Sei sulla pagina 1di 5

Powder Technology 196 (2009) 241–245

Contents lists available at ScienceDirect

Powder Technology
j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / p ow t e c

Synthesis of TiO2–Ag nanocomposite with sol–gel method and investigation of its


antibacterial activity against E. coli
Shahab Ansari Amin, Mohammad Pazouki ⁎, Azarmidokht Hosseinnia
Materials and Energy Research Center, MeshkinDasht, Karaj, Iran

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

Article history: TiO2–Ag nanocomposite was prepared by the sol–gel method and an azeotropic distillation with benzene
Received 19 November 2008 was used for dehydration of the gel. Because of gel dehydration by distillation method a nanopowder with a
Received in revised form 23 June 2009 surface area of 230 m2/g was produced which decreased to 80 m2/g after calcination. TEM micrographs and
Accepted 29 July 2009
XRD patterns showed that spherical nanosized Ag particles (≈10 nm) were deposited among TiO2 particles.
Available online 8 August 2009
The antibacterial activity of calcined powder at 300 and 500 °C was studied in the presence and in the
Keywords:
absence of UV irradiation against Escherichia coli as a model for Gram-negative bacteria. The antibacterial
TiO2–Ag tests confirmed the powder calcined at 300 °C possessed more antibacterial activity than the pure TiO2,
Sol–gel amorphous powder and the powder calcined at 500 °C under UV irradiation. In the absence of UV, the
Antibacterial activity reduction in viable cells was observed only with calcinated powder at 300 °C.
Surface area © 2009 Elsevier B.V. All rights reserved.
Nanocomposite

1. Introduction In recent years, many researchers have studied the effect of Ag on


antibacterial activity of TiO2 [25–28]. Keleher and others [26] studied
Titanium dioxide (TiO2) is a nontoxic material and has been the antibacterial activity of silver-coated TiO2 particles against
applied in environmental treatments such as water and air purifica- Escherichia coli and Staphylococcus aureus. They found that the
tion, water disinfection and sterilization because of its unique photoreduced Ag+ played a significant role against bacteria. Sondi
properties such as strong photocatalytic activity and chemical and Sondi [23] observed that the formation of ‘pits’ damaged the E.
stability [1–6]. The mechanism of photocatalytic activity of TiO2 is coli cells' membrane and resulted in cell death. While the silver
studied much in the literature [7–9]. The most widely accepted nanoparticles were found in the cells' membrane, silver titania
mechanism is the migration of valence electron to conduction band composite applies also as bioactive films. The resulted composite
and formation of hole-electron pairs. These hole-electron pairs react films grown by CVD method exhibited both self-regeneration and
with adsorbed molecules at semiconductor surface, resulting in destruction of bacteria [27].
degradation of adsorbates [9]. Titania is able to kill microorganisms; Bacterial growth inhibition in low concentration of silver and also
therefore it is used as a biocide [10–14]. However major limitations good distribution of silver on titania makes titania an appropriate
such as UV irradiation and fast recombination of hole-electron pairs matrix for silver titania antibacterial agent [29,30]. Production of
within nanoseconds, lead in application of additives such as Pt, Pd and silver particles in nanoranges results in the generation of high surface
Au for improving the photocatalytic efficiency of TiO2 [15–18]. These area which is an important property to inhibit bacterial growth. Also,
additives capture electrons to prevent the fast recombination of hole- titania nanosize particles create more hydroxyl groups than micro-
electron pairs. Silver is a suitable and nontoxic element which sized particles. The goal of this research work was to prepare TiO2–Ag
improves the TiO2 bioactivity because of its inborn antibacterial composite in nanorange by sol–gel method to a feasible maximum
activity against different microorganisms [19–21]. Old Egyptians and antibacterial activity.
Greeks discovered the antibacterial activity of silver. It is also well-
known that silver is a toxic material to most microorganisms. It is
believed that silver ions interact with sulfur, oxygen and nitrogen in 2. Experimental
the molecules of microorganisms and inactivate the cellular proteins
and finally cause cell death [22–24]. 2.1. Materials

Titanium tetraisopropoxide (TTIP, Merck, 98% pure), silver nitrate


(AgNO3, Riedel-de Haën, 99.8% pure), ascorbic acid (C6H8O6, Merck),
⁎ Corresponding author. Tel.: +98 261 6280036; fax: +98 261 6280030. 2-propanol (Merck, 99.8%) and benzene (Merck, 99.8%) were used
E-mail addresses: mpazouki@merc.ac.ir, mpaz6@yahoo.com (M. Pazouki). without any further purification. E. coli ATCC 25922 was used as a

