You are on page 1of 7

Journal of Photochemistry and Photobiology A: Chemistry 294 (2014) 130–136

Contents lists available at ScienceDirect

Journal of Photochemistry and Photobiology A:


Chemistry
journal homepage: www.elsevier.com/locate/jphotochem

Synthesis and visible light photocatalytic antibacterial activity of


nickel-doped TiO2 nanoparticles against Gram-positive and
Gram-negative bacteria
Hemraj M. Yadav a , Sachin V. Otari a , Raghvendra A. Bohara a , Sawanta S. Mali c ,
Shivaji H. Pawar a , Sagar D. Delekar a,b,∗
a
Center for Interdisciplinary Research, D.Y. Patil University, Kolhapur 416006, M.S., India
b
Department of Chemistry, Shivaji University, Kolhapur 416004, M.S., India
c
Advanced Chemical Engineering Department, Chonnam National University, Gwangju 500757, South Korea

a r t i c l e i n f o a b s t r a c t

Article history: Nanocrystalline anatase titanium dioxide (TiO2 ) nanoparticles doped with nickel ions (1.0–3.0 mol%) were
Received 7 June 2014 synthesized by sol–gel method. XRD and XPS showed the proper substitutions of the few sites of Ti4+ ions
Received in revised form 25 July 2014 by Ni2+ ions in titania host lattice. Particle size was estimated from TEM analysis and found in the range of
Accepted 31 July 2014
10–12 nm. UV–vis diffuse reflectance absorption measurement of doped titania nanoparticles shows the
Available online 9 August 2014
optical absorption in the visible region; which also confirms the incorporation of nickel ions in TiO2 crystal
lattice. For photocatalytic inactivation four common bacterial pathogens, Staphylococcus aureus, Bacillus
Keywords:
subtilis, Escherichia coli, and Salmonella abony were illuminated with nickel doped-TiO2 nanoparticles.
Ni-TiO2 nanoparticles
XPS
This shows a substantial decrease in bacterial numbers. The decrease in photoluminescence intensity
Antibacterial activity with increasing dopant content reveals the higher photocatalytic inactivation and lower recombination
E. coli rate of photogenerated charge carriers. The survival number of all bacteria species is not affected in dark
S. aureus with nanoparticles and in light condition without nanoparticles.
© 2014 Elsevier B.V. All rights reserved.

1. Introduction photocatalytic antimicrobial effect of TiO2 . They demonstrated the


photokilling mechanism for microbial cell using titania based mate-
Antibacterial agents are used extensively in hospitals, health rials. Upon irradiation of photon with energy higher than its optical
care settings, water and air purifications. TiO2 photocatalyst is energy band gap, the electron(e−)/hole(h+) pairs are generated and
attracting increasing interest for the decomposition of organic pol- react with O2 and H2 O to form superoxide anion radicals (O2 •− ) and
lutants and inhibition of microorganisms present in air and water hydroxyl radicals (• OH). These oxidative species (• OH, and O2 •− )
[1–4]. Modified TiO2 based photocatalytic disinfection systems are are all highly reactive, which are considered to be the dominant
a promising and alternative technology as compared with conven- oxidative species contributing to the mineralization of microorgan-
tional methods. Conventional methods are time intensive, short ism cells [2]. The other species, such as perhydroxyl radicals (HOO• )
term effective, and cannot be standardized. TiO2 is used due to its and hydrogen peroxide (H2 O2 ) are less effective against bacteria
high photocatalytic activity, stability, and non-toxicity. [6,7].
In 1985, Matsunaga et al. [5] first reported photochemical steril- Most of the TiO2 based photocatalytic disinfection studies are
ization of Saccharomyces cerevisiae (yeast), Lactobacillus acidophilus under UV light irradiations due to its wide band gap energy of
and Escherichia coli (bacteria), and Chlorella vulgaris (green algae) 3.2 eV. However, problem in the application of bare TiO2 photocat-
in water using a Pt-TiO2 photocatalyst under metal halide lamp alyst is the low efficiency and high electron–hole recombination
irradiation. Since then, much research has been performed on the rate. For utilization of solar energy efficiently and enhancing pho-
toactivity in the visible region of light band gap narrowing is an
effective way. Recently, many studies have been done to improve
∗ Corresponding author: Department of Chemistry, Shivaji University, Kolhapur
photocatalytic properties of TiO2 by doping with transition metal
416004, M.S., India. Tel.: +91 9890291575.
elements. Transition metal ions can provide additional energy lev-
E-mail addresses: sddelekar7@rediffmail.com, hemrajy@gmail.com els within the band gap of a semiconductor. Electron transfer from
(S.D. Delekar). one of these levels to the conduction band requires lower photon

