You are on page 1of 8

Journal of Colloid and Interface Science 336 (2009) 117124

Contents lists available at ScienceDirect

Journal of Colloid and Interface Science


www.elsevier.com/locate/jcis

Lasting antibacterial activities of AgTiO2/Ag/a-TiO2 nanocomposite thin lm photocatalysts under solar light irradiation
O. Akhavan *
Department of Physics, Sharif University of Technology, P.O. Box 11155-9161, Tehran, Iran

a r t i c l e

i n f o

a b s t r a c t
Photodegradation of Escherichia coli bacteria in presence of AgTiO2/Ag/a-TiO2 nanocomposite lm with an effective storage of silver nanoparticles was investigated in the visible and the solar light irradiations. The nanocomposite lm was synthesized by solgel deposition of 30 nm AgTiO2 layer on $200 nm anatase(a-)TiO2 lm previously doped by silver nanoparticles. Both Ag/a-TiO2 and AgTiO2/Ag/a-TiO2 lms were transparent with a SPR absorption band at 412 nm. Depth prole X-ray photoelectron spectroscopy showed metallic silver nanoparticles with diameter of 30 nm and fcc crystalline structure were self-accumulated on the lm surface at depth of 5 nm of the TiO2 layer and also at the interface of the AgTiO2 and a-TiO2 lms (at depth of 30 nm). Both OH bounds and H2O contents were concentrated on the lm surface and at the interface, as a prot in releasing more ionic (not metallic) silver nanoparticles. Antibacterial activity of the nanocomposite lm against E. coli bacteria was 5.1 times stronger than activity of the a-TiO2, in dark. Photo-antibacterial activity of the nanocomposite lm exposed by the solar light was measured 1.35 and 6.90 times better than activity of the Ag/a-TiO2 and a-TiO2, respectively. The main mechanism for silver ion releasing was inter-diffusion of water and silver nanoparticles through pores of the TiO2 layer. Durability of the nanocomposite lm was at least 11 times higher than the Ag/a-TiO2 lm. Therefore, the AgTiO2/Ag/a-TiO2 photocatalyst can be nominated as one of the effective and long-lasting antibacterial nanocomposite materials. 2009 Elsevier Inc. All rights reserved.

Article history: Received 15 December 2008 Accepted 20 March 2009 Available online 31 March 2009 Keywords: Silver nanoparticles Titanium oxide photocatalyst Solar light Photodegradation Escherichia coli bacteria

1. Introduction Nanocomposite heterogeneous photocatalysts working based on the interaction of light with the dispersed metallic nanoparticles have attracted great interest due to their high and broad photocatalytic activities. Among various metal-oxide materials being developed for photocatalytic applications, titanium dioxide (TiO2) has received great attentions because of its chemical stability and high reactivity under UV light irradiation (k < 390 nm) [15]. When ana tase TiO2 is exposed to UV light, holes hvb and excited electrons e are generated. The holes are trapped by water (H2O) or hydroxyl cb groups (OH) adsorbed on the surface to generate hydroxyl radicals (OH) [6,7] which is a powerful and indiscriminate oxidizing agent for degrading a wide range of organic pollutants [810]. After the rst study on bactericidal properties of anatase TiO2 [11], many works on inactivation of bacteria, molds, viruses, and even cancer cells have been reported [1217]. Since the post-separation of the powder catalyst in a slurry system is a main disadvantage [4,18], the immobilized TiO2 with high surface area (e.g., porous TiO2 thin lm) is more favorable in the antimicrobial applications [17,1922].
* Fax: +98 21 66022711. E-mail address: oakhavan@sharif.edu 0021-9797/$ - see front matter 2009 Elsevier Inc. All rights reserved. doi:10.1016/j.jcis.2009.03.018

On the other hand, due to the wide band gap of the TiO2 ($3.2 eV), it can only be applied in the UV region of the solar spectrum ($5% of the solar energy), while the visible light contains about 45% of the solar energy. Metal doping has long been known to be one of the most effective ways to change the intrinsic band structure of TiO2, and consequently, to improve its visible light sensitivity [2329] as well as increase its photocatalytic activity under UV irradiation [3032]. Among various dopants, noble metals (especially Ag) have received much attention for this purpose. However, before anything, silver itself is known as one of the most interesting antibacterial materials [17,3337]. It is generally believed Ag+ can bind to bacterial cell wall membrane (slightly negative), damage it and so alter its functionality [3840]. Ag+ can interact with thiol groups in proteins, resulting in inactivation of respiratory enzymes and leading to the production of reactive oxygen species [41]. In addition, because of interaction between the Ag+ and DNA structure of bacteria their multiplication may be prevented [38,39]. Ag particles of less than 10 nm are more toxic to bacteria such as Escherichia coli [42,43]. Therefore, researchers strongly aim to revive the bactericidal applications of silver (especially silver nanoparticles), due to evolution of new resistant bacteria against the common antibiotics [44]. As mentioned, silver particles enable to activate visible light excitation of TiO2 [45]. Recently, it was shown that Ag doped

