out to be a potential sink for electrons) demotes the
Jo

electron-hole recombination. The semiconductor gap in Graphene is said to be of 'zero-gap.'

Graphene electrons seemed to be almost neutral to the randomness and electron-electron

interactions and had been experimentally proved to have very long mean free paths [30,31].

The electrons present in metallic nanoparticles tend to excite easily in the presence of light

(sometimes even at higher wavelengths). The oscillating electric field of the light, which is

electromagnetic radiation, interacts with these metal atoms' conduction electrons. This

interaction leads to a strong oscillation of these electrons resonating with the incident photon

5
frequency. This resonance phenomenon is termed localized surface plasmon resonance (SPR) for

metallic nanoparticles. Modifications of these metallic nanoparticles regarding their sizes,

shapes, and electronic environment have malleably regulated their scattering and absorption

attributes [32]. Silver nanoparticles (Ag NPs) were widely acclaimed for their elite optical and

catalytic behaviour. Among the three most common metal nanoparticles (Silver, Gold, and

Copper) that display Plasmon Resonance in their visible light spectrum, Ag nanoparticles show

f
the signs of the maximum competence for plasmon excitation. Due to its high Plasmon

oo
resonances, optical excitation is the primary source of light interaction with matter. The doping

pr
of silver nanoparticles is crucial for tuning the visible range absorption in these nanoparticles

[27,33]. The Localized Surface Plasmon Resonance (LSPR) phenomenon present in Ag


e-
nanoparticles empowers them to exhibit intense and broad absorption in the visible part (400-800
Pr
nm) of the sunlight. Therefore, Ag nanoparticles lessen the recombination of photoelectron–hole

pairs through the electron transference to themselves. In addition to this, their wide-ranging
al

absorption of visible light induces a visible light hammered reaction for photocatalytic processes
n

in TiO2-rGO binary nanocomposite [5]. As in the present study, we have discussed the synthesis
ur

and characterization of Ag deposited Titania-rGO based ternary composite and their application
Jo

in the degradation of some of the aromatic organic compounds. The initial process involved

solvothermal treatment of Titania and Graphene (for Binary nanocomposite), followed by

microwave irradiation for Ag nanoparticles deposition over the surface of TiO2-rGO. Few

studies were conducted to determine the demineralization of targetted organic pollutants. The

synthesized ternary composite can be used for a wide range of recalcitrant environmental

pollutants. The synergistic interactions between TiO2, Reduced graphene oxide or rGO, and

silver (Ag) nanoparticles have resulted in enhancing the photocatalytic activity of semiconductor

6
Titania [32,33]. Hence a highly comprehensive visible light response was observed because of

high electron-hole separation efficiency. We have studied the photodegradation of some of the

aromatic organic compounds (benzene and derivatives) using ternary nanocomposite. The

present works also emphasize the effect of substituents on aromatic rings and hence will be

helpful to design the relevant materials for degradation. The inductive effect of functional groups

is correlated to the photodegradation of these compounds. This application can be a breakthrough

f
in the upcoming days to treat a range of organic pollutants, biological contamination, and

oo
macromolecules that act as a potent threat for the water bodies and sustain life for its survival.

pr
The work also stands for useful systems where compounds are not easily degradable, and hence

functional group modifications can help in the more accessible breakdown of complex organic
e-
compounds.
Pr
2. Materials and methods

2.1.Materials
al

The chemicals that were used in the reaction included Graphite powder (<20 μm, ≥99.99% trace
n

metals basis, Alfa Aesar), Titania (P-25, approximately around 21 nm of primary particle size,
ur

and ≥99.5% trace metals basis), concentrated H2SO4, Potassium peroxodisulfate (K2S2O8),
Jo

concentrated HNO3 (ACS reagent, 70%), KMnO4, Hydrogen peroxide (H2O2), silver nitrate

(ACS reagent, ≥99.0%) and sodium citrate (ACS reagent, ≥99.0%). All these chemicals that had

been used in synthesis had been purchased from Alfa Aesar and Sigma Aldrich. All the chemical

substances were, as such, used without any further treatment or purification.

2.2 Synthesis of Graphene Oxide (GO)

Graphene oxide (GO) for ternary nanocomposites was synthesized via the modified ''Hummers'

method.'' As in a typical method of synthesis, Graphite powder (3 g) had been taken in a mixture

7
of concentrated H2SO4 (15ml), Potassium peroxodisulfate (K2S2O8, 3g), and phosphoric acid (4

g) as mentioned in the reference [34,35]. The reaction mixture was heated up to 80oC using an

oil bath with constant mixing and stirring for about 24 hours on a magnetic stirrer. The product

obtained was then further diluted with Triple distilled water (300 ml), filtered thoroughly, and

repeatedly washed with distilled water to attain a neutral pH (constant monitoring with pH

paper). After this, the product was uniformly and properly dried under standard conditions

f
overnight. The preoxidized sample of graphite was then imperiled for further oxidation by

oo
adding to the mixture of a solution of 100 ml of concentrated Sulphuric Acid (H 2SO4) and 30 ml

pr
of concentrated Nitric Acid (HNO3) in a 10:3 ratio, under vigorous, followed by continuous

stirring and finally cooling it to 0oC. Potassium permanganate (KMnO4, 12 g) was then added
e-
very slowly and gradually under constant stirring and mixing with regularly maintaining and
Pr
controlling the reaction mixture temperature not more than 20 oC through cooling. Sequentially,

this product mixture was kept stirring at room temperature for around 96 hours. It was then
al

diluted with distilled water (500 ml) water on an ice bath holding the temperature below 50oC for
n

3 hours and treated with 30% H2O2 (20 ml) to obtain a brilliant yellow product. The crude
ur

product obtained was filtered and repeatedly washed with an aqueous hydrochloric acid solution
Jo

(HCl) in a 1:20 ratio (1 part acid and 20 parts water). Repeated washing was done to remove

residual metal ions or anions, followed by washing with triple distilled water to wipe out residual

acid completely. GO powder, hence obtained, was desiccated for 24 hours at 70 oC [22, 35]. The

summarized scheme of reaction method has been mentioned in Fig. S1 (Supplementary file).

