You are on page 1of 35

Accepted Manuscript

Enhanced photocatalytic and antibacterial activities of Ag-doped TiO2


nanoparticles under visible light

T. Ali, Ateeq Ahmed, Umair Alam, Imran Uddin, P. Tripathi, M. Muneer

PII: S0254-0584(18)30215-3

DOI: 10.1016/j.matchemphys.2018.03.052

Reference: MAC 20455

To appear in: Materials Chemistry and Physics

Received Date: 25 July 2017

Revised Date: 03 February 2018

Accepted Date: 16 March 2018

Please cite this article as: T. Ali, Ateeq Ahmed, Umair Alam, Imran Uddin, P. Tripathi, M. Muneer,
Enhanced photocatalytic and antibacterial activities of Ag-doped TiO2 nanoparticles under visible
light, Materials Chemistry and Physics (2018), doi: 10.1016/j.matchemphys.2018.03.052

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form.
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.
ACCEPTED MANUSCRIPT

Enhanced photocatalytic and antibacterial activities of Ag-doped


TiO2 nanoparticles under visible light
T. Ali1, Ateeq Ahmed1, Umair Alam2, Imran Uddin1, P. Tripathi*1, M. Muneer2
1Department of Applied Physics, Z.H.C.E.T., Aligarh Muslim University, Aligarh-202002, U.P, India
2Department of Chemistry, Aligarh Muslim University, Aligarh 202002, U.P, India

Abstract
In this paper, we report the photocatalytic and
antibacterial activities of nanosized Ti1-
xAgxO2 (0.00<x<0.08) photocatalyst
synthesized by sol-gel method. The X-ray
diffraction (XRD) and Raman spectra reveal
single phase anatase structure for all
substituted samples, ruling out the presence
of any secondary phase. This is well
supported by TEM micrographs which show
that shapes of the nanoparticles are non-
spherical. UV- visible spectrum illustrates
that an absorption edge shifts toward the
visible region. The 4.0 mole % of Ag-doped TiO2 shows 96% degradation of methylene blue
(MB) within 60 min under visible light irradiation and exhibit highest photocatalytic
performance. The active catalyst has superb stability and durability, maintaining its high
degradation efficiency as more as 89% after the five successive runs. An enhancement in
bactericidal activity is also observed against bacterial strain (Escherichia Coli, Pseudomonas
aeruginosa, Klebsiella pneumoniae and Enterobacter Cloacae) with increased Ag substitution.

*Author for Correspondence (P.Tripathi)


Email id: pushpendratripathi05@gmail.com
Mobile No: 08791472602
ACCEPTED MANUSCRIPT

1. Introduction
Water pollution is one of the most leading problem that badly affects the human as well as
aquatic life. The main cause of water contamination is the discharge of the industrial effluents
that contain mainly toxic chemicals and pose major threat to the living systems. It has been well
known that these pollutants are mutagenic as well carcinogenic in nature and their elimination
through the traditional techniques is a difficult task. Several techniques such as coagulation,
precipitation (heavy metals removal), flotation (oil separation), activated carbon adsorption, ion
exchange, membrane processes, reverse osmosis and electrodialysis have been employed for the
removal of organic compounds [1], but most of these are cumbersome and are not effective
enough to eliminate the contamination from the sewages. Therefore it is required to introduce
additional techniques to properly remove these compounds from wastewater. The advanced
oxidation processes (AOPs) involving heterogeneous semiconductor photocatalysts have
attracted enormous attention owing to their compatibility as a pollution mediator [2, 3]. AOPs
are defined as processes that produces highly oxidizing species, such as hydroxyl radicals and
other reactive oxygen species including superoxide anion radical, singlet oxygen and hydrogen
peroxide capable to degrade the target pollutants present in the wastewater [4]. Among all
reported semiconductor materials TiO2 is considered as a benchmark photocatalyst as well as an
antibacterial agent owing to its large surface area, high chemical stability, high catalytic activity
and low cost [5-7]. However, TiO2 possesses a wide band gap which restricts its practical
environmental application under visible light irradiation which comprises a wide range of the
solar spectrum. During photocatalysis, a semiconductor metal oxide such as TiO2 is irradiated
with light energy greater than the its band gap, resulting in photon absorption and excitation of
an electron from valence band to the conduction band, thereby generating a positively charged
hole in the valence band. These photoexcited charge carriers could participate in redox reaction
for degradation of pollutants by reacting with sorbed species [8]. The electron-hole charge
carriers may in turn undergo recombination which leads to decrease the overall performance of
the photocatalytic process. In order to solve the above mentioned problem, various techniques
have been employed such as metal and non-metal ion doping [9-12], semiconductors coupling
[13], noble metal deposition [14] and dye sensitization [15]. Among the entire techniques metal
ion doping is commonly used to obtain visible light-driven TiO2 photocatalysis owing to its high
efficiency, simplicity and facile synthesis. Noble metals display enhanced photocatalytic as well
ACCEPTED MANUSCRIPT

as antibacterial activity as they can absorb visible light due to their surface plasmon resonance
and can also act as electron traps to activate the reaction sites [16]. Thus to form an efficient
visible light-driven photocatalyst, TiO2 is doped with noble metals, such as Ag or Au. Silver ion
attracts considerable attention due to its amazing photocatalytic and antibacterial activity. It has
been reported that doping of Ag can enhance the photocatalytic as well as antibacterial activity
of TiO2 under visible light illumination by promoting electron–hole separation [17, 18]. This is
due to the fact that Ag can trap the excited electrons and therefore reducing the recombination
rate of photoinduced electron-hole pairs. Moreover, Ag NPs itself shows great potential in many
applications such as antimicrobial, drug delivery systems, biosensor and photonic devices [19-
22]. Furthermore, silver ions are capable of causing denaturation of proteins present in bacterial
cell walls and delay the bacterial growth [23]. The mechanism of the antibacterial activity of
silver NPs is associated with the formation of free radicals and resulting free-radical-induced
oxidative damage of the cell membranes of bacteria [24]. Another outstanding mechanism of
photocatalytic activity of Ag ions is that it may participate in photocatalytic oxidation reactions
between hydrogen atoms of thiol groups and oxygen molecules in the cell, that is two thiol
groups bonded together covalently by disulfide bonds (R–S–S–R), which result in blocking of
respiration and cell death of the bacteria [25]. A number of methods have been employed for the
synthesis of doped TiO2 NPs such as hydrothermal synthesis, pulsed laser deposition, diffusion
flame reactor and sol-gel etc. Among these, sol-gel [26, 27] is considered the most promising
technique owing to ease and simplicity of the process.
In this manuscript we synthesized the pure and Ag-doped TiO2 NPs using sol-gel method at the
calcinations temperatures of 400°C. The prepared samples were characterized using the various
standard analytical techniques such as X-ray diffraction (XRD), Scanning electron microscope
(SEM), Transmission electron microscope (TEM), Energy dispersive X-ray spectroscopy (EDS),
UV-visible spectroscopy, photoluminescence (PL) spectroscopy, Raman spectroscopy and
Fourier-transform infrared (FTIR). The photocatalytic activity of the prepared samples was
tested by the degradation of methylene blue under the visible light illumination. The antibacterial
activity of the synthesized materials examine against bacterial strain such as Escherichia Coli,
Pseudomonas aeruginosa, Klebsiella pneumoniae and Enterobacter Cloacae.
ACCEPTED MANUSCRIPT

