You are on page 1of 8

Journal of Photochemistry and Photobiology B: Biology 142 (2015) 178–185

Contents lists available at ScienceDirect

Journal of Photochemistry and Photobiology B: Biology


journal homepage: www.elsevier.com/locate/jphotobiol

Photocatalytic antibacterial activity of nano-TiO2 (anatase)-based


thin films: Effects on Escherichia coli cells and fatty acids
Urmas Joost a,b,⇑,1, Katre Juganson c,d,1, Meeri Visnapuu a,c,1, Monika Mortimer c, Anne Kahru c,
Ergo Nõmmiste a, Urmeli Joost a, Vambola Kisand a,b, Angela Ivask c,⇑
a
Institute of Physics, University of Tartu, Ravila 14c, 50411 Tartu, Estonia
b
Estonian Nanotechnology Competence Center, Ravila 14c, 50411 Tartu, Estonia
c
Laboratory of Environmental Toxicology, National Institute of Chemical Physics and Biophysics, Akadeemia tee 23, 12618 Tallinn, Estonia
d
Department of Chemistry, Tallinn University of Technology, Akadeemia tee 15, 12618 Tallinn, Estonia

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

Article history: Titanium dioxide is a photocatalyst with well-known ability to oxidise a wide range of organic
Received 15 July 2014 contaminants as well as to destroy microbial cells. In the present work TiO2 nanoparticles with high
Received in revised form 26 November 2014 specific surface area (150 m2/g) were used to prepare nanostructured films. The TiO2 nanoparticle-based
Accepted 1 December 2014
film in combination with UV-A illumination with intensity (22 W/m2) comparable to that of the sunlight
Available online 17 December 2014
in the UV-A region was used to demonstrate light-induced antibacterial effects. Fast and effective
inactivation of Escherichia coli cells on the prepared thin films was observed. Visualization of bacterial
cells under scanning electron microscopy (SEM) showed enlargement of the cells, distortion of cellular
membrane and possible leakage of cytoplasm after 10 min of exposure to photoactivated TiO2. According
to the plate counts there were no viable cells as early as after 20 min of exposure to UV-A activated TiO2.
In parallel to effects on bacterial cell viability and morphology, changes in saturated and unsaturated
fatty acids – important components of bacterial cell membrane-were studied. Fast decomposition of
saturated fatty acids and changes in chemical structure of unsaturated fatty acids were detected. Thus,
we suggest that peroxidation and decomposition of membrane fatty acids could be one of the factors
contributing to the morphological changes of bacteria observed under SEM, and ultimately, cell death.
Ó 2014 Elsevier B.V. All rights reserved.

1. Introduction can degrade different organic compounds [11–13], has antibacte-


rial [14,15] and photoinduced super-hydrophilic properties [9,10]
With increasing appearance of antibiotic-resistant bacteria, making it an ideal candidate for many applications where surfaces
new and more potent antibacterial agents and materials are are inaccessible for traditional cleaning and/or have to efficiently
needed. There are great hopes that nanotechnology could offer inhibit potentially pathogenic bacteria [16]. One of the advantages
new possibilities in this area [1,2]. Although nanoparticles (NPs) of using TiO2 in self-cleaning applications is that the material is
of silver, zinc oxide and copper oxide [3] are already used in several very stable, and that storage and repeated long term exposure to
antimicrobial applications [4], those nanomaterials may also pose UV has little or no effect on its antibacterial activity.
hazard to other organisms when released to the environment Antibacterial applications of TiO2 often use thin film coatings
[3,5]. Kahru and Dubourguier [6] showed that among seven most that can be deposited by various methods including sputtering
widespread nanomaterials (TiO2, ZnO, CuO, Ag, SWCNTs, MWCNTs [17], chemical solution deposition [18], pulsed laser deposition
and C60-fullerenes) TiO2 NPs proved to be the most environmen- [19] or sol–gel method [20–23].
tally benign. TiO2 has been studied extensively as a promising pho- The mechanism of TiO2 photocatalytic oxidation and conse-
tocatalyst [7], solar cell material [8], anti-fogging and self-cleaning quently, also antibacterial activity involves reactive oxygen species
coating material [9,10]. Due to its photocatalytic properties, titania (ROS); TiO2 in anatase phase has been shown as the most potent
form of TiO2 to produce ROS [24]. ROS production by TiO2 is induced
⇑ Corresponding authors at: Institute of Physics, University of Tartu, Ravila 14c, after absorption of a high-energy photon and subsequent excitation
50411 Tartu, Estonia (U. Joost). Laboratory of Environmental Toxicology, National of an electron from the TiO2 valence band to conduction band [24].
Institute of Chemical Physics and Biophysics, Akadeemia tee 23, 12618 Tallinn, The next and rate determining step in formation of ROS is photoex-
Estonia (A. Ivask).
cited electron transfer from conduction band of TiO2 to oxygen
E-mail addresses: urmas.joost@ut.ee (U. Joost), angela.ivask@kbfi.ee (A. Ivask).
1 molecule during which superoxide anions (O2 ) are produced [24].
Equal participation.

http://dx.doi.org/10.1016/j.jphotobiol.2014.12.010
1011-1344/Ó 2014 Elsevier B.V. All rights reserved.
U. Joost et al. / Journal of Photochemistry and Photobiology B: Biology 142 (2015) 178–185 179

