Вы находитесь на странице: 1из 9

Journal of Photochemistry & Photobiology, B: Biology 178 (2018) 395–403

Contents lists available at ScienceDirect

Journal of Photochemistry & Photobiology, B: Biology


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

Photo-induced toxicity of tungsten oxide photochromic nanoparticles T


a b b b b a
A.L. Popov , N.M. Zholobak , O.I. Balko , O.B. Balko , A.B. Shcherbakov , N.R. Popova ,

O.S. Ivanovac, A.E. Baranchikovc, V.K. Ivanovc,d,
a
Institute of Theoretical and Experimental Biophysics, Russian Academy of Sciences, Pushchino, Moscow region 142290, Russia
b
Zabolotny Institute of Microbiology and Virology, National Academy of Sciences of Ukraine, Kyiv D0368, Ukraine
c
Kurnakov Institute of General and Inorganic Chemistry, Russian Academy of Sciences, Moscow 119991, Russia
d
National Research Tomsk State University, Tomsk 634050, Russia

A R T I C L E I N F O A B S T R A C T

Keywords: We synthesised a new type of photochromic tungsten oxide nanoparticles, analysed their photocatalytic activity
Tungsten oxide and carried out a thorough analysis of their effect on prokaryotic and eukaryotic organisms. Ultrasmall hydrated
Nanoparticles tungsten oxide nanoparticles were prepared by means of hydrothermal treatment of tungstic acid in the presence
Phototoxicity of polyvinylpyrrolidone as a template, stabiliser and growth regulator. Tungstic acid was synthesised through an
Apoptosis
ion-exchange method using sodium tungstate solution and a strongly acidic cation exchange resin.
UV irradiation
Upon illumination, photochromic nanoparticles of WO3 were shown to increase greatly their toxicity against
both bacterial (both gram-positive and gram-negative – P. aeruginosa, E. coli and S. aureus) and mammalian cells
(primary mouse embryonic fibroblasts); under the same conditions, fungi (C. albicans) were less sensitive to the
action of tungsten oxide nanoparticles. UV irradiation of primary mouse fibroblasts in the presence of WO3
nanoparticles demonstrated a time- and dose-dependent toxic effect, the latter leading to a significant decrease
in dehydrogenase activity and an increase in the number of dead cells. WO3 nanoparticles were photo-
catalytically active under both UV light and even diffused daylight filtered through a window glass, leading to
indigo carmine organic dye discolouration.
The obtained experimental data not only show good prospects for biomedical applications of tungsten tri-
oxide, but also demonstrate the need for clear control of biosafety when it is used in various household materials
and appliances.

1. Introduction tance levels depending on changing needs [1]. One of the most pro-
mising materials for photochromic films or coatings is nanocrystalline
The modern technological paradigm requires research and devel- tungsten oxide, an n-type wide-bandgap semiconductor with a chemical
opment into new materials with specific physical and chemical prop- type of chromism [2,3]. The electro-, photo- and chemochromic prop-
erties. Recently, much attention has been paid to the development of erties of tungsten oxide (WO3) are widely used in electronic displays,
various types of stimuli-responsive materials, which are able to change optical modulators, windows with adjustable light transmission, rear-
their characteristics in response to external factors. For instance, in the view mirrors in cars, etc. [4–7]. Under light irradiation, exposure to
context of energy-saving, smart photo- or electrochromic materials chemical reagents or the application of an electric field, stoichiometric
have been developed which are able to control the throughput of visible WO3 (faintly yellowish) undergoes a number of serial transformations,
light and solar radiation into buildings by having different transmit- reversibly forming brightly coloured products [8]:


Corresponding author.
E-mail address: van@igic.ras.ru (V.K. Ivanov).

https://doi.org/10.1016/j.jphotobiol.2017.11.021
Received 15 August 2017; Received in revised form 14 November 2017; Accepted 15 November 2017
Available online 20 November 2017
1011-1344/ © 2017 Elsevier B.V. All rights reserved.
A.L. Popov et al. Journal of Photochemistry & Photobiology, B: Biology 178 (2018) 395–403

Fig. 1. Schematic diagram of the interaction mechanisms between light and WO3 nanoparticles in semiconductor (A) and plasmonic (B) states. Photocatalysis, autophotoreduction and
plasmonic heating are demonstrated.

In addition to the above-mentioned applications, tungsten oxide is contact with living cells, these ROS can cause oxidative stress, resulting
one of the most thoroughly studied photocatalysts [9–15] for organic in cell death. Another specific tungsten oxide feature is that it possesses
dye degradation. WO3 nanoparticles of different morphologies, ob- photoactivity at a wide range of wavelengths. The most widely explored
tained by the hydrothermal-microwave method, have been shown to TiO2-photocatalysts are effective under UV light only, while WO3-based
possess photocatalytic properties with respect to the discolouration of ones can change their bandgap upon irradiation, and thus tungsten
rhodamine B, indigo carmine and tetracycline hydrochloride under trioxide could be an effective visible light photocatalyst, wherein the
UV–vis irradiation [11]. WO3 nanoparticles obtained by surfactant-as- absorbed efficiency of sunlight can be enhanced enormously.
sisted sonication have provided a high rate of rhodamine B and indigo The autophotoreduction of tungsten oxide is accompanied by for-
carmine degradation under xenon lamp irradiation [12]. Multi-phase mation of free charge carriers, so photochromic colouration of tungsten
WO3 samples possess improved photocatalytic activity, (as demon- oxide occurs due to the local surface plasmon resonance (LSPR) arising
strated by the bleaching of rhodamine B), that can be attributed to a from appreciable free carrier concentrations [17,18]. It is well known
decrease in the electron-hole recombination rate, owing to the forma- that the interaction of plasmonic particles and light (HxWO3 → HxWO3⁎
tion of interphase junctions [13]. The efficient degradation of methy- photoexcitation) leads not only to redox processes on their surface (left
lene blue dye by tungsten oxide nanoplates, synthesised using a hy- part of Fig. 1, B), but also to the partial conversion of electromagnetic
drothermal or microwave-hydrothermal method, has been shown under energy into heat (right part of Fig. 1, B). The volumetric generation of
UV light irradiation [14]. The photocatalytic activity of WO3 nano- heat within the plasmonic nanoparticle Q(r, t) is affected by the in-
particles (obtained by annealing (NH4)xWO3 − y at 500 °C in air) and tensity of light, the particle's internal electromagnetic field distribution,
nanorods (prepared using a hydrothermal method using Na2WO4, HCl, and the thermal and electrical conductivity of the particle's material
(COOH)2 and NaHSO4 precursors at 200 °C) has been confirmed by the [18]:
decomposition of methyl orange in an aqueous solution under UV light
1 4πnk
σE (̃ r , t ) ∙E ̃ (r, t ); σ =