2. Experimental
2.1. Material and methods
Pure and silver doped TiO2 (0, 2, 4, 6 and 8 mole %) photocatalyst were prepared by sol–gel
method. For experimental process Titanium tetraisopropoxide, silver nitrate, glacial acetic acid
and absolute ethanol all were of analytical reagent grade and purchased from commercial sources
and used without further purification.
A Stoichiometric amount of silver nitrate has been dissolved in 80 ml of deionized water with an
addition of 5 ml glacial acetic acid. Similarly a stoichiometric amount of titanium
tetraisopropoxide was mixed in 70 ml of absolute ethanol with constant stirring. After that the
two solutions were added drop-wise together during 30 minute under vigorous stirring.
Subsequently, the obtained sol was stirred constantly for two hours and was aged for 48 hours at
the room temperature. Meanwhile, the gel was formed and obtained precipitates were filtered
and washed several times with ethanol and distilled water and dried for 12 h at 100°C. The

obtained materials were ground and calcinated at 400°C for 4 h (at the heating rate of 2.5°C/min).
Finally the samples was ground again and used further for characterizations.
2.2. Structural and optical characterization
The X-ray diffraction analyses (XRD) were carried out with Bruker D2 Phaser X-ray
diffractometer using Cu-Kα radiation (λ=1.5406 Aº). The crystallite size has been determined
using Williamson–Hall formula. Raman spectroscopy has measured by means of Jobin Yvon
Horibra LABRAM-HR visible (400-1100 nm) by employing the source of argon 488 nm. The
UV-visible spectra were recorded in the range of 250-700 nm using a UV-VIS
Spectrophotometer (Perkin Elmer Lambda 35). In order to visualize the morphology and purity
of the samples TEM, SEM and EDS analyses were performed by using JEOL-JEM 2100 and
JEOL JSM-6510LV respectively. Photoluminescence spectra were measured by Fluorescence
Spectrometer (Perkin Elmer LS 55).
2.3 Photocatalytic activity assessment

The photocatalytic efficiency of as prepared catalysts has been investigated by measuring the
decomposition of MB as a model pollutant under visible light irradiation. The light used for
photocatalytic activity was supplied by a 500 W halogen lamp, which was vertically placed
inside the photo reactor. In order to prevent the IR radiation and maintain constant temperature,
ACCEPTED MANUSCRIPT

cooled water was continuously circulated through the double-walled jacket of photoreactor
equipped with refrigerator. The procedure involved in the dye preparation and its degradation
rate determination were adopted from the previous studies [28, 29]. For each test, reaction
suspensions were prepared by adding 180 mg of catalysts into a quartz tube containing 180 mL
of 10 mg/L MB aqueous solution. Prior to light illumination, the suspension was allowed to stir
for 60 min to establish the adsorption desorption equilibrium between the dye molecule and
surface of photocatalysts. At the given time intervals, 5 mL of aliquot was withdrawn and
subsequently centrifuged at 8000 rpm, then filtered it to remove the photocatalysts. The
photocatalytic degradation of MB has been estimated from the reduction in absorption intensity
of MB at the characteristic lambda max 663 nm by employing UV−visible spectrophotometer
(Perkin Elmer lambda 35). The photocatalytic efficiency was estimated using the following
expression (1).
Ct
Degradation efficiency = [1 − ] × 100 % (1)
C0

Where C0 is the initial concentration of MB, obtained before illumination and Ct is the

concentration after irradiation time, respectively. To determine the reactive species generated
during photocatalysis, trapping experiments were conducted by dissolving different scavengers
such as 1,4-benzoquinone (BQ, used to trap O2•), isopropyl alcohol (IPA, used to scavenge •OH)
and Disodium ethylenediaminetetraacetate (Na2-EDTA used to trap hole) into aqueous
suspension of MB. In this experiment, same concentration of different scavengers were added
before addition of the catalyst and the rest of the experimental conditions were similar as used in
photocatalytic activity test and previous study [30].
2.4 Antibacterial studies
Antibacterial activity of all the samples was studied using Escherichia Coli, Pseudomonas
aeruginosa, Kleissella pneumoniae and Enterobacter Cloacae bacteria under visible light
irradiation by disc diffusion method. In these Pre-inoculums of all the bacterial strains were
inoculated separately in 10 mL of Luria Bertani (LB) medium and incubated at 37 °C for 24
hours. Approximately 106 CFU/mL (colony forming unit per mL) of above stated bacterial
strains was inoculated on Luria-Bertani agar plates. Filter papers sucked with nanoparticles were
placed on the surface of seeded agar plate respectively. After 24 hours assay of incubation at 37
ACCEPTED MANUSCRIPT

°C, diameters of the inhibition zones have been determined and optical images of the plates were
taken.
3. Results and Discussion
3.1. X-ray diffraction (XRD) Analysis

Fig. 1. The XRD patterns of the pure and Ag-doped TiO2 NPs

To investigate the crystal structural of pure and Ag-TiO2 NPs calcined at 400°C, XRD analysis
have been carried out in the range of 2θ = 20-80° as shown in Fig.1. All the diffraction peaks are
corresponding to the anatase phase of TiO2 (JCPSD Card: 21-1272). It point outs that any
diffraction peaks corresponding to impurity phases such as Ag or its oxides have not been
observed even at highest doping concentration, confirming the anatase phase is not disturbed
upon doping. Moreover, the absence of impurity phase revealed the successful incorporation of
silver ion into the structure of TiO2 matrix. However, major diffraction peaks shifts towards the
lower 2θ value and becomes broader with increasing doping ion concentration which may be due
to lattice strain present in the samples. The average crystallite size (D) and the lattice strain (  )
could be obtained from the full-widths at half-maximum (FWHM) of the diffraction peaks using
Williamson–Hall relation [31],
ACCEPTED MANUSCRIPT

 Cos 1  Sin
= + (2)
 D 

Where D,  and λ are the average crystallite size, lattice strain and wavelength of X-ray
radiation (Cu Kα =1.5418Å) respectively. The Williamson–Hall plot for all the samples are
displayed in Fig. 2(a). From the linear fit to the experimental data, lattice strain is estimated from
the slope of the line and the average crystallite size can be determined from intersection of the
line with the y- axis.

Fig. 2. Williamson-Hall plot for pure and Ag-doped TiO2 NPs.