0032-5910/$ – see front matter © 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.powtec.2009.07.021
242 S.A. Amin et al. / Powder Technology 196 (2009) 241–245

Fig. 1. The preparation steps of TiO2–Ag nanocomposite.

model microorganism for all inactivation studies. Doubly distilled and 2.4. Antibacterial activity measurement of TiO2–Ag nanocomposite
deionised water were also used in all experiments.
E. coli ATCC 25922 was grown overnight at 37 °C in a 250 mL
conical flask containing 10 mL Luria Bertani (LB) broth and placed on a
2.2. Characterization
rotary shaker with 180 rpm for 24 h.
A suspension of 1 mL LB, 1 mL deionised water and 10 µL E. coli
X-ray diffraction (XRD) analysis of TiO2–Ag powder was carried
culture (about 109 CFU/mL) was prepared in a 4 mL quartz cubic cell.
out by a Phillips diffractometer, employing Cu–Kα radiation of
1 mL of antibacterial suspension (1 mg/mL in sterilized water) was
wavelength 1.54 . The data were taken in the range of 5–80 (2 ),
added to quartz cell. The cell was placed in the UV spectrophotometer
with a step size of 0.02°. Transmission electron microscope (TEM)
chamber and irradiated with 270 nm wavelength to study the
images were taken using a Philips CM-200 FEG performed at 200 kV.
effectiveness of UV irradiation in bacterial growth. The sample
The specific surface areas were determined by the multipoint
analysis was replicated for more accuracy. A cell suspension was
Brunauer–Emmett–Teller (BET) method using a Micromeritics In-
irradiated under the same condition without addition of any anti-
strument (Gemini 2375 model).
bacterial sample for control.
After 40 min, the samples were taken out and a serial dilution from
2.3. Synthesis of nanosized TiO2–Ag 10− 2 to 10− 5 was prepared by addition of 100 µL cell slurry to the first
tube containing 10 mL LB broth. From each dilution tube 10 and
To prepare 80 mM of TiO2 sol, 5 mL TTIP in 10 mL 2-propanol was 100 µL were spread on LB agar plates. All the experiments were
added drop-wise to 200 mL solution of deionised water and nitric acid carried out in parallel without UV irradiation to compare the
with vigorous stirring and pH was fixed at 1.5. After 24 h, the antibacterial activity of TiO2–Ag nanocomposite in the absence of
transparent TiO2 sol was obtained. To arrive at 5 wt.% of Ag onto the UV irradiation. All the cultivated plates were placed in an incubator for
TiO2 nanoparticles, a required volume of AgNO3 (0.1 M) solution was 24 h and the cell colonies formed were counted manually. The
added to the existing sol. The sol was stirred roughly for 30 min to antibacterial assessment was indeed repeated three times and the
allow adsorption of Ag+ ions onto TiO2 surface. Reduction of Ag+ was
carried out by drop-wise addition of 25 mL ascorbic acid (0.02 M),
until a color change from colorless to brownish was noted. To increase
the surface area of final powder, the dehydrating of the gel was carried
out by azeotropic distillation with benzene [32]. In this method, the
colloidal TiO2–Ag nanoparticles were heated at 110 °C to get a viscose
gel. This was transferred to a 2-neck way flask containing a volume of
benzene. The flask was placed under a Dean–Stark trap with water
circulation and temperature risen to 90 °C while continuously stirring.
When the column was filled with water–benzene mixture, the water
was extracted from the outlet tap. The yellowish solution in the flask
was filtered after 24 h and finally the prepared powder was calcined at
300 and 500 °C for 2 h in air. Roughly, 20 mL of pure TiO2 sol also dried
at 110 °C to get TiO2 nanopowder without dehydration by azeotropic Fig. 2. The XRD pattern of TiO2–Ag of noncalcined, calcined at 300 °C and calcined at
distillation. The overall experimental procedure is presented in Fig. 1. 500 °C for 2 h in air atmosphere.
S.A. Amin et al. / Powder Technology 196 (2009) 241–245 243