irradiation [15]. Q (r , t ) =
2 λi μc
The common mechanism for organic dye degradation by photo-
active semiconductor (photocatalyst) is represented in the left part of where Ẽ(r, t) and Ẽ∗(r, t) are the generated electric field and its complex
Fig. 1, A: conjugate within the nanoparticles; σ – electrical conductivity at optical
When a photocatalyst absorbs light that has an energy that exceeds frequencies; λi – incident light wavelength; μ – relative magnetic per-
the semiconductor's bandgap, an electron of the valence band is pro- meability of the nanoparticle, and n and k – the real and imaginary
moted to the conduction band, thus creating an electron (e−)/hole (h+) parts of the refraction index of the nanoparticle, respectively. Plas-
pair. Due to the semiconductor's photoexcitation (WO3 → WO3⁎) and monic nanoparticles of non-stoichiometric tungsten oxide have been
the generation of electron/hole pairs, oxidation-reduction reactions used for photothermal cancer treatment upon IR irradiation [18–23].
take place at the surface: electrons reduce and holes oxidise the mole- Thus, volumetric heating of WO3 nanoparticles upon irradiation can
cules of the surrounding substrate (Sub). Upon contact between the dye bring an additional contribution to cytotoxicity, and thus should be
and the photoexcited nanoparticle, direct redox decomposition of the taken into account.
dye molecule occurs (Sub = Dye). More often, the first stage of pho- For this paper, we synthesised a new type of photochromic tungsten
tocatalysis is the redox transformation of water and/or oxygen oxide nanoparticles, analysed their photocatalytic activity and carried
(Sub = H2O, O2), forming the reactive oxygen species (ROS), which out a thorough analysis of their effect on prokaryotic and eukaryotic
then destroy other substances, (the indirect redox decomposition of the organisms. WO3 nanoparticles possess both dark and light cytotoxicity,
dye molecule). Obviously, the photocatalytic activity of the material which significantly distinguishes them from common photocatalysts,
can affect not only the decomposition of organic dyes, but also the and opens up new possibilities for their practical use.
biological components of living cells.
The mechanism of tungsten oxide photo-induced chromism (au-
tophotoreduction) is represented in the right part of Fig. 1, A. The ex- 2. Materials and Methods
cited electrons cause the reduction of tungsten (6 +) to (5 +) ions,
leading to the formation of coloured, non-stoichiometric, hydrated 2.1. Tungsten Oxide Nanoparticles
tungsten oxide (HxWO3, 0 < x < 1); the holes oxidise substrate (for
example, water [16]), forming ROS, such as hydroxyl radicals. Upon Ultrasmall tungsten oxide nanoparticles were synthesised by means
of hydrothermal treatment of tungstic acid in the presence of