118

O. Akhavan / Journal of Colloid and Interface Science 336 (2009) 117124

TiO2 highly improved photocatalytic inactivation of bacteria [4649]. For instance, only 1 wt% Ag in TiO2 reduced the reaction time required for complete removal of 107 cfu/ml E. coli from 65 to 16 min in UV-A light [47]. It is generally believed that Ag nanoparticles enhances photoactivity of TiO2 by lowering the recombination rate of its photo-excited charge carriers and/or providing more surface area for adsorption [50,51]. Visible light absorption by surface plasmon resonance of Ag nanoparticles is also thought to induce electron transfer to TiO2 resulting in charge separation and thus activation by visible light [45,52]. Therefore, AgTiO2 nanocomposites show great promise as efcient and visible light response photocatalytic materials in future. In practical applications, both high antibacterial activity and low silver release are two important characteristics for silverbased materials. High release level of silver, especially for silverbased bulk materials, leads to shortening the effective life of antibacterial activity [53]. If Ag nanoparticles and nanostructures with high antibacterial activities [5458] are immobilized on porous matrixes, the release time of silver can be delayed for a long time [59,60] so that these kinds of silver-supported materials will be of great potentials for bactericidal application. Recently, it was shown that the water contents of the solgel silica lms result in accumulation of majority of the metallic nanoparticles on surface of the aqueous and porous silica lm during a low-temperature ($100 C) heat treatment process [61]. Therefore, the same process can be used to synthesize surface accumulated Ag nanoparticles on aqueous TiO2 host layer to obtain a photocatalyst thin lm with a good sensitivity to the solar light and with strong and long life antibacterial activity. In this work, antibacterial activity and durability of solgel synthesized AgTiO2 nanocomposite layer deposited on Ag/a-TiO2 (anatase TiO2) thin lm was investigated against E. coli bacteria, in dark, visible light and the solar light. Self-accumulation of Ag nanoparticles on the lm surface was examined by depth prole X-ray photoelectron spectroscopy (XPS) and scanning electron microscopy (SEM) analyses. Moreover, the role of Ag nanoparticles in increasing the photocatalytic efciency of the AgTiO2/Ag/aTiO2 nanocomposite thin lm was investigated by using XPS and antibacterial analyses. The behavior and durability of silver ion release from the thin lms were studied in an acidic solution. A mechanism was also proposed for the ion release process.

UV irradiation. After drying the solutions containing the lms for 24 h, the as-prepared lms were washed several times with distilled water to remove the Ag ions adsorbed on the surface. Then, they were irradiated with UV light at wavelength of 255 nm for 24 h to convert the Ag ions to Ag metal. After this stage, color of the lms was light yellow. 2.3. Preparation of AgTiO2/Ag/a-TiO2 thin lms The AgTiO2 sol was prepared by dropwise adding TiCl4 to ethanol (25 ml) with a volume ratio of 1/10 and then adding DI water (25 ml) and silver nitrate (5 mg/ml AgNO3 in the solution) while the mixture was stirring. The prepared Ag/a-TiO2 thin lms were soaked in the AgTiO2 aqueous solution for 5 min and pulled up vertically at a speed of 0.2 mm/s, after cleaning with DI water, methanol and 2 h UV irradiation. Then, they were subsequently heated at 100 C for 1 h to convert the Ag ions to Ag metal. Thickness of the AgTiO2 lms grown by this method on a bare glass substrate was previously measured about 30 nm. The prepared lms were transparent with light yellow color and washed several times with DI water to remove the Ag ions adsorbed on the surface. It was also possible to convert the silver ions embedded in the TiO2 thin lm to metallic Ag particles by using the UV irradiation (similar to the last procedure). However, UV irradiation did not lead to a self-accumulation of Ag nanoparticles on surface of the TiO2 thin lm, in contrast with the heat treatment-based procedure (applied in this work) which resulted in a self-accumulation of metallic Ag nanoparticles on the lm surface. 2.4. Characterization of the thin lms Phase formation and crystalline properties of the lms were investigated by X-ray diffraction (XRD) obtained by using an XPert PRO MRD (PANalytical) analyzer with a Cu Ka radiation source. X-ray photoelectron spectroscopy (XPS) was employed to study the surface atomic composition and chemical state of the AgTiO2/ Ag/a-TiO2 thin lms. XPS data were obtained using a hemispherical analyzer with an Al Ka X-ray source (hm = 1486.6 eV) operating at a vacuum better than 107 Pa. All binding energy values were determined by calibration and xing the C(1s) line to 285 eV. Deconvolution of the XPS peaks was performed by using the SDP Ver. 4.0 software. Both survey scans and individual high-resolution scans for Ag(3d), Ti(2p), O(1s) and C(1s) peaks were recorded. XPS depth prole acquisitions were also acquired by Ar+ ion beam sputtering with incident energy of 1 keV and the sputtering rate of about 0.5 nm/min. Surface morphology of the thin lms and surface distribution of the Ag nanoparticles on the lms were studied by a Philips XL30 scanning electron microscopy (SEM). Size of the silver nanoparticles in the AgTiO2 thin lm was determined by transmission electron microscopy (TEM)-Philips CM200. The TEM sample was prepared by soaking a holey carbon coated Cu grid in the AgNO3 solution, dipping the dried grid in the AgTiO2 sol, and nally, heating the dip-coated grid at 100 C in air. The acceleration voltage of 200 kV was applied for the TEM observation. A Jasco V530 UVvis spectrophotometer was used to determine the optical absorption of the lms in the wavelength range of 300 1100 nm with 1 nm resolution. 2.5. Antibacterial test For antibacterial test of the a-TiO2, Ag/a-TiO2 and AgTiO2/Ag/aTiO2 thin lms, Escherichia coli (E. coli, ATCC 25922) was selected as a model for the Gram-negative bacteria. Before the microbiological experimentation, all glass wares and samples were sterilized by autoclaving at 120 C for 15 min. In each run, one of the synthe-