Photoholes generated into the valence band during photoexcitation the molar ratio of PTSA and Ti(OBu)4 was set to 0.2, the molar ratio
oxidize surface absorbed H2O, AOH and surface titanol groups of acac and Ti(OBu)4 was set to 3 and the molar ratio of water and
(>TiOH) into highly reactive hydroxyl radicals (OH). Although other Ti(OBu)4 was set to 10. The reaction was carried out overnight at
ROS also contribute to TiO2 photocatalytic activity, the majority can reflux conditions. The particles were centrifuged for 0.5–1.5 h at
be attributed to OH radicals [25–27], that not only act on TiO2 sur- 12,000g and subsequently washed 3 times with methanol. Finally,
face but can diffuse over short distances to the surrounding solution the particles were dispersed in 95% ethanol and the attained dis-
near the photocatalyst surface and effectively degrade organic com- persions were stable for months.
pounds that are not directly in contact with the photocatalyst Thin films were prepared by spin-coating colloidal solution of
[25,28,29]. Thus, hydroxyl radicals have been considered as the anatase TiO2 NPs on silicon (1 0 0) monocrystal substrates at ambi-
main ROS species to cause both, photocatalytic activity as well as ent atmospheric conditions. The substrates were cleaned prior to
antibacterial effects of TiO2 [30–32]. It is suggested that the main coating with 95% ethanol to remove small dust particles. The rota-
reason for high antibacterial activity of TiO2 particles is photooxida- tion frequency during spin-coating was 3000 rpm and coating time
tion of essential cellular components. Oxidation of one of the major was 0.5 min. The obtained precursor films were aged at ambient
components of bacterial plasma membrane-phospholipids-by conditions for 4 days to allow the remaining solvent to evaporate
photoactivated TiO2 has been reported in several studies [33,34]. slowly in order to prevent cracking of the films. After ageing the
However, it must be noted that studies presenting efficient and samples were annealed in a Nabertherm L5/11/S27 furnace at
rapid antibacterial activity of TiO2 have utilized short wavelength 400 °C and washed in deionized water for 10 min in ultrasonic bath
UV-light (UV-C and UV-B) [27,35]. Studies where UV-A light that to remove organic residue. The obtained films were characterized
has longer wavelength and, thus, is less harmful to living organisms using Raman spectroscopy, atomic force microscopy (AFM), scan-
per se, was used to photoactivate TiO2 have reported relatively poor ning electron microscopy (SEM) and X-ray photoelectron spectros-
antibacterial activity [16,36–39] or long exposure times needed for copy (XPS). The indirect optical band gap of the obtained anatase
achieving sufficient antibacterial effects [40,41]. thin films was evaluated using UV–Vis spectroscopy and Tauc plot
In the current study, we investigated the time-dependent anti- (Fig. S2). Characterization of the obtained nano-TiO2 thin films
bacterial effect of TiO2 NP (anatase)-based thin films (hereafter des- with high specific surface area has been reported in detail in our
ignated as nano-TiO2 thin films) under UV-A light (315–400 nm). E. previous work [43], including the description of the cleaning meth-
coli, a potential pathogen and a sanitary indicator bacterium was ods of nano-TiO2 thin films.
used as a model microorganism to evaluate the antimicrobial effi- Stearic, oleic and linoleic acid coatings for photoactivated deg-
cacy of light-activated TiO2 films. We used recombinant lumines- radation studies were prepared by spin-coating 0.5% solution of
cent E. coli which allowed easy luminescence-based assessment of respective fatty acid in methanol on nano-TiO2 thin film substrate
cell viability. At different exposure times both, the number of viable (approximately 1 cm2) for 0.5 min at 3000 rpm, so that the whole
bacterial cells as well as changes in bacterial morphology were substrate surface was uniformly covered.
determined by growth tests and scanning electron microscopy,
respectively. In parallel to the studies with bacterial cells the ability 2.3. UV treatment
of UV-A activated nano-TiO2 thin films to decompose saturated and
unsaturated fatty acids was evaluated by determining the changes UV-irradiation of nano-titania films with fatty acids added was
in fatty acid structure using X-ray photoelectron spectroscopy conducted in climate chamber Memmert CTC 256 at 25 °C at a rel-
(XPS). As these fatty acids are abundant in bacterial plasma mem- ative humidity of 70%. High relative humidity ensured the presence
brane the UV-light stimulated changes in fatty acid decomposition of water molecules on the nano-titania surface. When bacteria on
may yield further information on the antimicrobial mechanism of Si-monocrystal substrates (negative control) or nano-TiO2 thin
action of TiO2 thin films. films were exposed to UV in CTC 256 climate chamber the temper-
ature was set to 25 °C and the relative humidity to 90% to avoid
2. Materials and methods extensive drying of the droplet of bacterial suspension. The source
of UV-irradiation was a Philip Cleo 15 W fluorescent lamp with a
2.1. Materials maximum intensity at 355 nm, i.e. in the UV-A region. Light inten-
sity at the sample surface was measured with a lux meter Delta
Commercially available reagents: titanium(IV) butoxide Ohm HD 2302.0 equipped with a sensor LP 471 UV-A or LP 471
(Ti(OBu)4; Sigma–Aldrich, reagent grade), p-toluene sulfonic acid UV-B accordingly. It was found that the light intensity in the
(PTSA; Sigma–Aldrich, reagent plus), acetyl acetone (acac; UV-A region (315–400 nm) was 22 W/m2 which is comparable to
Sigma–Aldrich, reagent plus), butanol (Sigma–Aldrich), methanol the UV-light intensity of sunlight on the ground level according
(Sigma–Aldrich P99.8%) and deionized water (MilliQ) were used to the ASTMG 173-03 table [44], and in the UV-B region
for the preparation of TiO2 NPs. Glutaraldehyde (Sigma–Aldrich, (280–315 nm) less than 0.1 W/m2. Due to the dominance of UV-A
70% in H2O) and anhydrous ethanol were used in sample prepara- light in the UV-spectra of the used UV-lamp, the applied light is
tion for electron microscopy study. Stearic acid (C18:0), oleic acid further designated as UV-A.
(C18:1 cis-9; both from Cayman Chemical Company, purity >98%)
and linoleic acid (C18:2 cis-9,12; Sigma Aldrich, purity P99%) 2.4. Evaluation of antibacterial activity of the nano-TiO2 thin films
(Fig. S1) were used in fatty acid degradation experiments. Butanol
was dried using CaH2 and distilled before utilization; all other Recombinant constitutively bioluminescent gram-negative bac-
reagents were used as received. 0.9% NaCl solution that was pre- teria E. coli MC1061 (pSLlux) [45] were used as test organisms to
pared using NaCl from Sigma Aldrich and MilliQ water was used evaluate the antibacterial activity of nano-TiO2 thin films. The cells
in bacterial viability assays. were cultivated in Luria-Bertani (LB) broth supplemented with
100 lg/mL of ampicillin at room temperature (23 ± 2 °C) overnight
2.2. Synthesis of titania nanoparticles and preparation of nano-TiO2 without shaking. The cells were harvested in exponential growth
thin films phase by centrifugation at 3500g for 10 min. The cells were washed
twice with 0.9% NaCl or with MilliQ water and were resuspended
Titania NPs were synthesised using a method described by either in equal volume of 0.9% NaCl or in MilliQ. According to direct
Scolan and Sanchez [42] with slightly modified parameters. Briefly, plate counting, the final number of viable bacterial cells either in
180 U. Joost et al. / Journal of Photochemistry and Photobiology B: Biology 142 (2015) 178–185