http://dx.doi.org/10.1016/j.jphotochem.2014.07.024
1010-6030/© 2014 Elsevier B.V. All rights reserved.
H.M. Yadav et al. / Journal of Photochemistry and Photobiology A: Chemistry 294 (2014) 130–136 131

energy than in the situation of an unmodified TiO2 [8]. Our previous


work demonstrates photochemical transformations by Fe-doped
TiO2 nanoparticles [9] and photocatalytic bactericidal activity using
Cu-doped TiO2 nanoparticles [10] under visible light irradiations.
Several dopant ions have been investigated as potential dopants
for bactericidal effect including copper [10,11] vanadium [12], iron
[13], and tin [14].
Although the use of TiO2 in photocatalytic disinfection and
degradation of dyes have been studied extensively, but there is no
report on photocatalytic disinfection using Ni-doped TiO2 under
visible light irradiation. The objective of the present study is to syn-
thesize Ni-TiO2 nanoparticles by sol–gel method and their use in
efficient photocatalytic antibacterial activity against Gram-positive
Staphylococcus aureus (S. aureus), Bacillus subtilis (B. subtilis) and
Gram-negative Escherichia coli (E. coli), Salmonella abony (S. abony)
under fluorescent visible light irradiation.

2. Experimental

2.1. Preparation of Ni-TiO2 photocatalyst Fig. 1. XRD patterns of undoped TiO2 and Ni-TiO2 nanoparticles.

The Ni-TiO2 photocatalysts were prepared with the modified of Industrial Microorganisms), National Chemical Laboratory
sol–gel method, which was reported previously [9]. In the prepara- (NCL), Pune, India while B. subtilis is purchased form American
tion, titanium (IV) isopropoxide was mixed with glacial acetic acid Type Culture Collection (ATCC), India.
and magnetically stirred for 5 min, and then aqueous solution of For all experiments, a suspension of each bacterial species with a
sodium dodecyl sulfate was added and again stirred for 1 h. To get concentration of 1 × 106 cfu/mL (identified by UV–vis spectropho-
the desired concentration of Ni2+ ion as a dopant (1.0–3.0 mol%) tometer) was used. For photocatalytic inactivation experiment, a
in TiO2 , required stoichometric amount of aqueous solution of multilamp borosilicate glass reactor having eight fluorescent tubes
NiSO4 (H2 O)6 was then added to the mixture and stirred vigorously (Philips 8 W WW T5,  > 400 nm) was used [10]. All samples were
for 2 h. The pH of the resulting solution was adjusted to 10 by adding exposed to a light intensity of ∼0.5 mW cm−2 . For all experiments
ammonia. The solution was further stirred at 60 ◦ C for 3 h and then 1 mg Ni-TiO2 nanoparticles were suspended in 5 mL saline (0.9%
filtered. The catalyst was washed with 50 mL of ethanol. After dry- of NaCl at pH 7.0) containing bacterial suspension and exposed to
ing at 110 ◦ C in oven, the catalyst was calcinated in air at 500 ◦ C visible light. 100 ␮L of bacterial suspension was pipetted at regu-
for 5 h. The color of these Ni-TiO2 nanoparticles is yellowish. The lar time intervals and spread on freshly prepared Mueller–Hinton
various samples such as 1.0 mol% Ni-doped TiO2 , 2.0 mol% Ni-doped agar plates and incubated at 37 ◦ C for 24 h. Then the grown colonies
TiO2 , and 3.0 mol% Ni-doped TiO2 are denoted as Ni1-TiO2 , Ni2-TiO2 were counted. During all these experiments, control tests were