2.3 Synthesis of Titania-reduced Graphene oxide (Binary) Composite (TiO 2-rGO)

In this step, the hydrothermal method of synthesis was applied to synthesize the binary

nanocomposite of TiO2-rGO. TiO2 (2g) was ultra-sonicated using a probe sonicator (50kW) in a

8
Water Ethanol Mixture (80 ml, 3:1 ratio) for 2 hours and followed by the addition of Graphene

Oxide (0.05g). This dispersion was further probe-sonicated in a Water-Ethanol mixture (40 ml)

in a 3:1 ratio for 2 hours to exfoliate Graphene oxide and homogenization the reaction mixture.

The materials were then tightly sealed-packed in a thermally stable Teflon beaker packed in a set

up of stainless steel (as shown in Fig S2), and it was set at a temperature of about 120o C for 24

hours in a hot air oven. In the end, the as-synthesized product was centrifuged and rinsed thrice

f
by application of deionized water-ethanol mixture, and the final product was dried at 80 o C for

oo
around 12 hours [22,36,37].

pr
2.4 Ag NPs addition

The deposition of silver nanoparticles over the binary composite was done through the
e-
microwave irradiation method for the chemical reduction of silver nitrate. The microwave
Pr
method is one of the homogenous and rapid ways for such reductions [38]. An aqueous solution

of 3 mmol AgNO3 solutions (60 ml) was taken in a beaker, and TiO2-rGO (0.5 g) composite was
al

dispersed into the solution. Then, three mM sodium citrate solution (30 ml) was added into this
n

suspension and left for stirring for about 5 minutes. After this, the suspension was microwaved
ur

(100 W) for 5 minutes and allowed to cool at room temperature (schematically shown in fig. S3).
Jo

The product was then centrifuged and washed with triple distilled water thrice to remove surface

impurities. Finally, the product was kept in a hot air oven at 80o C for 12 hours [22].

2.5 Characterizations

Sample morphological studies were performed with specialized characterization techniques.

Transmission Electron Microscopy (TEM) and High-resolution Transmission Electron

Microscopy (HR-TEM) analysis were carried on a TECNAI G2 F20 S-TWIN Transmission

Electron Microscope Instrument. High-resolution Scanning Electron Microscopy (SEM) data has

9
been recorded over a ZEISS Supra 40 High-Resolution Scanning Electron Microscope for

microstructural characterizations. Powder X-ray Diffraction studies had been done on RIGAKU

High-Resolution Powder Diffractometer 12kW XRD between 2θ~10 to 80o angle of diffraction.

Diffused reflectance spectra UV-Visible absorption peak spectra were obtained with the help of

the SHIMADZU UV-Visible spectrophotometer. The X-ray Photoelectron Spectroscopy (XPS)

data for sample characterization was performed with AMICUS, Kratos Analytical, A Shimadzu,

f
serial number Cb40143/01. Raman spectra of Binary and ternary nanocomposites were collected

oo
on a Renishaw in Via Raman spectrometer, 785 nm solid-state diode laser, 0.5% intensity of 300

pr
mW, 20 sec exposure time. The Fourier-Transform Infra-Red spectroscopy (FT-IR) analysis was

done between a range of 4000–400 cm-1 using a Spectrum 100, Perkin Elmer spectrometer by
e-
KBr pellet method. The morphological attributes have been calculated from the TEM images
Pr
using ImageJ, and all the plots have been done using Origin 8.0 software. The XPS data were

plotted on XPSPeak software.


al

2.6 Photodegradation of methylene blue (MB) dye


n

Equal quantities (40 mg each) of TiO2, TiO2-rGO, and TiO2-rGO-Ag samples were dispersed
ur

into aqueous Methylene Blue (MB) dye solution with a concentration of 50 ppm (50 mg/L). The
Jo

reaction mixture was left unstirred in the dark for 12 hours to bring about an adsorption-

desorption equilibrium. It was followed by irradiation (300 W) of the visible light given for 7

hours under steady conditions. A fixed amount of reaction mixture was taken at a fixed time

interval from the photo-reactor for measuring the change in absorbance after centrifugation in

micro-centrifuge for 2 minutes at 8000 rpm. The dye concentration variation was acquired as per

the concentration-absorbance (at 664 nm) relationship and was plotted using Origin Pro

Software.

10
2.7 Photocatalytic degradation

Aqueous solutions of benzene, toluene, and phenol were prepared separately with a

concentration of 10 ppm using deionized water. The analytical samples have been prepared

considering the solubility of these organic compounds in water. The solubility of benzene,

toluene and phenol in water is 1.74, 0.53 and 82.4 gL-1, respectively [39,40,41,42]. Considering

these values, the solutions have been prepared in an optimized range of concentration. Reaction

f
conditions were optimized for different concentrations of catalysts. In the present study, 100

oo
mg/l catalyst concentration was taken to investigate these organic compounds for photocatalytic

pr
degradation. These solutions were kept aside in the dark for 12 hours with continuous stirring for

maintaining adsorption-desorption equilibrium. The solutions were then exposed to visible


e-
radiation (300 W) in a photoreactor. The change in the concentration during the photochemical
Pr
reaction was monitored at regular time intervals.

3 3. Results and discussions


al

3.1 X-ray diffraction (XRD) patterns


n

The Powder X-ray Diffraction analyses were performed for investigating crystal structure and
ur

crystalline nature of as-synthesized samples, as shown in Fig. S4. GO demonstrate one

prominent peak at 2θ~11.5o, which corresponds to the (001) plane along with the appearance of a
Jo

small peak at 2θ~42.5° [43]. Notably, no such observable diffraction patterns from carbon

species were identified after the composite fabrication of GO with TiO 2 (in TiO2-rGO). These

relevances could be attributed to the shielding of the main characteristic peak for rGO at 24.5o by

the main peak in anatase TiO2 at 25.3o [22,44,45]. Fig.S4, has clearly shown the X-ray diffraction

(XRD) patterns for Graphene oxide (GO), Binary nanocomposite (TiO2-rGO), and ternary

nanocomposite (TiO2-rGO-Ag). The peaks that are observed at 25.3o, 37.0o, 48.0o, 53.9o, 62.7o,

11
68.8o, 70.5o, and 75.0o are indexed as (101), (004), (200), (105), (204), (116), (220) and (215) in

the crystal planes of anatase TiO2, respectively (anatase phase prevailed due to the lower

synthesis temperature). Additionally, characteristic peaks of diffractions have also been observed

at 27.8o, 36.3o, and 41.4o, which were respectively assigned to the (110), (101), and (111) of the

faces in rutile TiO2 [22,46]. Three additional peaks of 44.0o, 64.6o, and 77.8o were found in TiO2-

rGO-Ag samples, which were related to (200), (220), and (311) planes, respectively, of metallic

f
Ag nanoparticles (Table S1). These peaks indicated successful decoration of Ag nanoparticles

oo
within TiO2-rGO. Another peak that appeared at 38.1o, which is to be attributed to the Ag (111)

pr
plane, is not displayed. This phenomenon is possibly due to overlapping with an intense peak of