0.9% NaCl or in MilliQ was about 108 colony forming units per [Gauss-Lorentz hybrid function (GL 70, Gauss 30%, Lorentz 70%)
millilitre (CFU/mL). Antibacterial effects of nano-TiO2 thin films was used for fitting and Tougaard background was used for back-
to E. coli were determined quantitatively and visually. ground removal]. The residual carbon on nano-TiO2 surfaces was
In order to quantitatively evaluate the antibacterial activity of taken into account by measuring the C 1s spectra of nano-TiO2 film
nano-TiO2 thin films, 10 lL of E. coli MC1061 (pSLlux) suspension before covering it with fatty acids and subtracting it afterwards
in 0.9% NaCl was added onto nano-TiO2 thin film with approximate from the spectra of samples covered with fatty acids. This was
surface area of 1 cm2. The actual contact area between the bacterial done based on our previous work [43] where we noticed that after
suspension and the surface depended on the irradiation time and washing the nano-TiO2 films in deionized water the composition
therefore was difficult to determine accurately. The surfaces with and the amount of residual carbon was always very similar. It
bacterial suspension were illuminated with UV-light as described was not possible to measure the effects of UV-A radiation on
above (see Section 2.3) for 0, 5, 10, 15 and 20 min. For control, bac- non-photocatalytic surface due to technical reasons (the fatty acids
teria on Si-monocrystals with no TiO2 were illuminated identically. evaporated from smooth Si(1 0 0) surface during pumping before
For non-UV control, the cells were incubated on the same surfaces XPS spectra could be measured). However, it has been shown that
for 20 min in the dark. After the exposure, the surfaces with bacte- photochemical effects potentially occurring on Si-monocrystal sur-
rial suspension were placed into a 15 mL round-bottom polypro- face are at least 10 times smaller than photocatalytic effects that
pylene tube containing 1 mL of 0.9% NaCl and the bacteria were take place on TiO2 surface and can be therefore neglected [48].
separated from the surface by vortexing for 5 min. Then 15 lL of Due to surface region deviation from chemical homogeneity in
the resultant bacterial suspension and its 10-, 100-, and 1000-fold the working range of photoelectron spectroscopy (surface region
dilutions were spread onto LB agar plates containing 100 lg/mL of with thickness up to three electron mean free paths) the absolute
ampicillin. The number of surviving luminescent bacterial colonies amounts of different chemical compounds were considered cau-
(CFU) was counted in the dark after overnight incubation at 37 °C. tiously and are presented in this study only to outline trends and
The experiment was conducted on two different days in three rep- estimates of processes occurring during photo-oxidation.
licates and the counts corresponding to a particular sample were
averaged.
3. Results and discussion
Visual observation of antibacterial effects of nano-TiO2 thin
films was done using SEM (FEI Helios NanoLab 600). For SEM imag-
3.1. Characterization of nano-TiO2 thin films
ing, 10 lL of E. coli MC1061 (pSLlux) suspension in MilliQ water
was added drop-wise onto the aforementioned surfaces. The expo-
Characterization of nano-TiO2 thin films has been reported in
sure conditions were the same as for the quantitative evaluation. In
our previous work [43]. Colloidal suspension of anatase TiO2 NPs
addition, 40 and 60 min exposure times were applied. After the
that was spin-coated onto the Si-monocrystal surface consisted
exposure, the samples were moved to a chamber filled with water
mainly of monodisperse particles with hydrodynamic diameter
vapour (see Fig. S3) and 20 lL of fixative (2.5% glutaraldehyde solu-
less than 10 nm (Fig. S5). Taking into account the spherical nature
tion in ethanol) was pipetted onto each sample. The samples were
of the particles, the calculated surface area of the TiO2 particles was
fixated for 2 h and then placed to a chamber filled with ethanol
approximately 150 m2/g. After deposition, the anatase nano-TiO2
vapours to initiate dehydration. After 15 min the samples were
particles (Fig. S6) were annealed at 400 °C and subsequently
allowed to dry at ambient conditions for 5 min under a cover.
washed with deionised water. The annealing step was necessary
The dehydration cycle was repeated twice. Next, the samples were
for removal of organic material used during NP synthesis. This thin
carefully rinsed with anhydrous ethanol and allowed to dry at
film preparation methodology is suitable for most of the materials
ambient conditions under a cover for three days before SEM imag-
that can be annealed at temperatures up to 400 °C. The thickness of
ing. SEM images were acquired with an accelerating voltage of
resulting TiO2 thin film was 115 nm (measured with AFM) [43]. It
1.0 kV. Prior to SEM imaging all samples were treated identically
is likely that the aggregation of 10 nm TiO2 particles during depo-
(including glutaraldehyde treatment), therefore ‘‘bulged’’ bacteria
sition and annealing took place. Yet, the structural appearance of
(see Fig. S4) which appeared after 10 min UV treatment were con-
thin film surface indicated high surface area of TiO2 also in the
sidered severely affected or dead. A number of images from differ-
form of a thin film after annealing (Fig. S7). High surface area of
ent places on the sample were taken and one image representing
the TiO2 nanomaterial ensured the presence of surface adsorbed
the whole sample was presented.
species that can act as electron and/or hole traps [24] increasing
thus the lifetime of electron/hole pairs and increasing the effi-
2.5. Analysis of structural changes in fatty acids on nano-TiO2 thin
ciency of the photocatalyst, making this material an efficient pho-
films using X-ray photoelectron spectroscopy (XPS)
tocatalyst. Optical band gap of prepared nano-TiO2 thin films was
3.5 eV (Fig. S2). This is wider than the 3.2 eV band gap of bulk
XPS was used to investigate the chemical changes occurring in
TiO2 and is characteristic to nano-sized TiO2 material [49].
stearic, oleic and linoleic acid coatings on nano-TiO2 thin films
upon exposure to UV-A illumination. XPS measurements were
conducted using a surface station equipped with an electron 3.2. Photoinduced antibacterial activity of nano-TiO2 thin films
energy analyser (SCIENTA SES 100) and a non-monochromatic twin
anode X-ray tube (Thermo XR3E2), with characteristic energies of Antibacterial effects of UV-treated nano-TiO2 thin films were
1253.6 eV (Mg Ka1,2 FWHM 0.68 eV) and 1486.6 eV (Al Ka1,2 FWHM evaluated by determining the ability of exposed bacteria to form
0.83 eV). All XPS measurements were conducted in ultra-high vac- bioluminescent colonies on agarized growth media (viability
uum conditions. The binding energy scales for the XPS experiments assay) and also by visual inspection of bacterial morphology
were referenced to the binding energy of Ti 2p3/2 (458.6 eV) photo- (SEM images). Exposure of bacteria to UV-A light on nano-TiO2 thin
emission line. To estimate overall atomic concentrations of differ- films had a distinct effect on their viability already after 5 min:
ent compounds and elements Average Matrix Relative Sensitivity the number of colony forming bacteria decreased about 4 times,
Factors (AMRSF) procedure [46] and transmission function of the from 7.9 ⁄ 107 to 2 ⁄ 107 CFU/mL (Fig. 1). Exposure for 10 min
instrument were used. Raw data were processed using Casa XPS caused a decrease of colony forming bacteria from 7.9 ⁄ 107 to
[47] software. Data processing involved removal of Ka and Kb 1.3 ⁄ 106 CFU/mL and no colony forming bacterial cells remained
satellites, removal of background and fitting of components after 20 min of exposure. To analyse the effect of UV-A light alone,
U. Joost et al. / Journal of Photochemistry and Photobiology B: Biology 142 (2015) 178–185 181