Table 1 the difficulty of silver characterization by XRD. Therefore, disappear-


The properties of synthesized TiO2–Ag nanocomposite. ing Ag peak at 2 = 38.1° is because of TiO2 peaks at 2 = 37.8° and
Material Name Calcination Crystallite size Surface area 38.6°.
temperature (°C) (nm) (m2/g) The mean crystallite sizes of calcined powder at 300 and 500 °C
Pure TiO2 T500 500 12 31.7 were evaluated by the Scherer equation:
TiO2–Ag CB – Amorphous 230
q CB300 300 7 80 kλ
q CB500 500 11 78 d= ð1Þ
β cosθ

Where λ is 1.54 corresponding to irradiation wavelength, k is 0.9,


results were reported as an average with error bars as percentage β is the full width at half maximum of strongest line and θ is the Bragg
difference for each data point. angle of the most intense peak at a specific phase. The evaluated
crystallite sizes (d) and surface areas are shown in Table 1.
3. Results and discussion
3.2. The dehydration effect of benzene on surface area of powder
3.1. X-ray diffraction analysis
One of the advantages of the sol–gel method is the preparation of
The crystalline phase of the prepared TiO2–Ag nanoparticles was powder that has a high surface area. It's been known that the
analyzed by XRD and the image is shown in Fig. 2. The samples were azeotropic distillation in sol–gel method has extraordinarily increased
calcined at 300 and 500 °C respectively. the surface area of synthesized powder via dehydration process. In
In noncalcined XRD pattern, a peak at 2 = 25.3° matching to (101) previous published research works, the final powder was often
of anatase and 2 = 55° corresponding to (211) is produced; however, prepared by drying gel in oven [29–31]; however, we used the
the major phase of noncalcined powder is amorphous because of azeotropic distillation to extract the water molecules from the gel to
having broad base separation. The peaks assigned to Ag in non- hold a high surface area powder. This dehydration was performed by
calcined powder at 2 = 38.1° (111), 2 = 44.5° (200) and a peak at mixing benzene with gel while stirring the mixture and increasing the
2 = 64.4° (220), suggest successful reduction of Ag metal at TiO2 temperature until the benzene–water azeotrop vaporized. The vapors
surfaces. After calcinations of the sample at 500 °C, the anatase peaks were cooled by water circulation and the water was extracted from
became sharper and the crystallinity of particles improved from the outlet tap after condensation.
amorphous TiO2 to anatase phase (Fig. 2). Because of the nanosized After mixing the powder with benzene, the surface area of
silver covered by TiO2 particles, the Ag metal peaks at higher synthesized TiO2–Ag powder was increased to 230 m2 g− 1 which is
temperatures did not appear [28]. Also, the low amount of Ag to considered a high surface area. This is expected to have a special
titania (5 wt.%) and closeness of 2 peaks of Ag and titania, increased property which needs to be studied in more detail. The surface area of

Fig. 3. Transmission electron micrograph (a) and EDX pattern (b) of noncalcined powder and TEM micrograph of calcined powder at 500 °C for 2 h (c).
244 S.A. Amin et al. / Powder Technology 196 (2009) 241–245

calcined TiO2–Ag nanocomposite dehydrated by azeotropic distilla-


tion shows around 140% increase compared to the gel dried in oven.

3.3. TEM analysis

Fig. 3a, b and c shows the monograms taken by TEM from the TiO2–
Ag powder of noncalcinated and calcinated at 500 °C after 2 h exposure
in the air. Since no crystalline rings were observed in electron diffraction
of CB in comparison with observed crystalline rings after CB calcination,
it is resulted that the major phase of CB is amorphous which is similar to
XRD pattern of CB. With consideration to silver peaks which are
observed in EDX pattern of CB (Fig. 3b) and the intense peaks were
detected in XRD pattern of CB (Fig. 2), it is suggested the reduction of
silver in metal phase in TiO2 amorphous matrix. It is shown from these
figures that Ag spherical particles (dark points) were deposited with
12 nm average sizes in the TiO2 matrix (Fig. 3a and c). The crystallite size
of TiO2 increased from 2–5 nm (before calcination) to 10 nm after Fig. 5. Number of E. coli colonies grown on LB agar plates as a percentage of the number
calcination at 500 °C. The TEM crystallite sizes are very close to the of colonies grown on nanocomposite-free control plates under 40 min UV irradiation.
results calculated from the XRD patterns using the Scherer equation. (Initial colony ≈ 107 CFU/mL). Coded samples are shown in Table 1.