396
A.L. Popov et al. Journal of Photochemistry & Photobiology, B: Biology 178 (2018) 395–403

polyvinylpyrrolidone (PVP K-30, average mol. wt. 40,000) as a tem- Microorganisms were cultured in a slant meat-peptone agar medium
plate, stabiliser and growth regulator. Tungstic acid was prepared at 37°C for 24 h. Then they were suspended in a physiological isotonic
through an ion-exchange method using sodium tungstate (Na2WO4) solution (0.85% NaCl) and adjusted to the McFarland turbidity stan-
solution and a strongly acidic cation exchange resin (Amberlite® dard no.4 (1.2 × 109 CFU/ml). Obtained suspensions were diluted, by
IR120). Briefly, ion exchange resin (in a hydrogen form) was swollen in saline, to the required concentrations and further used to determine the
water and loaded into a glass column of 200 ml volume. Then, 100 ml antimicrobial properties of WO3 nanoparticles.
of 0.05 M sodium tungstate solution was passed through the column,
dropwise, 4 g of PVP was added to the obtained eluent, and the solution 2.3.2. Bacterial Viability Assays
was transferred to a flask and stirred for 4 h under reflux. During Two antimicrobial susceptibility tests were used: disc diffusion and
heating, the clear sol of tungsten trioxide was formed, as evidenced by dilution assay. The agar overlay assay was carried out as described
the appearance of a UV absorption band at 325 nm and a Tyndall cone. elsewhere [24,25]; it was shown that this technique is not suitable for
Transmission electron microscopy (TEM) and selected area electron nanoparticle toxicity study, due to their low diffusion capacity in agar.
diffraction (SAED) measurements were performed using a Leo912 AB The suspension dilution assay was carried out as described elsewhere
Omega analytical transmission electron microscope operating at an [26,27]; this technique proved to be adequate for the evaluation of
accelerating voltage of 100 kV. Energy dispersive X-ray (EDX) spectra bactericidal activity of the WO3 nanoparticles.
were recorded using a Carl Zeiss NVision 40 scanning electron micro- Briefly, the experiments were carried out in 1.5 ml Eppendorf tubes;
scope equipped with an Oxford Instruments X-Max detector (80 mm2) 0.1 ml of WO3 nanoparticle sols of different concentrations (test group),
at an accelerating voltage of 20 kV. or 0.1 ml of distilled water (control), were added to 0.9 ml of a sus-
For the toxicological experiments, the obtained sol was diluted with pension of microorganisms having a total titer of about 105 CFU/ml. All
water to produce 0.1–25.0 mg/ml colloid solutions. the experiments were carried out in two parallel sets, under light and
dark conditions.
2.2. Photocatalytic Activity Measurements The number of viable microorganisms was measured both in the
control samples, and in the test tubes after incubation with nano-
2.2.1. Dye particles in the dark and under UV irradiation for various intervals of
Indigo carmine was purchased from Sigma-Aldrich (Aldrich, time. Bacterial suspensions incubated under the same conditions, but
#57000) and used as received, without further purification. without nanoparticles, were used as a control. To determine the titer of
microorganisms, 0.1 ml of bacterial suspension was sampled, then im-
2.2.2. Photocatalytic Activity Test mediately diluted using the method of serial tenfold dilutions, and then
The photocatalytic activity of the obtained nanoparticles was eval- 10 μl aliquots were plated on nutrient agar. Cells were cultivated at
uated in the reaction of photodecomposition of indigo carmine in an 37°С for 24 h and the number of cell colonies on agar plates was
aqueous solution under UV irradiation, without a detailed study of the counted. The number of grown colonies was extrapolated to 1 ml of
reaction intermediates. The source of ultraviolet radiation was a ver- suspension and measured in CFU/ml.
tically installed DELUX T8 36 W G13 UV lamp (analogue of Philips TUV
G15 T8) with a maximum emission at 253.7 nm. The reaction con- 2.3.3. UV Irradiation
tainers were 1 × 1 cm cuvettes (standard quartz cells for spectro- Test tubes were exposed to ultraviolet irradiation through UV-A
photometer), located around an UV lamp at a distance of 10 cm. (λmax = 340 nm) Wood's lamp (Philips, 15 W). Samples were placed at
Cuvettes were filled with 0.25 mM of indigo carmine aqueous solution a distance of 30 cm from the lamp for 30 or 60 min; the control samples
and colourless tungsten oxide sols with various concentrations in the were kept under similar conditions, at room temperature and in the
range of 0–0.18 mM. The temperature of the environment was held at dark. Control tubes not containing nanoparticles were kept under both
25 ± 0.5 °C. The concentration of the dye in the cuvettes was de- light and dark conditions.
termined every 4–16 min, and then the cuvettes were shaken and im-
mediately returned back to their location near the lamp. The con- 2.4. Mammalian Cells Toxicity Study
centration change of the dye was determined using an Agilent
Technologies Cary 5000 UV–Vis spectrophotometer at the adsorption 2.4.1. Cell Culture
maximum of the dye λ = 600–620 nm. The removal of indigo carmine The culture of primary mouse embryonic fibroblasts was obtained
(%) was calculated according to the equation: from embryos of SHK white mice on day 13 of pregnancy, according to
the previously published protocol [28]. The cells were cultured in
C0 − Ct ∗
Dye removal (%) = 100 DMEM/F12 (1:1) medium, with the addition of 10% fetal calf serum
C0
and 100 U/ml penicillin/streptomycin, under 5% СО2 at 37°С. All ex-
where C0 and Ct are the initial concentration of indigo carmine and the periments were conducted using cultures taken from passages 1–2.
concentration of indigo carmine at a given irradiation duration (t), re-
spectively. A similar series of experiments was performed using natural 2.4.2. UV Irradiation
scattered illumination. For this, the cuvettes were exposed on a win- Primary mouse embryonic fibroblasts of the experimental group
dowsill and illuminated by diffused summer sunlight at noon (average were irradiated for 10, 15 or 20 mins with the UV light (pulsed UV lamp
European latitude), which passed through two conventional 3 mm si- with a continuous spectrum of radiation in the range of 200–700 nm
licate glasses. Melitta Alpha 05, 50 W cm− 2); the cells in the control group were not
irradiated. In the next step, cells were grown for an additional 24 h in
2.3. Bacterial Toxicity Study the CO2-incubator at 37 °C, under a humidified atmosphere of 5% CO2
(v/v), and were subsequently sampled for morphology determination,
2.3.1. Cell Cultures live/dead viability or MTT assay.
Four reference strains from the Ukrainian collection of micro-
organisms (UCM, Zabolotny Institute of Microbiology and Virology, 2.4.3. MTT Assay
National Academy of Sciences of Ukraine) were tested: Staphylococcus Determination of activity of mitochondrial and cytoplasmic dehy-
aureus UCM B-904 (ATCC 25923), Escherichia coli UCM B-906 (ATCC drogenases in living cells was performed using the MTT assay, which is
25922), Pseudomonas aeruginosa UCM B-907 (ATCC 27853) and based on the reduction of a colourless tetrazolium salt (3-[4,5-di-
Candida albicans UCM Y-2681 (ATCC 10231). methylthiazol-2-yl]-2,5-diphenyltetrazolium bromide, MTT). After an