Fig. 1. Colony forming potential of Escherichia coli (pSLlux) in different exposure conditions. The effect of UV-irradiation length on colony forming potential of E. coli (pSLlux)
applied onto Si-monocrystal substrates (blue bars) or nano-TiO2 thin films (red bars). After 20 min of exposure on nano-TiO2 thin film to UV-A light no viable cells able to
grow on agar plate were observed (marker with ⁄). The colony forming potential of E. coli on surfaces kept in the dark (black and dark grey bars) was determined only in the
beginning (0 min) and at the end of the experiment (20 min). (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of
this article.)

E. coli cells were illuminated for 20 min on Si-monocrystal surface other hand, as this high UV-A intensity is not realistic in real con-
not coated with TiO2: 5, 10, 15 and 20 min illumination resulted in ditions, no conclusions could be made whether this material would
a slight decrease of bacterial number (from 108 to 8.4 ⁄ 107, be also effective in practice.
5.9 ⁄ 107, 4.7 ⁄ 107 and 1.3 ⁄ 107 CFU/mL, respectively; Fig. 1). Examination of photoactivated nano-TiO2 film exposed bacte-
Since silicon is considered one of the most biocompatible materials rial cells by SEM (Fig. 2) showed that as bacterial number
[50] and silicon surface stored in atmosphere is completely passiv- decreased, morphology of bacterial cells changed. The shape of
ated in the sense of photocatalytic action with amorphous SiO2 the bacteria became more discursive and expanded, the diameter
layer [51], this slight decrease was most likely not due to any anti- and the length of the bacteria increased and the structure of the
bacterial effect of the silicon material but could be attributed to bacterial cell membrane was distorted. In addition, a halo sur-
drying of bacterial droplet. rounding the bacteria was observed that could be associated with
We also studied the potential antibacterial effects of nano-TiO2 the leakage of organic material or ions for example K+ from the
surface in dark conditions: incubation of bacteria for 20 min on bacterial cell what has also been observed in earlier studies
TiO2 surface decreased the bacterial number from 7.9 ⁄ 107 to [14,40,55]. In general, imaging results correlated well with quan-
3.2 ⁄ 107 CFU/mL (Fig. 1). Similar decrease of viable bacteria (from titative analysis of antibacterial activity, showing increasing num-
108 to 4.3 ⁄ 107 CFU/mL) was observed on Si(1 0 0)-monocrystal ber of bacteria with damaged cell wall when the length of UV-A
surface in dark conditions, i.e. no TiO2 present (Fig. 1; Fig. S8). illumination was increased. 10-min exposure of bacteria to photo-
Again, we proposed that this slight decrease could be attributed activated TiO2 surface, where first distorted cells were seen under
to drying of 10 lL bacterial droplet on the surface of Si-monocrys- SEM, decreased the number of viable bacteria by two orders of
tal and the TiO2 thin film. Very small proportion of non-viable cells magnitude (Fig. 1). 20-min exposure distorted most of the cells
on nano-TiO2 thin films in dark conditions and on Si-monocrystal as visualized by SEM (Fig. 2) and resulted in 100% lethality of bac-
surface under UV-A light suggest that high killing efficiency of teria according to viable colony counts (Fig. 1). Although all the
UV-irradiated TiO2 thin films was indeed due to photoactivation bacteria were killed on photoactivated TiO2 thin films in 20 min,
of nano-TiO2 film. Therefore, we concluded that the TiO2 thin film i.e., no colonies were formed on agarized growth media, cellular
developed in this study was effective in killing bacterial cells after debris was still clearly visible under SEM and complete destruc-
20 min exposure to environmentally relevant UV-A light (22 W/m2 tion of bacterial cells was not achieved even in 60 min (Fig. S9).
corresponds to UV-A in solar spectrum [44]). It is somewhat diffi- Therefore, our data showed that considerably longer time is
cult to compare the results of this study with published literature needed to completely decompose and degrade bacterial cells on
due to variable conditions used in each study. Only in two studies TiO2 surface than the time that is needed to affect bacterial
TiO2 thin films were excited using UV-A light at intensity similar to viability.
this study. Shiraishi et al excited anatase thin films using UV-A
(20 W/m2) with peak intensity at 352 nm and in these conditions, 3.3. Chemical changes in fatty acids on photoactivated nano-TiO2 thin
90 min was required to completely inactivate bacterial cells [52]. film
Yeung et al used UV-A light (26.5 W/m2) with peak intensity at
365 nm and showed that 30 min was required to inhibit 99,7% of Changes in chemical structure of stearic (C18:0), oleic (C18:1
B. subtilis and E. coli cells [37]. Thus, the thin films developed in cis-9) and linoleic acid (C18:2 cis-9,12) – the most abundant
both of these previous studies were remarkably less effective in fatty acids in bacterial plasma membrane [56,57] -induced by
bacterial inactivation than the thin film design in this study. A UV-illuminated nano-TiO2 films were analysed by XPS. The assay
number of published studies have utilized lower intensity UV-A was performed by adding above-mentioned fatty acids on TiO2 thin
light (5–10 W/m2) and therefore, in these studies 60–90 min of films and irradiating with UV-light. XPS spectra obtained from the
UV-irradiation has been applied for total inactivation of bacteria fatty acids before treatment corresponded well to their structure
even when rather diluted bacterial suspension was used [40,53]. and the number of carbon atoms in each different chemical state.
In case of small volume of concentrated bacterial suspension and C 1s XPS band components from sp2 carbon (two single bonds
very high UV-A light intensity (200 W/m2) 90% bacterial inhibition and one double bond) at 284.1 ± 0.1 eV [58], sp3 carbon (four single
on TiO2 thin film was reached very quickly, in 2 min [54]. On the bonds) at 284.8 ± 0.1 eV [59] and carboxylic group at 288.2 ± 0.1 eV
182 U. Joost et al. / Journal of Photochemistry and Photobiology B: Biology 142 (2015) 178–185

Fig. 2. Images of surviving luminescent bacteria that were able to yield colonies after exposure to UV-light on Si-monocrystal substrates (upper block of panels) or on nano-
TiO2 thin films (lower block of panels). Different sectors of agar plates show the colonies of luminescent bacteria in 15 lL of 102, 103 and 104-fold dilutions of UV-exposed
bacterial suspension. The photos were taken in the dark. Inset on each image is an SEM image of the bacteria after respective exposure conditions. Scale bars on SEM images
correspond to 2 lm. SEM images show considerable morphological changes of bacterial cells already after 10-min UV-irradiation on nano-TiO2 thin film. See also Fig. S9 for
absence of analogous cellular damage in dark conditions.