3.4. Antibacterial activity same in 2 different concentrations of E. coli cells. This is contradictory
with the results obtained by Sondi and Sondi [23]. Whereas antibacterial
To study the bacterial inactivation characteristics of synthesized results of sample CB300 are similar to their findings. This can be
powder, dilution method was used against E. coli bacterium. Fig. 4 attributed to the better participation of Ag and TiO2 at 300 °C. When
displays the number of survived E. coli colonies as the number of 107 CFU is in the plate, CB300 prevented bacterial growth. It is believed
colonies grew on nanocomposite-free control plates under 40 min UV that by increasing the calcination temperature from 300 to 500 °C, the
irradiation (first colony count ≈109 CFU/mL). All the samples were crystallinity of TiO2 improves, resulting in increasing the effectiveness of
dehydrated by azeotropic distillation and calcined at 300 and 500 °C TiO2 in killing bacteria; however it followed that the CB300 had more
for 2 h in air atmosphere. antibacterial activity than CB500. It was thought that by increasing the
Following Fig. 4, it is obvious that the calcined powder at 300 °C calcination temperature, the TiO2 crystals grew and covered the Ag
(CB300) had the highest antibacterial activity which inhibited bacterial nanoparticles. This finally causes the reduction of Ag contribution in
growth by 90%. In cultured plates with noncalcined powder (CB), the inhibition of bacterial growth and decreases the efficiency of killing
sizes of bacterial colonies were significantly reduced; however no bacteria. Hence, the calcined sample's temperature at 300 °C was
significant decrease in the number of colonies was noted. Because of sufficiently optimized for more antibacterial activity.
amorphous structure of CB, it is reasonable to say that the growth delay To study the antibacterial activity of samples without UV irradiation,
of E. coli is due to the presence of silver nanoparticles and not due to the all the nanocomposite samples (i.e., CB, CB300 and CB500) and TiO2
TiO2 nanoparticles. The calcinated TiO2-free silver nanoparticles at 500 °C nanoparticles without Ag (i.e., T500) were experimented against E. coli
inhibited the bacterial growth by 30%. Although the duration of UV in the daylight. Table 2 shows that in all of the samples except CB300 no
irradiation was partially shortened (40 min in 270 nm), the efficiency of reduction of initial E. coli colonies, i.e. 1.2 × 109 CFU/mL was observed
antibacterial activity of pure TiO2 was retained at acceptable levels. This without UV irradiation, whereas CB300 was able to inhibit bacterial
can be due to the high surface area of TiO2 which results in generation of growth more than 70%, i.e. 6.3 × 108 CFU/mL compare to 1.2 × 109 CFU/
more hydroxyl radicals to decompose the E. coli cells [33,34]. Fig. 5 shows mL. In the presence of UV irradiation the percentage of CB300
the antibacterial activity of samples in less concentration of E. coli antibacterial activity was increased to 95%, i.e. 1.05 × 108 CFU/mL
(≈107 CFU/mL) which was prepared by a routine dilution. It can be seen compare to 1.2 × 109 CFU/mL. It is also indicated that since the period of
that the antibacterial behavior of all the samples except CB300 is the UV irradiation was short (40 min) therefore no significant antibacterial
effect was observed just by UV on E. coli growth in this study. The
antibacterial behavior of CB300 regarding to rest of the samples
indicated that by addition of Ag to TiO2 and then calcination at 300 °C
the inhibition of bacterial growth can be secured even in the absence of
the UV irradiation.

4. Conclusions

In this research work, the TiO2–Ag nanocomposite powder was


synthesized by sol–gel method using a reducing agent. The antibac-
terial activity of the powder was examined against E. coli as a model of

Table 2
Inactivation of E. coli (2.1 × 109 CFU/mL) in the presence and in the absence of UV
irradiation.