Table: 1 Average crystallite sizes and band gap energy values of Ag-doped TiO2 nanoparticles.
S.No. % Ag-doping D (nm) Lattice a (Å) c (Å) c/a V(Å3)
concentration strain
1 0 13.00 0.02285 3.783 9.534 2.520 136.44

2 2 12.20 0.02543 3.782 9.513 2.515 136.07

3 4 12.50 0.01788 3.783 9.521 2.516 136.25

4 6 9.25 0.02393 3.781 9.517 2.517 136.05


5 8 9.50 0.01750 3.780 9.513 2.516 135.92
ACCEPTED MANUSCRIPT

On the basis of these XRD data, we calculated the average crystallite size (D), lattice strain,
lattice parameters (a, b and c) and lattice volume for all samples which are displayed in table 1. It
has been observed that with increasing Ag content (i.e., from 0 to 2.0 mole %), lattice parameters
and average crystallite size decreases, whereas the lattice strain (  ) increases with decreasing
average crystallite size. Between 4.0 mole% and 8.0 mole% a small variation is observed. The
doping of metal ions in optimal concentration generally hinders the growth of crystallites [32].
The influence of Ag substitution on TiO2 particle size decline attributed to grain-boundary
restraining, by dopant ions, which limits the grain growth by the symmetry-breaking effects of
the dopant at the boundary, consequently particle sizes decreases [33]. Moreover, the repulsion
between the crystallites owing to presence of Ag ions in the crystallites may possibly responsible
for controlling crystallites growth, resulting in smaller particle size. The experimental results
suggest that 2.0 mole% doping may be the optimal doped amount of silver, which is effective to
control the crystallite growth. The excess amount of Ag ions on the surface may stimulate the
growth of crystallite of TiO2 during calcination by coalescence of neighboring crystallites
forming larger particles. Furthermore, the increase in lattice strain with decreasing crystallite size
may be occurs due to the difference in ionic radius of Ag+ (1.26 °A) and Ti4+ (0.60 °A). Owing to
this, the strain in the TiO2 nanoparticles increases at lower Ag doping concentration (i.e., 2.0
mole %) which results in decrease in crystallite size. However, the lattice constant and strain
decreases at higher Ag doping that could be due to the increase of number of interstitial Ag+ ions
which results in increase in crystallite size.
ACCEPTED MANUSCRIPT

3.2. Raman Spectroscopy Analysis

Fig.3. Raman spectra of pure and Ag-doped TiO2 NPs.

Raman spectroscopy were performed for further investigation of structural phases of pure and
Ag-doped TiO2 NPs as shown in Fig. 3. It appears that mainly five bands related to the six
Raman active modes have been observed for the anatase TiO2 (3Eg + 2B1g + A1g) with A1g mode
overlapped by the B1g peak at 515.7 cm-1 [34-36]. The Raman bands were observed at 143,
193.6, 394.2, 515.7 and 637.2 cm-1 for pure TiO2 and Ag-doped TiO2 NPs showed a similar
Raman peak pattern to that of anatase TiO2. Any peaks corresponding to silver oxide were not
observed at even highly doped sample. This means that Raman spectral results are good
agreement with XRD observation. Moreover, Ag-doped TiO2 NPs preserved the anatase structure
which suggests that Ag dopants are substitutionally incorporated into the structure of TiO2
framework. However it appears that Raman bands at 143 cm-1 (inset in figure 3) slightly shifts
towards the longer wavelength as the silver ion content increases. In general, the shift in Raman
peak occurs due to the alternation in the structure, the particle size, the nature of defects and so
on [37]. However boarding of the peaks has been noticed which may be due to the effect of
particle size on the force constants and vibrational amplitudes [38].
ACCEPTED MANUSCRIPT

3.3. UV- Visible Analysis

Fig. 4. (a) UV–visible spectra of pure and Ag-doped TiO2 NPs

In order to investigate the optical properties of Ag-doped TiO2 nanoparticles the UV-visible
absorption spectra of all the samples were measured and results are displayed in the inset of Fig.
4 (a). The incorporation of Ag ion into the TiO2 lattice result in a reduction of band gap energy.
As the amount of doping concentration of Ag ion increases, more number of photons will be
absorbed in the higher wavelength range. As a consequence, the utility range of light will be fall
in visible range, which results in the considerable improvement in photocatalytic activity of Ag-
doped TiO2 NPs. The band gap energy of synthesized materials can be expressed by the
following equation:
hν = A (hν - Eg)n /α (3)
Where α is the absorption coefficient, Eg is the band gap energy, A is a constant, which depends
on the transition probability, n corresponds to the nature of the transition, i.e., direct or indirect.
ACCEPTED MANUSCRIPT

Fig. 4. (b) Optical band gap energy plot of pure and Ag-doped TiO2 NPs

The optical band gap energy of pure and Ag-doped TiO2 NPs has been calculated by
extrapolating the linear region of the plot of hν verses (αhν)2 as shown in Fig. 4 (b). The
variations of energy band gap with doping concentration (Fig. 4(a)) demonstrates that after
doping with silver ions the band gap energy decreases, which shows the red shift. Because of this
red shift the recombination rate of photoinduced electrons and holes decreases which improved
photocatalytic activity.
ACCEPTED MANUSCRIPT

3.4. Fourier-Transform Infrared (FTIR) Spectroscopy Analysis

Fig. 5. FTIR spectra of pure and Ag-doped TiO2 NPs recorded at room temperature.

To investigate the behavior of reaction intermediate and functional groups present in the
synthesized photocatalyst FTIR Spectra were recorded in the range of 4000 – 400 cm-1 using
KBr as a medium as shown in Fig. 5. A broad band appeared at 3443 cm−1 attributed to the O-H
stretching vibration and band at 1633 cm−1 ascribed to O-H bending vibration which occurs due
to chemically adsorbed water molecules. The presence of hydroxyl group play key role in
enhancement of photocatalytic activity because the OH groups acts as main scavenger of
photogenerated electron and hole which leads the formation of hydroxyl radical (OH●) required
for the degradation of MB dye. The band in region of 2925-2855 cm-1 represents the symmetric
and antisymmetric vibrations which arise due to the presence of surfactants. The broad band at
450–850 cm-1 refers to Ti–O bending mode of vibrations which confirms the presence of metal
oxygen bonding [39].
ACCEPTED MANUSCRIPT

3.5. Scanning Electron Microscope (SEM) and Energy dispersive X-ray spectroscopy
(EDX) Analysis

Fig. 6. SEM micrographs of (a) pure TiO2 NPs and (b) 2.0 % Ag-doped TiO2 NPs (c) EDS spectrum of Ag-
doped TiO2 NPs. (d) SAED pattern of Ag-doped TiO2 NPs.