and Ni3-TiO2 , respectively. performed under the same irradiation conditions without nanopar-
ticles and a dark experiment with nanoparticles without light
2.2. Characterizations exposer was carried out simultaneously. Each set of experiments
was performed in triplicate.
The X-ray diffraction patterns (XRD) of the samples were
recorded on a Bruker AXS D8-Advance X-ray diffractometer with 3. Results and discussion
Cu-K␣ radiation of wavelength 1.5406 Å in 2␪ range from 20◦ to 80◦ .
X-ray photoelectron spectroscopy (XPS) was used for determining 3.1. X-ray diffraction
the surface compositions of the photocatalysts, using a Physical
Electronics 5600 Multi-technique System with monochromatic Al Fig. 1 shows the XRD patterns of undoped TiO2 and Ni-TiO2
K ␣ radiation. UV–vis diffuse reflectance absorbance spectra of nanoparticles. The diffraction peaks at 2 = 25.07, 37.59, 47.98,
all samples were obtained using a UV–visible spectrophotome- 53.72, 54.86, 62.58, 68.54, 70.05, 75.05, and 82.65 corresponds to
ter (UV3600, Shimadzu, Japan) in the range of 200–800 nm. TEM the (101), (004), (200), (105), (211), (204), (116), (220), (215), and
image of the samples were recorded on a Tecnai F30 field emission (224), respectively for plane of tetragonal anatase TiO2 (JCPDS 21-
transmission electron microscope operating at 300 kV. Elemental 1272). There were no peaks corresponding to rutile phase or oxide
composition was determined from EDS analysis. The photolumi- of the dopant metal ions. The crystallite size was calculated using
nescence (PL) spectra of the sample were recorded by using JASCO Debye–Scherrer formula:
F.P.-750 Model, (Japan) spectrofluorometer.
0.9
D= (1)
ˇ cos 
2.3. Photocatalytic bactericidal activity
where D is the crystallite size,  is the wavelength of the X-ray
All solutions and reagents were prepared with distilled and radiation (CuKa = 0.15418 nm),  is the diffraction angle and ˇ is the
deionized water. All glassware was washed with distilled water full width at half maximum (FWHM). The calculated crystallite sizes
and autoclaved at 121 ◦ C for 20 min prior to use. All experiments are 11.47 nm, 9.67 nm, 8.59 nm, and 7.96 nm for 0.0 mol%, 1.0 mol%,
were performed under sterile conditions. 2.0 mol% and 3.0 mol% doping of Ni2+ in TiO2 , respectively. It can
For the photocatalytic inactivation test, four common be seen that, with increasing dopant concentration the crystallite
pathogenic bacteria, Gram-positive Staphylococcus aureus NCIM size decreases. The decrease in crystallite size can be correlated to
2654, Bacillus subtilis ATCC 6633 and Gram-negative Escherichia coli increase in structural defects that prevent particle growth [15]. The
NCIM 2066, Salmonella abony NCIM 6017 were used. The S. aureus, ionic radius of Ni2+ ion is different from that of Ti4+ ion. Doping of
E. coli and S. abony were purchased from NCIM (National Collection Ni2+ generates oxygen vacancies in the lattice of TiO2 to maintain
132 H.M. Yadav et al. / Journal of Photochemistry and Photobiology A: Chemistry 294 (2014) 130–136

Fig. 2. XPS for the Ni3-TiO2 (a) survey; (b) Ti 2p; (c) O 1s configuration; and (d) Ni 2p configuration.

