Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Environmental Pollution 227 (2017) 39e48

Contents lists available at ScienceDirect

Environmental Pollution
journal homepage: www.elsevier.com/locate/envpol

Toxicity of TiO2, in nanoparticle or bulk form to freshwater and marine


microalgae under visible light and UV-A radiation*
M. Sendra a, *, I. Moreno-Garrido a, M.P. Yeste b, J.M. Gatica b, J. Blasco a
a
Department of Ecology and Coastal Management, Institute of Marine Sciences of Andalusia (CSIC), Campus Río S. Pedro, 11510, Puerto Real, Ca diz, Spain
b diz,
Department of Material Science, Metallurgical Engineering and Inorganic Chemistry, Faculty of Sciences, University of Cadiz, E-11510, Puerto Real, Ca
Spain

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

Article history: Use of titanium dioxide nanoparticles (TiO2 NPs) has become a part of our daily life and the high
Received 13 December 2016 environmental concentrations predicted to accumulate in aquatic ecosystems are cause for concern.
Received in revised form Although TiO2 has only limited reactivity, at the nanoscale level its physico-chemical properties and
19 April 2017
toxicity are different compared with bulk material. Phytoplankton is a key trophic level in fresh and
Accepted 20 April 2017
Available online 25 April 2017
marine ecosystems, and the toxicity provoked by these nanoparticles can affect the structure and
functioning of ecosystems. Two microalgae species, one freshwater (Chlamydomonas reinhardtii) and the
other marine (Phaeodactylum tricornutum), have been selected for testing the toxicity of TiO2 in NP and
Keywords:
Toxicity
conventional bulk form and, given its photo-catalytic properties, the effect of UV-A was also checked.
Phytoplankton Growth inhibition, quantum yield reduction, increase of intracellular ROS production, membrane cell
TiO2 nanoparticles damage and production of exo-polymeric substances (EPS) were selected as variables to measure.
Freshwater TiO2 NPs and bulk TiO2 show a relationship between the size of agglomerates and time in freshwater
Seawater and saltwater, but not in ultrapure water. Under two treatments, UV-A (6 h per day) and no UV-A
exposure, NPs triggered stronger cytotoxic responses than bulk material. TiO2 NPs were also associ-
ated with greater production of reactive oxygen species and damage to membrane. However, microalgae
exposed to TiO2 NPs and bulk TiO2 under UV-A were found to be more sensitive than in the visible light
condition. The marine species (P. tricornutum) was more sensitive than the freshwater species, and
higher Ti internalization was measured. Exopolymeric substances (EPS) were released from microalgae in
the culture media, in the presence of TiO2 in both forms. This may be a possible defense mechanism by
these cells, which would enhance processes of homoagglomeration and settling, and thus reduce
bioavailability.
© 2017 Elsevier Ltd. All rights reserved.

1. Introduction science, and the production of engineered nanomaterials (ENMs)


has increased exponentially in the last decade. As a result, ENMs are
Phytoplankton is a key trophic level in aquatic ecosystems inevitably released into aquatic systems including estuaries, coastal
because these organisms are responsible for primary production waters and oceans, which are the ultimate sink for nanoparticles
(Behrenfeld et al., 2006). The presence of legacy and emergent (NPs) (Canesi et al., 2010, 2012; Ratnasekhar et al., 2015). The effect
pollutants in aquatic ecosystems can affect their structure and of those NPs on biological systems is still not well understood
functioning. Among contaminants, engineered nanomaterials (Maurer-Jones et al., 2013; Wiesner et al., 2006).
(ENMs) (Miller et al., 2012) represent a new input of contaminants Titanium dioxide in the form of nanoparticles (TiO2 NPs) has
and their impact needs to be assessed. become a part of our daily life, and these NPs are now widely used
There has been increasing application of nanotechnology to in drug delivery systems, therapeutics and biosensors, cosmetics,
molecular biology, biomaterials, semiconductors and surface production of paints, coatings, plastics, skin care products, foods,
water remediation devices and pharmaceuticals, inter alia (Vance
et al., 2015). Predicted environmental concentrations (PEC) of
*
This paper has been recommended for acceptance by B. Nowack. TiO2 NPs show the highest values for NPs in surface water, with a
* Corresponding author. mode of 3 and 0.30 ng L1 for freshwater and seawater respectively,
E-mail address: marta.sendra@icman.csic.es (M. Sendra).

http://dx.doi.org/10.1016/j.envpol.2017.04.053
0269-7491/© 2017 Elsevier Ltd. All rights reserved.
40 M. Sendra et al. / Environmental Pollution 227 (2017) 39e48