The SEM images and EDX spectrum of synthesized photocatalyst have been used to characterize
the information on surface morphology and elemental composition. The SEM micrographs of
pure and Ag-doped TiO2 nanoparticles are displayed in Figures 6(a), 6(b) and 6(c) respectively.
It has been clearly shown that the shapes of the synthesized nanoparticles are spherical with
aggregation of tiny crystals. Generally, all the Ag-doped TiO2 NPs consist of somewhat smaller
particle size than that of pure TiO2. Fig. 6(d) gives the typical EDX spectra of pure and Ag-
doped TiO2 NPs. The chemical analysis carried out by EDX shows the presence of Ag in the
doped sample along with the main constituent Ti and O. The occurrence of separate signal of
these elements suggesting that Ag ion is successfully incorporated into the TiO2 host structure.
ACCEPTED MANUSCRIPT

3.6. Transmission Electron Microscope (TEM) and Selected Area Electron Diffraction
(SAED) Analysis

Fig.7. TEM images of (a) pure TiO2 (b) 2% Ag-doped TiO2 NPs (c) 4% Ag-doped TiO2 and histograms of (d)
pure TiO2 (e) 2% Ag-doped TiO2 NPs (f) 4% Ag-doped TiO2 NPs

The surface morphology and particle size of pure and Ag-doped TiO2 photocatalyst were further
investigated by TEM as shown in Figs. 7 (a), (b) and (c) respectively. The TEM analysis reveals
the presence of non-spherical particles for both pure and doped TiO2 catalyst. However, these
micrographs also suggest that the synthesized NPs have small in size but well crystalline in
nature. The TEM results are agreed with average crystallite size obtained from XRD pattern and
the size distribution of them is represented by the histograms in Figs. 7(c), 7(d) and 7(f). From
these histograms it appears that the average particle size reduced for 2.0 mole% and 4.0 mole%
Ag-doped TiO2 NPs as compared to pure TiO2. The crystalline nature has also been studied from
SAED pattern which is displayed in Fig. 6(e) in which the ring pattern indicates the
polycrystalline nature of TiO2 NPs. Also from the electron diffraction pattern it can be
ACCEPTED MANUSCRIPT

demonstrated that ring consists of well distinct spots due to crystalline nature of TiO2 which is an
important prerequisite for the improvement of photocatalytic activity.

3.7. Photoluminescence (PL) spectroscopy Analysis

Fig. 8. Photoluminescence spectra of pure and Fe doped TiO2 nanoparticles.

Photoluminescence spectroscopy (PL) is a powerful technique to explore the electronic structure,


the transfer behavior and recombination rate of photoexcited electron–hole pairs in
semiconductors [40]. The photoluminescence (PL) spectra of pure and Ag-doped TiO2 NPs with
an excitation wavelength of 320 nm are shown in Fig. 8. It can be observed that emission peaks
of pure TiO2 are almost similar to those of Ag-doped TiO2, but the PL intensity of the Ag-doped
TiO2 is lower than that of pure TiO2. The intensity of photoluminescence spectra are directly
related to the recombination of electrons and holes, therefore lower PL intensity suggest the
delay in recombination rate which also implies that large number of photogenerated electrons
and holes are participate in the photochemical transformation, thereby enhancing the
photocatalytic activity of Ag-doped TiO2 NPs.
ACCEPTED MANUSCRIPT

3.8. Photocatalytic activity evaluation

The photocatalytic activities of Ag-doped TiO2 nanoparticles with different mole % of Ag were

investigated by decomposing MB as a model pollutant under visible light illumination. MB is a

typical cationic organic pollutant and is being extensively used in textile industry. Temporal

changes in the concentration of MB were monitored by examining the reduction in the intensity

of the MB absorption peak at λ = 664 nm as depicted in Fig. 9(a). The time courses of

photocatalytic degradation of MB in an aqueous suspension of different Ag-doped TiO2

nanoparticles under visible light irradiation were represented in the Fig. 9(b). In order to

ascertain the self-degradation and role of catalysts in the degradation of MB, a control

experiment was conducted in the absence of photocatalyst under analogous condition. The inset

of the Fig. 9(a) shows that in the absence of catalyst absorbance of MB is hardly reduced,

manifesting the stability of the dye. However, little degradation was observed in the presence of

the catalyst under dark condition, which is due to the adsorption of dye on the surface of the

catalyst. Aforementioned results indicate that light and catalyst are necessary for the effective

degradation of MB. A significant reduction in the concentration of MB with time is realized in

the presence of 4.0 mole % Ag-doped TiO2, which is displayed in Fig. 9(b). This interesting

result of MB degradation (96%) has been achieved in 60 min of irradiation time. On the other

hand, all Ag-doped TiO2 samples show the highest photocatalytic activity compared to undoped

TiO2 which suggests 30% of degradation under investigation. The reason for the 30% of MB

degradation by undoped TiO2 can be explained on the basis of self-sensitized mechanism [41].

According to this mechanism, the surface adsorbed dye molecule could excite its electrons from

HOMO to LUMO level which would subsequently transfer to the conduction band of TiO2 and

then further participate in the photocatalytic degradation reaction after converting into reactive

oxygen species. It is interesting to notice that all doped TiO2 showed enhanced photocatalytic
ACCEPTED MANUSCRIPT

activity and could degrade MB 72-96% during the course of irradiation as shown in Fig. 8 (b).

This remarkable enhanced photocatalytic activity of Ag-doped TiO2 nanoparticles could be due

to the combined effect of the dye-sensitized and synergetic action of Ag-doped TiO2 [42]. While

in the case of bare TiO2 only dye sensitization was played the major in degradation owing to

large band gap energy of TiO2 which requires only UV light to excite its electron. The extension

of optical response of TiO2 upon Ag incorporation may be due to the formation of Ti3+, which

might have formed during the oxidation of ethanol with the help of Ag or Ag+ ion as reported in

previous studies [43, 44]. During the preparation, an electron can be extracted by Ag+ ion from

ethanol and then transferred to the Ti4+ to form Ti3+ species. The reason for the formation of Ti3+

species by Ag+ doping is the large ionic radius of Ag+ (1.26 Å) compared to Ti4+ (0.68 Å) [43].

When Ag replaces Ti from its lattice, a great amount of energy is required to replace thus only a

small fraction of Ti4+ would be replaced by Ag+, leading to only small amounts of Ti3+

formation. The speculation over Ti3+ formation is confirmed by the shift in spectral response and

photocatalytic performance of doped TiO2 under visible light source. A similar phenomenon was

also observed in the previous literature where authors have stated the same reason for the

enhanced photocatalytic performance of Ag-doped TiO2 under visible light source. It is pertinent

to discuss here that the photocatalytic activity of Ag-doped TiO2 nanoparticles for degradation of

MB increases with the increase of Ag content from 2.0 mole % to 4.0 mole %. Any incorporation

after 4.0 mole % led to decrease in photocatalytic activity and a significant reduction was

observed with 8.0 mole % of Ag-doped TiO2 NPs because more silver content could be

detrimental to photonic efficiency [42, 45]. The effective charge separation and strong light

absorption capability of 4 mole % Ag- doped TiO2 NPs render it more efficient and account for

the observed increase in the degradation of MB as compared to all modified samples. It can be
ACCEPTED MANUSCRIPT

concluded that 4.0 mole % of Ag doping is effectively suppressed the recombination of charge

carriers on the surface of the catalyst in order that a large number of substrates are adsorbed on

the surface of catalysts thereby enhanced the photocatalytic activity. In addition, it is well

accepted that the reduction potential of silver ion is suitably positioned to reduce photocatalytic

Ag+ ion to Ag0, creating metallic silver on the surface of TiO2 nanoparticles [46]. The instant

color change of the catalyst during photocatalysis clearly indicates that reduction of Ag+ ions

could take place after getting electron during TiO2 excitation and self-sensitization of MB.