TiO2 at 37.0o. These peaks have been indexed in accordance with silver nanoparticles cubic
e-
structure (JCPDS No. 65-2871) [22,32,37].
Pr

3.2 Scanning electron microscopy (SEM)


al

Fig. 1 represents SEM images for Graphene (GO) samples, Titania (TiO 2), TiO2-rGO, and TiO2-
n

rGO-Ag composites. SEM image of GO represents a typical disordered layered structure of GO


ur

than the ordered arrangement of graphite powder as visible in the first image folded at places.
Jo

The sheets had individual edges with kinks and wrinkled much in accordance with the earlier

reported works [35,47]. This creasing of Graphene sheets could be assigned to various oxygen-

containing functional groups over the surface and along the edges of GO. Flakes of rGO were

visible over TiO2 nanoparticles in the binary composites. The ternary composite containing Ag

nanoparticles visible over the binary composites has been observed in the TEM images. Ag

nanoparticles were uniformly dispersed on the binary composite (on the scale of 100 nm for the

images captured through TEM). SEM image of TiO2 (mixed anatase and rutile phase) revealed

12
the particle nature of irregular agglomerates, which seemed to appear as rough surfaces in the

pictures. In binary TiO2-rGO, TiO2 anatase and rutile particles appeared to be composed of the

fused irregular sphere of different sizes and shapes. However, after Ag deposition, noticeable

changes were evident in the surface morphology of ternary composite (TiO 2-rGO-Ag). Irregular

and varying shapes and sizes of the composite are observed to be ordered spherical shapes

composed of Ag deposited over TiO2-rGO. No appearance of graphitic material in Binary and

f
ternary composite could be attributed to a relatively higher concentration of TiO 2 than GO as

oo
mentioned in the synthesis process [22,48].

pr
e-
Pr
n al
ur
Jo

Fig. 1. SEM images of Graphene (GO), Titania (TiO2), TiO2-rGO, TiO2-rGO-Ag

3.3 Transmission electron microscopy (TEM)/High-Resolution transmission electron

microscopy (HR-TEM)

13
For validating the SEM analysis, TEM images were taken to identify detailed structures. HR-

TEM was also performed to calculate d-spacing. The sizes of Ag nanoparticles were estimated to

be lying in the range of 20-30 nm. These investigations through Transmission Electron

microscopy (TEM) and High-Resolution Transmission Electron Microscopy (HR-TEM) of

Graphene GO, TiO2, TiO2-rGO, and TiO2-rGO-Ag nanocomposites displayed below in Figures 2

and 3, respectively. The synthesized composite revealed a layered structure with a micron range

f
that has creases after the hydrothermal reduction [22,49,50]. Graphene oxide (GO) sheets are

oo
clearly visible in Fig. 2 A and TiO2 nanoparticles are seen in Fig. 2B. The agglomerated particles

pr
of TiO2 (Tetragonal bipyramidal (of anatase) and tetragonal prism structure of Rutile) [51] were

decorated on the graphitic layer as shown in Figure 2.C. TiO2 nanoparticles seem to be dispersed
e-
inhomogeneously over rGO sheets over the edges and wrinkles [52,53]. There is a co-existence
Pr
of TiO2 and Ag nanoparticles, which was confirmed in Fig. 2 D. Fig. 3.D showed that the Titania

(TiO2) and Ag nanoparticles attained layered structures with the lattice spacings of 0.34 nm, 0.32
al

nm, and 0.24 nm. These lattice spacings correspond to (101) in the Anatase plane of TiO 2, (101)
n

in the Rutile plane of TiO2, and (111) in the plane of Ag nanoparticles, respectively. From Fig.
ur

3A-D, there have been visible changes in microstructures from sheet to the composite materials.
Jo

The images indicated formation of homogenous and intricate patterns of TiO2 and Ag

nanoparticles over the r-GO sheets. Furthermore, HR-TEM results also indicated Ag, TiO2, and

rGO that create intimate interfaces within the nanocomposite samples. Here, we can also see that

the Ag nanoparticles were dispersed much evenly over the TiO 2 and rGO without apparent

clustering of Ag nanoparticles on the graphene nanosheets. These nanosheets act as two-

dimensional “mat”-like structures with which the Ag nanoparticles interact and thereby obstruct

the clustering [11,22,54].

14
f
oo
pr
e-
Pr
n al

Fig. 2. TEM images of A. Graphene Oxide (GO) sheets; B: TiO2 TEM images (TiO2); C: TiO2
ur

stacked over GO sheets after the solvothermal treatment; D. TiO 2-rGO-Ag


Jo

15
f
oo
pr
e-
Pr
n al

Fig. 3. HR-TEM images of A. Graphene (GO) sheets; B: TiO2; C. TiO2 stacked over Graphene
ur

sheets after the hydrothermal reaction; D. Ag nanoparticles deposited over the binary composite.
Jo

3.4 Energy-dispersive X-ray Spectroscopy (EDX)

The energy-dispersive X-ray spectroscopy (EDX) has been carried out for determining the

elemental compositions of the nanocomposites (Fig. 4). The peak for the carbon (C) element was

generated due to the graphitic material. The presence of Ti was related to the TiO2 nanoparticles

in the nanocomposites. The Oxygen (O) element corroborated the oxygen-functional groups on

Graphene as well as TiO2. However, TiO2 nanoparticles prevented restacking and agglomeration

of rGO nanosheets. The intricate contact of atoms between both components enables the easy

16
migration of electrons from TiO2 into rGO during photoexcitation of semiconductors. This

charge transfer ultimately hinders the charge (electron-hole) recombination processes and

enhances the photocatalytic activity of the material [55]. The ternary composites show a tiny

percent of Ag metal atoms in the composite, i.e., 0.02%.

f
oo
pr
e-
Pr
n al
ur
Jo

Fig. 4. Energy-dispersive X-ray spectroscopy of Binary and Ternary nanocomposites (EDX)

3.5 Fourier transform infra-red spectra (FT-IR)

FT-IR investigation studies of the ternary nanocomposite TiO 2-rGO-Ag were done before

(Black) and after the methylene blue (Red). The peak of about 3400 (broad, intense) and 1640

(very weak) cm-1 was assigned for the surface adsorbed water and hydroxyl groups on the

nanosheets [56]. The absorption band present at 650 cm-1 was corresponding to the Ti-O

vibration mode of the Ti3+-O-Ti4+ framework. Carbon-Hydroxyl bond (C-OH) and epoxide bond

17
(C-O-C) peaks were also observed at 1380 and 1260 cm-1, respectively [57,58,59]. Similarly, the

peaks near 2912 cm-1 and 2842 cm-1 could be ascribed to the asymmetric and symmetric CH 2

stretching modes of graphitic carbon, respectively. The peak around 1588 cm-1 signifies that

there is C=C which stretches from the unoxidized graphitic domain of the nanocomposite [60].