397
A.L. Popov et al. Journal of Photochemistry & Photobiology, B: Biology 178 (2018) 395–403

interval of 24 h after UV irradiation, a standard MTT assay was per- small WO3 nanoparticles demonstrated a very high rate of photo-
formed. chromic colouration, which takes place especially easily in an aqueous
medium. Our preliminary data indicate that the bandgap of thus syn-
2.4.4. Live/Dead Viability Assay thesised tungsten oxide was size-dependent. The optical bandgap cal-
To evaluate the cytotoxic effects of WO3 nanoparticles in combi- culated using the absorbance spectra (Fig. 1, C) was Eg ≈ 3.2 eV, while
nation with UV light irradiation, a Live/Dead Viability Kit (Invitrogen, for the bulk material Eg0 = 2.6 eV [29].
Life Technologies) was used. Cells attached to the 12-well plates were
processed according to the manufacturer's protocol and visualised and 3.2. Photocatalytic Activity
photographed 25 min after adding the kit, using an Axiovert 200
fluorescence-light microscope (Carl Zeiss, Germany), and recorded Indigo carmine (C16H8N2Na2O8S2) dye is prone to photodegradation
using a Canon A620 digital camera (Canon, USA). A green signal (SYTO under illumination, and is often used in connection with the photo-
9, λ = 485/498 nm) characterised the live cells and a red signal (pro- catalytic activity of nanoparticles tests [30–33], including tungsten
pidium iodide, λ = 535/617 nm) characterised the dead cells. For each oxide nanoparticles [11,12,32,33]. Our data indicate that the synthe-
cell group, four fields in each well were examined. sised WO3 sols did possess photocatalytic activity in the processes of
To evaluate the cytotoxic effect of the WO3 nanoparticles in com- indigo carmine discolouration under illumination (Fig. 3). According to
bination with UV irradiation, primary mouse embryonic fibroblasts UV–vis measurements, no absorption bands intrinsic to WO3 nano-
were incubated for 24 h in the cell growth medium, supplemented with particles in the range of 600–800 nm were observed until complete
different concentrations of WO3 nanoparticles. Next, the medium was indigo carmine discolouration (Fig. 1S).
changed for a fresh one without the WO3 nanoparticles, and the cell Indigo carmine is light-sensitive dye and can be easily decomposed
cultures were irradiated for 10, 15 and 20 min. After 24 h, the cultures under intense UV irradiation (Fig. 3, A, curve “0”). Photocatalytically
were sampled for the morphological study, the live/dead viability and active tungsten oxide nanoparticles drastically accelerated the rate of
MTT assay. dye decomposition in a concentration-dependent manner (Fig. 3, A,
curves “1”–“5”). Moreover, the photocatalytic activity of WO3 nano-
3. Results and Discussion particles remains at nearly the same high level (Fig. 3, A, curves “1”-
“5”) even upon illumination with diffused daylight filtered through a
3.1. Tungsten Oxide Nanoparticles common window, which contained only a small fraction of UV com-
ponent and could not directly destroy the dye molecules (Fig. 3, B,
Highly photochromic, ultra-small, WO3 nanoparticles were synthe- curve “0”). This fact should be taken into account by the users of “smart
sised using the protocol proposed, based on the hydrothermal treatment windows” containing tungsten oxide nanoparticles. In turn, according
of tungstic acid in the presence of polyvinylpyrrolidone (Fig. 2). Ultra- to our observations, WO3 sols illuminated by UV/yellow-filtered light

Fig. 2. A – TEM image and SAED pattern; B – EDX spectrum of WO3 nanoparticles. C, D – absorbance of WO3 sols upon exposure to direct sunlight; the spectra were taken in the dark
every 2 min. C – UV–Vis-NIR spectra, D – dynamics of the relative change of the optical density of the UV–Vis-NIR bands of WO3 sol when exposed in the dark.

398
A.L. Popov et al. Journal of Photochemistry & Photobiology, B: Biology 178 (2018) 395–403

Fig. 4. Antibacterial activity of WO3 nanoparticles to various bacterial strains and yeasts
after 30 min of exposure without UV irradiation.

protozoa and bacteria (including S. aureus and E. coli) was recently


tested [34]: for most of the test species, tungsten oxide nanoparticles
were found to be non-toxic at concentrations below 0,1 mg/ml. Simi-
larly, antibacterial activity of some metal and metal oxide nanoparticles
against peri-implantitis pathogens Prevotella intermedia, Porphyromonas
gingivalis, Fusobacterium nucleatum and Aggregatibacter actinomyce-
temcomitans was studied [35]. Tungsten oxide had the lowest cytoxo-
city, and the antibacterial activity of the nanoparticles decreased in
the following sequence: Ag > Ag + CuO > Cu2O > CuO > Ag
+ ZnO > ZnO > TiO2 > WO3. On the other hand, a number of re-
searchers reported the high antibacterial activity of tungsten oxide.
Moreover, tungsten species have been shown to damage bacterial
Fig. 3. Relative dynamics of indigo carmine dye decomposition by WO3 nanoparticles
genomic material [36]. For example, the genotoxic effects of nano-WO3
upon UV (A) and diffused daylight (B) illumination. Indigo carmine concentration – were assessed in bacteria using a reverse mutation assay in S. typhi-
0.25 mM, tungsten oxide concentrations: “0” – 0 mM (control); “1” – 0.015 mM; “2” – murium [37], or by analysing the direct interaction of WO3 nanoplates
0.03 mM; “3” – 0.06 mM; “4” – 0.09 mM; “5” – 0.18 mM. (For interpretation of the re- with naked DNA plasmid, which was further transferred into E. coli cells
ferences to colour in this figure legend, the reader is referred to the web version of this for a mutation study [38]. WO3 prepared from peroxotungstic acid
article.)
showed high antibacterial activity against Staphylococcus aureus and
Klebsiella pneumonia, even in the dark [39]. Ghasempour et al. [40]
showed negligible photocatalytic activity (Fig. S2). studied the antibacterial activity of tungsten oxide nanorods/microrods
against E. coli under visible light irradiation and in the dark. The na-
norods (with diameters of 50–90 nm) of WO3 could inactivate ~58% of
3.3. Toxicity of WO3 to Bacterial Cells the bacteria in the dark and > 92% after 24 h of visible light irradia-
tion. The stronger antibacterial activity of WO3 samples synthesised at
3.3.1. Antibacterial Activity of Tungsten Oxide Nanoparticles without UV the lowest temperature was assigned to the higher density of the pho-
Irradiation togenerated electron-hole pairs on the surface and high bound water
Tungsten oxide nanoparticles were found to be toxic for most of the content, as well as O–H groups on the surface of the tungsten oxide
bacteria strains studied, even without UV irradiation. Thus, at a con- films. Disinfection of E. coli was also investigated [41] using an im-
centration of 25 mg/ml, WO3 nanoparticles drastically decreased E. coli mobilised thin film tungsten trioxide photocatalyst in a visible light-
viability after 5 min of contact; longer exposures (30 and 60 min) driven photoelectrocatalytic batch cell. Optimal disinfection efficiency
completely sterilised the cultural media. Diluted solutions of WO3 sol (> 99% within 15 min) was registered when the WO3 catalyst was il-
decreased bacterial viability in a concentration-dependent manner luminated under closed circuit conditions. Thus, the irradiation dras-
(Fig. 4). The same effect was registered for S. aureus and P. aeruginosa. tically enhanced the microbiocidal ability of WO3 nanoparticles.
The latter strain was the most sensitive to tungsten oxide nanoparticles:
an abrupt decline (by two to three orders of magnitude) in bacterial
viability was observed during the first 5 min of contact with nano- 3.3.2. Antibacterial Activity of WO3 Nanoparticles under UV-Irradiation
particles in the concentration range of 10.0–25.0 mg/ml, whereas, after Since WO3 nanoparticles have a significant bactericidal effect in the
30 min, there were no viable cells in the medium (Fig. 4). C. albicans concentration range of 5.0–25.0 mg/ml, a study of UV-enhanced anti-
yeasts were found to be non-sensitive to WO3 nanoparticles, and no bacterial activity was carried out at a minimal concentration of 0.5 mg/
changes in viability were found in the whole range of concentrations ml.
studied (Fig. 4). Thus, the sensitivity of the studied microorganisms to Preliminary studies on the bactericidal activity of ultraviolet irra-
the bactericidal action of WO3 nanoparticles decreased in the following diation under the experimental conditions chosen were performed; for
sequence: P. aeruginosa > S. aureus > E. coli > C. albicans. the most sensitive strain, P. aeruginosa, the decrease in viability after
Existing data on the antibacterial activity of tungsten oxide nano- 60 min of UV exposure was about 10–15% (4.6 × 108 CFU/ml before,
particles are somewhat contradictory. Some studies have shown low and 4.0 × 108 CFU/ml after irradiation). The same UV dose slightly
dark cytotoxicity of nanoparticles to microorganisms. For instance, the reduced the viability of S. aureus (by 5%), and virtually did not affect E.
toxicity of 12 metal-based nanoparticles (including WO3) to algae, coli (Table 1).