Fig. 3. Decomposition and changes in chemical composition of stearic (C18:0), oleic (C18:1 cis-9) and linoleic (C18:2 cis-9,12)-acid after their exposure to UV-illumination on
nano-TiO2 films. a, b, c – X-ray photoelectron spectra; d, e, f, – percentages of different carbon species.

[58,60,61] could be identified. In oleic and linoleic acid C 1s spectra group is suppressed due to the tendency of fatty acids to form ori-
contributions from both, sp3 and sp2 carbons as well as from car- ented monolayers [62,63] where the hydrocarbon chain is located
boxylic group could be detected (see division of C 1s experimental perpendicular to substrate plane, carboxylic group is located on the
spectra into the sub-bands, Fig. 3b and c); in stearic acid the con- substrate and the length of molecule is comparable to the escape
tributions from sp3 carbon and carboxylic group could be identified depth of photoelectrons [64]. Chemical shifts in the O 1s spectra
(Fig. 3a, 0 min condition). In case of stearic acid C 1s spectra, posi- are smaller than respective shifts in C 1s spectra, also O 1s signal
tion of the contribution from the carboxylic group is at slightly connected with fatty acids was mixed with O 1s signal coming
higher energy at 288.6 ± 0.1 eV (Fig. 3a). In the C 1s spectra of stea- from TiO2. Therefore O 1s signal was not possible to use for similar
ric and oleic acid the intensity of the contribution from carboxylic analysis.
U. Joost et al. / Journal of Photochemistry and Photobiology B: Biology 142 (2015) 178–185 183

As XPS measurements of photo-oxidized fatty acids were con- This statement is in line with our observation on damaged surfaces
ducted in ultra-high vacuum (pressure in the order of 10 10 mbar), of bacterial cells treated with photo-activated TiO2 thin films
photo-oxidation products with low molecular mass and molecules (Fig. 2). We also propose that a halo surrounding the damaged bac-
that were not absorbed on the nano-titania surface were removed terial cells that was visible on SEM images (Fig. 2) may reveal the
leaving behind fatty acids directly linked to the surface of nano- leakage of intracellular components. Thus, our study suggested that
TiO2 films. The latter enabled the monitoring of chemical changes the final result of antibacterial effect of photo-activated nano-TiO2
that occurred to fatty acids without the interference from other thin film was loss of the integrity of the main cellular outer barriers
photo-oxidation products. and bacterial cell envelope.
XPS spectra of the fatty acids after their exposure to
UV-illumination on nano-TiO2 films (Fig. 3) suggested that 4. Conclusions
photo-oxidation of unsaturated and saturated fatty acids was dif-
ferent. Photo-oxidation of stearic acid did not induce any changes This study investigated time-dependent antibacterial properties
in the structure of the fatty acid. Only decrease in sp3 carbons and of photo-activated nano-TiO2 thin films using two different biolog-
carbon in carboxylic groups (marked as OAC@O) was observed ical systems: viable bacterial cells (in vivo approach) and crucial
(Fig. 3a). Thus, in case of this saturated fatty acid no other chemical components of bacterial plasma membrane-fatty acids.
changes than shortening of the alkyl chain resulting finally in total Total inactivation of bacteria on nano-TiO2 thin films after
mineralization of the molecule was detected during photo- 20 min of UV-A irradiation confirmed the remarkable photoactive
oxidation. nature of the nano-TiO2 and developed thin films. SEM investiga-
However, photo-oxidation of oleic (Fig. 3b) and linoleic acid tions confirmed the results of the viability tests, showing that dam-
(Fig. 3c) was different due to the radical reactions associated with age to bacterial cell envelope occurred in the same time frame as
carbon double bond. In the C 1s spectra of oleic and linoleic acid a inactivation of bacteria. XPS analysis indicated that both, saturated
shoulder related to CAO bond [58,60] appeared at 286.2 ± 0.1 eV as well as unsaturated fatty acids commonly present in bacterial
already after 1 min of UV-illumination. The appearance of CAO plasma membrane, decomposed within 10 min of exposure to
bonds during photo-oxidation of unsaturated fatty acids can be photoactivated nano-TiO2 thin films. In addition, before complete
linked to the formation of peroxides, as is proposed by several mineralization peroxidation of unsaturated fatty acids, as evi-
authors [32,34,65]. Formation of peroxides in oleic and linoleic denced by the formation of CAO bonds, took place. Our results
acids was most likely driven by OH radicals attacking a hydrogen indicated that fast and effective peroxidation and decomposition
atom in RAH and by that, creating a carbon radical R. In the next of membrane fatty acids could be one of the factors contributing
step, molecular oxygen is added to R creating a peroxyl radical to the morphological changes and leakage of the bacterial cell com-
ROO, peroxyl radical abstracts a hydrogen from the RAH bond cre- ponents as observed under SEM, and ultimately, bacterial death.
ating a lipid hydroperoxide ROOH. Each lipid hydroperoxide con- Though in membranes fatty acids are bound to form hydrophobic
tains one CAO bond between carbon and oxygen [34]. Formation tails of phospholipids, their decomposition should be similar to
of CAO bond was relatively fast: during the first three minutes of the decomposition of free fatty acids. As fatty acids are present
photo-oxidation the relative number of CAO bonds in oleic and lin- in most biological membranes it is likely that the UV-light-acti-
oleic acid layers increased (Fig. 3(e) and (f)), and then started to vated nano-TiO2 thin films described in this study are applicable
decrease as the total carbon composition decreased due to for removal of most of undesired organisms. Considering the fast
photo-oxidation. After 3-min exposure the amount of CAO carbon and efficient oxidative damage mediated inactivation of bacteria
became similar to the remaining amount of sp2 carbon for both on the nano-TiO2 thin films demonstrated in this study, our data
oleic and linoleic acid layers indicating that the change in the supports further development and application of nanosized TiO2
chemical composition of unsaturated fatty acids was quite exten- based novel surface coating materials.
sive. One OH radical can initiate a process resulting in peroxida-
tion of several fatty acids (radical chain reaction); thus, even very Acknowledgements
low concentrations of OH radicals can cause significant oxidative
damage to the components of bacterial cellular membrane. The Financial support by the Estonian Ministry of Education and
time required for total photo-mineralization was similar for all Research (target-financed themes IUT2-25 and IUT23-5), Estonian
three fatty acids. After 10 min of exposure to UV-illumination on Science Foundation (Grants ETF8561 and 8737), Estonian
nano-TiO2 thin film the peaks of all carbon compounds had disap- Nanotechnology Competence Center (EU29996), ERDF projects
peared in the XPS spectra suggesting total mineralization of fatty (‘‘IRGLASS’’ 3.2.1101.12-0027, ‘‘TRIBOFILM’’ 3.2.1101.12-0028,
acids. ‘‘Nano-Com’’ 3.2.1101.12-0010, ‘‘Mesosystems: Theory and
As stearic, oleic and linoleic acids are the main components of Applications’’ TK114, ‘‘High-technology Materials for Sustainable
bacterial membranes [57,66] it can be suggested that the chemical Development’’ TK117), Graduate School ‘‘Functional Materials
changes that were observed in those fatty acids upon exposure to and Technologies’’ (European Social Fund project 1.2.0401.09-
photo-activated nano-TiO2 thin film were also taking place in E. coli 0079), Development Fund of University of Tartu, is gratefully
cells. Thus, it is reasonable to suggest that already after short-time acknowledged. Markus Laars is gratefully acknowledged for
exposure of E. coli to photo-activated nano-TiO2, unsaturated fatty technical assistance.
acids first change their chemical composition followed by mineral-
ization whereas saturated fatty acids just mineralize. Different
Appendix A. Supplementary material
behaviour of saturated and unsaturated fatty acids during photo-
oxidation process has also been shown earlier; Leung et al. [14]
Supplementary data associated with this article can be found, in
and Kiwi and Nadtochenko [34] showed that cell membrane’s sus-
the online version, at http://dx.doi.org/10.1016/j.jphotobiol.2014.
ceptibility to TiO2 photo-oxidation has been linked to the degree of
12.010.
unsaturation of fatty acid chains of phospholipids.
A recent report by Kubacka et al. [27] showed that the cell wall
References
and cell membrane composition and integrity were one of the main
biological processes that were affected by photo-activated nano- [1] R.Y. Pelgrift, A.J. Friedman, Nanotechnology as a therapeutic tool to combat
TiO2 thin film in pathogenic bacterium Pseudomonas aeruginosa. microbial resistance, Adv. Drug Deliver. Rev. 65 (2013) 1803–1815.
184 U. Joost et al. / Journal of Photochemistry and Photobiology B: Biology 142 (2015) 178–185