and in sediments, 1200 and 390 mg kg1, for freshwater and marine 2.1. Nanoparticle solutions
environments, respectively (Boxall et al., 2007; Gottschalk et al.,
2015). Stock of TiO2 NPs and bulk TiO2 were prepared in ultrapure
The properties, photo-stability and photo-reactivity of TiO2 NPs water immediately before the experiments. Suspensions were
have led researchers to pay considerable attention to character- sonicated with a tip sonicator (UP 200S Dr. Hielscher GmbH) for
ization of the resulting risk for ecosystems and organisms (Coll 10 min with cycle of 0.5 and frequency of 50 Hz.
et al., 2015; Li et al., 2014; Wallis et al., 2014). Due to its photo-
catalytic properties, TiO2, under UV-A radiation, can produce 2.2. Characterization of TiO2 NPs and TiO2 bulk
various increased adverse effects in an organism, including cell
damage, unregulated cell signaling, change of cell motility, Textural properties were characterized by measuring the
apoptosis, genotoxicity, pulmonary inflammation, abnormal im- adsorption/desorption of N2 at 196  C, employing a Micromeritics
mune response, DNA damage and fibrosis (Fu et al., 2014; Montiel- ASAP 2020 automatic device. Before measurements, samples were
Davalos et al., 2012; Sayes et al., 2006; Shi et al., 2013). submitted to a surface cleaning pre-treatment under high vacuum
TiO2 NPs have been classified as “harmful”, with LC50 values at 200  C during 2 h. The isotherms obtained were used to calculate
between 10 and 100 mg mL1, to organisms such as yeast, bacteria, the specific surface area (SBET) as well as the total pore volume of
algae, crustaceans, nematodes and fish (Kahru and Dubourguier, the samples studied.
2010). In microalgae, the cytotoxicity seems to be due to mem- The initial hydrodynamic size of TiO2 NPs and the zeta potential
brane damage, impairment of the effective quantum yield of PS II, of TiO2 in both forms (NPs and bulk) were studied in ultrapure
and disrupted cell cycle (Dalai et al., 2013; Navarro et al., 2008). In water, freshwater and artificial marine water through Dynamic
many studies, the role of UV-A radiation has not been taken into Light Scattering (Zetasizer Nano ZS90, Malvern Instruments, Mal-
account, so the toxicity of a photo-reactive material is under- vern, UK, and its software version 7.10) at 1 mg L1. The dispersion
estimated. Additionally, NPs can be adsorbed onto algal cell sur- of TiO2 NPs was prepared in ultrapure water, following the CEINT/
faces, and this can also reduce growth by physical shading effects NIST 1200-3 and 1200-4 Standard Protocol (Taurozzi et al.,
(Hund-Rinke and Simon, 2006); and the additional weight of the 2012a,b).
NPs can force the sedimentation of algae (Huang, 2005). Particle size distributions and images revealing the particle
The extent and type of damage depend on physicochemical shape, and the micro- and nano-structure of the samples, were
characteristics of TiO2 NPs (e.g. size, charge, crystalline forms and obtained by means of Transmission Electron Microscopy (TEM). The
coating) and environmental factors (e.g. ionic strength (IS), pH and microscope used was a JEOL2010F model, working at 200 kV. This
dissolved organic materials which govern their bioavailability and instrument has a structural resolution of 0.19 nm. The number of
reactivity) (Gonzalez et al., 2008; Liu et al., 2014; Zhao et al., 2014). images recorded allowed the measurement of 130 particles to
The bioavailability of TiO2 NPs is related to their dispersion in the determine average size and mode data.
aquatic environment and their greater tendency to agglomerate Agglomeration of TiO2 NPs and bulk TiO2 was assessed over time
(NP-NP). Recent ecotoxicological investigations have shown that (0, 0.5, 1, 3, 6, 24 and 48 h) in three different media (ultrapure water,
TiO2 NPs are toxic to marine organisms such as invertebrates, artificial freshwater and artificial marine water) (ASTM, 1975;
cyanobacteria and polychaetes (Barmo et al., 2013; Canesi et al., F
abregas et al., 2000). The initial concentration in the samples
2010; Cherchi and Gu, 2010; Galloway et al., 2010; Zhu et al., was 250 mg L1 for freshwater and marine water, but for ultrapure
2011). Only a few of these studies previously cited are focused on water concentrations of 250 mg L1 and 500 mg L1 were evalu-
microalgae; the great majority deal mainly with freshwater species ated. Changes in agglomeration states were measured by a Malvern
(Miller et al., 2010, 2012). Specific mechanisms need to be taken Master Sizer 2000, with software version 5.61.
into account when assessing toxicity. Phytoplankton cells can Crystallization of TiO2 NPs and bulk TiO2 were checked by XRD
produce exopolymeric substances (EPS) as a response when to confirm the information given by suppliers (see Fig. SI5).
exposed to potentially dangerous substances. EPS can enhance the
aggregation and flocculation of NPs, thus altering their bioavail- 2.3. Test organisms
ability and reactivity (Dalai et al., 2013; Soldo et al., 2005). To date
those mechanisms have not been adequately studied. Two microalgae species were selected, one found in freshwater
In this work, the toxicity of TiO2 NPs and bulk TiO2 has been environments (Chlamydomonas reinhardtii P.A., (Dangeard, 1888),
assessed for two phytoplanktonic species, one from the marine CHLOROPHYCEAE) and the other in marine environments (Phaeo-
environment (Paeodactylum tricornutum, and the other found in dactylum tricornutum, (Bohlin, 1897), BACILLARIOPHYCEAE). Both
freshwater systems (Chlamydomonas reinhardtii). The joint effect of were obtained from the ICMAN Marine Microalgae Culture
exposure to TiO2 and UV-A radiation during 6 h per day, simulating Collection (IMMCC). Cells were grown in filtered (0.2 mm) fresh-
the natural condition, has been taken in account in order to water culture medium and F/2 marine medium without EDTA (see
improve knowledge of the mechanisms involved in the toxicity of supplementary information), respectively, for two weeks prior to
TiO2 NPs and bulk TiO2 in environmental conditions. the experiment (Fabregas et al., 2000; Guillard and Ryther, 1962).
Artificial marine water used was Substitute Ocean Water D1141-75
from the ASTM (ASTM, 1975).
2. Material and methods
2.4. Toxicity bioassays
Reagents: TiO2 NPs (CAS number, 13463-67-7, as nano-powder,
79% anatase, 21% rutile, particle size <100 nm) and bulk TiO2 (CAS Bioassays were carried out using either TiO2 NPs or bulk TiO2.
number, 1317-80-2, as pure rutile, particle size >100 nm) were Two light treatments were also applied: continuous visible light
obtained from Sigma Aldrich, together with DCFH-DA (20 -7’- (1.6 mWcm2) and the same light regime plus 6 h of UV-A
dichlorofluorescein diacetate), and propidium iodide. All glassware (0.20 mW cm2, Multiple Ray Lamp) per day, simulating two nat-
was washed with diluted nitric acid (10%) and rinsed several times ural environments. The intensity of UV-A was measured with a
with de-ionized water (Milli-Q) before use. All reagents for media digital UVX radiometer (UVP, Analytik Jena AG).
were of analytical grade. Growth inhibition bioassays were carried out following OECD
M. Sendra et al. / Environmental Pollution 227 (2017) 39e48 41