399
A.L. Popov et al. Journal of Photochemistry & Photobiology, B: Biology 178 (2018) 395–403

Table 1 least two major mechanisms that can be involved in the antimicrobial
WO3 nanoparticles' effect on the viability of microorganisms. effect of nanoparticles, namely direct action and/or indirect interaction
[43]. In different microorganisms, nanoparticles are piled up differently
Initial concentration The time of incubation – 60 min The fold of
inhibition in the vicinity of a bacterial cell. In most cases, the nanoparticles are
Without NP 0.5 mg/ml WO3 located directly on the membrane; under stress conditions, C. albicans
NP forms exopolysaccharide capsules, which prevent contact between the
particles and the cell. In this case, the direct damaging action of the
P. aeruginosa 9.0 × 105 1.8 × 106 9.5 × 105 1.89
P. aeruginosa + 60 min UV 1.2 × 106 3.6 × 105 3.33 particles on the cell membrane is absent. In turn, an indirect me-
(340 nm) chanism, involving the nanoparticle-mediated formation of ROS (Fig. 1)
The fold of inhibition 1.5 2.64 in the microenvironment, with subsequent damaging of the cell mem-
E. coli 2.0 × 105 1.6 × 105 1.2 × 105 1.33 brane, makes probably only a small contribution to the bactericidal
E. coli + 60 min UV 1.6 × 105 3.5 × 104 4.57
action of photoactive WO3 nanoparticles.
(340 nm)
The fold of inhibition 1.0 3.43 The observed differences in the number of viable microorganisms
S. aureus 5.7 × 104 6.6 × 104 6.2 × 104 1.06 may depend on their structure and physiology: for example, the sensi-
S. aureus + 60 min UV 6.3 × 104 1.7 × 102 370.59 tivity of microorganisms to the phototoxic effect of WO3 nanoparticles
(340 nm)
(gram-positive > gram-negative > eukaryotes) correlates with the
The fold of inhibition 1.04 364.71
cell wall structure of the investigated microorganisms.