[2] D. Campoccia, L. Montanaro, C.R. Arciola, A review of the biomaterials [30] F. Özyildiz, M. Güden, A. Uzel, I. Karaboz, O. Akil, H. Bulut, Antimicrobial
technologies for infection-resistant surfaces, Biomaterials 34 (2013) 8533–8554. activity of TiO2-coated orthodontic ceramic brackets against Streptococcus
[3] A. Ivask, K. Juganson, O. Bondarenko, M. Mortimer, V. Aruoja, K. Kasemets, I. mutans and Candida albicans, Biotechnol. Bioproc. Eng. 15 (2010) 680–685.
Blinova, M. Heinlaan, V. Slaveykova, A. Kahru, Mechanisms of toxic action of [31] J.C. Ireland, P. Klostermann, E.W. Rice, R.M. Clark, Inactivation of Escherichia
Ag, ZnO and CuO nanoparticles to selected ecotoxicological test organisms and coli by titanium dioxide photocatalytic oxidation, Appl. Environ. Microbiol. 59
mammalian cells in vitro: a comparative review, Nanotoxicology 8 (S1) (2014) (1993) 1668–1670.
57–71. [32] O.K. Dalrymple, E. Stefanakos, M.A. Trotz, D.Y. Goswami, A review of the
[4] A. Ivask, S. George, O. Bondarenko, A. Kahru, Metal-containing nano- mechanisms and modeling of photocatalytic disinfection, Appl. Catal. B-
antimicrobials: differentiating the impact of solubilized metals and particles. Environ. 98 (2010) 27–38.
in: N. Cioffi, M. Rai (Ed.), Nano-Antimicrobials: Progress and Prospects, [33] P-C. Maness, S. Smolinski, D.M. Blake, Z. Huang, E.J. Wolfrum, W.A. Jacoby,
Springer, 2012, pp. 253–290 Bactericidal activity of photocatalytic TiO2 reaction: toward an understanding
[5] O. Bondarenko, K. Juganson, A. Ivask, K. Kasemets, M. Mortimer, A. Kahru, of its killing mechanism, Appl. Environ. Microbiol. 65 (1999) 4094–4098.
Toxicity of Ag, CuO and ZnO nanoparticles to selected environmentally [34] J. Kiwi, V. Nadtochenko, New evidence for TiO2 photocatalysis during bilayer
relevant test organisms and mammalian cells in vitro: a critical review, Arch. lipid peroxidation, J. Phys. Chem. B 108 (2004) 17675–17684.
Toxicol. 87 (2013) 1181–1200. [35] L. Rizzo, A.D. Sala, A. Fiorentino, G.L. Puma, Disinfection of urban wastewater
[6] A. Kahru, H-C. Dubourguier, From ecotoxicology to nanoecotoxicology, by solar driven and UV lamp – TiO2 photocatalysis: effect on a multi drug
Toxicology 269 (2010) 105–119. resistant Escherichia coli strain, Water Res. 53 (2014) 145–152.
[7] M. Keshmiri, M. Mohseni, T. Troczynski, Development of novel TiO2 sol–gel- [36] L. Armelao, D. Barreca, G. Bottaro, A. Gasparotto, C. Maccato, C. Maragno, E.
derived composite and its photocatalytic activities for trichloroethylene Tondello, U.L. Štangar, M. Bergant, D. Mahne, Photocatalytic and antibacterial
oxidation, Appl. Catal. B-Environ. 53 (2004) 209–219. activity of TiO2 and Au/TiO2 nanosystems, Nanotechnology 18 (2007) 375709.
[8] G.K. Kiema, M.J. Colgan, M.J. Brett, Dye sensitized solar cells incorporating [37] R-M. Wang, B-Y. Wang, Y-F. He, W-H. Lv, J-F. Wang, Preparation of composited
obliquely deposited titanium oxide layers, Sol. Energy Mat. Sol. C 85 (2005) Nano-TiO2 and its application on antimicrobial and self-cleaning coatings,
321–331. Polym. Adv. Technol. 21 (2010) 331–336.
[9] R. Wang, K. Hashimoto, A. Fujishima, M. Chikuni, E. Kojima, A. Kitamura, M. [38] K.L. Yeung, W.K. Leung, N. Yao, S. Cao, Reactivity and antimicrobial properties
Shimohigoshi, T. Watanabe, Light-induced amphiphilic surfaces, Nature 388 of nanostructured titanium dioxide, Catal. Today 143 (2009) 218–224.
(1997) 431–432. [39] R. Aminedi, G. Wadhwa, N. Das, B. Pal, Shape-dependent bactericidal activity of
[10] R. Wang, K. Hashimoto, A. Fujishima, M. Chikuni, E. Kojima, A. Kitamura, M. TiO2 for the killing of Gram-negative bacteria Agrobacterium tumefaciens under
Shimohigoshi, T. Watanabe, Photogeneration of highly amphiphilic TiO2 UV torch irradiation, Environ. Sci. Pollut. Res. 20 (2013) 6521–6530.
surfaces, Adv. Mater. 10 (1998) 135–138. [40] K. Sunada, T. Watanabe, K. Hashimoto, Studies on photokilling of bacteria on
[11] J. Chen, L. Eberlein, C.H. Langford, Pathways of phenol and benzene TiO2 thin film, J. Photochem. Photobiol. A 156 (2003) 227–233.
photooxidation using TiO2 supported on a zeolite, J. Photochnol. Photobiol. A [41] T.P.T. Cushnie, P.K.J. Robertson, S. Officer, P.M. Pollard, R. Prabhu, C. McCullagh,
148 (2002) 183–189. J.M.C. Robertson, Photobactericidal effects of TiO2 thin films at low
[12] M. Anpo, T. Shima, S. Kodama, Y. Kubokawa, Photocatalytic hydrogenation of temperatures – a preliminary study, J. Photochem. Photobiol. A 216 (2010)
CH3CCH with H2O on small-particle TiO2: size quantization effects and 290–294.
reaction intermediates, J. Phys. Chem. 91 (1987) 4305–4310. [42] E. Scolan, C. Sanchez, Synthesis and characterization of surface-protected
[13] M. Stylidi, D.I. Kondarides, X.E. Verykios, Pathways of solar light-induced nanocrystalline titania particles, Chem. Mater. 10 (1998) 3217–3223.
photocatalytic degradation of azo dyes in aqueous TiO2 suspensions, Appl. [43] U. Joost, A. Saarva, M. Visnapuu, E. Nõmmiste, K. Utt, R. Saar, V. Kisand,
Catal. B-Environ. 40 (2003) 271–286. Purification of titania nanoparticles thin films: triviality or a challenge?, Ceram
[14] T.Y. Leung, C.Y. Chan, C. Hu, J.C. Yu, P.K. Wong, Photocatalytic disinfection of Int. 40 (2014) 7125–7132.
marine bacteria using fluorescent light, Water Res. 42 (2008) 4827–4837. [44] Reference Solar Spectral Irradiance: Air Mass 1.5 <http://rredc.nrel.gov/solar/
[15] Y.W. Cheng, R.C.Y. Chan, P.K. Wong, Disinfection of Legionella pneumophila by spectra/am1.5/>.
photocatalytic oxidation, Water Res. 41 (2007) 842–852. [45] A. Ivask, T. Rõlova, A. Kahru, A suite of recombinant luminescent bacterial
[16] K.P. Kühn, I.F. Chaberny, K. Massholder, M. Stickler, V.W. Benz, H.-G. Sonntag, strains for the quantification of bioavailable heavy metals and toxicity testing,