charge neutrality. In this work, the nickel dopant is incorporated by in the gas phase [19]. The Ni 2p3/2 peak position at 853.69 eV is
replacing few Ti4+ ions in TiO2 crystal lattice. These results clearly quite different from that of metallic Ni (852.7 eV), NiO (853.8 eV)
confirm the successful substitution of few Ti4+ ions by Ni2+ ions. and Ni2 O3 (856.7 eV). The binding energy difference between Ni
2p3/2 and Ni 2p1/2 core level is 18.65 eV, which is different from
3.2. X-ray photoelectron spectroscopy the value of metallic Ni (17.27 eV) and NiO (17.49 eV) [16,20,21].
The change in the Ni2p peak and a shift of Ti2p peaks corresponds
The valance state, substitution and contents of all these nickel- to the rearrangement of Ti4+ ions and Ni2+ ions. These results give
doped TiO2 nanoparticles were examined by XPS analysis as shown evidence that a few of the Ti4+ ions are successfully substituted by
in figure. The XPS survey scan spectrum of the Ni3-TiO2 sample Ni2+ ions without forming any detectable impurity phase, such as
shows the existence of Ni, O and Ti ions in an almost stoichio- Ni, Ni2 O3 and NiO. Moreover, XPS results are correlated with the
metric composition (Fig. 2a). Fig. 2b shows the core level Ti 2p EDS results, confirming Ni2+ ions are doped up to 3.0 mol% into TiO2
spectra of Ni-TiO2 sample. For TiO2 , Ti 2p3/2 and Ti 2p1/2 , peaks host lattice.
are observed at 457.21 and 462.94 eV, respectively. The splitting
between the Ti 2p1/2 and Ti 2p3/2 is 5.73 eV, demonstrating a nor- 3.3. UV–vis diffuse reflectance spectroscopy
mal state of Ti4+ ion in the sample [16]. Fig. 2c shows the XPS spectra
of O 1s configuration. The binding energies at 531.47 eV, 529.44 eV Fig. 3 shows the UV–vis absorption spectra in the range of
reveal the existence of surface hydroxyl groups and O 1s electron 300–700 nm for undoped TiO2 and Ni-TiO2 nanoparticles. The
binding energy arising from titanium lattice, respectively [17]. The absorption spectra of the doped samples show a stronger visi-
XPS spectrum shows complex structure 853.69 eV, 861.08 eV for Ni ble light absorption indicating the band gap was decreased upon
2p3/2 , and 872.34 eV, 877.96 eV for Ni 2p1/2 , respectively and these doping. This visible light absorption ascribed the fact that there is
are attributed due to the multiplet splitting (Fig.2d). The peak at a new energy level for Ni2+ /Ni+ below the conduction band and
861.08 eV is due to the O 2p → Ni 3d charge transfer transitions above the valance band edge of TiO2 [22]. Such new energy lev-
[18]. The binding energy of a Ni atom is deduced to be 853.69 eV els and oxygen vacancies generated by metal doping, induce the
on TiO2 host lattice and which is 7.39 eV smaller than atomic Ni bathochromic shift in the band gap transition and the visible light
H.M. Yadav et al. / Journal of Photochemistry and Photobiology A: Chemistry 294 (2014) 130–136 133

of a new energy level in TiO2 semiconductor. This causes the color


change in the doped samples from white to pale yellow.

3.4. Transmission electron microscopy and EDS analysis

TEM image of the Ni3-TiO2 nanoparticles and the corresponding


SAED pattern are shown in Fig. 4a and b, respectively. It is evident
from TEM micrograph that the synthesized Ni-TiO2 nanoparticles
have uneven and non-spherical particles. The particle size obtained
from TEM micrograph is within the range of 8–10 nm. This is in close
agreement with the average crystallite size obtained from XRD
analysis. The homogeneous distribution of nickel in TiO2 matrix
was confirmed from TEM micrograph. EDS analysis was performed
to insure the existence of nickel in TiO2 matrix. Fig. 4c shows the
EDS spectrum of Ni3-TiO2 nanoparticles which indicates that nickel
metal ions have been successfully integrated into TiO2 , in an atomic
ratio very close to that mentioned in the experimental section for
synthesis of Ni-TiO2 nanoparticles.