Guidelines (OECD, 1994), in order to determine the effective values shaken for 30 s in order to remove extracellular TiO2 (Zhou et al.,
for 50% inhibition after 72 h (EC50) of microalgae cellular concen- 2012). After this, the process was repeated three times. The
tration (initial cell density, 104 cells mL1). Cells were counted required number of repetitions was determined by a previous study
under microscope using a Neubauer counting chamber. The growth using flow cytometry, in order to detect the presence of TiO2 ag-
inhibition values were fitted to the Hampel model (Hampel et al., gregates (by both forward scattering and side scattering) and no
2001), with different concentrations: 0.1, 1, 10, 100, 200 and concentration of Ti with MSFIA-LWCC was measured. The final
400 mg L1 under continuous white light; and 0.1, 1, 2, 5 and pellet was stored at 80  C until the intracellular metal content was
10 mgL1 under the visible light plus UV-A treatment. quantified by MSFIA-LWCC.
The effective quantum yield of photosynthetic energy conver-
sion in PSII in darkness was measured by fluorometry using a 2.6. Quantification of exo-polymeric substances (EPS)
Phyto-PAM (Heinz Walz GmbH) equipped with an ED-101 US/MP
Optical Unit. This parameter shows the efficiency of the photo- Concentrations of EPS were determined after 72 h in samples
chemical energy conversion process (Schreiber et al., 1995). Values from the toxicity bioassays, according to Dubois et al., 1956 (Dubois
of EC50 of effective quantum yield were also fitted to the same et al., 1956). Algal cultures (30 mL) were centrifuged at 10,000 g for
model (Hampel et al., 2001). Effective quantum yields were 10 min, and then the supernatant and pellet were separated. The
measured at the same concentration as those for EC50 growth supernatant fraction was chosen to measure dissolved EPS. Equal
inhibition. volumes of pre-chilled ethanol were added to the samples, and the
Production of intracellular ROS (reactive oxygen species: su- resulting mixtures were kept at 4  C overnight, in order to pre-
peroxide, hydroxyl and hydrogen peroxide) was quantified using cipitate dissolved EPS. After a second centrifugation at 10,000 g for
the DCFH-DA (20 -7’-dichlorofluorescein diacetate) method (He 30 min, supernatants were discarded and pellets were washed
et al., 2002; Stachowski-Haberkorn et al., 2013). ROS were twice at 15,000 g for 15 min with ultrapure water to obtain pure
measured at 3, 8, 24, 48 and 72 h at 2 and 5 mg L1 for both TiO2 NPs dissolved EPS. The quantification of EPS was carried out using the
and bulk TiO2 under visible light, and at 2 mg L1 under UV-A. For phenol-sulfuric acid method (Dubois et al., 1956).
this bioassay a negative control (without treatment) and a positive
control with 0.1 mM of H2O2 (to see the signal correctly in the flow 2.7. Statistical analysis
cytometer) were employed.
After 30 min in darkness, at room temperature conditions, ROS All bioassays were performed in triplicate, controls were
were measured by green fluorescence (FL1, 533/30 nm) of stained developed for both conditions (UV-A and visible regime) in order to
cells using a FACSCalibur Flow Cytometer (Becton-Dickinson®). assess the effect of UV-A by itself and effect of particles tested. Data
Finally, we calculated the percentage of oxidative stress as the are shown as average ± standard deviation between replicates.
number of cells that fluoresce green (as stained with DCFH-DA) in Statistical analyses were carried out using the IBM SPSS, Statistics
relation to the total number of cells counted by the flow cytometer 23 program. To assess the responses, i.e. growth, effective quantum
(equation (1)). yield, % of ROS, % membrane damage and EPS two General Linear
Models (GLM; SPSS, 2005) were constructed (McCullagh and
Cells Dyed with DCFH  DA Nelder, 1989; SPSS, 2005). The first GLM was a general one in
ROS ¼ *100 (1)
Total Cells which all the factors are in competition, and the second model is
Membrane integrity was also quantified, following the Propi- segmented by medium (fresh and marine) and exposure condition
dium Iodide (PI) method (Xiao et al., 2011). This parameter was (UVA þ visible light and visible light only). The fixed factors were
measured at 3, 8, 24, 48 and 72 h for 2 and 5 mg L1 under visible medium (freshwater and marine water), exposure condition
light, and at 2 mg L1 under UV-A treatment. Samples were incu- (UVA þ visible light and visible light only, treatment (TiO2 in NP
bated for 20 min at PI dosage (10 mg mL1 for 2$106 cells). Fluo- and bulk form), concentration and time.
rescence of PI inside the cells was detected with a FL2 detector by Data were checked by homogeneity of variance (Levene test), a
flow cytometry (the same equipment as that used in the ROS one-way Anova test and a Bonferroni post hoc test at p < 0.05. If the
analysis). The percentage of membrane damage was calculated data did not pass the Levene test for homogeneity, a post hoc
following equation (2). Methanol was used as positive control. Tamhane T2 test was applied.