L. Erdinger, Disinfection of surfaces by photocatalytic oxidation with titanium BMC Biotechnol. 9 (2009) 41.
dioxide and UV-A light, Chemosphere 53 (2003) 71–77. [46] M.P. Seah, I.S. Gilmore, S.J. Spencer, Quantitative XPS: I. analysis of X-ray
[17] C.M. Visinescu, R. Sanjines, F. Lévy, V.I. Pârvulescu, Photocatalytic degradation photoelectron intensities from elemental data in a digital photoelectron
of acetone by Ni-doped titania thin films prepared by dc reactive sputtering, database, J. Electron Spectrosc. 120 (2001) 93–111.
Appl. Catal. B-Environ. 60 (2005) 155–162. [47] N. Fairley CasaXPS version 2.3.12. www.casaxps.com.
[18] Y-M. Lim, J-H. Jeong, J-H. An, Y-S. Jeon, K-O. Jeon, K-S. Hwang, B-H. Kim, [48] L. Peruchon, E. Puzenat, A. Girard-Egrot, L. Blum, J.M. Herrmann, C. Guillard,
Nickel-doped titanium oxide films prepared by chemical solution deposition, J. Characterization of self-cleaning glasses using Langmuir–Blodgett technique to
Ceram. Process. Res. 6 (2005) 302–304. control thickness of stearic acid multilayers: importance of spectral emission
[19] C.E.R. Torres, A.F. Cabrera, L.A. Errico, S. Duhalde, M. Rentería, F. Golmar, F.A. to define standard test, J. Photochem. Photobiol. A 197 (2008) 170–176.
Sánchez, XAS study of the local environment of impurities in doped TiO2 thin [49] A.B. Panda, S.K. Mahapatra, P.K. Barhai, A.K. Das, I. Banerjee, Understanding of
films, Physica B 398 (2007) 219–222. gas phase deposition of reactive magnetron sputtered TiO2 thin films and its
[20] U. Joost, R. Pärna, M. Lembinen, K. Utt, I. Kink, M. Visnapuu, V. Kisand, Heat correlation with bactericidal efficiency, Appl. Surf. Sci. 258 (2012) 9824–9831.
treatment and substrate dependent properties of titania thin films with high [50] Y. He, Y. Su, Silicon Nano-biotechnology SpringerBriefs in Molecular Science,
copper loading, Phys. Status Solidi A 210 (2013) 1201–1212. VIII 2014 p. 109.
[21] R. Pärna, U. Joost, E. Nõmmiste, T. Käämbre, A. Kikas, I. Kuusik, M. Hirsimäki, I. [51] J. Su, H. Yu, X. Quan, S. Chen, H. Wang, Hierarchically porous silicon with
Kink, V. Kisand, Effect of cobalt doping and annealing on properties of titania significantly improved photocatalytic oxidation capability for phenol
thin films prepared by sol–gel process, Appl. Surf. Sci. 257 (2011) 6897–6907. degradation, Appl. Catal. B-Environ. 138 (2013) 427–433.
[22] R. Pärna, U. Joost, E. Nõmmiste, T. Käämbre, A. Kikas, I. Kuusik, I. Kink, M. [52] K. Shiraishi, H. Koseki, T. Tsurumoto, K. Baba, M. Naito, K. Nakajama, H. Shindo,
Hirsimäki, V. Kisand, Effect of different annealing temperature and SiO2/ Antibacterial metal implant with a TiO2-conferred photocatalytic bactericidal
Si(100) substrate on the properties of nickel containing titania thin sol–gel effect against Staphylococcus aureus, Surf. Interface Anal. 41 (2009) 17–22.
films, Phys. Status Solidi A 209 (2012) 953–965. [53] T.C. Cheng, C.Y. Chang, C.I. Chang, C.J. Hwang, H.C. Hsu, D.Y. Wang, K.S. Yao,
[23] V. Kisand, U. Joost, V. Reedo, R. Pärna, T. Tätte, J. Shulga, A. Saar, L. Matisen, A. Photocatalytic bactericidal effect of TiO2 film on fish pathogens, Surf. Coat.
Kikas, I. Kink, Influence of the heating temperature on the properties of nickel Technol. 203 (2008) 925–927.
doped TiO2 films prepared by sol–gel method, Appl. Surf. Sci. 256 (2010) [54] M. Lilja, J. Forsgren, K. Welch, M. Astrand, H. Enqvist, M. Stromme,
4538–4542. Photocatalytic antimicrobial properties of surgical implant coatings of
[24] O. Carp, C.L. Huisman, A. Reller, Photoinduced reactivity of titanium dioxide, titanium dioxide deposited though cathodic arc evaporation, Biotechnol.
Prog. Solid State Chem. 32 (2004) 33–177. Lett. 34 (2012) 2299–2305.
[25] S. Kim, W. Choi, Kinetics and mechanisms of photocatalytic degradation of [55] H.A. Foster, I.B. Ditta, S. Varghese, A. Steele, Photocatalytic disinfection using
(CH3)nNH+4 n (0 6 n 6 4) in TiO2 suspension: the role of OH radicals, Environ. titanium dioxide: spectrum and mechanism of antimicrobial activity, Appl.
Sci. Technol. 36 (2002) 2019–2025. Microbiol. Biotechnol. 90 (2011) 1847–1868.
[26] M. Cho, H. Chung, W. Choi, J. Yoon, Linear correlation between inactivation of [56] A.P. Day, J.D. Oliver, Changes in membrane fatty acid composition during entry
E. coli and OH radical concentration in TiO2 photocatalytic disinfection, Water of Vibrio vulnificus into the viable but nonculturable state, J. Microbiol. 42
Res. 38 (2004) 1069–1077. (2004) 69–73.
[27] A. Kubacka, M.S. Diez, D. Rojo, R. Bargiela, S. Ciordia, I. Zapico, J.P. Albar, C. [57] M.M. Or-Rashid, N.E. Odongo, B.W. McBride, Fatty acid composition of ruminal
Barbas, V.A.P. Martinos dos Santos, M. Fernández-García, M. Ferrer, bacteria and protozoa, with emphasis on conjugated linoleic acid, vaccenic
Understanding the antimicrobial mechanism of TiO2-based nanocomposite acid, and odd-chain and branched-chain fatty acids, J. Animal Sci. 85 (2007)
films in a pathogenic bacterium, Sci. Rep. 4 (2014) 4134. 1228–1234.
[28] M.C. Lee, W. Choi, Solid phase photocatalytic reaction on the soot/TiO2 [58] V. Datsyuk, M. Kalyva, K. Papagelis, J. Parthenios, D. Tasis, A. Siokou, I. Kallitsis,
interface: the role of migrating OH radicals, J. Phys. Chem. B 106 (2002) C. Galiotis, Chemical oxidation of multiwalled carbon nanotubes, Carbon 46
11818–11822. (2008) 833–840.
[29] C.S. Turchi, D.F. Ollis, Photocatalytic degradation of organic water [59] J. Diaz, G. Paolicelli, S. Ferrer, F. Comin, Separation of the sp3 and sp2
contaminants: mechanisms involving hydroxyl radical attack, J. Catal. 122 components in the C1s photoemission spectra of amorphous carbon films,
(1990) 178–192. Phys. Rev. B 54 (1996) 8064–8069.
U. Joost et al. / Journal of Photochemistry and Photobiology B: Biology 142 (2015) 178–185 185