Without UV irradiation Under UV irradiation (40 min, 270 nm)

Pure TiO2 No inactivation 30% inactivation (± 5%)


CB No inactivation No significant inactivation
Fig. 4. Number of E. coli colonies grown on LB agar plates as a percentage of the number
CB300 70% inactivation (±5%) 95% inactivation (± 5%)
of colonies grown on nanocomposite-free control plates under 40 min UV irradiation.
CB500 No inactivation 45% inactivation (± 5%)
(Initial colony ≈ 109 CFU/mL). Coded samples are shown in Table 1.
S.A. Amin et al. / Powder Technology 196 (2009) 241–245 245

Gram-negative bacteria. To improve the surface area of synthesized [17] L. Zang, W. Macyk, C. Lange, et al., Visible-light detoxification and charge
generation by transition metal chloride modified titania, Chem. Eur. J. 6 (2000)
powder, the azeotropic distillation method for dehydration of the gel 379–384.
was used. The antibacterial results of the nanocomposite sample [18] A. Dawson, P.V. Kamat, Semiconductor-metal nanocomposites. Photoinduced
calcined at 300 °C showed that 300 °C is an ideal temperature to reach fusion and photocatalysis of gold-capped TiO2 (TiO2/Gold) nanoparticles, J. Phys.
Chem. B 105 (2001) 960–966.
the maximum E. coli inhibition growth. The calcined TiO2–Ag powder [19] M.S. Lee, S.S. Hong, M. Mohseni, Synthesis of photocatalytic nanosized TiO2–Ag
at 300 °C was able to inhibit bacterial growth even in the absence of particles with sol–gel method using reduction agent, J. Mol. Catal., A Chem 242
UV irradiation; however, this antibacterial activity was not noted for (2005) 135–140.
[20] V. Rupa, D. Manikandan, D. Divakar, T. Sivakumar, Effect of deposition of Ag on
the amorphous sample and sample calcined at 500 °C. TiO2 nanoparticles on the photodegradation of reactive yellow-17, J. Hazard.
Mater. 147 (2007) 906–913.
References [21] S.X. Liu, Z.P. Qu, X.W. Han, C.L. Sun, A mechanism for enhanced photocatalytic
activity of silver-loaded titanium dioxide, Catal. Today 93 (2004) 877–884.
[1] M.R. Hoffmann, S.T. Martin, W. Choi, D.W. Bahnemann, Environmental applica- [22] Q.L. Feng, J. Wu, G.Q. Chen, F.Z. Cui, T.N. Kim, J.O. Kim, A mechanistic study of the
tions of semiconductor photocatalysis, Chem. Rev. 95 (1995) 69–96. antibacterial effect of silver ions on Escherichia coli and Staphylococcus aureus,
[2] A. Fujishima, T.N. Rao, D.A. Tryk, Titanium dioxide photocatalysis, J. Photochem. J. Biomed. Mater. Res. 52 (2000) 662–668.
Photobiol. C: Photochem. Rev. 1 (2000) 1–21. [23] I. Sondi, B.S. Sondi, Silver nanoparticles as antimicrobial agent: a case study on E.
[3] R.J. Watts, S. Kong, M.P. Orr, G.C. Miller, B.E. Henry, Photocatalytic inactivation of coli as a model for Gram-negative bacteria, J. Colloid Interface Sci. 275 (2004)
coliform bacteria and viruses in secondary waste water effluent, Water Res. 29 177–182.
(1995) 95–100. [24] P.K. Stoimenov, R.L. Klinger, G.L. Marchin, K.J. Klabunde, Metal oxide nanoparticles
[4] A.L. Linsebigler, G. Lu, J.T. Yates Jr., Photocatalysis on TiO2 surfaces: principles, as bactericidal agents, Langmuir 18 (2002) 6679–6686.
mechanisms, and selected results, Chem. Rev. 95 (1995) 735–758. [25] K.D. Kim, D.N. Han, J.B. Lee, H.T. Kim, Formation and characterization of Ag-
[5] B. Ohtani, Y. Ogawa, S. Nishimoto, Photocatalytic activity of amorphous anatase deposited TiO2 nanoparticles by chemical reduction method, Scr. Mater. 54 (2006)
mixture of titanium(IV) oxide particles suspended in aqueous solutions, J. Phys. 143–146.
Chem. B 101 (1997) 3746–3752. [26] H.M. Sung, J.R. Choi, H.J. Hah, S.M. Koo, Y.C. Bae, Comparison of Ag deposition