Cells Dyed with PI 3. Results


Membrane damage ¼ *100 (2)
Total Cells
3.1. Characterization of TiO2 NPs and bulk TiO2

Physical characteristics of the dry powdered TiO2 NPs and bulk


2.5. Quantification of intracellular and extracellular TiO2 NPs TiO2 together with their properties in different media are shown in
Fig. SI1 (TEM study) and the supplementary information
Concentration of Ti in four fractions (total, supernatant, extra- (Table SI1). Porosity and the smaller size of particles of the TiO2 NPs
cellular and intracellular) was quantified by a multi-syringe flow sample in relation to TiO2 bulk (38 ± 12 nm for TiO2 NPs and
injection analysis system using a liquid waveguide capillary cell 423 ± 154 nm for bulk; and weighted by surface, 46 nm for TiO2 NPs
(MSFIA-LWCC) (Sa nchez-Quiles et al., 2013), after prior digestion and 534 nm for bulk) are consistent with their greater specific
with potassium peroxodisulfate. Titanium recovery was 87.7 ± 1.5%. surface area BET in comparison to bulk TiO2 (18 times higher in the
Accumulation bioassays were performed in flasks with a volume NPs sample).
of 10 mL, containing 10 mL of treated algal cultures at an initial cell The size of the agglomerates of TiO2 NPs measured by DLS was
density of 106 cells$mL1 and concentration of 0.1 mg L1 of TiO2 largest in ultrapure water and smallest in saltwater; zeta potential
NPs. Samples for Ti measurements were collected at 6, 24 and 48 h. was less electronegative in saltwater than in ultrapure water.
For control and treated algal cells the whole volume was (shown in Fig. SI2); the polydispersity index (PDI, indicative of the
centrifuged at 4000 g for 20 min. Supernatants were stored and heterogeneity of hydrodynamic radii of particles)) is greater than in
pellets re-suspended in 5 mL of EDTA (0.02M) dissolved in PBS, and ultrapure water, outside of acceptable limits (PDI<0.7 when the
42 M. Sendra et al. / Environmental Pollution 227 (2017) 39e48

samples were measured after 2 h). Due to this high PDI the samples Table 1
were assessed over time with the Mastersizer 2000 to study the EC50 for cellular growth of C. reinhardtii and P. tricornutum under UV-A regime and
non-UV-A regime (visible light only) exposed to TiO2 NPs and bulk TiO2.
agglomeration. Two different concentrations in ultrapure water
(250 and 500 mg L1) were studied. Neither TiO2 NPs nor bulk TiO2 EC50 of growth inhibition (mg$L1)
showed larger agglomerate sizes at higher concentrations (Fig. SI3). UV-A þ Visible light Visible light only
With respect to differences in the agglomeration of TiO2 NPs and
TiO2 NPs Bulk TiO2 TiO2 NPs Bulk TiO2
bulk TiO2 between different media, agglomerate sizes were quite
C. reinhardtii 2.30 ± 1.76 1.35 ± 3.06 551.7 ± 163.79 423.70 ± 18.74
similar between freshwater and saltwater media but there were
P. tricornutum 1.98 ± 0.09 6.58 ± 0.67 132.0 ± 7.0 185.0 ± 26.0
statistically significant differences in size between ultrapure water
and the other two media (freshwater and marine media) (p < 0.05)
(Table SI2 in supplementary information). Both TiO2NPs and bulk
TiO2 showed a relationship between size of agglomerates and time P. tricornutum was observed to be more sensitive than C. reinhardtii
in freshwater and saltwater but not in ultrapure water (Fig. SI4). In under both exposure conditions and with both forms of TiO2 ma-
artificial marine water, the agglomeration occurs in the first 30 min. terial (p < 0.05).
However, in freshwater NPs were stable for at least the first hour.
3.4. Effective quantum yield of PII
3.2. Cytotoxicity of TiO2 NPs and bulk TiO2
With respect to EC50 values calculated for the effective quantum
Data of the controls from the UV-A and visible light regimes yield of photosynthetic energy conversion in PSII (Fig. 1), it was not
were analyzed using the GLM for each of the cytotoxic responses possible to calculate this parameter under UV-A conditions,
studied with the object of determining the effect of these condi- (0.1e10 mg L1) since the highest concentrations used did not reach
tions (UV-A and visible light regimes) on the microalgae popula- 50% inhibition and no clear dose-response trend was found for
tion. These conditions were significant for the particular cytotoxic either microalgae. For C. reinhardtii, under the non-UV-A regime,
responses, growth and percentage of ROS (p < 0.05); however they 50% inhibition was not reached, and predicted EC50 values were
were not significant for the effective quantum yield of PII and the 462 ± 147 and 487 ± 105 mg L1 for TiO2 NPs and TiO2 bulk
percentage of cell membrane damage (p > 0.05). However, when respectively. In the case of P. tricornutum those values were