Fig. 3. UV–vis diffuse reflectance spectra of undoped TiO2 and Ni-TiO2 nanoparti- 3.5. Photoluminescence spectroscopy
cles.
The PL spectra show broad emission peaks in the region of
440–600 nm. These emission signals are due to the surface defects
and the charge transfer transition from an oxygen vacancy trapped
absorption through a charge transfer between a dopant and con- electron [24]. Interestingly, the PL spectra of Ni-TiO2 show a gradual
duction or valance band or a d–d transition in the crystal field decrease in the peak intensity with increasing nickel concentration
according the energy level [3,8,23]. The calculated band gap energy in TiO2 . The surface defects reduce the recombination rate of pho-
for TiO2 ,Ni1–TiO2 , Ni2–TiO2 , and Ni3–TiO2 , are 3.21, 2.95, 2.58, and togenerated charge carriers. The excited electrons are trapped by
2.36 eV, respectively. After doping with nickel metal ions in TiO2 , the oxygen vacancies and holes are trapped by dopant metal ions.
the absorption shifted the in visible light region due to introduction Additionally, the excited electrons can migrate from the valance

Fig. 4. (a) TEM image; (b) SAED pattern; and (c) EDS spectrum of Ni3-TiO2 nanoparticles.
134 H.M. Yadav et al. / Journal of Photochemistry and Photobiology A: Chemistry 294 (2014) 130–136

Fig. 5. The PL spectra of Ni-TiO2 nanoparticles. Fig. 6. Inactivation of S. aureus as function of time.

band to the new energy levels introduced nearer to the conduction


band by nickel doping, which also reduces the PL intensity [25].
Our PL results reveal that lower the PL intensity, a little bit higher
photocatalytic inactivation activity (Fig. 5).

4. Photocatalytic antibacterial activity

In the previous work, we reported the highest photocatalytic


antibacterial activity of copper-doped TiO2 nanoparticles under
visible light irradiation against S. aureus and E. coli bacteria [10].
Metal modified TiO2 photocatalyst attract more attention because
of its extended visible light absorption. For this reason, in this work,
the inactivation of different bacterium with nickel-doped TiO2 was
investigated in presence of visible light.
We characterized the photocatalytic inactivation of the differ-
ent Gram strain bacteria with Ni-doped TiO2 (Ni1-TiO2 , Ni2-TiO2
and Ni3-TiO2 ) nanoparticles. The photocatalytic inactivation of the
bacterial strains was increased with increased nickel dopant con-
centration in TiO2 host lattice. The survival number of all bacteria
species is not affected in dark (with nanoparticles) and light condi-
Fig. 7. Inactivation of B. subtilis as function of time.
tion (without nanoparticles). This reveals that antibacterial activity
testing method itself did not inactivate bacteria during experiment.
This characteristic pattern is observed in all tested bacterial strains.
For each of the species under this investigation the reduction in
colony forming units following irradiation in the presence of Ni-
TiO2 are displayed in Figs. 6–9. Bacterial inactivation was observed
for all four bacterial samples in presence of Ni-TiO2 with light only.
For all four bacteria, Ni3-TiO2 nanoparticles show efficient pho-
tocatalytic inactivation than Ni2-TiO2 and Ni1-TiO2 nanoparticles.
This might be due the band gap narrowing as a result of new gen-
erated energy levels within the TiO2 band gap which induces more
visible light absorption.
Fig. 6 shows the inactivation of Gram-positive S. aureus by the
photocatalytic antibacterial activity of Ni-TiO2 nanoparticles after
visible light exposer. In dark test, S. aureus cells were not inactivated
by contact with Ni-TiO2 within the experimental time scale. No S.
aureus inactivation was observed under visible light alone within
the experimental time scale. However, S. aureus inactivation was
observed when visible light was irradiated in presence of nanopar-
ticles. After 2 h light irradiation, a significant level of S. aureus
inactivation was observed. After 3 h few colonies were still found,
while for complete inactivation of S. aureus 4 h of light irradiation
Fig. 8. Inactivation of E. coli as function of time.
is required. Similar experiment was performed for Gram-positive
H.M. Yadav et al. / Journal of Photochemistry and Photobiology A: Chemistry 294 (2014) 130–136 135

membrane [31,32]. Consequently, a higher number of hydroxyl rad-


ical attacks for Gram-negative bacteria are needed for complete
bacterial inactivation. Photocatalytic inactivation is a consequence
of the combination of various parameters including band gap
energy, absorption of light, intensity of incident light, formation
of electron/hole pairs, generation of ROS, surface area, crystallinity,
phase purity, particle size, etc. Studies on the influence of several
parameters on photocatalytic inactivation of bacteria with metal
modified TiO2 nanoparticles are currently in progress.