2. Materials and methods 2.1. Preparation of anatase TiO2 thin lms Solgel technique was utilized to prepare anatase TiO2 (a-TiO2) thin lms on glass substrates with dimension of 10 25 mm2. TiCl4 was added dropwise to ethanol (50 ml) with a volume ratio of 1/10 while it was stirring. TiO2 lms were obtained by dip-coating method. At rst, the substrates were cleaned by abluent, deionized (DI) water (18 MX) and acetone in turns. Then, they were immersed in the sol for about 1 min and pulled up vertically at a speed of 1 mm/s. After drying the samples at room temperature for 24 h, they were subsequently heated at 100 C for 1 h. The crystallization of the lms was occurred by heat treatment at 500 C for 60 min. Thickness of the lms was measured about 200 nm by a prolometer. The prepared a-TiO2 thin lm was nally immersed in H2SO4 solution (0.05 M) for 5 min, for roughening its surface. 2.2. Preparation of Ag/a-TiO2 thin lms To deposit Ag nanoparticles in the prepared a-TiO2 thin lm, the lm was immersed in a beaker containing AgNO3 aqueous solution (5 mg/ml), after carefully cleaning with DI water, methanol and 2 h

O. Akhavan / Journal of Colloid and Interface Science 336 (2009) 117124

119

sized thin lms was immersed in 10 ml nutrient broth inoculated by about 105 cfu (colony forming units)/ml of E. coli. The solutions containing the different samples were kept in the dark or under visible or the solar light. The visible light was obtained by a white regular uorescent lamp (10 W, wavelength peak at 425 nm) with an optical high-pass lter (k > 410 nm, L41, Kenko Co.). The light power was $1 mW/cm2 at the position of the catalyst. The bacterial photodegradation of the photocatalysts was also checked during the summer in Tehran (IRAN) at around noon under the solar light irradiation. The average solar intensity was measured $1 mW/cm2. For easily and correctly counting the grown bacterial colonies, 0.1 ml of the treated solution was diluted with DI water to a suitable volume. Then, the diluted solution was spread on a nutrient agar plate and incubated at 37 C for 48 h for counting the bacterial colonies. The reported data were the average value of three parallel runs or the values showed the best tting for an exponential reduction. The dispersion of each data point was about 50% of that point which is not so considerable in analysis of a logarithmic diagram. 2.6. Silver ions release from the AgTiO2/Ag/a-TiO2 thin lm For silver ion release detection, at rst, 10 samples of the Ag/aTiO2 or the AgTiO2/Ag/a-TiO2 nanocomposite thin lms were immersed in a ask containing 10 ml HNO3 solution (0.1 M) at room temperature. Then, the concentrations of Ag+ released were measured by anodic stripping voltammetry (ASV) of 1 ml of the prepared solution, as also previously described by others [17,62]. The ASV measurements for silver ions release were carried out using an Autolab PGSTAT100 Potentiostat/Galvanostat, using a conventional three-electrode electrochemical cell containing a glass carbon electrode and a Pt wire as the working and counter electrodes, respectively. In addition, Ag/AgCl electrode with saturated KCl solution was employed as reference electrode. The discharge potential was 0.55 V, and the stripping scan was carried out in the potential range of 00.6 V with a scan rate of 100 mV/ s. It should also to be noted that the current of released silver ions was calibrated by measuring the currents of AgNO3 solutions with various concentrations, in our experimental conditions. 3. Results and discussion Fig. 1 shows XRD patterns of the synthesized Ag/a-TiO2 and Ag TiO2/Ag/a-TiO2 thin lms. Both the patterns show diffraction peaks relating to (1 0 1), (0 0 4), (2 0 0), (1 0 5), (2 1 1), and (2 0 4) planes of anatase phase of TiO2 (a-TiO2) which is the favorable structure with better photocatalytic functional properties than its rutile phase. The mean crystalline size of the TiO2 lm was calculated $11 nm, based on the peak broadening analysis described by Scherrers equation for the (1 0 1) peak. Fig. 1a shows no peak related to crystallization of silver particles in the Ag/a-TiO2 thin lm. On the other hand, XRD pattern of the AgTiO2/Ag/a-TiO2thin lm (Fig. 1b) clearly exhibits the (2 0 0) and (2 2 0) diffraction peaks of the metallic silver. Since these peaks are attributed to face centered cubic (fcc) structure of metallic silver, the Ag(1 1 1) peak can be overwhelmed in the stronger TiO2(0 0 4) peak. The mean crystalline size of the Ag nanoparticles, determined using the Scherrers equation on (2 0 0) diffraction peak, was calculated about 12 nm. Since the AgTiO2/Ag/a-TiO2 thin lms were prepared to be applied in the antibacterial purposes, their surface analysis and so understanding their surface properties are necessary. Fig. 2 shows the high-resolution XPS peaks of Ti(2p), Ag(3d) and O(1s) core levels of the AgTiO2/Ag/a-TiO2 thin lm. The Ti(2p) XPS peak has been presented in Fig. 2a. Two peaks in the Ti(2p) binding energy region were observed. The peak located at binding energy of

a)
Intensity (arb. unit)

TiO 2 (101)

TiO 2 (004)

TiO 2 (200)

TiO 2 (105) TiO 2 (211)

10

20

30

40

50

60

TiO 2 (204) 70 60 TiO 2 (204) Ag(220) 70

2 (degree)

b)
Intensity (arb. unit)

TiO 2 (101)

TiO 2 (004) +Ag(111)

TiO 2 (200) Ag(200) 50

10

20

30

40

2 (degree)
Fig. 1. XRD patterns of (a) the Ag/a-TiO2 and (b) the AgTiO2/Ag/a-TiO2 nanocomposite thin lms.