The spectra of the photocatalyst of before and after the dye degradation was recorded, and there

was no observable change in the spectra (Fig. 5), which indicated no modification of the

f
functionality of the composite during the degradation process. Besides, the absence of a new

oo
peak confirmed the non-possibility of any chemical reaction between photocatalyst and the

pr
substrate, thereby ensuring the catalytic nature of the material.

e-
Pr
n al
ur
Jo

Fig. 5. FT-IR Spectra of the ternary photocatalyst (before (black) and after (red) the degradation.

18
3.6 Solid-state UV-Visible spectroscopy

Solid-state UV-Visible absorbance spectra of Graphene (Red), Titania (Pink), TiO 2-rGO (Black),

and TiO2-rGO-Ag (Blue) nanocomposites were recorded on a solid-state UV-Visible absorption

spectrophotometer. Spectra had been plotted for wavelength against absorbance (Fig. 6). The

visible range of radiation (400-800) nm was taken. This range is the Plasmon resonance range,

which is mainly due to the Ag nanoparticles. The spectra showed a hump around 562 nm in

f
TiO2-rGO-Ag composite, thereby indicating its absorbance in the visible region due to Ag

oo
nanoparticles incorporation into the system. Furthermore, to confirm the origin of a new peak

pr
and monitor the change, spectra were normalized at 562 nm. The addition of Ag nanoparticles

over the binary composite certainly adds to the enhanced absorption and reduction in the
e-
bandgap. The reduced bandgap is the primary reasons for absorption in the visible region of
Pr
radiation, hence the photoactivity of the nanocomposites for organic compounds in this region.
n al
ur
Jo

Fig. 6. Solid-state UV-Visible spectra of Graphene, Titania, TiO2-rGO, and TiO2-rGO-Ag

19
3.7 Raman spectra study

Raman spectroscopy is an illustrious and non-destructive analytical means for detailed analysis

and characterization of structural and electronic properties of the Graphene (in terms of both

disorder and defect structures) and the effect of doping on these layered structures [35]. The

Raman scattering spectra of Binary and ternary composites were recorded. A peak at 1354

f
corresponds to the first order D mode. D mode is also known as the breathing mode of aromatic

oo
rings, which arises from the defect in the sample doubly resonant disorder-induced mode).

pr
Another peak at 1600 cm-1 is because of the G mode, which is due to the zone center optical E 2g

phonon induced mode. The phonon induction is at the Brillouin zone center that results from the
e-
bond stretching of sp2 carbon pairs in both the rings and chains as well [61]. Bands at 143, 192,
Pr
394, 514, and 638 cm-1were assigned for Eg, Eg, B1g, B1g, and Eg Raman active modes of TiO2

(Anatase) as shown in Fig. S5 [35,49]. The enhancement in the ID/IG in the ternary composite
al

was attributed to the phenomenon called 'Surface-Enhanced Raman Scattering or SERS arising
n

due to the introduction of Ag nanoparticles [28,35].


ur

3.8 X-ray photoelectron spectroscopy (XPS)


Jo

As shown in Fig. 7, the XPS spectral analysis relates to the chemical binding energies of TiO 2-

rGO-Ag nanocomposite. The peaks at 286, 530, 460, and 367 eV show the chemical binding

energies of 1s (C), 1s (O), 2p3/2 (Ti), and 3d5/2 (Ag), respectively. The peak corresponding to 286

eV signaled because of the sp2 carbon atoms present in rGO, and the peaks ranging from 280-

300 eV represent oxygenated carbon species (C-OH and COOH). These peaks indicated a partial

reduction of Graphene oxide (GO) while carrying out the solvothermal reaction [22,49]. Peak

around 287.88 eV indicated a C-O bond, and also, there are a substitution of lattice Titanium (Ti)

20
atoms to create a Ti-O-C structure in the material. The peaks at the 459.6 and 465.2 eV were

related to the Ti 2p3/2 and Ti 2p1/2 for the Ti core level XPS spectrum. The splitting between

these two points was 5.7 eV. Core-level XPS plots for Ag 3d showed two peaks, which are

concentrated at 367.5 for Ag 3d5/2 and 373.6 eV for Ag 3d3/2 [46,62].

f
oo
pr
e-
Pr
n al
ur
Jo

Fig. 7. X-ray Photoelectron Spectrum A. Survey Spectra of TiO2-rGO-Ag; B. C 1s; C. Ti 2p; D.

Ag 3d of TiO2-rGO-Ag

4. Degradation of methylene blue

The photodegradation pattern of aqueous methylene blue solution was monitored through

solution state UV-Visible spectral studies in the presence of photocatalysts (TiO 2, TiO2-rGO, and

TiO2-rGO-Ag nanocomposites). While in the previous two, the dye was feebly decolorized due

21
to the adsorption on Graphene (TiO2-rGO) and photosensitization of the TiO2, the later ternary

composite shows enhanced photoactivity even in the visible light range. A slow visible change

was observed, and the color of the solution changed gradually from dark blue to colorless. The

entire change was monitored through UV-Visible spectroscopy over a time range. UV-Visible

spectral data for the change in the absorbance (wavelength vs. absorbance) and the ratio of

concentration change (C/Co vs. Time) were plotted (Fig. 8). Here 'C' is the dye concentration

f
changing with time (min), and '''Co' is the initial concentration. The Surface Plasmon Resonance

oo
(SPR) effect of Ag nanoparticles is stimulated under irradiation by visible light and creating a

pr
superior local electronic field in the system. A Schottky Junction slowly develops in the ternary

nanocomposite. The junction forms between Ag atoms and TiO2 coming in contact because of
e-
the higher Fermi level of TiO2. Therefore, this Schottky barrier hinders the motion of electrons
Pr
across the boundary, transferring from Ag to TiO2. These electrons are generated due to the

photo-oxidation of oxidants and facilitated the methylene blue dye degradation [31,32,63].
n al
ur
Jo