[6] S. Sakthivel, B. Neppolian, M.V. Shankar, B. Arabindoo, M. Palanichamy, V. effects on the photocatalytic activity of nanoparticulate TiO2 under visible and UV
Murugesan, Solar photocatalytic degradation of azo dye: comparison of photo- light irradiation, J. Photochem. Photobiol., A Chem. 163 (2004) 37–44.
catalytic efficiency of ZnO and TiO2, Sol. Energy Mater. Sol. Cells 77 (2003) 65–82. [27] J. Keleher, J. Bashant, N. Heldt, L. Johnson, Y. Li, Photo-catalytic preparation of silver-
[7] A. Fujishima, K. Honda, Electrochemical photolysis of water at a semiconductor coated TiO2 particles for antibacterial applications, W. J. Microbiol. Biotechnol. 18
electrode, Nature 238 (1972) 37–38. (2002) 133–139.
[8] S. Turchi, D.F. Ollis, Photocatalytic degradation of organic water contaminants: [28] L.A. Brook, P. Evans, H.A. Foster, M.E. Pemble, A. Steele, D.W. Sheel, H.M. Yates,
mechanisms involving hydroxyl radical attack, J. Catal. 122 (1990) 178–192. Highly bioactive silver and silver/titania composite films grown by chemical
[9] P.V. Kamat, D. Meisel, Nanoparticles in advanced oxidation processes, Curr. Opin. vapour deposition, J. Photochem. Photobiol., A Chem. 187 (2007) 53–63.
Colloid Interface Sci. 7 (2002) 282–287. [29] V. Subramanian, E.E. Wolf, P.V. Kamat, Influence of metal/metal ion concentration
[10] Y. Kikuchi, K. Sunada, T. Iyoda, K. Hashimoto, A. Fujishima, Photocatalytic on the photocatalytic activity of TiO2–Au composite nanoparticles, Langmuir 19
bactericidal effect of TiO2 thin films: dynamic view of the active oxygen species (2003) 469–474.
responsible for the effect, J. Photochem. Photobiol., A Chem. 106 (1997) 51–56. [30] H.Y. Ki, J.H. Kim, S.C. Kwon, S.H. Jeong, A study on multifunctional wool textiles
[11] K. Sunada, T. Watanabe, K. Hashimoto, Studies on photokilling of bacteria on TiO2 treated with nano-sized silver, J. Mater. Sci. 42 (2007) 8020–8024.
thin film, J. Photochem. Photobiol., A Chem. 156 (2003) 227–233. [31] S. Pal, Y.K. Tak, J.M. Song, Does the antibacterial activity of silver nanoparticles
[12] R. Tomada Matsunaga, T. Nakajima, H. Wake, Photochemical sterilization of microbial depend on the shape of the nanoparticle? A study of the Gram-negative bacterium
cells by semiconductor powders, FEMS Microbiol. Lett. 29 (1985) 211–214. Escherichia coli, Appl. Environ. Microbiol. 73 (2007) 1712–1720.
[13] H.N. Pham, T. McDowell, E. Wikins, Photocatalytically-mediated disinfection [32] A. Hosseinnia, M. Keyanpour-Rad, M. Kazemzad, M. Pazouki, A novel approach for
of water using TiO2 as a catalyst and spore-forming Bacillus pumilus as a model, preparation of highly crystalline anatase TiO2 nanopowder from the agglomerates,
J. Environ. Sci. Health. Part A 30 (1995) 627–636. Powder Technol. 190 (2009) 390–392.
[14] P.A. Christensen, T.P. Curtis, T.A. Egerton, S.A.M. Kosa, J.R. Tinlin, Photoelec- [33] Z. Huang, P.-C. Maness, D.M. Blake, E.J. Wolfrum, S.L. Smolinski, W.A. Jacoby,
trocatalytic and photocatalytic disinfection of E. coli suspensions by titanium Bactericidal mode of titanium dioxide photocatalysis, J. Photochem. Photobiol., A
dioxide, Appl. Catal., B Environ. 41 (2003) 371–386. Chem. 130 (2000) 163–170.
[15] H. Gerischer, A. Heller, The role of oxygen in photooxidation of organic molecules [34] C. Maneerat, Y. Hayata, Antifungal activity of TiO2 photocatalysis against Penicil-
on semiconductor particles, J. Phys. Chem. 95 (1991) 5261–5267. lium expansum in vitro and in fruit tests, Int. J. Food Microbiol. 107 (2006) 99–103.
[16] G. Rothenberger, J. Moser, M. Gratzel, N. Serpone, D.K. Sharma, Charge carrier
trapping and recombination dynamics in small semiconductor particles, J. Am.
Chem. Soc. 107 (1985) 8054–8059.

Potrebbero piacerti anche