464.7 eV corresponds to the Ti(2p1/2) and another one located at 459.1 eV is assigned to the Ti(2p3/2). The slitting between Ti(2p1/ 2) and Ti(2p3/2) core levels is 5.6 eV, indicating a normal state of Ti4+ in the anatase TiO2 lm. High-resolution XPS spectrum of Ag(3d) core level has been shown in Fig. 2b. The Ag(3d5/2) and Ag(3d3/2) peaks were found at binding energy of 368.5 eV and 374.5 eV, respectively. Moreover, the slitting of the 3d doublet of Ag is 6.0 eV, indicating formation of metallic silver nanoparticles [63,64] on surface of the AgTiO2/Ag/a-TiO2 thin lm. To further study the chemical state of the silver atoms accumulated on the surface, a detailed deconvolution of the Ag(3d) peak was also performed. The binding energy of Ag(3d5/2) core level for Ag, Ag2O and AgO is 368.6, 368.2 and 367.8 eV, respectively. Hence, the Ag(3d5/2) peak was deconvoluted into three Gaussian components with identical full width at half maximum (FWHM) after a Shirley background subtraction. Based on the deconvolution analysis, we have found that about 91% of the silver atoms accumulated on the surface were in the Ag0 (metallic) state, while only about 1% and 8% of the silver atoms were in Ag+ and Ag2+ chemical states, respectively. The O(1s) photoelectron peak of the AgTiO2/Ag/a-TiO2 thin lm has been presented in Fig. 2c. This peak was also deconvoluted into three Gaussian components with identical FWHM after a Shirley background subtraction. The rst component located at 530.7 eV

TiO 2 (105) TiO 2 (211)

120

O. Akhavan / Journal of Colloid and Interface Science 336 (2009) 117124

c)

O2OH H2O Fitted peak Exp. Peak

368.5 eV

b)

Ag (91%) Ag+ (1 %) Ag2+ (8 %) Fitted peak Exp. Peak

374.5 eV

6.0 eV

Intensity (arb. unit)

528 530 532 534 536

365 367

369

371

373

375

377

379

5.6 eV a)

459.1 eV
453 455 457 459 461 463

464.7 eV
465 467 469

Binding Energy (eV)


Fig. 2. XPS spectra of (a) Ti(2p); (b) deconvolution of Ag(3d) and (c) deconvolution of O(1s) core levels of the AgTiO2/Ag/a-TiO2 nanocomposite thin lm.

was assigned to the lattice oxygen of the TiO2 thin lm. The second component at 532.0 eV corresponds to the oxygen of surface OH bound in Ti(OH)OTi(OH), which in photo generation of electronhole changes into OH free radicals that is a promoting agent in the photocatalysis processes [2,64]. The third deconvoluted O(1s) peak, at 533.0 eV, corresponds to oxygen of water molecules included in the lm structure or adsorbed on the lm surface. The Ag/TiO2, OH/TiO2 and H2O/TiO2 molar ratios of the Ag TiO2/Ag/a-TiO2 thin lm at the different depths of the lm was calculated from the area ratio of the deconvoluted XPS peaks, as shown in Fig. 3. It is seen that the Ag/TiO2 molar ratio on the lm surface was calculated to be 0.69 which is considerably (one order in magnitude) higher than the nominal value considered in the sol (0.07). This indicates accumulation of the silver atoms on surface of the TiO2 thin lms, which can be an advantage in processes in which interaction with surface Ag nanoparticles are important. A similar observation was recently reported for self-accumulation of silver nanoparticles on surface of aqueous silica thin lm [61]. Here, we also measured the Ag/TiO2 ratio at the different depths of the thin lm. The slight increase in the Ag/TiO2 ratio at the depth of 5 nm shows that more silver nanoparticles were embedded at near the surface of the TiO2 layer. It was found that value of the Ag/TiO2 ratio substantially decreased, as the sputtering depth increased up to 20 nm, indicating accumulation of the silver nanoparticles on the surface (from surface down to $10 nm). At depth of 30 nm (corresponding to the interface of AgTiO2 and Ag/aTiO2 layers), a peak for the Ag/TiO2 ratio was observed which can be attributed to accumulation of the Ag particles at the interface. This accumulation at the interface (depth of 30 nm) can act as an auxiliary and lasting storage of silver atoms, especially during a silver ion releasing process. By increasing the depth of the thin lm from 30 to 45 nm, a considerable decrease in the Ag/TiO2 ratio

1.2
Ag/TiO2 OH/TiO2 H2O/TiO2

1.0

0.8

Molar ratio

0.6

0.4

0.2

0.0 0 10 20 30 40 50

Depth (nm)
Fig. 3. Variation of Ag/TiO2, OH/TiO2 and H2O/TiO2 molar ratios of the AgTiO2/Ag/ a-TiO2 nanocomposite thin lm calculated from the XPS analysis at the different sputtering depths.

was observed. The variation of OH/TiO2 molar ratio, calculated from the area ratio of the second and rst deconvoluted peaks of the O(1s) core level, has also been presented in Fig. 3. A high value molar ratio of the OH/TiO2 (1.17) was observed on the lm surface, due to more photocatalytic activity of the lm surface. By increasing the sputtering depth, the OH/TiO2 molar ratio substantially decreased to 0.06 at 25 nm. But, at 30 nm (location of the interface), the OH/TiO2 molar ratio showed a small peak. This