22
f
oo
pr
e-
Pr
n al
ur
Jo

Fig. 8 A. Methylene blue degradation plot B. C/Co vs. Time (min) plot for catalytic efficiency

measurement

23
5. Photodegradation of benzene and its derivatives

There has been a reduction in the concentration of targeted organic compounds with time. The

ratio of concentration change vs. time (min) was plotted for benzene, toluene, and phenol (Figure

S6). From the C/Co plot of benzene, the slow and steep degradation of benzene has been

observed. The photodegradation of toluene and phenol, however, showed an irregular pattern. In

the case of phenol, however, we can see a slower activation but faster rate of the

f
photodegradation phenomenon. The rate of photocatalytic degradation of benzene was less as

oo
compared to phenol and toluene, which may be accounted for due to the presence of active

pr
methyl and hydroxyl groups and their effects. To establish a correlation between the substituents

and photodegradation rates, we can count on the inductive effect of H, -OH, and –CH3 (σI = 0.00,
e-
0.19, and 0.04 respectively) attached to the benzene ring. The OH group has the highest positive
Pr
inductive effect among all three, which clearly indicates the higher electron-donating tendency

[16]. The photodegradation plots further impact that the ring containing the hydroxyl group
al

shows a linear degradation pattern as compared to the other two compounds. There may be a
n

presence of different catalytic sites for these processes, as mentioned in some older literature
ur

works. While –OH group tends to be more polar will attached to sites with higher polarity and
Jo

benzene and toluene will show hydrophobicity and hence adsorb at a different site [64]. These

groups tend to increase the generation of oxidants that facilitates the rate of photodegradation.

Conclusions

The ternary nanocomposite (TiO2-rGO-Ag) was synthesized successfully using the hydrothermal

method of binary nanocomposite followed by microwaved assisted deposition of silver

nanoparticles as per the methods mentioned above. The homogeneity has been achieved through

24
the probe sonication, and creased structures have been noticed over the sheets on hydrothermal

treatment. Complete reduction of Ag+ to Ag has been attained through microwave treatment,

which is quite visible in X-ray diffraction patterns. The characterizations were successfully done

using advanced characterization techniques. Ag NPs were homogenously spread over the surface

of TiO2-rGO and were visible as dark spots in the TEM images. Methylene Blue

photodegradation experiment proved the enhancement in the photocatalytic efficiency of the

f
material after the addition of Ag nanoparticles. The photocatalyst was highly active in the visible

oo
spectral region and was also applicable in the degradation of benzene, phenol, and toluene. The

pr
rate of degradation of phenol was highest and followed a linear plot. The degradation of benzene

and toluene was less and followed an irregular pattern. The higher photoresponse for phenol is
e-
certainly due to the presence of the OH group. The material holds good potential to examine for
Pr
a range of organic chemicals to be investigated for understanding the phenomenon of

photoactivity. The ternary nanocomposites have advancements over the binary nanocomposites.
al

Acknowledgment:
n

The author (A. Ojha) thanks UGC for Senior Research Fellowship. Thanking is also given to
ur

CIF, IIT (BHU) and Department of Chemistry and Physics, BHU for providing support in
Jo

instrumentation facilities. One of the authors (A.A. Ghfar) acknowledges the support by the

Distinguished Scientist Fellowship Program (DSFP-2021), King Saud University, Riyadh, Saudi

Arabia.

25
REFERENCES

[1]. A. Bernabeu, R.F. Vercher, L. Santos-juanes, P.J. Simón, C. Lardín,, M.A. Martínez, J.A.
Vicente, R. González, C. Llosá, A. Arques, A.M. Amat, Solar photocatalysis as a tertiary
treatment to remove emerging pollutants from wastewater treatment plant effluents.
Catalysis Today 161 1 (2019) 235–240.
[2]. I. Salvadó-Estivill, A. Brucato, G.L. Puma, Two-dimensional modeling of a flat-plate
photocatalytic reactor for oxidation of indoor air pollutants, Ind. Eng. Chem. Res. 46

f
(2007) 7489–7496. https://doi.org/10.1021/ie070391r.

oo
[3]. T. Pernyeszi, I. Dékány, Photocatalytic degradation of hydrocarbons by bentonite and
TiO 2 in aqueous suspensions containing surfactants, Colloids Surfaces A Physicochem.

pr
Eng. Asp. 230 (2003) 191–199. https://doi.org/10.1016/j.colsurfa.2003.09.023.
[4]. B.R. Cruz-Ortiz, J.W.J. Hamilton, C. Pablos, L. Díaz-Jiménez, D.A. Cortés-Hernández,
e-
P.K. Sharma, M. Castro-Alférez, P. Fernández-Ibañez, P.S.M. Dunlop, J.A. Byrne,
Mechanism of photocatalytic disinfection using titania-graphene composites under UV
Pr
and visible irradiation, Chem. Eng. J. 316 (2017) 179–186.
https://doi.org/10.1016/j.cej.2017.01.094.
al

[5]. P. Singh, A. Ojha, A. Borthakur, R. Singh, D. Lahiry, Emerging trends in


photodegradation of petrochemical wastes : a review, Environ. Sci. Pollut. Res. (2016)
n

22340–22364. https://doi.org/10.1007/s11356-016-7373-y.
ur

[6]. S. Jafarinejad, Environment, Recent developments in the application of sequencing batch


reactor (SBR) technology for the petroleum industry wastewater treatment, Chem. Int. 3
Jo

(2017) 342–350. https://doi.org/10.5281/zenodo.1473343.


[7]. D. Li, Perspective for smart factory in petrochemical industry, Comput. Chem. Eng. 91
(2016) 136–148. https://doi.org/10.1016/j.compchemeng.2016.03.006.
[8]. J. He, A. Kumar, M. Khan, I.M.C. Lo, Critical review of photocatalytic disinfection of
bacteria: from noble metals- and carbon nanomaterials-TiO2 composites to challenges of
water characteristics and strategic solutions, Sci. Total Environ. 758 (2021) 143953.
https://doi.org/10.1016/j.scitotenv.2020.143953.