Above process offers the advantage of inhibition of charge carrier and favors the oxidation of the

substrate. In continuation of the above discussion, the better separation of charge carriers in case

of Ag-doped TiO2 is confirmed by PL emission spectrum of TiO2 and the results clearly verify

that 4.0 mole % Ag-doped TiO2 could effectively separate the charge carriers. The PL results are

in good agreement with photocatalytic activity results where 4.0 mole % of Ag-doped TiO2

showed the excellent photocatalytic activity, which could be attributed to the low recombination

of charge carriers.

Fig. 9. (a) Absorption spectra of MB at 60 min of irradiation time (4.0 mole % Ag-TiO2 NPs)
ACCEPTED MANUSCRIPT

Fig. 9. (b) Change in the concentration of MB in the presence of different Ag-doped TiO2 nanoparticles.

3.9 Comparison of photocatalytic activity with Degussa P25

Fig. 10. (a) Change in UV-Vis absorption spectra of MB in the presence of Degussa P25 and (b) degradation
% of MB in the presence of different samples under visible light irradiation.
The photocatalytic performance of Ag doped TiO2 was compared with Degussa P25, which is a
benchmark photocatalytic material and widely used due to its relatively high photocatalytic
activity in the degradation of organic pollutants. The excellent photocatalytic activity of Degussa
ACCEPTED MANUSCRIPT

P25 could be ascribed due to the co-presence of anatase and rutile phase in TiO2 that induces an
efficient separation of charge carriers thereby leading to higher photodegradation rate. The
photocatalytic activity of Degussa P25 was conducted under the similar conditions as used in the
case of Ag-doped TiO2. The comparison result of photocatalytic activity is shown in Fig. 10. The
activity of Degussa P25 is comparatively larger than that of the pure TiO2 but does not exceed
above than the Ag modified TiO2. The results confirm that Ag-doped TiO2 is an excellent
photocatalyst for the removal of organic pollutant under visible light irradiation.

3.10. Reusability of Catalyst

100

80
Degradation

60

40

20

0
First Second Third Fourth Fifth
Number of

Fig.11. Cycling runs of MB degradation using 4.0 mole % Ag-doped TiO2 NPs under visible light
illumination.

The reusability test was conducted with most efficient catalyst (4.0 mole % Ag-TiO2) for MB

degradation in order to check the stability of catalyst for practical application. In this experiment,

the removed catalyst after photodegradation experiment was washed with water and acetone for

multiple times followed by drying at 100 0C and then used for the further experiment. In each

cycle, the reaction time was fixed for 60 min and the results showed no observable loss of

degradation efficiency even after five cycles as depicted in Fig 11. The aforementioned results

suggest the good stability of photocatalyst which remains unchanged even after five successive

runs.
ACCEPTED MANUSCRIPT

4.0 Mechanism of Photocatalytic Processes

Fig.12. Mechanistic schematics of Ag-doped TiO2 irradiated with visible light.

In order to understand the reaction mechanism in depth and reactive species involved in the

degradation of MB, trapping experiments were carried out using different scavengers in an

aqueous solution of MB as shown in Fig.13. The 4.0 mole % Ag doped TiO2 is the best catalyst

that was used in the scavenger analysis. Before addition of catalyst, three scavengers such as

isopropyl alcohol (IPA), disodium ethylenediaminetetraacetic acid (EDTA-2Na) and 1,4-

benzoquinone (BQ) were dissolved individually into the solution of MB and were used to trap

hydroxyl radical (●OH), hole (h+) and superoxide radical (●O2-), respectively. The concentration

of all the scavengers used in this study was fixed as 2 mM and all the reaction conditions were

remained same. It can be seen (figure 13) that highest photodegradation of MB was observed in

the absence of any scavengers, due to that all reactive species generate during illumination,

participated in photodegradation of MB. When the EDTA and BQ were added into the solution,
ACCEPTED MANUSCRIPT

the photocatalytic activity is slightly reduced which manifest that hole and superoxide play a

little role for degradation of MB. Moreover, the addition of IPA made the significant reduction in

the photodegradation of MB. These results indicate that hydroxyl radicals are the predominant

active species instead of superoxide and hole in the degradation of MB over 4.0 mole% Ag-

doped TiO2 nanoparticles under visible light irradiation.

Fig. 13. Effect of different scavengers on degradation % of MB over 4 mole % Ag doped TiO2 under visible

light irradiation.

Combined with the above discussions and especially through the trapping experiments, we

propose a photocatalytic mechanism (figure 12) and explain the enhancement of photocatalytic

activity of Ag-doped TiO2 NPs under visible light source. It is well known that MB can be

degraded either by intrinsic photocatalysis or photosensitization under visible light [47, 48]. The

intrinsic photodegradation of MB is directed by the formation of charge carriers in modified


ACCEPTED MANUSCRIPT

semiconductors, while self-sensitization process could be initiated by illuminating MB dye

molecules adsorbed on the catalyst surface with wavelength larger than its λmax 665 nm, which

would transfer its electron from LUMO to CB of TiO2 NPs. The electrons present in the CB of

TiO2 after sensitization process would react with sorbed species to generate ●O2− radicals for the

oxidation of dye molecules. In our case, we have observed both the effects such as intrinsic

photocatalysis and photosensitization processes for MB degradation. The intrinsic photocatalysis

of Ag-doped TiO2 is tentatively proposed. As is well known, effective separation and efficient

absorption of visible light are crucial for enhancement of photocatalytic activity [49]. In this

regard, Ag is known to have the ability to fit in both of the above conditions and has been proven

to be an effective dopant to shift the optical response and suppress the recombination of charge

carriers [30]. As shown in Fig. 12, when Ag-doped TiO2 is exposed to visible light, an electron is

excited from the VB of TiO2 to the localized states created by Ti3+ and then to the CB, which

may in turn transfer to surface deposited Ag nanoparticles. The effective separation of charge

carriers is achieved through the above processes. The accumulated electrons in the Ag

nanoparticles would react with adsorbed oxygen to form ●O2− radicals and holes present in the

VB of TiO2 can form hydroxyl radicals by reacting with water molecules. The electron present in

Ag nanoparticles could either react with molecular oxygen or surface Ti4+ to generate ●O2− and

Ti3+ ion, respectively [50]. In both the conditions, an effective separation of charge carriers is

attained. The separated charge carriers would react with sorbed species to form oxy-radicals

which could ultimately decompose the organic pollutants. The proposed photocatalysis

mechanism for the degradation of methylene blue could be summarized as follows:

Ag-TiO2 + hν Ag-TiO2 (CB e-) + Ag-TiO2 (VB h+) (4)

Ti3+ Ti4+ + e- (5)


ACCEPTED MANUSCRIPT

h+ + H2O OH● (6)

e- + Ag Ag- (7)

Ag- + O2 ●O −
2 (8)

●O − + e- + 2H+ H2O2 (9)


2

H2O2 + e- OH● + -OH (10)

MB/MB* + OH● Degradation products (11)