O. Akhavan / Journal of Colloid and Interface Science 336 (2009) 117124

121

peak indicated further presence of the aqueous pores at the boundary surfaces of the AgTiO2 and Ag/a-TiO2 layers in the AgTiO2/ Ag/a-TiO2 thin lm. In addition, the possibility in production of OH bound at both the interface and lm surface (where the silver atoms accumulated) can cause the released silver atoms and/or particles are in their positive ionic state (not in the metallic state). The H2O/TiO2 molar ratio of the lm at the different depths was also calculated from the area ratio of the third and rst deconvoluted peaks of the O(1s) core level. From Fig. 3 it is seen that by increasing the depth, the water ratio decreased. This means that surface of the AgTiO2/Ag/a-TiO2 thin lm was still aqueous and so porous, which can be an advantage in surface interactions and controlling the silver ion releasing processes. The weak peak located at 30 nm, shows that a relative more water can be found at the interface due to presence of further pores and/or grooves at there, in consistency with the observed small peak for the OH/ TiO2 molar ratio. It should also be noted the water included in the lm is cause of the out-diffusion of the metallic nanoparticles, initially dispersed in the lm, into the surface, during a low temperature ($100 C) annealing process, as previously indicated elsewhere [61]. Surface morphology of the AgTiO2/Ag/a-TiO2 thin lm was studied by SEM, as shown in Fig. 4. The XPS analysis showed that the silver particles were accumulated on the surface. Thus, the particles observed on the surface were attributed to the Ag nanoparticles on the surface of aqueous TiO2 lm. The diameter histogram of the surface particles (shown as the bottom inset of Fig. 4) indicated that average diameter of these surface nanoparticles are about 35 nm. TEM image of the Ag nanoparticles in the AgTiO2 lm has also been presented as the top inset of Fig. 4. It shows the silver nanoparticles with diameters in the range of 1331 nm. The UVvis measurement of the Ag/a-TiO2 and AgTiO2/Ag/aTiO2 thin lms exhibited a single surface plasmon resonance (SPR) absorption peak centered at 412 nm, as shown in Fig. 5. It is well-known that, the UVvis absorption spectra of the semispherical Ag nanoparticles with sizes ranged from 3 to 20 nm exhibit a single absorption peak at $410 nm [6567]. Therefore, consistent with the XPS results, SPR band showed that the silverbased nanoparticles in the Ag/a-TiO2 and also in the air-annealed AgTiO2/Ag/a-TiO2 thin lms were metallic silver nanoparticles. The narrower and lower intensity SPR absorption peak of the Ag/ a-TiO2 thin lm than the AgTiO2/Ag/a-TiO2 can be due to smaller

c)
Absorption (arb. unit)

b)

a)

300

500

700 900 Wavelength (nm)

1100

Fig. 5. UVvis absorption spectra of (a) the a-TiO2, (b) the Ag/a-TiO2, and (c) the Ag TiO2/Ag/a-TiO2 photocatalyst thin lms.

size and lower concentration of the silver nanoparticles in the former thin lm structure, respectively. For a better comparison, the optical absorption spectrum of the a-TiO2 thin lm was also shown in Fig. 5, indicating the transparency of the synthesized a-TiO2 thin lm. The bactericidal activity of the a-TiO2, Ag/a-TiO2 and AgTiO2/ Ag/a-TiO2 thin lms, as compared to a blank glass substrate (control sample), were studied against E. coli bacteria in dark and in the visible and the solar light irradiations, as shown in Fig. 6. It is seen that the viable bacteria reduced in an exponential manner. The slope of the tted line yields relative rate of reduction of the number of viable bacteria. The a-TiO2 thin lms showed a weak antibacterial activity with the relative rate of reduction of about 0.7 102 min1 in dark. The photo-antibacterial activity of the a-TiO2 thin lms was also checked in the visible and solar light irradiations. No change was observed for antibacterial activity of the a-TiO2 thin lms in the visible light, as compared to its activity in dark. But, the a-TiO2 improved the rate of reduction of the viable

Fig. 4. SEM image of the AgTiO2/Ag/a-TiO2 nanocomposite thin lm. The bottom inset shows diameter histogram of the particles accumulated on the surface. The top inset shows TEM image of Ag nanoparticles embedded in the TiO2 thin lm.

122

O. Akhavan / Journal of Colloid and Interface Science 336 (2009) 117124

1000000

Blank samples

Cell forming unit (cell/ml)

100000
a-TiO2

10000

1000

100
Ag/a-TiO2 Ag-TiO2/Ag/a-TiO2

10

1 0 20 0 40 60 80 100 120 140 160 180 200 0

Time (min)
Fig. 6. CFU of E. coli cultured for various periods in a medium containing the a-TiO2, Ag/a-TiO2 and AgTiO2/Ag/a-TiO2 photocatalyst thin lms as compared to blank (control) sample, in dark (j) and under irradiation of the visible light (N) and the solar light (d).

bacteria up to about 20% in the solar light irradiation. By deposition of the Ag nanoparticles on the a-TiO2 thin lm (the Ag/a-TiO2 thin lm), the rate of reduction of viable bacteria highly increased to 2.7 102 min1 corresponding to 190 min for killing all the bacteria in dark. The photocatalytic activity of the Ag/a-TiO2 thin lms was also checked in the visible and the solar light irradiations. In the visible light, the relative rate of reduction of the viable bacteria was found 3.4 102 min1 corresponding to 150 min to kill all the bacteria, indicating an improvement than the dark condition. This improvement showed that the Ag/a-TiO2 thin lm can also work better in the solar light irradiation, due to its UV spectrum. In fact, it was observed that the antibacterial activity of the Ag/aTiO2 thin lms was improved in the solar light irradiation with the rate of reduction of 4.3 102 min1 ($60% improvement than the dark condition) and the required time of 120 min to kill all the E. coli bacteria. By adding the AgTiO2 nanocomposite layer onto the Ag/a-TiO2 thin lm, the rate of reduction of viable bacteria increased to 3.6 102 min1 corresponding to 140 min for killing all the bacteria in dark. For the AgTiO2/Ag/a-TiO2 thin lm, the relative rate of reduction of the viable bacteria was improved to 4.7 102 and 5.8 102 min1 corresponding to 110 and 90 min for killing all the bacteria, in the visible and solar light irradiations, respectively. Therefore, the AgTiO2/Ag/a-TiO2 thin lm structure can efciently improve the photocatalytic activity of TiO2 and Ag doped TiO2 thin lms in dark and in the solar light irradiation. It was recently reported that bactericidal efciency of silver ions in AgNO3 can signicantly be enhanced by UV-A or visible light irradiation [68,69]. Hence, we also checked the effect of light irradiation of a mercury lamp (peak wavelengths at 275, 350 and 660 nm) on the bactericidal efciency of Ag nanoparticles in an AgSiO2 thin lm prepared similar to the AgTiO2 thin lm used at this work. Details of the AgSiO2 thin lm preparation can be found elsewhere [61,66]. Anyway, no considerable change in the antibacterial activity of the AgSiO2 thin lm irradiated by the light was observed as compared to its activity in dark (of course, after elimination of the irradiation effect for a blank sample). Therefore, the silver nanoparticles in the AgSiO2 thin lm and also in the Ag