26
[9]. W. Te Liew, A. Adhitya, R. Srinivasan, Sustainability trends in the process industries: A t
ext mining-based analysis, Comput. Ind. 65 (2014) 393–400.
https://doi.org/10.1016/j.compind.2014.01.004.
[10]. A.U. Israel, I.B. Obot, S.A. Umoren, V. Mkpenie, G.A. Ebong, Effluents and solid waste
analysis in a petrochemical company - A case study of Eleme Petrochemical Company
Ltd, Port Harcourt, Nigeria, E-Journal Chem. 5 (2008) 74–80.
https://doi.org/10.1155/2008/805957.
[11]. J. Mangalam, M. Kumar, M. Sharma, M. Joshi, High adsorptivity and visible light

f
assisted photocatalytic activity of silver/reduced graphene oxide (Ag/rGO)

oo
nanocomposite for wastewater treatment, Nano-Structures and Nano-Objects. 17 (2019)
58–66. https://doi.org/10.1016/j.nanoso.2018.11.003.

pr
[12]. M.G. Bird, H. Greim, Kaden, Benzene -Health effects & mechanisms of bone marrow
toxicity: implications for t-AML & the mode of action framework, Chem. Biol. Interact.
e-
184 (1–2) (2010) 3–6.
[13]. M.H. Fulekar, Microbial degradation of petrochemical waste-polycyclic aromatic
Pr
hydrocarbons, Bioresour. Bioprocess. 4 (2017). https://doi.org/10.1186/s40643-017-
0158-4.
al

[14]. S.K. Bhunia, N.R. Jana, Reduced graphene oxide-silver nanoparticle composite as visible
light photocatalyst for degradation of colorless endocrine disruptors, ACS Appl. Mater.
n

Interfaces. 6 (2014) 20085–20092. https://doi.org/10.1021/am505677x.


ur

[15]. S.N. Sarma, J.M. Blais, H.M. Chan, Neurotoxicity of alkylated polycyclic aromatic
compounds in human neuroblastoma cells, J. Toxicol. Environ. Heal. - Part A Curr.
Jo

Issues. 80 (2017) 285–300. https://doi.org/10.1080/15287394.2017.1314840.


[16]. C. Bougheloum, A. Messalhi, Photocatalytic Degradation of Benzene Derivatives on
TiO2 Catalyst, Phys. Procedia. 2 (2009) 1055–1058.
https://doi.org/10.1016/j.phpro.2009.11.062.
[17]. S. Mazhar, A. Ditta, L. Bulgariu, I. Ahmad, M. Ahmed, A.A. Nadiri, Sequential
treatment of paper and pulp industrial wastewater: Prediction of water quality parameters
by Mamdani Fuzzy Logic model and phytotoxicity assessment, Chemosphere. 227 (2019)
256–268. https://doi.org/10.1016/j.chemosphere.2019.04.022.

27
[18]. G Sharma, M Naushad, Adsorptive removal of noxious cadmium ions from aqueous
medium using activated carbon/zirconium oxide composite: Isotherm and kinetic
modelling, Journal of Molecular Liquids 310, 113025, 2020,
https://doi.org/10.1016/j.molliq.2020.113025
[19]. B. Long, J. Huang, X. Wang, Photocatalytic degradation of benzene in gas phase by
nanostructured BiPO4 catalysts, Prog. Nat. Sci. Mater. Int. 22 (2012) 644–653.
https://doi.org/10.1016/j.pnsc.2012.11.007.
[20]. F.J. Beltrán. Ozone Reaction Kinetics for Water & Wastewater Systems. CRC Press,

f
Florida. (2008) ISBN 1-56670-629-7.

oo
[21]. P. Singh, A. Borthakur, A review on biodegradation and photocatalytic degradation of
organic pollutants: A bibliometric and comparative analysis, J. Clean. Prod. 196 (2018)

pr
1669–1680. https://doi.org/10.1016/j.jclepro.2018.05.289.
[22]. Y. Yang, E. Liu, H. Dai, L. Kang, H. Wu, J. Fan, X. Hu, H. Liu, Photocatalytic activity of
e-
Ag-TiO2-graphene ternary nanocomposites and application in hydrogen evolution by
water splitting, Int. J. Hydrogen Energy. 39 (2014) 7664–7671.
Pr
https://doi.org/10.1016/j.ijhydene.2013.09.109.
[23]. Z. Zhu, X. Li, Q. Zhao, Z. Qu, Y. Hou, L. Zhao, S. Liu, G. Chen, FTIR study of the
al

photocatalytic degradation of gaseous benzene over UV-irradiated TiO2 nanoballs


synthesized by hydrothermal treatment in alkaline solution, Mater. Res. Bull. 45 (2010)
n

1889–1893. https://doi.org/10.1016/j.materresbull.2010.09.007.
ur

[24]. A. Fujishima, X. Zhang, D.A. Tryk, TiO2 photocatalysis and related surface phenomena,
Surf. Sci. Rep. 63 (2008) 515–582. https://doi.org/10.1016/j.surfrep.2008.10.001.
Jo

[25]. Z. Xiong, L.L. Zhang, J. Ma, X.S. Zhao, Photocatalytic degradation of dyes over
graphene-gold nanocomposites under visible light irradiation, Chem. Commun. 46 (2010)
6099–6101. https://doi.org/10.1039/c0cc01259a.
[26]. A.T. Besha, Y. Liu, C. Fang, D.N. Bekele, R. Naidu, Assessing the interactions between
micropollutants and nanoparticles in engineered and natural aquatic environments, Crit.
Rev. Environ. Sci. Technol. 50 (2020) 135–215.
https://doi.org/10.1080/10643389.2019.1629799.
[27]. J. Song, Z. Yin, Z. Yang, P. Amaladass, S. Wu, J. Ye, Y. Zhao, W.Q. Deng, H. Zhang,
X.W. Liu, Enhancement of photogenerated electron transport in dye-sensitized solar cells

28
with introduction of a reduced graphene oxide-TiO 2 junction, Chem. - A Eur. J. 17
(2011) 10832–10837. https://doi.org/10.1002/chem.201101263.
[28]. W. Gao, M. Wang, C. Ran, X. Yao, H. Yang, J. Liu, D. He, J. Bai, One-pot synthesis of
Ag/r-GO/TiO2 nanocomposites with high solar absorption and enhanced anti-
recombination in photocatalytic applications, Nanoscale 6(10) (2014) 5498–5508.
https://doi.org/10.1039/c3nr05466g.
[29]. D. Akinwande, C.J. Brennan, J.S. Bunch, P. Egberts, J.R. Felts, H. Gao, R. Huang, J.S.
Kim, T. Li, Y. Li, K.M. Liechti, N. Lu, H.S. Park, E.J. Reed, P. Wang, B.I. Yakobson, T.