4.1 Antibacterial Activity by zone inhibition assays

The antibacterial activity of Degussa P25, pure and Ag-doped TiO2 NPs were carried out by zone
inhibition assays on luria-bertani-agar, against bacterial strain Escherichia Coli, Pseudomonas
aeruginosa, Klebsiella pneumoniae and Enterobacter Cloacae as displays in Fig.14. It appears
that the as-synthesized NPs showing very nice clearance zones around the paper disc. The
commercially available P25 TiO2 showing very narrow zone of inhibition in all bacterial strains
in comparision to as synthesized pure TiO2. A well-defined zone of inhibition has been observed
in Ag-doped TiO2 NPs at lower concentration of Ag (2.0 mole%) and zone of inhibition further
increase as the concentration increases (figure14). It is evident that Ag nanoparticle carrying
antimicrobial property, enhances the antimicrobial effect and required fewer dosses for a variety
of antimicrobial applications. In contrast, silver is known to be a broad-spectrum anti-
bactericidal agent [51]. The difference in bactericidal activity could be related with the cell
membrane differences amongst Gram-positive and Gram-negative bacterial strains.
ACCEPTED MANUSCRIPT

Fig.14. Antibacterial Activity of bacterial strains by disc diffusion method for P25 TiO2; Pure TiO2 (0%); 2
% Ag-TiO2; 4 % Ag-TiO2; 6 % Ag-TiO2 and 8 % Ag-TiO2 nanoparticles.
ACCEPTED MANUSCRIPT

The bacterium was exposed to Degussa P25, pure TiO2 and Ag-doped TiO2 nanoparticles and
zone of clearance measured manually after 24 hours of incubation at 37 ºC. Fig. 15 shows the
percentage of zone of inhibition versus different doping concentration of Ag along with Degussa
P25 for different bacterial strains and results are tabulated in table 2 with standard deviation. It is
clear that at lower concentrations of Ag NPs and pure TiO2, percentage of inhibition of bacterial
cell is very less and maximum inhibition about 16% were taken place on Escherichia Coli and
Pseudomonas aeroginosa. Also, exposure of Ag doping (2.0 mole %) in TiO2 NPs showed
almost similar cell survival with little decrease. Further, increase the percentage doping of Ag
NPs showed decrease in cell survival and zone of inhibition increased and showing very nice
clearance zones around the paper disc. It is clear from Fig.15, that 8.0 mole% doping of Ag
nanoparticles showed enhanced bactericidal activity against all bacterial strains and percentage
of zone of inhibition reached almost twice in comparison to pure TiO2 nanoparticles. Thus, from
this experiment, it is clear that Ag-doped TiO2 show enhanced antibacterial activity against all
bacterial strains.
Table 2: Zone of inhibition of bacterial strains with standard deviation.

Zone of inhibition (%) ± standard deviation

Bacterial Nanoparticles (Concentration)


Strains
P25 TiO2 0% 2.0 % 4.0 % 6.0 % 8.0 %

Escherichia 12.5 ± 0.19 16.8 ±0.25 18.7 ± 0.28 21.5± 0.33 24.5 ± 0.37 26.4 ± 0.43
Coli
Pseudomonas 13.7 ± 0.41 17.7 ± 0.54 19.6 ± 0.59 22.4± 0.68 25.3 ± 0.77 27.0 ± 0.82
aeruginosa
Klebsiella 10.3 ± 0.19 14.5 ± 0.27 17.3 ± 0.32 21.4 ± 0.40 26.0 ± 0.48 30.5 ± 0.57
pneumoniae
Enterobacter 6.5± 0.20 8.5± 0.26 21.6 ± 0.68 24.5 ± 0.77 28.3 ± 0.89 29.8 ± 0.93
Cloacae
ACCEPTED MANUSCRIPT

Fig.15. Bar graph for zone of inhibition versus Ag doping concentration.

The enhancement in the antibacterial activity of the as-synthesized samples against tested
bacterial strain also supported by photocatalytic activity which is responsible for possible killing
mechanisms of bacteria. It has been shown (Fig. 4 (a)) that doping of Ag into the TiO2 lattice,
shifts the absorption peaks towards longer wavelength and due to this fact the Ag-doped TiO2
can be excited using visible irradiation. Moreover, the Ag NPs act as electron traps which
suppress the recombination of photoinduced electrons and holes and produces more hydroxyl
radicals consequently enhanced the antibacterial activity.
Conclusions
The photocatalytic activity of pure and Ag-doped TiO2 photocatalyst was examined by
performing the photodegradation of methylene blue dye. The experimental result exhibited that
Ag-doped TiO2 photocatalyst can effectively degrade MB under visible light irradiation and 4.0
mole % Ag-doped TiO2 showed the highest photocatalytic activity among all the samples
synthesized by sol-gel route. The antibacterial activity was tested against bacterial strain such as
Escherichia Coli, Pseudomonas aeruginosa, Klebsiella pneumoniae and Enterobacter Cloacae.
ACCEPTED MANUSCRIPT

In 24 hours of incubation time period, Ag-doped TiO2 NPs displayed improved antibacterial
activity against all bacterial strains. Parallel analysis by various characterization tools (XRD,
Raman, SEM, EDS, TEM, UV- visible, FTIR and PL) also demonstrated that the Ag-doped TiO2
photocatalyst showed enhanced photocatalytic and antibacterial activity. TEM analysis confirms
the presence of anatase phase structure without any other impurity phases. SEM micrographs
showed the uniform morphology in the form of tiny crystals. EDS spectrum revealed the
presence of Ag, along with to Ti and O. The FTIR spectrum shows the coordination of O2 ions
surrounding the Ti ion. UV- visible spectroscopy shows a systematic variation in absorption edge
with increasing Ag substitution.

Acknowledgments
Authors are thankful to the University Grant Commission (UGC), New Delhi, India for financial
support. We acknowledge Dr. Vasant Sathe (UGC-DAE-CSR, Indore) for providing the facility
of Micro Raman spectrometer.

REFERENCES
1. Vinod Kumar Gupta, Imran Ali, Tawfik A. Saleh, Arunima Nayak and Shilpi Agarwal,
Chemical treatment technologies for waste-water recycling—an overview, RSC Adv. 2
(2012) 6380–6388.
2. Prem Singh Saud, Bishweshwar Pant , Al-Mahumnur Alam, Zafar Khan Ghouri, Mira
Park, Hak-Yong-Kim, Carbon quantum dots anchored TiO2 nanofibers: effective
photocatalyst for waste water treatment, Ceram. Int. 41 (2015) 11953-11959.
3. Friedler, E.; Gilboa, Y., Performance of UV disinfection and the microbial quality of grey
water effluent along a reuse system for toilet flushing, Sci. Total Environ., 208 ( 9)
(2010) 2109– 2117.
4. Alex Omo Ibhadon and Paul Fitzpatrick, Heterogeneous Photocatalysis: Recent
Advances and Applications Catalysts 3 (2013) 189-218.
5. Y. Ma, X. Wang, Y. Jia, X. Chen, H. Han and C. Li, Titanium Dioxide-Based
Nanomaterials for Photocatalytic Fuel Generations, Chem. Rev. 114 (2014) 9987.
6. M. Zimbone, M. A. Buccheri, G. Cacciato, R. Sanz, G. Rappazzo, S. Boninelli, R.
Reitano, L. Romano, V. Privitera, M. G. Grimaldi, Photocatalytical and antibacterial
ACCEPTED MANUSCRIPT