[60] B.P. Payne, M.C. Biesinger, N.S. McIntyre, X-ray photoelectron spectroscopy [64] J. Zemek, S. Hucek, A. Jablonski, I.S. Tilinin, Photoelectron escape depth, J.
studies of reactions on chromium metal and chromium oxide surfaces, J. Electron. Spectrosc. Relat. Phenom. 76 (1995) 443–447.
Electron. Spectrosc. Relat. Phenom. 184 (2011) 29–37. [65] O.K. Dalrymple, W. Isaacs, E. Stefanakos, M.A. Trotz, D.Y. Goswami, Lipid
[61] S. Beverly, S. Seal, S. Hong, Identification of surface chemical functional groups vesicles as model membranes in photocatalytic disinfection studies, J.
correlated to failure of reverse osmosis polymeric membranes, J. Vac. Sci. Photochem. Photobiol. A 221 (2011) 64–70.
Technol. A 18 (2000) 1107–1113. [66] J. Shumann, A. Leichtle, J. Thiery, H. Fuhrmann, Fatty acid and peptide profiles
[62] J.B. Brzoska, I.B. Azouz, F. Rondelez, Silanization of solid substrates: a step in plasma membrane and membrane rafts of PUFA supplemented RAW264.7
toward reproducibility, Langmuir 10 (1994) 4367–4373. macrophages, Plos One 6 (2011) e24066.
[63] J. Rathousky, V. Kalousek, M. Kolár, J. Jirkovsky, P. Barták, A study into the self-
cleaning surface properties – the photocatalytic decomposition of oleic acid,
Catal. Today. 161 (2011) 202–208.

You might also like