5. Conclusions

This study demonstrates visible light assisted photocatalytic


inactivation using Ni-TiO2 nanoparticle which represents the
most effective way to remove pathogenic bacteria. Nanocrytalline
anatase TiO2 nanoparticles doped with nickel metal ions were pre-
pared by sol–gel method. The nickel metal ions were successfully
incorporated in TiO2 matrix. The UV–vis diffuse reflectance mea-
surement confirms the optical absorption enhancement in visible
region of TiO2 nanoparticles with nickel-doping. It was found that
Fig. 9. Inactivation of S. abony as function of time. the Ni-TiO2 nanoparticles exhibited higherphotocatalytic antibac-
terial property toward Gram-positive, S. aureus, B. Subtilis and to
B. subtilis inactivation under visible light irradiations using Ni-TiO2 a lesser extent of Gram-negative E. coli and S. abony. The nickel-
samples (Fig. 7). In dark and control experiments, the viability of B. modified TiO2 nanoparticles which have been tested under weak
subtilis did not seem to be affected within the experimental time visible light are a promising way for future environmental and
scale. More than 70% bacteria were inactivated within 120 min of biomedical photocatalytic disinfection applications.
irradiation by Ni3-TiO2 sample.
Additionally, photocatalytic inactivation of two Gram-negative Acknowledgements
species was also evaluated in this study. Fig. 8 shows the survival
of E. coli as a function of time. The 100% reduction time for E. coli at The authors thank the Department of Science and Technology
initial cell concentration of 2.7 × 104 cfu/mL was 300 min (Fig. 8). (DST-No.SR/FT/CS-137/2010) and University Grant Commission
Photocatalytic inactivation of S. abony under visible light irradiation (UGC-F. No. 39-728/2010-SR), New Delhi, India for financial sup-
is shown in Fig. 9. In case of S. abony, complete inactivation was port under major research schemes.
observed within 360 min of light irradiation. The time required for
complete inactivation of S. abony was higher than the time required References
for E. coli inactivation.
There is some controversy in the literature regarding the pho- [1] A. Mills, S. Le Hunte, An overview of semiconductor photocatalysis, J. Pho-
tocatalytic inactivation of the microorganisms due to different tochem. Photobiol. A: Chem. 108 (1997) 1–35.
[2] Y. Kikuchi, K. Sunada, T. Iyoda, K. Hashimoto, A. Fujishima, Photocatalytic bac-
experimental settings [26]. It is well known that the photokilling of tericidal effect of TiO2 thin films: dynamic view of the active oxygen species
bacteria is caused by the attack of reactive oxygen species [7]. It is responsible for the effect, J. Photochem. Photobiol. A: Chem. 106 (1997) 51–56.
proposed that the photocatalytic killing mechanism first damages [3] M.R. Hoffmann, S.T. Martin, W. Choi, D.W. Bahnemannt, Environmental appli-
cations of semiconductor photocatalysis, Chem. Rev. 95 (1995) 69–96.
the bacterial cells membrane. The internal bacterial components
[4] T. Matsunaga, R. Tomoda, T. Nakajima, N. Nakamura, T. Komine, Continuous-
then leak from the cells and ultimately, the photocatalytic reac- sterilization system that uses photosemiconductor powders, Appl. Environ.
tion oxidizes the cell components [6,27,28]. For a semiconductor Microbiol. 54 (1988) 1330–1333.
[5] T. Matsunaga, R. Tomoda, T. Nakajima, H. Wake, Photoelectrochemical steril-