TiO2/Ag/a-TiO2 thin lm structure did not any contribution in the photodegradation (not in the usual degradation) of the bacteria. This result might also be expected from the XRD pattern of the AgTiO2/Ag/a-TiO2 thin lm (Fig. 1b) which shows no peak related to the crystalline AgNO3, and from its XPS analysis (Fig. 2b) and optical absorption spectrum (Fig. 5c) which show formation of metallic silver nanoparticles, in contrast to the results shown by Liu et al. [69] and Kim et al. [68] who showed formation of silver ions (not silver metals) and their role in the photodegradation of bacteria, respectively. The silver ion releasing of the Ag/a-TiO2 and AgTiO2/Ag/a-TiO2 thin lms was also investigated. Fig. 7 shows the amounts of the released silver ions from the Ag/a-TiO2 (Fig. 7a) and AgTiO2/Ag/ a-TiO2 (Fig. 7b) thin lms as a function of the elapsing time in the HNO3 solution, after three sequent measurements. For the Ag/a-TiO2 thin lm, it is seen that silver ions were initially released with a high rate in the rst ve days. This time can be called saturation time dened as the time required for reaching the amount of silver ions to 80% of its saturation amount in the solution at a long time. After the saturation time, however, the process of releasing silver ions was controlled in a relatively slower release rate. The initial stage of silver ion release process is generally controlled by the water diffusion on surface of the matrix. For both the Ag/ a-TiO2 and the AgTiO2/Ag/a-TiO2 thin lms with silver nanoparticles accumulated on the surface, easy water diffusion on the surface could be resulted in a sharp release of the silver ions at the initial stage. On the other hand, the considerable reduction in the release rate of silver ions at the latter stage (the saturation stage) can be attributed to the slow diffusion of water in the pores of the TiO2 lm, and so change in the silver ion releasing mechanism. This change in the release behavior of the silver ions from both the Ag/a-TiO2 and the AgTiO2/Ag/a-TiO2 thin lms completely distinguishes them from the bulk silver-based materials exhibiting a sharp ion release after a minimum releasing at initial stages [62]. Moreover, it was found that after the three sequent measurements on the Ag/a-TiO2 thin lm, the saturation amount of the silver ions in the solution reduced from 9.4 to 4.0 nM/ml, corresponding to 57% reduction in the initial saturation amount. This indicates a high consumption of the accumulated silvers on surface of the a-

10
Concentration (M/ml)

1st Measurement tsat = 5 days


15 10 5 0 0 10 20 30 40 50 60

9 8

Ag+ release (nM/ml)

7 6 5 4 3 2 1 0 0 5

a)
Current (A)

3rd Measurement 1st Measurement

3rd Measurement t sat = 8 days

b)

10

15

20

25

Time (day)
Fig. 7. The silver ion release curves of (a) the Ag/a-TiO2 and (b) the AgTiO2/Ag/aTiO2 nanocomposite thin lms, after three sequent measurements in HNO3 solution. The inset shows the obtained standard curve of silver ions for calibration of the ion current.

O. Akhavan / Journal of Colloid and Interface Science 336 (2009) 117124

123

TiO2 thin lm. For the Ag/a-TiO2 thin lm covered by the AgTiO2 nanocomposite layer (i.e., AgTiO2/Ag/a-TiO2 nanocomposite thin lm), however, the situation is completely different. Fig. 7b shows that, this time, the saturation time is eight days which means a slower silver ion release rate at the rst stage and consequently a lower saturation amount of the silver ions in the solution, as compared to the obtained characteristics of the Ag/a-TiO2. The reduction observed in the release rate of the AgTiO2/Ag/a-TiO2 thin lm also conrms that the silver particles accumulated on the lm surface were effectively protected by a very thin ($ nm) titania layer (as also resulted based on the depth prole XPS analysis) which controls further the releasing process. In addition, it was observed that even after the three sequent ion release measurements, the saturation amount only decreased from 2.8 to 2.5 nM/ml, corresponding to only 10% reduction in the initial saturation amount. This means that by covering the Ag/a-TiO2 thin lm with the AgTiO2 nanocomposite (barrier) layer, further controlling the silver ion release is provided so that durability of the AgTiO2/Ag/a-TiO2 nanocomposite thin lm (of course with higher (about 2 times) silver storage) increases to at least 11 times of the Ag/a-TiO2 thin lm. Therefore, the AgTiO2/Ag/a-TiO2 thin lm can be considered as efcient and also lasting antibacterial nanocomposite materials for future antibacterial and bio-medical applications. 4. Summary The AgTiO2/Ag/a-TiO2 nanocomposite thin lm photocatalyst sensitive to the solar light with an efcient storage of Ag nanoparticles both on the lm surface and at interface of the AgTiO2 and a-TiO2 thin lms was simply synthesized by solgel. XRD analysis conrmed formation of a-TiO2 and crystalline metallic silver grains with fcc structure. The SPR absorption band at 412 nm in wavelength was attributed to formation of spherical Ag nanoparticles in the Ag doped thin lms. The depth prole XPS analysis showed that the Ag nanoparticles (average diameter of $35 nm) were effectively embedded at depths of 5 nm of the aqueous TiO2 layer and at depth of 30 nm (at the interface) of the nanocomposite lm. In addition, the OH bounds and also water contents of the lm were concentrated both on the lm surface and at the interface, as an advantage for release of more ionic (not metallic) silver nanoparticles. The photodegradation of the thin lms was studied by testing its antibacterial activity against E. coli bacteria, in dark and in the visible and the solar light irradiations. The relative rate of reduction of the viable bacteria for the nanocomposite lm (3.6 102 min1) was 5.1 times greater than the corresponding value for the a-TiO2 thin lm, in dark. In the solar light irradiation, the rate of reduction of the viable bacteria for the AgTiO2/Ag/aTiO2 lm (5.8 102 min1) was measured 1.35 and 6.90 times greater than the corresponding rate for the Ag/a-TiO2 and a-TiO2 thin lms, respectively. The ion release measurement indicated that the Ag/a-TiO2 and AgTiO2/Ag/a-TiO2 thin lms show a saturation behavior at long times (5 and 8 days in HNO3 solution, respectively), which is a completely different behavior from the general behavior of ion release in silver-based bulk materials. Hence, the ion releasing process was mainly controlled by interdiffusion of water and silver nanoparticles through the pores of the lms, not by easy surface diffusion of water on the Ag nanoparticles. Consistent with the XPS result, the ion release measurements also showed the surface Ag nanoparticles in the AgTiO2/ Ag/a-TiO2 lm were protected by a nanometric TiO2 barrier layer against an easy ion release. Three sequent ion release measurements of the Ag/a-TiO2 and AgTiO2/Ag/a-TiO2 lms showed that durability of the latter lm (of course with 2 times silver storage) is at least 11 times greater than the former one. Therefore, the