3.4. Mammalian Cells

Tungsten oxide nanoparticles are typically adjudged to be of low


toxicity to mammalian cells. Recently, the toxicity of 11 metal oxide
nanoparticles (including WO3) in the concentration range of 3–100 μg/
ml to three mammalian cell types in vitro was tested [44]; tungsten
oxide nanoparticles were found to be non-toxic in concentrations below
100 μg/ml. Chitosan-capped WO3 nanoparticles demonstrated no sig-
nificant dark cytotoxicity to U266 cell line at concentrations up to
5 mg/ml after 24 h of incubation [45]. Chitosan- [45] and poly-ε-ca-
prolactone-coated [46] tungsten trioxide nanoparticles are low-toxic
and can be used as a contrast agent for X-ray computer tomography. In
mammalian systems, the genotoxic potential of WO3 has been eval-
uated using a transformation assay in BALB/c cells [47]. The interac-
tion of human genomic DNA and WO3 nanoparticles was registered and
characterised by Fourier transform infrared spectroscopy, photo-
Fig. 5. The influence of WO3 nanoparticles (0,5 mg/ml), UV irradiation (340 nm, 60 min) luminescence spectroscopy, agarose gel-electrophoresis and polymerase
and their combination on the viability of various bacterial and yeast strains. chain reaction [48].
According to the MTT test, pretreatment of primary mouse fibro-
The sensitivity of the studied microorganisms to the bactericidal blast cells with WO3 nanoparticles at concentrations of 5, 10 and
action of WO3 nanoparticles under UV irradiation was found to de- 25 mg/ml led to a significant decrease in their viability (up to 50% of
crease in the following sequence: S. aureus > E. coli > P. the control) after 24 h of incubation (Fig. 6). Lower WO3 nanoparticle
aeruginosa > C. albicans (Fig. 5). concentrations (1, 0.5 and 0.1 mg/l) did not cause a significant de-
Thus, concentrations in excess of 5.0 mg/ml of tungsten oxide na- crease in dehydrogenase activity, and cell viability was maintained at
noparticles demonstrated a significant bactericidal effect on P. aerugi- the level of control values. UV irradiation of primary mouse fibroblasts
nosa, E. coli and S. aureus strains (Fig. 4). UV irradiation of bacteria in enhanced the toxic effect of WO3 nanoparticles, depending on exposure
the presence of the minimally studied concentration of WO3 nano- time, whereas incandescent light illumination (colour temperature of
particles (0.5 mg/ml) provided an increase in their bactericidal activity 2700 K) showed no significant effect (Fig. S3). The strongest decrease in
by 3–5 times for P. aeruginosa and E. coli, whereas S. aureus was the viability (23% of the control) was observed with 20 min UV irradiation
most sensitive to the phototoxic effect of WO3 nanoparticles: the
number of viable organisms decreased > 350 times after 60 min of UV
irradiation in the presence of 0.5 mg/ml WO3 nanoparticles. Possibly,
this very specific action of tungsten oxide nanoparticles on S. aureus is
due to the presence of teichoic acids located within the cell walls of
these bacteria. These phosphate-containing bacterial copolymers can
inhibit the penetration of WO3 nanoparticles through a membrane,
decreasing the dark toxicity of nanoparticles. On the other hand, in-
teraction of WO3 with teichoic acid can result in the formation of
tungsten phosphates on the surface of nanoparticles, which increases
their catalytic activity. The irradiation of the particles leads to the re-
duction of W6 + to W5 + and W4 +; phosphate species stabilise tungsten
ions in a low valence state, preserving their stronger catalytic activity
[42], and so the phototoxicity of WO3 nanoparticles becomes higher for
this strain.
WO3 nanoparticles had virtually no toxic effect on C. albicans, in the
Fig. 6. Influence of WO3 nanoparticles under UV irradiation (200–700 nm, 10–20 min)
dark or under UV irradiation, which may have been due to its dense, on the dehydrogenase activity of primary mouse embryonic fibroblasts after 24 h of in-
chitin-containing cellular envelope. It is well known that there are at cubation (MTT assay).

400
A.L. Popov et al. Journal of Photochemistry & Photobiology, B: Biology 178 (2018) 395–403

Fig. 7. Morphology and viability analysis of primary mouse embryonic fibroblasts upon UV irradiation in the presence of different concentrations of WO3 nanoparticles (DIC-Syto 9/PI).
Scalebar: 20 μm.

and the highest WO3 nanoparticle concentration (25 mg/ml). At a nanoparticles. Additionally, it should be noted that WO3 nanoparticles
concentration of 10 and 5 mg/ml, WO3 nanoparticles, upon 10 and inhibited the growth and proliferation of primary mouse fibroblasts in a
15 min' UV irradiation, also significantly reduced the viability of the dose-dependent manner, even without exposure to UV radiation
cells, increasing the number of dead cells (Fig. 7). It is also worth noting (Fig. 8). The mechanism of the WO3 nanoparticles cytostatic effect can
that UV irradiation of the cells in the presence of WO3 nanoparticles led be explained by the direct interaction of nanoparticles with phosphate
to the detachment of cells from the substrate and their scalding, which groups in the DNA molecule, which subsequently led to the disruption
was most likely due to structural disturbances in the focal contacts of of its structure and the replication process as a whole. There was also
cells and the cytoskeleton, as a result of active ROS generation by WO3 indirect evidence of the ability of WO3 nanoparticles to inhibit enzyme