these data were compared for each condition (UV-A and visible 145 ± 27 and 167 ± 26 mg L1 for TiO2 NPs and bulk TiO2,
light regime) and culture media (freshwater and marine micro- respectively, showing a clearer dose-response dependent
algae), other factors such as treatment (NPs and bulk)*concentra- cytotoxicity.
tion were significant (p < 0.05) for all cytotoxic responses. Each Taking together the results for both freshwater and marine
cytotoxic responses are explained below: microalgae in function of time, at 6 h from the start of the experi-
ments, inhibition of quantum yield was higher than at 24, 48 or
3.3. Microalgae growth population 72 h. Neither of the species (marine and freshwater) showed sig-
nificant differences between treatments (NPs and TiO2 bulk) under
TiO2 NPs and bulk TiO2 showed a dose-exposure dependent UV-A and non-UV-A regimes (p > 0.05).
cytotoxicity to C. reinhardtii and P. tricornutum under the UV-A light
regime and visible light (see Table SI4). Nevertheless, when they 3.5. Reactive oxygen species (ROS) and membrane damage
were exposed to visible light at low concentrations of both sub- endpoints
stances, microalgae populations showed a hormetic response, as
cultures grew up to 5% more than the controls. The GLM values Fig. 2 shows the percentage of cells with ROS signals in relation
(Table SI3) show, in most of the responses, how the exposure to the total number of cells for both microalgae. Results for TiO2 NPs
condition (UV-A and visible light only regimes), treatment, con- and bulk TiO2 are compared under both regimes, UV-A radiation
centration, and time resulted in significant differences (p < 0.05) in and visible light only. Under non-UV-A conditions, C. reinhardtii did
the toxicity variables. However, in the GLM by species (freshwater not show great differences with respect to ROS production. In the
and marine microalgae) and exposure condition (UV-A and visible case of P. tricornutum, under the same exposure conditions, treat-
light) the fit to the model is improved (Table SI4). ments showed clear differences (p < 0.05) with respect to the
When cells were subjected to the UV-A regime, toxicity was controls. There is also a dose-response trend in the case of the bulk
higher than in experiments with visible light only; and thus EC50 TiO2, which is not so clear for the TiO2 NPs treatments. In any case,
values were lower for both materials (TiO2NPs and bulk TiO2, with when treatments (TiO2 NPs and bulk TiO2) are compared, NPs al-
significant differences p < 0.05). ways provoke greater intracellular production of ROS.
In the UV-A regime, for C. reinhardtii, EC50 values for growth Under the UV-A regime, concentrations of 2 mg L1 for each
inhibition were 2.6 ± 1.7 mg L1 and 1.3 ± 3.0 mg L1 for TiO2 NPs treatment (TiO2 NPs and bulk TiO2) were compared with controls
and bulk TiO2 respectively (p > 0.05). C. reinhardtii results indicated for both microalgae species. In this case, for both species, the bulk
that the material (NPs or bulk) with different size did not show TiO2 treatment showed no difference from the controls, but TiO2
differences in the EC50; therefore in this case we can conclude that NPs clearly provoke greater production of intracellular ROS. This
the photocatalytic property of the materials is key in the TiO2 difference is quite clear in the case of P. tricornutum at 72 h
toxicity under UV-A regardless of their size. When cells were (3.3± 1.6% of affected cells for controls, and 59 ± 6 for TiO2 NPs).
incubated without UV-A exposure, inhibition did not reach 50% With respect to membrane cell damage (Fig. 2), results under
even at the highest concentrations used and the calculated (ex- the non-UV-A regime for both microalgae showed significant dif-
pected) EC50 values for TiO2 NPs and bulk TiO2 were 552 ± 164 and ferences between controls and TiO2 NPs, and between both treat-
424 ± 19 mgL1 respectively (p > 0.05). For P. tricornutum, EC50 ments (TiO2 NPs and bulk) (p < 0.05). Differences in function of
values were 2 ± 0.09 and 6.6 ± 0.7 mg L1 for TiO2 NPs and bulk interaction between exposure time with concentration were only
TiO2 respectively under UV-A (p < 0.05), and 132 ± 7 and found for P. tricornutum (Data analyzed by short cases GLM,
185 ± 26 mgL1 for TiO2 NPs and bulk TiO2 respectively under Table SI4). As can be noted in Fig. 2, membrane cell damage, for the
visible light only (non-UV-A regime) (p < 0.05) (Table 1). non-UV-A regime, is clearly concentration-dependent. Under UV-A,
M. Sendra et al. / Environmental Pollution 227 (2017) 39e48 43