photocatalyst, with energy provided larger than the band gap, the
ization of microbial cells by semiconductor powders, FEMS Microbiol. Lett. 29
electron/hole pairs are generated and react with O2 and H2 O to (1985) 211–214.
form superoxide anion radicals (O2 •− ) and hydroxyl radicals (• OH). [6] P.C. Maness, S. Smolinski, D.M. Blake, Z. Huang, E.J. Wolfrum, W.A. Jacoby, Bac-
These oxidative species (h+ , • OH, and O2 •− ) are all highly reactive, tericidal activity of photocatalytic TiO2 reaction: toward an understanding of
its killing mechanism, Appl. Environ. Microbiol. 65 (1999) 4094–4098.
and are considered to be the dominant oxidative species contribut- [7] D.M. Blake, P.C. Maness, Z. Huang, E.J. Wolfrum, J. Huang, W.A. Jacoby, Appli-
ing to the mineralization of microorganism cells [3]. cation of the photocatalytic chemistry of titanium dioxide to disinfection and
The results obtained in the current study show that the rate of the killing of cancer cells, Sep. Purif. Methods 28 (1999) 1–50.
[8] J. Choi, H. Park, M.R. Hoffmann, Effects of single metal–ion doping on the visible
inactivation of Gram-positive species using Ni-TiO2 nanoparticles light photoreactivity of TiO2 , J. Phys. Chem. C 114 (2010) 783–792.
in the presence of visible light is greater than that of Gram-negative [9] S.D. Delekar, H.M. Yadav, S.N. Achary, S.S. Meena, S.H. Pawar, Structural refine-
species. In all four species, for Gram-negative S. abony species much ment and photocatalytic activity of Fe-doped anatase TiO2 nanoparticles, Appl.
Surf. Sci. 263 (2012) 536–545.
higher time was required. The Gram-positive species are more [10] H.M. Yadav, S.V. Otari, V.B. Koli, S.S. Mali, C.K. Hong, S.H. Pawar, S.D. Delekar,
susceptible than the Gram-negative species and this is in agree- Preparation and characterization of copper-doped anatase TiO2 nanoparticles
ment with previously published reports [29–31]. The difference with visible light photocatalytic antibacterial activity, J. Photochem. Photobiol.
A: Chem. 280 (2014) 32–38.
is commonly ascribed to the difference in cell wall structure. The
[11] C. Karunakaran, G. Abiramasundari, P. Gomathisankar, G. Manikandan, V.
thick cell wall in Gram-positive bacteria composed of a many lay- Anandi, Cu-doped TiO2 nanoparticles for photocatalytic disinfection of bacteria
ers of peptidoglycan and teichoic acids whereas the cell wall of under visible light, J. Colloid Interface Sci. 352 (2010) 68–74.
[12] G. Fu, P.S. Vary, C.T. Lin, Anatase TiO2 nanocomposites for antimicrobial
Gram-negative bacteria is a relatively thin with an outer mem-
coatings, J. Phys. Chem. B 109 (2005) 8889–8898.
brane containing lipopolysaccharides and lipoproteins bilayers. [13] T.A. Egerton, S.A.M. Kosa, P.A. Christensen, Photoelectrocatalytic disinfection
Also this may relate to different affinities for photocatalyst and cell of E. coli suspensions by iron doped TiO2 , Phys. Chem. Chem. Phys. 8 (2006)
wall of bacteria [30]. Gram-negative bacteria are relatively more 398–406.
[14] F. Sayilkan, M. Asiltürk, N. Kiraz, E. Burunkaya, E. Arpaç, H. Sayilkan, Pho-
resistant because of the nature of their cell wall, which restricts tocatalytic antibacterial performance of Sn4+ -doped TiO2 thin films on glass
absorption of many molecules to movements through the cell substrate, J. Hazard. Mater. 162 (2009) 1309–1316.
136 H.M. Yadav et al. / Journal of Photochemistry and Photobiology A: Chemistry 294 (2014) 130–136