401
A.L. Popov et al. Journal of Photochemistry & Photobiology, B: Biology 178 (2018) 395–403

[5] H. Zheng, J.Z. Ou, M.S. Strano, R.B. Kaner, A. Mitchell, K. Kalantar-Zadeh,
Nanostructured tungsten oxide – properties, synthesis, and applications, Adv. Funct.
Mater. 21 (12) (2011) 2175–2196.
[6] C.G. Granqvist, Electrochromic tungsten oxide films: review of progress 1993–1998,
Sol. Energy Mater. Sol. Cells 60 (3) (2000) 201–262.
[7] C. Guo, S. Yin, T. Sato, Tungsten oxide-based nanomaterials: morphological-control,
properties and novel applications, reviews in advanced sciences and, Engineering 1
(3) (2012) 235–263.
[8] M. Weil, W.D. Schubert, The beautiful Colours of tungsten oxides, ITIA, Newsletter
(2013) 1–9.
[9] C. Di Valentin, F. Wang, G. Pacchioni, Tungsten oxide in catalysis and photo-
catalysis: hints from DFT, Top. Catal. 56 (15–17) (2013) 1404–1419.
[10] G.R. Bamwenda, H. Arakawa, The visible light induced photocatalytic activity of
tungsten trioxide powders, Appl. Catal. A Gen. 210 (1) (2001) 181–191.
[11] D.B. Hernandez-Uresti, D. Sánchez-Martínez, A. Martínez-de la Cruz, S. Sepúlveda-
Guzmán, L.M. Torres-Martínez, Characterization and photocatalytic properties of
hexagonal and monoclinic WO3 prepared via microwave-assisted hydrothermal
synthesis, Ceram. Int. 40 (3) (2014) 4767–4775.
[12] D. Sanchez-Martinez, C. Gomez-Solis, L.M. Torres-Martinez, CTAB-assisted ultra-
Fig. 8. The growth rate of primary mouse fibroblasts in the presence of WO3 nano- sonic synthesis, characterization and photocatalytic properties of WO3, Mater. Res.
particles. The cells were plated in a 96-well plate at a density of 30,000/cm2; 6 h later, Bull. 61 (2015) 165–172.
WO3 nanoparticles were added and, each day, the living cells were counted using a he- [13] Y. Lu, G. Liu, J. Zhang, Z. Feng, C. Li, Z. Li, Fabrication of a monoclinic/hexagonal
junction in WO3 and its enhanced photocatalytic degradation of rhodamine B, Chin.
mocytometer (Goryaev chamber).
J. Catal. 37 (3) (2016) 349–358.
[14] J. Sungpanich, T. Thongtem, S. Thongtem, Photocatalysis of WO3 nanoplates syn-
thesized by conventional-hydrothermal and microwave-hydrothermal methods and
activity (e.g. Taq polymerase activity) via surface adsorption, sug-
of commercial WO3 nanorods, J. Nanomater. (2014) 131.
gesting that the toxicity of WO3 nanoparticles has a complex nature. [15] S.I. Boyadjiev, T. Nagyné Kovács, I. Lukács, I.M. Szilágyi, Photocatalytic properties
The obtained experimental data not only show good prospects for of h-WO3 nanoparticles obtained by annealing and h-WO3 nanorods prepared by
hydrothermal method, AIP Conf. Proc. 1722 (2016) 1–5.
biomedical applications of tungsten trioxide, but also demonstrate the
[16] J. Kim, C.W. Lee, W. Choi, Platinized WO3 as an environmental photocatalyst that
need for clear control of biosafety when it is used in various household generates OH radicals under visible light, Environ. Sci. Technol. 44 (17) (2010)
materials and appliances. 6849–6854.
[17] K. Manthiram, A.P. Alivisatos, Tunable localized surface plasmon resonances in
tungsten oxide nanocrystals, J. Am. Chem. Soc. 134 (9) (2012) 3995–3998.
4. Conclusions [18] B.E. Smith, P.B. Roder, X. Zhou, P.J. Pauzauskie, Nanoscale materials for hy-
perthermal theranostics, Nano 7 (16) (2015) 7115–7126.
A method for tungsten oxide nanoparticles synthesis is proposed. [19] Z. Zhou, B. Kong, C. Yu, X. Shi, M. Wang, W. Liu, Y. Sun, Y. Zhang, H. Yang, S. Yang,
Tungsten oxide nanorods: an efficient nanoplatform for tumor CT imaging and
Upon UV irradiation, WO3 nanoparticles possess high photocatalytic photothermal therapy, Sci. Rep. 4 (2014) 3653.
activity in an indigo carmine dye photodecomposition reaction. [20] S.M. Sharker, S.M. Kim, J.E. Lee, K.H. Choi, G. Shin, S. Lee, K.D. Lee, J.H. Jeong,
Tungsten oxide nanoparticles showed both light and dark cytotoxic H. Lee, S.Y. Park, Functionalized biocompatible WO3 nanoparticles for triggered
and targeted in vitro and in vivo photothermal therapy, J. Control. Release 217
effects against prokaryotic microorganisms and eukaryotic cells. We (2015) 211–220.
have identified their different sensitivity to the effects of WO3 nano- [21] J. Qiu, Q. Xiao, X. Zheng, L. Zhang, H. Xing, D. Ni, Y. Liu, S. Zhang, Q. Ren, Y. Hua,
particles, which, apparently, is associated with the morphological fea- K. Zhao, W. Bu, Single W18O49 nanowires: a multifunctional nanoplatform for
computed tomography imaging and photothermal/photodynamic/radiation sy-
tures of their cell membrane structure and different metabolic activ-
nergistic cancer therapy, Nano Res. 8 (11) (2015) 3580–3590.
ities. Under UV illumination, the toxicity of tungsten oxide [22] J. Liu, J. Han, Z. Kang, R. Golamaully, N. Xu, H. Li, X. Han, In vivo near-infrared
nanoparticles has been shown to increase notably to both bacterial and photothermal therapy and computed tomography imaging of cancer cells using
novel tungsten-based theranostic probe, Nano 6 (11) (2014) 5770–5776.
mammalian cells. The toxic effect of tungsten oxide nanoparticles is
[23] L. Wen, L. Chen, S. Zheng, J. Zeng, G. Duan, Y. Wang, Z. Chai, Z. Li, M. Gao,
probably due to their high photocatalytic activity, which leads to cel- Ultrasmall biocompatible WO3-x nanodots for multi-modality imaging and com-
lular membrane damage, disruption of metabolism and cell death. This bined therapy of cancers, Adv. Mater. 28 (25) (2016) 5072–5079.
fact is to be taken into account during the development of new pho- [24] R.D. Waite, M.A. Curtis, Pseudomonas Aeruginosa PAO1 pyocin production affects
population dynamics within mixed-culture biofilms, J. Bacteriol. 191 (4) (2009)
tosensitive materials, coatings and end products with photosensitive 1349–1354.
properties. Possibly, it will be useful for auto-cleaning and disinfecting [25] S. Iwatani, T. Zendo, F. Yoneyama, J. Nakayama, K. Sonomoto, Characterization
the outer surfaces of smart windows. and structure analysis of a novel bacteriocin, lacticin Z, produced by Lactococcus
lactis QU 14, Biosci. Biotechnol. Biochem. 71 (8) (2007) 1984–1992.
[26] S. Dalai, S. Pakrashi, R.S. Kumar, N. Chandrasekaran, A. Mukherjee, A comparative
Acknowledgment cytotoxicity study of TiO2 nanoparticles under light and dark conditions at low
exposure concentrations, Toxicol. Res. 1 (2) (2012) 116–130.
[27] Y.H. Tsuang, J.S. Sun, Y.C. Huang, C.H. Lu, W.H. Chang, C.C. Wang, Studies of
The procedure for WO3 nanoparticles synthesis was elaborated with photokilling of bacteria using titanium dioxide nanoparticle, Artif. Organs 32 (2)
support from the Russian Science Foundation (project 16-13-10399). (2008) 167–174.
[28] A. Popov, N. Popova, I. Selezneva, A. Akkizov, V. Ivanov, Cerium oxide nano-
particles stimulate proliferation of primary mouse embryonic fibroblasts in vitro, J.
Appendix A. Supplementary data
Mater. Sci. Eng. C 68 (2016) 406–413.
[29] H. Watanabe, K. Fujikata, Y. Oaki, H. Imai, Band-gap expansion of tungsten oxide
Supplementary data to this article can be found online at https:// quantum dots synthesized in sub-nano porous silica, Chem. Commun. 49 (76)
(2013) 8477–8479.
doi.org/10.1016/j.jphotobiol.2017.11.021.
[30] M. Vautier, C. Guillard, J.M. Herrmann, Photocatalytic degradation of dyes in
water: case study of indigo and of indigo carmine, J. Catal. 201 (1) (2001) 46–59.
References [31] B.V. Kumar, C.P. Sajan, K.M. Rai, K. Byrappa, Photocatalytic activity of TiO2:
AlPO4-5 zeolites for the degradation of indigo caramine dye, Indian J. Chem.
Technol. 17 (3) (2010) 191–197.
[1] C.G. Granqvist, Electrochromics for smart windows: oxide-based thin films and [32] D. Sánchez-Martínez, A. Martínez-de la Cruz, E. López-Cuéllar, Synthesis of WO3
devices, Thin Solid Films 564 (2014) 1–38. nanoparticles by citric acid-assisted precipitation and evaluation of their photo-
[2] T. He, J. Yao, Photochromic materials based on tungsten oxide, J. Mater. Chem. 17 catalytic properties, Mater. Res. Bull. 48 (2) (2013) 691–697.
(43) (2007) 4547–4557. [33] A. Martínez-de la Cruz, D.S. Martínez, E.L. Cuéllar, Synthesis and characterization
[3] S.K. Deb, Opportunities and challenges in science and technology of WO3 for of WO3 nanoparticles prepared by the precipitation method: evaluation of photo-
electrochromic and related applications, Sol. Energy Mater. Sol. Cells 92 (2008) catalytic activity under vis-irradiation, Solid State Sci. 12 (1) (2010) 88–94.
245–258. [34] V. Aruoja, S. Pokhrel, M. Sihtmäe, M. Mortimer, L. Mädler, A. Kahru, Toxicity of 12
[4] C.G. Granqvist (Ed.), Handbook of Inorganic electrochromic Materials, 1995 metal-based nanoparticles to algae, bacteria and protozoa, Environ. Sci. Nano 2 (6)
9780080532905eBook. (2015) 630–644.