Fig. 1. EC50 for effective quantum yield for C. reinhardtii and P. tricornutum under UV-A regime and non-UV-A regime (visible light only) exposed to TiO2 NPs and bulk TiO2.

both C. reinhardtii and P. tricornutum showed differences in mem- experiment, the concentrations of TiO2 NPs in the supernatant were
brane cell damage between controls and TiO2 NPs, and between 86± 0.4% and 69± 3% of the total fraction for C. reinhardtii and
TiO2 NPs and bulk TiO2 (p < 0.05) at 72 h; C. reinhardtii showed P. tricornutum, respectively (the initial concentration of TiO2 NPs
18± 1% of damaged cells and P. tricornutum showed 34± 2%, when was 0.1 mg L1). The first wash - of the three washes performed in
exposed to 2 mg$L-1 of TiO2 NPs (control values for the two species each sample - removed 5± 2% and 17 ± 1% of the TiO2 NPs for
under the same conditions were 7± 2% and 1.4± 0.2%, respectively). C. reinhardtii and P. tricornutum, respectively. The second wash
removed 6.5± 0.4% of TiO2 NPs for P. tricornutum, but no further Ti
3.6. Quantification of intracellular and extracellular TiO2 NPs was removed for C. reinhardtii. After three consecutive washes no
further Ti was removed from cellular surfaces in either case.
Plotted in Fig. 3 is the concentration in the different fractions at However, after 24 h incubation, different results were obtained.
different time intervals (6, 24 and 48 h). At 6 h from the start of the The concentrations of Ti from TiO2 NPs measured in supernatants
44 M. Sendra et al. / Environmental Pollution 227 (2017) 39e48

Fig. 2. Percentage of cells containing measurable ROS and cell membrane-damaged (in relation to total number of cells) for C. reinhardtii and P. tricornutum under UV-A regime
(panel A) and non-UV-A regime (visible light only, panel B) exposed to TiO2 NPs and bulk TiO2.

before washings were lower than values after 6 h, with 50± 3% and P. tricornutum. Intracellular Ti (measured in pellet) was also lower
30± 0.7% of the total fraction of TiO2 NPs being removed for at 0.07± 0.5% for C. reinhardtii and 6± 1% for P. tricornutum
C. reindhardtii and P. tricornutum, respectively. Nevertheless, the (Table SI5).
first wash removed more Ti when compared with the samples after
6 h (35± 0.8% for C. reinhardtii and 30± 1% for P. tricornutum).
After 48 h, the concentration of Ti in the supernatant was again 3.7. Quantification of exo-polymeric substances
higher, at 60± 3% for C. reinhardtii and 47± 2% for P. tricornutum. But
the percentage of the total fraction of Ti removed by the first wash Concentrations of exo-polymers (EPS) in supernatants were
was lower, at 18± 0.1% for C. reindhartii and 18± 1% for measured after 72 h of incubation, for both TiO2 NPs and bulk TiO2
(Fig. 4). Greater production of EPS was observed at high
M. Sendra et al. / Environmental Pollution 227 (2017) 39e48 45

Fig. 3. Concentration of TiO2 NPs in different compartments (supernatants of consecutive centrifugations and in pellets) after 6, 24 and 48 h for C. reinhardtii and P. tricornutum, for
an initial NP concentration of 0.1 mgL1.