activity of TiO2 nanoparticles obtained by laser ablation in water, Appl. Catal., B 165
(2015) 487-494.
7. Nandagopal S., Robin Augustine, Soney C. Georgea, Jayachandran V. P., Nandakumar
Kalarikkal, and Sabu Thomas, Gentamicin Loaded Electrospun Poly(ε-
Caprolactone)/TiO2 Nanocomposite Membranes with Antibacterial Property against
Methicillin Resistant Staphylococcus aureus, Polymer-plastics technology and
engineering 17 (2016) 1785–1796.
8. Shabina Kappadan, Tesfakiros Woldu Gebreab, SabuThomas, Nandakumar Kalarikkal,
Tetragonal BaTiO3 nanoparticles: An efficient photocatalyst for the degradation of
organic pollutants, Mater. Sci. Semicond. Process 51 (2016) 42-47.
9. Matiullah Khan, Peng Jiang, Jing Li, and Wenbin Cao, Enhanced photoelectrochemical
properties of TiO2 by codoping with tungsten and silver, J. Appl. Phys. 115 (2014)
153103.
10. Junqing Yan, Yunxia Zhang, Shengzhong Liu, Guangjun Wu, Landong Li and Naijia
Guan, Facile synthesis of an iron doped rutile TiO2 photocatalyst for enhanced visible-
light-driven water oxidation, J. Mater. Chem. A 3 (2015) 21434.
11. Sajid Ali Ansari, Mohammad Mansoob Khan, Mohd Omaish Ansari and Moo Hwan Cho
Nitrogen-doped titanium dioxide (N-doped TiO2) for visible light photocatalysis, New J.
Chem. 40 (2016) 3000.
12. Qiang Sun, Jing Zhang, Peiqing Wang, Jun Zheng, Xunni Zhang, Yanzheng Cui,
Jiangwei Feng, and Yuejin Zhu, Sulfur-doped TiO2 nanocrystalline photoanodes for dye-
sensitized solar cells, J. Renew. Sustain. Energy 4 (2012) 023104.
13. Rawal, S.B.; Bera, S.; Lee, W. I.; Jang, Du-J.; Lee, W. I., Design of Visible-Light
Photocatalysts by Coupling of Narrow Bandgap Semiconductors and TiO2: Effect of their
Relative Energy Band Positions on the Photocatalytic Efficiency, Catal. Sci. Technol. 3
(2013) 1822−1830.
14. Karácsonyi, É. Baia, L. Dombi, A. Danciu, V. Mogyorósi, K. Pop, L. C. Kovács, G.
Coşoveanu, V. Vulpoi, A. Simon, S. Pap, Zs., The Photocatalytic Activity of
TiO2/WO3/Noble Metal (Au or Pt) Nanoarchitectures Obtained by Selective
Photodeposition, Catal. Today 208 (2013) 19−27.
15. Ronca, E.; Pastore, M.; Belpassi, L.; Tarantelli, F.; Angelis, F. D., Influence of the Dye
Molecular Structure on the TiO2 Conduction Band in Dye-Sensitized Solar Cells:
ACCEPTED MANUSCRIPT

Disentangling Charge Transfer and Electrostatic Effects. Energy Environ. Sci. 6 (2013)
183−193.
16. Peng, Wang.; Baibiao, Huang.; Ying, Dai.; Whangbo, M. H., Plasmonic Photocatalysts:
Harvesting Visible Light with Noble Metal Nanoparticles. Phys. Chem. Chem. Phys. 14
(2012) 9813−9825.
17. Lidiaine M. Santos, Werick A. Machado, Marcela D. Franca, Karen A. Borges, Roberto
M. Paniago, Antonio O. T. Patrocinio and Antonio E. H. Machado, Structural
characterization of Ag-doped TiO2 with enhanced photocatalytic activity, RSC Adv. 5
(2015) 103752-103759.
18. E. Albert, P. A. Albouy, A. Ayral, P. Basa, G. Csık, N. Nagy, S. Roualdes, V. Rouessac,
G. Safran, A. Suhajda, Z. Zolnai and Z. Horvolgyi, Antibacterial properties of Ag–TiO2
composite sol-gel coatings, RSC Adv. 5 (2015) 59070.
19. Robin Augustine, Nandakumar Kalarikkal, and Sabu Thomas, A facile and rapid method
for the black pepper leaf mediated green synthesis of silver nanoparticles and the
antimicrobial study, Appl Nanosci. 4 (2014) 809–818.
20. Paige K. Brown, Ammar T. Qureshi, Alyson N. Moll, Daniel J. Hayes, and W. Todd
Monroe, Silver Nanoscale Antisense Drug Delivery System for Photoactivated Gene
Silencing, ACS Nano, 7 (4) (2013) 2948–2959
21. Lifei Chen, Huaqing Xie and Jing Li, Electrochemical glucose biosensor based on silver
nanoparticles/multiwalled carbon nanotubes modified electrode, J Solid State
Electrochem 16 (2012) 3323–3329.
22. Hyosung Choi, Seo-Jin Ko, Yuri Choi, Piljae Joo, Taehyo Kim, Bo Ram Lee, Jae-Woo
Jung, Hee Joo Choi, Myoungsik Cha, Jong-Ryul Jeong, In-Wook Hwang, Myoung Hoon
Song, Byeong-Su Kim & Jin Young Kim, Versatile surface plasmon resonance of
carbon-dot-supported silver nanoparticles in polymer optoelectronic devices, Nature
Photonics 7 (2013) 732–738.
23. Robin Augustine, Nandakumar Kalarikkal, Sabu Thomas, Electrospun PCL membranes
incorporated with biosynthesized silver nanoparticles as antibacterial wound dressings,
Appl Nanosci. 6 (2016) 337-344.
ACCEPTED MANUSCRIPT