AgTiO2/Ag/a-TiO2 photocatalyst is an excellent and long lasting antibacterial nanocomposite material appropriate for future biomedical and bactericidal applications. Acknowledgments The authors would like to thank Research Council of Sharif University of Technology and also Iranian Nanotechnology Initiative for nancial support of the project. References
[1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] [26] [27] [28] [29] [30] [31] [32] [33] [34] [35] [36] [37] [38] [39] [40] [41] [42] [43] [44] [45] [46] [47] [48] [49] [50] [51] A. Fujishima, K. Honda, Nature 238 (1972) 37. M.S. Hoffmann, T. Martin, W. Choi, D.W. Bahnemann, Chem. Rev. 95 (1995) 69. G.A. Somorjai, Chem. Rev. 96 (1996) 1223. A. Mills, S.L. Hunte, J. Photochem. Photobiol. A 108 (1997) 1. C. Burda, X. Chen, R. Narayanan, M.A. El-Sayed, Chem. Rev. 105 (2005) 1025. R. Hoffmann, T. Martin, Y. Choi, W. Bahnemann, Chem. Rev. 95 (1995) 69. J. Robertson, P. Robertson, L. Lawton, J. Photochem. Photobiol. A 175 (2005) 51. J. Carey, J. Lawrence, H. Tosine, Bull. Environ. Contam. Toxicol. 16 (1976) 697. V. Nadtochenko, A. Rincon, S. Stanca, J. Kiwi, J. Photochem. Photobiol. A 169 (2005) 131. A. Fujishima, N.R. Tata, A. Tryk, J. Photochem. Photobiol. C 1 (2000) 1. T. Matsunaga, R. Tomoda, T. Nakajima, H. Wake, FEMS Microbiol. Lett. 29 (1985) 211. V. Nadtochenko, N. Denisov, O. Sarjusiv, D. Gumy, C. Pulgarin, J. Kiwi, J. Photochem. Photobiol. A 181 (2006) 401. G. Rincon, C. Pulgarin, Appl. Catal. B 51 (2004) 283. J.C. Yu, W. Ho, J. Lin, H. Yip, P.K. Wong, Environ. Sci. Technol. 37 (2003) 2296. J.S. Hur, Y. Koh, Biotechnol. Lett. 24 (2002) 23. K. Sunada, T. Watanabe, K. Hashimoto, Environ. Sci. Technol. 37 (2003) 4785. Y. Liu, X. Wang, F. Yang, X. Yang, Micropor. Mesopor. Mater. 114 (2008) 431. M. Bellantone, H.D. Williams, L.L. Hench, Antimicrob. Agents Chemother. 16 (2002) 1940. A. Henglein, J. Phys. Chem. 97 (1993) 5457. A.P. Alivisatos, J. Phys. Chem. 100 (1996) 13226. Y. Murakami, J. Matsumoto, Y. Takasu, J. Phys. Chem. B 103 (1999) 1836. G. Fu, P.S. Vary, C.T. Lin, J. Phys. Chem. B 109 (2005) 8889. W.O. Williamson, Nature 143 (1939) 279. M. Anpo, S. Kishiguchi, Y. Ichihashi, M. Takeuchi, H. Yamashita, K. Ikeue, B. Morin, A. Davidson, M. Che, Res. Chem. Intermediates 27 (2001) 459. J.C. Yu, W. Ho, J. Yu, H. Yip, P.K. Wong, J. Zhao, Environ. Sci. Technol. 39 (2005) 1175. R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki, Y. Taga, Science 293 (2001) 269. C. Burda, Y. Lou, X. Chen, A.C.S. Samia, J. Stout, J.L. Gole, Nano Lett. 3 (2003) 1049. S.U.M. Khan, M. Al-Shahry, W.B. Ingler Jr., Science 297 (2002) 2243. C.H. Park, S.B. Zhang, S.H. Wei, Phys. Rev. B 66 (2002) 073202. W. Choi, A. Termin, M. Hoffmann, J. Phys. Chem. 98 (1994) 13669. W. Mu, J.M. Herrmann, P. Pichat, Catal. Lett. 3 (1989) 73. D. Duonghong, E. Borgarello, M. Gratzel, J. Am. Chem. Soc. 103 (1981) 4685. J. Husheng, H. Wensheng, W. Liqiao, X. Bingshe, L. Xuguang, Dent. Mater. 24 (2008) 244. X. Bingshe, N. Mei, W. Liqiao, H. Wensheng, L. Xuguang, J. Photochem. Photobiol. A 188 (2007) 98. Q. Cheng, C. Li, V. Pavlinek, P. Saha, H. Wang, Appl. Surf. Sci. 252 (2006) 4154. S.-D. Oh, S. Lee, S.-H. Choi, I.-S. Lee, Y.-M. Lee, J.-H. Chun, H.-J. Park, Colloids Surf. A 275 (2006) 228. O. Akhavan, E. Ghaderi, Sci. Technol. Adv. Mater. 10 (2009) 015003. T. Yuranova, A.G. Rincon, A. Bozzi, S. Parra, J. Photochem. Photobiol. A 161 (2003) 27. Q.L. Feng, J. Wu, G.Q. Chen, F.Z. Cui, T.N. Kim, J.O. Kim, J. Biomed. Mater. Res. 52 (2000) 662. I. Sondi, B. Salopek-Sondi, J. Colloid Interface Sci. 275 (2004) 177. Y. Matsumura, K. Yoshikata, S. Kunisaki, T. Tsuchido, Appl. Environ. Microbiol. 69 (2003). X.H. Xu, W.J. Brownlow, S.V. Kyriacou, Q. Wan, J.J. Viola, Biochemistry 43 (2004) 10400. S.K. Gogoi, P. Gopinath, A. Paul, A. Ramesh, S.S. Ghosh, A. Chattopadhyay, Langmuir 22 (2006) 9322. S.V. Kyriacou, W.J. Brownlow, X.H. Xu, Biochemistry 43 (2004) 140. M.K. Seery, R. George, P. Floris, S.C. Pillai, J. Photochem. Photobiol. A 189 (2007) 258. K. Page, R.G. Palgrave, I.P. Parkin, M. Wilson, S.L.P. Savin, A.V. Chadwick, J. Mater. Chem. 17 (2007) 95. M.P. Reddy, A. Venugopal, M. Subrahmanyam, Water Res. 41 (2007) 379. E.V. Skorb, L.I. Antonouskaya, N.A. Belyasova, D.G. Shchukin, H. Mhwald, D.V. Sviridov, Appl. Catal. B 84 (2008) 94. C. Hu, Y. Lan, J. Qu, X. Hu, A. Wang, J. Phys. Chem. B 110 (2006) 4066. H.M. Sung-Suh, J.R. Choi, H.J. Hah, S.M. Koo, Y.C. Bae, J. Photochem. Photobiol. A 163 (2004) 37. A. Sclafani, M.N. Mozzanega, J.M. Herrmann, J. Catal. 168 (1997) 117.