f
Zhang, Y.W. Zhang, Y. Zhou, Y. Zhu, A review on mechanics and mechanical properties

oo
of 2D materials—Graphene and beyond, Extrem. Mech. Lett. 13 (2017) 42–77.
https://doi.org/10.1016/j.eml.2017.01.008.

pr
[30]. V.B. Mohan, K. tak Lau, D. Hui, D. Bhattacharyya, Graphene-based materials and their
composites: A review on production, applications and product limitations, Compos. Part
e-
B Eng. 142 (2018) 200–220. https://doi.org/10.1016/j.compositesb.2018.01.013.
[31]. A.H. Castro Neto, F. Guinea, N.M.R. Peres, K.S. Novoselov, A.K. Geim, The electronic
Pr
properties of graphene, Rev. Mod. Phys. 81 (2009) 109–162.
https://doi.org/10.1103/RevModPhys.81.109.
al

[32]. Y. Wen, H. Ding, Y. Shan, Preparation and visible light photocatalytic activity of Ag/TiO
2/graphene nanocomposite, Nanoscale. 3 (2011) 4411–4417.
n

https://doi.org/10.1039/c1nr10604j.
ur

[33]. D.D. Evanoff, G. Chumanov, Synthesis and optical properties of silver nanoparticles and
arrays, ChemPhysChem. 6 (2005) 1221–1231. https://doi.org/10.1002/cphc.200500113.
Jo

[34]. N.I. Zaaba, K.L. Foo, U. Hashim, S.J. Tan, W.W. Liu, C.H. Voon, Synthesis of Graphene
Oxide using Modified Hummers Method: Solvent Influence, Procedia Eng. 184 (2017)
469–477. https://doi.org/10.1016/j.proeng.2017.04.118.
[35]. L. Shahriary, A. a. Athawale, Graphene Oxide Synthesized by using Modified Hummers
Approach, Int. J. Renew. Energy Environ. Eng. 02 (2014) 58–63.
[36]. C.P.P. Wong, C.W. Lai, K.M. Lee, S.B. Abd Hamid, Advanced chemical reduction of
reduced graphene oxide and its photocatalytic activity in degrading reactive black 5,
Materials (Basel). 8 (2015) 7118–7128. https://doi.org/10.3390/ma8105363.

29
[37]. E.H. Alsharaeh, T. Bora, A. Soliman, F. Ahmed, G. Bharath, M.G. Ghoniem, K.M. Abu-
Salah, J. Dutta, Sol-gel-assisted microwave-derived synthesis of anatase Ag/TiO2/Go
nanohybrids toward efficient visible light phenol degradation, Catalysts. 7 (2017) 1–11.
https://doi.org/10.3390/catal7050133.
[38]. P. Periyat, N. Leyland, D.E. McCormack, J. Colreavy, D. Corr, S.C. Pillai, Rapid
microwave synthesis of mesoporous TiO2 for electrochromic displays, J. Mater. Chem.
20 (2010) 3650–3655. https://doi.org/10.1039/b924341k.
[39]. P. Huyskens, J. Mullens, A. Gomez, J. Tack, Solubility of Alcohols, Phenols and

f
Anilines in Water. Bulletin Des Sociétés Chimiques Belges 84(3) (1975) 253–262.

oo
https://doi.org/10.1002/bscb.19750840318.
[40]. R.L. Bohon, W.F. Claussen, The Solubility of Aromatic Hydrocarbons in Water, J. Am.

pr
Chem. Soc. 73 (1951) 1571–1578. https://doi.org/10.1021/ja01148a047.
[41]. L.D. Scheikunde, Alcohols, phenols and anilines in water, 64 (1975).
e-
[42]. Björkman, Thermische Klärschlammbehandlung, Schweizerische Zeitschrift Für Hydrol.
31 (1969) 632–645. https://doi.org/10.1007/BF02543692.
Pr
[43]. Y.L.F. Musico, C.M. Santos, M.L.P. Dalida, D.F. Rodrigues, Surface Modi fi cation of
Membrane Filters Using Graphene and Graphene Oxide-Based Nanomaterials for
al

Bacterial Inactivation and Removal, (2014).


[44]. Y.J. Xu, Y. Zhuang, X. Fu, New insight for enhanced photocatalytic activity of TiO2 by
n

doping carbon nanotubes: A case study on degradation of benzene and methyl orange, J.
ur

Phys. Chem. C. 114 (2010) 2669–2676. https://doi.org/10.1021/jp909855p.


[45]. P. Singh, M.C. Vishnu, K.K. Sharma, A. Borthakur, P. Srivastava, D.B. Pal, D. Tiwary,
Jo

P.K. Mishra, Photocatalytic degradation of Acid Red dye stuff in the presence of
activated carbon-TiO2 composite and its kinetic enumeration, J. Water Process Eng. 12
(2016) 20–31. https://doi.org/10.1016/j.jwpe.2016.04.007.
[46]. P. Wang, L. Han, C. Zhu, Y. Zhai, S. Dong, Aqueous-phase synthesis of Ag-TiO2-
reduced graphene oxide and Pt-TiO2-reduced graphene oxide hybrid nanostructures and
their catalytic properties, Ceram. Int. 4 (2011) 1153–1162.
https://doi.org/10.1007/s12274-011-0165-2.
[47]. E. Vasilaki, I. Georgaki, D. Vernardou, M. Vamvakaki, N. Katsarakis, Ag-loaded TiO 2
/reduced graphene oxide nanocomposites for enhanced visible-light photocatalytic

30
activity, Appl. Surf. Sci. 353 (2015) 865–872.
https://doi.org/10.1016/j.apsusc.2015.07.056.
[48]. C. Berberidou, G.Z. Kyzas, I. Paspaltsis, T. Sklaviadis, I. Poulios, Photocatalytic
disinfection and purification of water employing reduced graphene oxide/TiO2
composites, J. Chem. Technol. Biotechnol. 94 (2019) 3905–3914.
https://doi.org/10.1002/jctb.6188.
[49]. X. Pan, Y. Zhao, S. Liu, C.L. Korzeniewski, S. Wang, Z. Fan, Nanoparticle Composite
Photocatalysts, ACS Appl. Mater. Interfaces. 4 (2012) 3944–3950.