24. Hwang, E. T., Lee, J. H.; Chae, Y. J., Kim, Y. S.; Kim, B. C., Sang, B.-I., Gu, M. B.,
Analysis of the toxic mode of action of silver nanoparticles using stress-specific
bioluminescent bacteria, small 4 (2008) 746–750.
25. Kim, J. S., Kuk, E., Yu, K. N., Kim, J.-H., Park, S. J., Lee, H. J., Kim, S. H., Park, Y. K.,
Park, Y. H., Hwang, C.-Y., Kim, Y.-K., Lee, Y.-S., Jeong, D. H., Cho, M.-H., Nanomed.:
Nanotech. Biol. Med. 3 (2007) 95–101.
26. S. Valencia, X. Vargas, L. Rios, G. Restrepo, J.M. Marín, Sol-gel and low-temperature
solvothermal synthesis of photoactive nano-titanium dioxide, J. Photochem.
Photobio. A Chem. 251 (2013) 175–181.
27. Ateeq Ahmed, T. Ali, M. Naseem Siddique, Abid Ahmad, and P. Tripathi, Enhanced
room temperature ferromagnetism in Ni doped SnO2 nanoparticles: A comprehensive
study, J. Appl. Phys. 122 (2017) 083906.
28. Umair Alam, Azam Khan, Waseem Raza, Abuzar Khan, Detlef Bahnemann, M. Muneer,
Highly efficient Y and V co-doped ZnO photocatalyst with enhanced dye sensitized
visible light, Catal. Today 284 (2016) 169-178.
29. U. Alam, A. Khan, D. Bahnemann, M. Muneer, Synthesis of iron and copper cluster-
grafted zinc oxide nanorod with enhanced visible-light-induced photocatalytic activity, J.
Colloid Interface Sci. 509 (2018) 68–72.
30. U. Alam, M. Fleisch, I. Kretschmer, D. Bahnemann, M. Muneer, One-step hydrothermal
synthesis of Bi-TiO2 nanotube/graphene composites: An efficient photocatalyst for
spectacular degradation of organic pollutants under visible light irradiation, Appl. Catal.
B 218 (2017) 758–769.
31. M. Naseem Siddique, Ateeq Ahmed, P. Tripathi, Electric transport and enhanced
dielectric permittivity in pure and Al doped NiO nanostructures, J. Alloys Compd.735
(2018) 516-529.
32. Ahmad, A.; Thiel, J.; Ismat, S. Structural Effects of Niobium and Silver Doping on
Titanium Dioxide Nanoparticles. J. Phys.: Conf. Ser. 61 (2007) 11−15.
33. Sajid I. Mogal, Vimal G. Gandhi, Manish Mishra, Shilpa Tripathi, T. Shripathi,
Pradyuman A. Joshi, and Dinesh O. Shah, Single-Step Synthesis of Silver-Doped
Titanium Dioxide: Influence of Silver on Structural, Textural, and Photocatalytic
Properties, Ind. Eng. Chem. Res. 53 (2014) 5749−5758.
ACCEPTED MANUSCRIPT

34. M. C. Mathpal, A. K. Tripathi, M. K. Singh, S. P. Gairola, S. N. Pandey and A. Agarwal,


Effect of annealing temperature on Raman spectra of TiO2 nanoparticles Chem. Phys.
Lett. 555 (2013) 182-186.
35. Lidiaine M. Santos, Werick A. Machado, Marcela D. França, Karen A. Borges, Roberto
M. Paniago, Antonio O. T. Patrocinio, Antonio E. H. Machado, Structural
characterization of Ag-doped TiO2 with enhanced photocatalytic activity, RSC Adv.
2015 DOI: 10.1039/C5RA22647C.
36. H. Fang, C. X. Zhang, L. Liu, Y. M. Zhao and H. J. Xu, Recyclable three-dimensional Ag
nanoparticle-decorated TiO2 nanorod arrays for surface-enhanced Raman scattering
Biosens. Bioelectron. 64 (2015) 434-441.
37. Z. V. Popovic, Z. Dohcevic-Mitrovic, M. Scepanovic, M. Grujic-Brojcin and S.
Askrabic, Raman scattering on nanomaterials and nanostructures, Annalen der Physik,
523 (2011) 62-74.
38. T. Ali, P Tripathi, Ameer Azam, Waseem Raza, Arham S Ahmed, Ateeq Ahmed and M
Muneer, Photocatalytic performance of Fe-doped TiO2 nanoparticles under visible light
irradiation, Mater. Res. Express 4 (2017) 015022.
39. A.N. Murashkevich, A.S. Lavitskaredrafya, T.I. Barannikova, I.M. Zharskii, Infrared
absorption spectra and structure of TiO2–SiO2 composites, J. Appl. Spectrosc. 75 (2008)
730–734.
40. Hurum, D. C., Agrios, A. G., Gray, K. A., Rajh, T., Thurnauer, M. C., Explaining the
enhanced photocatalytic activity of Degussa P25 mixed-phase TiO2 using EPR, J. Phys.
Chem. B. 107 (2003) 4545–4549.
41. Lun Pan, Ji-Jun Zou, Xiangwen Zhang, Li Wang, Water-mediated promotion of dye
sensitization of TiO2 under visible light, J. Am. Chem. Soc. 133 (2011) 10000–10002.
42. N. Sobana, M. Muruganadham1, M. Swaminathan, Nano-Ag particles doped TiO2 for
efficient photodegradation of Direct azo dyes, J. Mol. Catal. A: Chem. 258 (2006) 124–
132.
43. T.-D. Pham, B.-K. Lee, Effects of Ag doping on the photocatalytic disinfection of E. coli
in bioaerosol by Ag–TiO2/GF under visible light, J Colloid Interface Sci. 428 (2014) 24–
31
ACCEPTED MANUSCRIPT

44. S. Rengaraj, X.Z. Li, Enhanced photocatalytic activity of TiO2 by doping with Ag for
degradation of 2, 4, 6-trichlorophenol in aqueous suspension, J. Mol. Catal. Chem. 243
(2006) 60–67.
45. Selim Demirci, Tuncay Dikici, Metin Yurddaskal, Serdar Gultekin, Mustafa Toparli
Erdal Celik, Synthesis and characterization of Ag doped TiO2 heterojunction films and
their photocatalytic performances, Appl. Surf. Sci. 390 (2016) 591–601.
46. C. Sahoo, A.K. Gupta, Anjali Pal, Photocatalytic degradation of Crystal Violet (C.I.
Basic Violet 3) on silver ion doped TiO2, Dyes Pigm 66 (2005) 189-196.
47. L. Pan, J.-J.Zou, X. Zhang, and Li Wang, Water-Mediated Promotion of Dye
Sensitization of TiO2 under Visible Light, J. Am. Chem. Soc. 133 (2011) 10000–10002.
48. J. Yu, G. Dai, Q. Xiang and M. Jaroniec, Fabrication and enhanced visible-light
photocatalytic activity of carbon self-doped TiO2 sheets with exposed {001} facets, J.
Mater. Chem. 21 (2011) 1049–1057.
49. T.-D. Pham, B.-K. Lee, Photocatalytic comparison of Cu- and Ag-doped TiO2/GF for
bioaerosol disinfection under visible light, J. Solid State Chem. 232 (2015) 256–263.
50. Y. Cao, H. Tan, T. Shi, T. Tang and J. Li, Preparation of Ag-doped TiO2 nanoparticles
for photocatalytic degradation of acetamiprid in water, J. Chem Technol Biotechnol. 83
(2008) 546–552.
51. Panacek, A., Kvıtek, L., Prucek, R., Kolar, M., Vecerova, R., Pizurova, N., Sharma, V.
K., Nevecna, T., Zboril R., Silver Colloid Nanoparticles:  Synthesis, Characterization,
and Their Antibacterial Activity, J. Phys. Chem. B 110 (2006) 16248-16253.
ACCEPTED MANUSCRIPT

Highlights

 Ag- doped TiO2 nanoparticles were synthesized and characterized.


 The structural properties were analyzed by XRD and TEM.
 The photocatalytic experimental results indicate that Ag-doped TiO2 nanoparticles
effectively degrade MB under the visible light irradiation.
 An enhancement in bactericidal activity is also observed against bacterial strain.

You might also like