124

O. Akhavan / Journal of Colloid and Interface Science 336 (2009) 117124 [63] J.F. Moulder, W.F. Stickle, P.E. Sobol, K.D. Bomben (Eds.), Hand Book of X-ray Photoelectron Spectroscopy, PerkinElmer Corporation Physical Electronics Division, 1992. [64] C.D. Wagner, A.V. Naumkin, A. Kraut-Vass, J.W. Allison, C.J. Powell, J.R. Rumble Jr., NIST Standard Reference Database 20, Ver. 3.4, 2006. Available from: <http://srdata.nist.gov/xps/index.htm>. [65] W. Li, S. Seal, E. Megan, J. Ramsdell, K. Scammon, G. Lelong, L. Lachal, K.A. Richardson, J. Appl. Phys. 93 (2003) 9553. [66] A. Babapour, O. Akhavan, R. Azimirad, A.Z. Moshfegh, Nanotechnology 17 (2006) 763. [67] H. Jia, W. Xu, J. An, D. Li, B. Zhao, Spectrochim. Acta A 64 (2006) 956. [68] J.Y. Kim, C. Lee, M. Cho, J. Yoon, Water Res. 42 (2008) 356. [69] T.-X. Liu, X.-Z. Li, F.-B. Li, Environ. Sci. Technol. 42 (2008) 4540.

[52] Y. Tian, T. Tatsuma, Chem. Commun. 16 (2004) 1810. [53] Y.L. Wang, Y.Z. Wan, X.H. Dong, G.X. Cheng, H.M. Tao, T.Y. Wen, Carbon 36 (1998) 1567. [54] H.J. Lee, S.Y. Yeo, S.H. Jeong, J. Mater. Sci. 38 (2003) 2199. [55] C.N. Lok, C.M. Ho, R. Chen, Q.Y. He, W.Y. Yu, H. Sun, P.K.H. Tam, J.F. Chiu, C.M. Che, J. Proteome. Res. 5 (2006) 916. [56] J.R. Morones, J.L. Elechiguerra, A. Camacho, K. Holt, J.B. Kouri, J.T. Ramirez, M.J. Yacaman, Nanotechnology 16 (2005) 2346. [57] T. Maneerung, S. Tokura, R. Rujiravanit, Carbohydr. Polym. 72 (2008) 43. [58] S.S. Mahapatra, N. Karak, Mater. Chem. Phys. 112 (2008) 1114. [59] S. Zhang, R. Fu, D. Wu, W. Xu, Q. Ye, Z. Chen, Carbon 42 (2004) 3209. [60] Y.Z. Wan, Y.L. Wang, G.X. Cheng, H.L. Luo, X.H. Dong, Carbon 39 (2001) 1607. [61] O. Akhavan, R. Azimirad, A.Z. Moshfegh, J. Phys. D 41 (2008) 195305. [62] R. Kumar, H. Munstedt, Biomaterials 26 (2005) 2081.

You might also like