Fig. 4. Exo-polymeric substances released by C. reinhardtii and P. tricornutum, exposed to TiO2 NPs and bulk TiO2 at different concentrations (0.1, 1, 10, 100, 200 and 400 mgL1), after
72 h exposure.

concentrations of both TiO2 NPs and bulk TiO2. Production of EPS exposure to bulk TiO2, EPS production for the same species at the
was greater in cells treated with TiO2 NPs than with bulk TiO2 same concentration was 4.3 ± 2.7$105 mg of EPS$ 104 cells (con-
(p < 0.05). For instance, C. reinhardtii, exposed to TiO2 NPs at trols produced 2 ± 0.1$105 mg$104 cells after the same exposure
10 mg L1, produced 23 ± 1.2$105 mg of EPS$104 cells; while for time). At the same concentration and after 72 h of incubation,
46 M. Sendra et al. / Environmental Pollution 227 (2017) 39e48

P. tricornutum produced 62 ± 4.3, 20.1 ± 0.9 and 7.8 ± 0.9$105 mg weighted by surface, than the bulk form, so the surface available for
of EPS$104 cells for exposure to TiO2 NPs, to bulk TiO2 and in con- interaction between cells and NPs is greater than that between cells
trols, respectively. and the bulk material. Under a regime without UV-A, the marine
microalgae showed greater sensitivity than the freshwater species.
4. Discussion This may be due to the effect caused by salt ions, which would act to
compress the electro-double layer of NPs and cells, and thus
4.1. Behavior of particles in different culture media impede the electrostatic repulsion force between them, facilitating
the heteroagglomeration of NPs and cells (Ma et al., 2015). Ac-
The aggregation or agglomeration behavior in solution of TiO2 cording to Ma et al. (2015), homoagglomeration of anatase NPs
NPs and bulk TiO2 can affect the availability and the form of assessed by individual settling experiments were practically the
interaction with microalgae, and can provoke toxic effects. The same when the IS value was in the range between 50 and 500 mM;
most important processes in relation to TiO2 NPs are homoag- however, between 5 and 50 mM the differences were significant. In
glomeration by NPs with other NPs, and heteroagglomeration by our media the IS value was 30 and 700 mM for freshwater and
NPs with cells (Lin et al., 2014). These processes are controlled by saltwater media, respectively. This could be the reason for the
the physicochemical properties of NPs, (size and surface charge), similar agglomerate size measured in both media, although the size
and by solution conditions (ionic strength, pH, and concentration of was significantly different in comparison with that in ultrapure
natural organic matter) (Jiang et al., 2009). water (IS ¼ 0 mM). Heteroagglomeration of anatase NPs and cells
studied in co-settling experiments with different IS values, showed
4.2. Importance of radiation in toxicological test with that heteroagglomeration was greater when the IS value of the
photocatalytic substances media was 500 mM than it was when the IS value was 50 mM (Ma
et al., 2015). This could explain the greater sensitivity of the marine
Controls for each of the cytotoxic responses for each condition microalgae species compared with the freshwater species selected.
(UV-A and visible light regime) are essential for studying the effect Nevertheless, concentrations which imply a reduction of 50% of the
of UV radiation on the microalgae population. In the general GLM parameters measured are, in general, far removed from those found
model, which takes into account all factors, significant differences in realistic scenarios, according to results reported by other authors
for microalgae population growth and percentage of ROS were (Velzeboer et al., 2008; Warheit et al., 2007).
observed between the UV-A and visible light regime controls.
However, in the single GLM devised exclusively for conditions and 4.3. Importance of microalgae species in toxicological test with TiO2
culture media, statistically significant differences in all cytotoxic NPs
responses were reported when the interaction between treatment
and concentration was considered. Results obtained for a single species are often of limited
In general, UV-A exposure increases the toxicity of TiO2 to comparability with another species from a different taxonomic
microalgae. The toxicity of TiO2 under UV-A radiation is due to the group. For our study, a Chlorophyceae found in freshwater media
anatase form of Ti, which is more photo-reactive than the rutile and a Bacillariophyceae found in marine environments were
form (Sayes et al., 2006; Uchino et al., 2002). The use of commercial selected as representative species from the most important taxo-
formulations including anatase is increasing, as photocatalytic nomic groups of microalgae in their respective biogenesis. More
properties of this form are often required in new applications. comparative studies of the sensitivity of similar species in both
According to our results, at least one of the mechanisms media need to be undertaken. C. reindhartii cells present a glyco-
responsible for the toxicity of TiO2 (in both forms, as NPs or as bulk protein cell wall, while P. tricornutum cells are covered by a silica
material) could be the generation of reactive oxygen species (ROS) frustule. Contact between the surrounding media and the cyto-
inside the cells. This production is increased by exposure to UV-A plasm across the space between the valves could be easier than
radiation. In this respect, when active photocatalytic compounds contact between the media and cytoplasm of Chlorophyceae across
(such as TiO2) are involved in toxicity tests, classical approaches the cellulose cell wall. Piccapietra et al. (2012), studying the accu-
could be missing part of the toxicity for not considering ROS species mulation of silver from Ag NPs, demonstrated that metal accumu-
generates by tested substances in the culture media. In this study lation is greater in mutants of C. reindhartii lacking a cell wall when
the results of EC50 of growth inhibition under the UV-A regime compared with wild-type cells (Piccapietra et al., 2012).
applied were two orders of magnitude smaller than the result with It has been demonstrated previously that TiO2 NPs are more
no UV-A regime, calculated by other authors (Dalai et al., 2013; Lee toxic to microalgae than TiO2 in bulk form (Aruoja et al., 2009).
and An, 2013). Recent research demonstrated different drifts in the Differences in EC50 found in the literature can be attributed to
phytoplankton population when the microalgae community was differences in the stability and reactivity of NPs and differences in
exposed to photocatalytic compounds under different scenarios the experimental matrix, exposure media and concentration
(UV radiation and solar radiation without UV radiation) (Sendra (Handy et al., 2012). The present study has demonstrated, for both
et al., 2017). NPs and bulk material, a dose-dependent but non time-dependent
Heteroagglomeration (NPs-Cells) related to ionic strength in the reduction of effective quantum yield, and increases in ROS pro-
culture media. duction, membrane damage and bioaccumulation. Additionally,
In bioassays carried out in a regime of light without UV-A, other increased production of EPS by cells has been demonstrated in the
cytotoxicity mechanisms must be relevant. Interaction between the presence of TiO2 NPs and bulk, pointing to possible mechanisms of
cell surface and the TiO2 NP may be the first and most important detoxification (Nielsen et al., 2008). EPS excreted by cells may
interaction between NPs and cells. Hydrogen bonding and/or other inhibit the attachment of NPs to cells through steric repulsion or
polar interactions could also contribute to the interaction between may facilitate NP-cell agglomeration by bridging between NPs and
a NP and a cell (EU, 2011). In other studies, anatase in crystalline cells through adsorption on the surface of NPs (Khan et al., 2011).
form showed much stronger interactions with cells than the rutile EPS could also trigger flocculation of NPs, and thus fewer NPs would
form of TiO2, which confirms the greater toxicity of this chemical be bioavailable to microalgae, reducing their accumulation or
form (Braydich-Stolle et al., 2009; Ji et al., 2011). Furthermore, TiO2 changing their distribution inside the cells (Dong et al., 2007; Miao
NPs present a greater surface area and smaller size of particles, et al., 2009; Soldo et al., 2005). This could be the reason for the
M. Sendra et al. / Environmental Pollution 227 (2017) 39e48 47