402
A.L. Popov et al. Journal of Photochemistry & Photobiology, B: Biology 178 (2018) 395–403

[35] M.A. Vargas-Reus, K. Memarzadeh, J. Huang, G.G. Ren, R.P. Allaker, Antimicrobial Catal. 130 (2000) 3807–3812.
activity of nanoparticulate metal oxides against peri-implantitis pathogens, Int. J. [43] N. Zholobak, V. Ivanov, A. Scherbakov, Interaction of nanoceria with micro-
Antimicrob. Agents 40 (2) (2012) 135–139. organisms. In book: Nanobiomaterials in Antimicrobial Therapy, Elsevier (12)
[36] R. Lemus, C.F. Venezia, An update to the toxicological profile for water-soluble and 419–450.
sparingly soluble tungsten substances, Crit. Rev. Toxicol. 45 (5) (2015) 388–411. [44] A. Ivask, T. Titma, M. Visnapuu, H. Vija, A. Kakinen, M. Sihtmae, S. Pokhrel,
[37] G. Hasegawa, M. Shimonaka, Y. Ishihara, Differential genotoxicity of chemical L. Madler, M. Heinlaan, V. Kisand, R. Shimmo, A. Kahru, Toxicity of 11 metal oxide
properties and particle size of rare metal and metal oxide nanoparticles, J. Appl. nanoparticles to three mammalian cell types in vitro, Curr. Top. Med. Chem. 15
Toxicol. 32 (1) (2012) 72–80. (18) (2015) 1914–1929.
[38] P. Thongkumkoon, K. Sangwijit, C. Chaiwong, S. Thongtem, P. Singjai, L.D. Yu, [45] M. Firouzi, R. Poursalehi, H. Delavari, F. Saba, M.A. Oghabian, Chitosan coated
Direct nanomaterial-DNA contact effects on DNA and mutation induction, Toxicol. tungsten trioxide nanoparticles as a contrast agent for X-ray computed tomography,
Lett. 226 (1) (2014) 90–97. Int. J. Biol. Macromol. 98 (2017) 479–485.
[39] K. Sayama, H. Hayashi, Y. Konishi, T. Gunji, H. Sugihara, Photocatalytic and an- [46] A. Jakhmola, N. Anton, H. Anton, N. Messaddeq, F. Hallouard, A. Klymchenko,
tibacterial activities over WO3 on glass filters, Chem. Lett. 39 (8) (2010) 884–885. T.F. Vandamme, Poly-ε-caprolactone tungsten oxide nanoparticles as a contrast
[40] F. Ghasempour, R. Azimirad, A. Amini, O. Akhavan, Visible light photoinactivation agent for X-ray computed tomography, Biomaterials 35 (9) (2014) 2981–2986.
of bacteria by tungsten oxide nanostructures formed on a tungsten foil, Appl. Surf. [47] G. Hasegawa, M. Shimonaka, Y. Ishihara, Differential genotoxicity of chemical
Sci. 338 (2015) 55–60. properties and particle size of rare metal and metal oxide nanoparticles, J. Appl.
[41] E. Scott-Emuakpor, G.I. Graeme, M.J. Todd, D.E. MacPhee, Disinfection of E. coli Toxicol. 32 (2012) 72–80.
contaminated water using tungsten trioxide-based photoelectrocatalysis, Environ. [48] V.B. Kumar, C.E. Sawian, D. Mohanta, S. Baruah, N.S. Islam, Physical and bio-
Eng. Manag. J. 15 (4) (2016) 899–903. physical characteristics of nanoscale tungsten oxide particles and their interaction
[42] V.V. Lesnyak, N.S. Slobodyanik, V.K. Yatsimirsky, N.A. Boldyreva, Catalytic activity with human genomic DNA, J. Nanosci. Nanotechnol. 11 (6) (2011) 4659–4666.
of tungsten phosphate (IV), (V), (VI) at carbon monoxide oxidation, Stud. Surf. Sci.

403

Вам также может понравиться