f
https://doi.org/10.1021/am300772t.

oo
[50]. Y. Shimizu, M. Ateia, M. Wang, D. Awfa, C. Yoshimura, Disinfection mechanism of E.
coli by CNT-TiO2 composites: Photocatalytic inactivation vs. physical separation,

pr
Chemosphere. 235 (2019) 1041–1049.
https://doi.org/10.1016/j.chemosphere.2019.07.006.
e-
[51]. T. Ohno, K. Sarukawa, M. Matsumura, Crystal faces of rutile and anatase TiO2 particles
and their roles in photocatalytic reactions, New J. Chem. 26 (2002) 1167–1170.
Pr
https://doi.org/10.1039/b202140d.
[52]. D.M. Tobaldi, D. Dvoranová, L. Lajaunie, N. Rozman, B. Figueiredo, M.P. Seabra, A.S.
al

Škapin, J.J. Calvino, V. Brezová, J.A. Labrincha, Graphene-TiO2 hybrids for


photocatalytic aided removal of VOCs and nitrogen oxides from outdoor environment,
n

Chem. Eng. J. 405 (2021). https://doi.org/10.1016/j.cej.2020.126651.


ur

[53]. Y. Zhang, H. Wu, J. Zhang, H. Wang, W. Lu, Enhanced photodegradation of


pentachlorophenol by single and mixed cationic and nonionic surfactants, J. Hazard.
Jo

Mater. 221–222 (2012) 92–99. https://doi.org/10.1016/j.jhazmat.2012.04.005.


[54]. A.H. Qusti, R.M. Mohamed, M. Abdel Salam, Photocatalytic synthesis of aniline from
nitrobenzene using Ag-reduced graphene oxide nanocomposite, Ceram. Int. 40 (2014).
https://doi.org/10.1016/j.ceramint.2013.10.144.
[55]. L.L. Tan, W.J. Ong, S.P. Chai, B.T. Goh, A.R. Mohamed, Visible-light-active oxygen-
rich TiO2 decorated 2D graphene oxide with enhanced photocatalytic activity toward
carbon dioxide reduction, Appl. Catal. B Environ. 179 (2015) 160–170.
https://doi.org/10.1016/j.apcatb.2015.05.024.

31
[56]. V. Štengl, D. Popelková, P. Vláčil, TiO2-graphene nanocomposite as high performace
photocatalysts, J. Phys. Chem. C. 115 (2011) 25209–25218.
https://doi.org/10.1021/jp207515z.
[57]. F.F. Bentley, L.D. Smithson, A.L. Rozek, Infrared Spectra, & Characteristic Frequencies
∼700–300 cm–1. Interscience Publishers (New York) (1968) 689-731.
[58]. V. Štengl, S. Bakardjieva, T.M. Grygar, J. Bludská, M. Kormunda, TiO2-graphene oxide
nanocomposite as advanced photocatalytic materials, Chem. Cent. J. 7 (2013) 1–12.
https://doi.org/10.1186/1752-153X-7-41.

f
[59]. E. Lam, J.H. Chong, E. Majid, Y. Liu, S. Hrapovic, A.C.W. Leung, J.H.T. Luong,

oo
Carbocatalytic dehydration of xylose to furfural in water, Carbon N. Y. 50 (2012) 1033–
1043. https://doi.org/10.1016/j.carbon.2011.10.007.

pr
[60]. T.F. Emiru, D.W. Ayele, Controlled synthesis, characterization and reduction of
graphene oxide: A convenient method for large scale production, Egypt. J. Basic Appl.
e-
Sci. 4 (2017) 74–79. https://doi.org/10.1016/j.ejbas.2016.11.002.
[61]. Tuinstra, F., Koenig, J.L., Raman Spectrum of Graphite, J. Chem. Phys. 53 (1970) 1126–
Pr
1130. https://doi.org/10.1063/1.1674108.
[62]. H. Wang, X. Qiao, J. Chen, S. Ding, Preparation of silver nanoparticles by chemical
al

reduction method, Colloids Surfaces A Physicochem. Eng. Asp. 256 (2005) 111–115.
https://doi.org/10.1016/j.colsurfa.2004.12.058.
n

[63]. R.M. Mohamed, UV-assisted photocatalytic synthesis of TiO2-reduced graphene oxide


ur

with enhanced photocatalytic activity in decomposition of sarin in gas phase, Desalin.


Water Treat. 50 (2012) 147–156. https://doi.org/10.1080/19443994.2012.708560.
Jo

[64]. C. Bouquet-Somrani, F. Fajula, A. Finiels, P. Graffin, P. Geneste, J.L. Olivé,


Photocatalytic degradative oxidation of Diuron in organic and semi-aqueous systems over
titanium dioxide catalyst, New J. Chem. 24 (2000) 999–1002.
https://doi.org/10.1039/b006024k.

32
Graphical Abstract

f
oo
pr
e-
Pr
nal
ur
Jo

33
CRediT authorship contribution statement
Ankita Ojha: Conceptualization, Writing - original draft, Writing - review & editing,
Pardeep Singh: Conceptualization, Writing - original draft, Artwork, Writing - review &
editing,
Ramesh Oran: Writing - original draft, Writing - review & editing
Dhanesh Tiwary: Writing - original draft, Writing - review & editing

Ajay K. Mishra: Writing - original draft, Writing - review & editing

f
Ayman A. Ghfar: Writing - review & editing

oo
Mu. Naushad: Writing - review & editing

pr
Tansir Ahamad: Writing - review & editing

Binota Thokchom: Writing - original draft, Writing - review & editing


e-
K. Vijayaraghavan: Writing - original draft, Writing - review & editing
Pr
Rangabhashiyam S: Writing - original draft, Writing - review & editing
n al
ur
Jo

34
Declaration of interests

☑The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

f
oo
pr
e-
Pr
n al
ur
Jo

35
Highlights:
 TiO2-rGO-Ag was synthesized via hydrothermal method assisted with microwave

 The hybrid catalyst was characterized by HR-XRD, HRTEM, HRSEM, FTIR, EDX

and XPES

 Photocatalytic degradation studies of benzene, toluene and phenol were performed

f
oo
pr
e-
Pr
n al
ur
Jo

36

You might also like