[15] B. Choudhury, A. Choudhury, Structural, optical and ferromagnetic properties [24] J. Yu, L. Qi, M. Jaroniec, Hydrogen production by photocatalytic water splitting
of Cr doped TiO2 nanoparticles, Mater. Sci. Eng. B 178 (2013) 794–800. over Pt/TiO2 nanosheets with exposed(001) facets, J. Phys. Chem. C 114 (2010)
[16] C.D. Wagner, W.M. Riggs, L.E. Davis, J.F. Moulder, G.E. Muilenberg, Handbook 13118–13125.
of X-Ray Photoelectron Spectroscopy, Perkin-Elmer Corp., Physical Electronics [25] H. Tang, H. Berger, P.E. Schmid, F. Lévy, G. Burri, Photoluminescence in TiO2
Division, USA, 1979. anatase single crystals, Solid State Commun. 87 (1993) 847–850.
[17] L. Li, C. Liu, Y. Liu, Study on activities of vanadium (IV/V) doped TiO2 (R) [26] S. Josset, N. Keller, M.C. Lett, M.J. Ledoux, V. Keller, Numeration methods
nanorods induced by UV and visible light, Mater. Chem. Phys. 113 (2009) for targeting photoactive materials in theUV-A – a photocatalytic removal of
551–557. microorganisms, Chem. Soc. Rev. 37 (2008) 744–755.
[18] K.S. Kim, R.E. Davis, Electron spectroscopy of the nickel–oxygen system, J. Elec- [27] K. Sunada, T. Watanabe, K. Hashimoto, Bactericidal activity of copper-deposited
tron Spectrosc. Relat. Phenom. 1 (1972) 251–258. TiO2 thin film under weak UV light illumination, Environ. Sci. Technol. 37 (2003)
[19] K. Godehusen, T. Richter, P. Zimmermann, M. Martins, 2p Photoionization of 4785–4789.
atomic Ni: a comparison with Ni metal and NiO photoionization, Phys. Rev. [28] Y.S. Kim, L.T. Linh, E.S. Park, S. Chin, G.N. Bae, J. Jurng, Antibacterial performance
Lett. 88 (2002) 217601. of TiO2 ultrafine nanopowder synthesized by a chemical vapor condensation
[20] G.H. Yu, L.R. Zeng, F.W. Zhu, C.L. Chai, W.Y. Lai, Magnetic properties and X-ray method: effect of synthesis temperature and precursor vapor concentration,
photoelectron spectroscopy study of NiO/NiFe films prepared by magnetron Powder Technol. 215–216 (2012) 195–199.
sputtering, J. Appl. Phys. 90 (2001) 4039–4043. [29] L.K. Adams, D.Y. Lyon, P.J.J. Alvarez, Comparative eco-toxicity of nanoscale TiO2 ,
[21] Z. Yin, N. Chen, F. Yang, S. Song, C. Chai, J. Zhong, H. Qian, K. Ibrahim, Structural, SiO2 , and ZnO water suspensions, Water Res. 40 (2006) 3527–3532.
magnetic properties and photoemission study of Ni-doped ZnO, Solid State [30] H.A. Foster, I.B. Ditta, S. Varghese, A. Steele, Photocatalytic disinfection using
Commun. 135 (2005) 430–433. titanium dioxide: spectrum and mechanism of antimicrobial activity, Appl.
[22] K. Ubonchonlakate, L. Sikong, T. Tontai, F. Saito, P. aeruginosa inactivation with Microbiol. Biotechnol. 90 (2011) 1847–1868.
silver and nickel doped TiO2 film coat on glass fiber roving, Adv. Mater. Res. [31] L. Caballero, K.A. Whitehead, N.S. Allen, J. Verran, Inactivation of Escherichia coli
150–151 (2010) 1726–1731. on immobilized TiO2 using fluorescent light, J. Photochem. Photobiol. A: Chem.
[23] S.D. Sharma, D. Singh, K.K. Saini, C. Kant, V. Sharma, S.C. Jain, C.P. Sharma, 202 (2009) 92–98.
Sol–gel-derived super-hydrophilic nickel doped TiO2 film as active photo- [32] J.G. Tortora, R.B. Funke, L.C. Case, Microbiology: An Introduction, 10th ed., Per-
catalyst, Appl. Catal. A: Gen. 314 (2006) 40–46. son Education Inc., New York, 2010.

You might also like