increased Ti concentration in the supernatant and wash fractions and immune function of the marine bivalve Mytilus galloprovincialis. Aquat.
Toxicol. 132, 9e18.
after 48 h compared with that after 24 h: there may be greater
Behrenfeld, M.J., O/'Malley, R.T., Siegel, D.A., McClain, C.R., Sarmiento, J.L.,
adsorption between the TiO2 NPs and the released exopolymer Feldman, G.C., Milligan, A.J., Falkowski, P.G., Letelier, R.M., Boss, E.S., 2006.
particles, and this process could reduce the reactivity of the NPs and Climate-driven trends in contemporary ocean productivity. Nature 444,
the resulting oxidative stress, leading to decreased levels of ROS 752e755.
Bohlin, K.H., 1897. Zur morphologie und biologie einzelliger algen.
(Miao et al., 2009). On the other hand, EPS could adsorb NPs onto Boxall, A., Chaudhry, Q., Sinclair, C., Jones, A., Aitken, R., Jefferson, B., Watts, C., 2007.
the surface of cells. Phaeodactylum tricornutum produced more EPS Current and Future Predicted Environmental Exposure to Engineered Nano-
than Chlamydomonas reinhardtii (even in the controls), and thus particles. Central Science Laboratory, York, UK.
Braydich-Stolle, L.K., Schaeublin, N.M., Murdock, R.C., Jiang, J., Biswas, P.,
another possibility to consider is that the interaction between NPs Schlager, J.J., Hussain, S.M., 2009. Crystal structure mediates mode of cell death
and cells producing agglomeration may be facilitated by the pres- in TiO2 nanotoxicity. J. Nanoparticle Res. 11, 1361e1374.
ence of EPS surrounding the cell. Thus, the role of these EPS in the Canesi, L., Ciacci, C., Fabbri, R., Marcomini, A., Pojana, G., Gallo, G., 2012. Bivalve
molluscs as a unique target group for nanoparticle toxicity. Mar. Environ. Res.
toxicity of NPs to microalgae is still not well elucidated. The 76, 16e21.
reduced internalization of NPs by cells over time may also be Canesi, L., Fabbri, R., Gallo, G., Vallotto, D., Marcomini, A., Pojana, G., 2010. Bio-
explained by detoxification mechanisms; therefore these micro- markers in Mytilus galloprovincialis exposed to suspensions of selected nano-
particles (Nano carbon black, C60 fullerene, Nano-TiO2, Nano-SiO2). Aquat.
algae showed tolerance at certain doses of this emergent pollutant Toxicol. 100, 168e177.
(Maeda and Sakaguchi, 1990). Coll, C., Notter, D., Gottschalk, F., Sun, T., Som, C., Nowack, B., 2015. Probabilistic
environmental risk assessment of five nanomaterials (nano-TiO2, nano-Ag,
nano-ZnO, CNT, and fullerenes). Nanotoxicology 1e9.
5. Conclusions Cherchi, C., Gu, A.Z., 2010. Impact of titanium dioxide nanomaterials on nitrogen
fixation rate and intracellular nitrogen storage in Anabaena variabilis. Environ.
For microalgae the toxicity mechanisms for TiO2, in NP and bulk Sci. Technol. 44, 8302e8307.
Dalai, S., Pakrashi, S., Joyce Nirmala, M., Chaudhri, A., Chandrasekaran, N.,
form, are related to the interaction between nanoparticles and cell Mandal, A.B., Mukherjee, A., 2013. Cytotoxicity of TiO2 nanoparticles and their
surface. The marine microalgae, P. tricornutum, is more sensitive to detoxification in a freshwater system. Aquat. Toxicol. 138e139, 1e11.
TiO2 than the freshwater C. reinhdartii. Effective quantum yield was Dangeard, P.-A., 1888. Recherches sur les algues infe rieures.
Dong, J., Mao, W., Zhang, G., Wu, F., Cai, Y., 2007. Root excretion and plant tolerance
the most sensitive toxicity parameter, with the most rapid time of to cadmium toxicity-a review. Plant Soil Environ. 53, 193.
response. The characteristics of the exposure media are relevant for Dubois, M., Gilles, K.A., Hamilton, J.K., Rebers, P., Smith, F., 1956. Colorimetric
the TiO2 NPs assay, especially ionic strength, which affects both method for determination of sugars and related substances. Anal. Chem. 28,
350e356.
homo- and heteroagglomeration: a higher IS can impede the
EU, 2011. (European community). European commission recommendation of 18
internalization processes and resulting toxicity. EPS are involved October 2011 on the definition of nanomaterial. In: Publ. Office Eur. Union.
both in the protection mechanism and in heteroagglomeration Offical Journal of the European Union (20.10.2011):L 275/38.
Fabregas, J., Domínguez, A., Regueiro, M., Maseda, A., Otero, A., 2000. Optimization
processes; however, further studies are necessary to clarify their
of culture medium for the continuous cultivation of the microalga Haemato-
role. coccus pluvialis. Appl. Microbiol. Biotechnol. 53, 530e535.
Exposure to UV-A radiation is not usually taken into account in Fu, P.P., Xia, Q., Hwang, H.-M., Ray, P.C., Yu, H., 2014. Mechanisms of nanotoxicity:
classical (i.e. standard) bioassays, but the effect of this variable generation of reactive oxygen species. J. Food Drug Analysis 22, 64e75.
Galloway, T., Lewis, C., Dolciotti, I., Johnston, B.D., Moger, J., Regoli, F., 2010. Suble-
needs to be measured when photocatalytic substances, such as thal toxicity of nano-titanium dioxide and carbon nanotubes in a sediment
TiO2, are tested, because the sensitivity of organisms can be two dwelling marine polychaete. Environ. Pollut. 158, 1748e1755.
orders of magnitude greater. However, the primary mechanism of Gonzalez, L., Lison, D., Kirsch-Volders, M., 2008. Genotoxicity of engineered nano-
materials: a critical review. Nanotoxicology 2, 252e273.
toxicity revealed in bioassays performed under visible light seems Gottschalk, F., Lassen, C., Kjoelholt, J., Christensen, F., Nowack, B., 2015. Modeling
to be the interaction between the NPs and the cell surface, resulting flows and concentrations of nine engineered nanomaterials in the Danish
in cell membrane damage. environment. Int. J. Environ. Res. Public Health 12, 5581e5602.
Guillard, R.R.L., Ryther, J.H., 1962. Studies of marine planktonic diatoms: I. Cyclotella
nana Hustedt, and Detonula confervacea (cleve) Gran. Can. J. Microbiol. 8,
Acknowledgements 229e239.
Hampel, M., Moreno-Garrido, I., Sobrino, C., Lubia n, L.M., Blasco, J., 2001. Acute
toxicity of LAS homologues in marine microalgae: esterase activity and inhi-
This research has been funded by the Junta de Andalucía bition growth as endpoints of toxicity. Ecotoxicol. Environ. Saf. 48, 287e292.
(PE2011-RNM-7812 project and FQM-110 group) and the Spanish Handy, R.D., van den Brink, N., Chappell, M., Mühling, M., Behra, R., Dusinska , M.,
National Research Plan (CTM2012-38720-C03-03) andFEDER Simpson, P., Ahtiainen, J., Jha, A.N., Seiter, J., Bednar, A., Kennedy, A.,
Fernandes, T.F., Riediker, M., 2012. Practical considerations for conducting
fundings (MAT2013-40823-R). I would like to thank Dra. Catalina ecotoxicity test methods with manufactured nanomaterials: what have we
Fernandez and Eugenia Zuasti for the support in DLS analysis. We learnt so far? Ecotoxicol. Lond. Engl. 21, 933e972.
also thank the SC-ICYT of Cadiz University (UCA) for use of its He, Y.Y., Klisch, M., Ha €der, D.P., 2002. Adaptation of cyanobacteria to UV-B stress
correlated with oxidative stress and oxidative damage. Photochem. Photobiol.
Electron Microscopy division facilities. 76, 188e196.
This research did not receive any specific grant from funding Huang, C.-P., 2005. Short-Term chronic toxicity of photocatalytic nanoparticles to
agencies in the public, commercial, or not-for-profit sectors. bacteria, algae, and zooplankton, Nanotechnology and the Environment: Ap-
plications and Implications Progress Review Workshop III, p. 89.
Hund-Rinke, K., Simon, M., 2006. Ecotoxic effect of photocatalytic active nano-
Appendix A. Supplementary data particles (TiO2) on algae and daphnids (8 pp). Environ. Sci. Pollut. Res. 13,
225e232.
Ji, J., Long, Z., Lin, D., 2011. Toxicity of oxide nanoparticles to the green algae
Supplementary data related to this article can be found at http:// Chlorella sp. Chem. Eng. J. 170, 525e530.
dx.doi.org/10.1016/j.envpol.2017.04.053. Jiang, J., Oberdo € rster, G., Biswas, P., 2009. Characterization of size, surface charge,
and agglomeration state of nanoparticle dispersions for toxicological studies.
J. Nanoparticle Res. 11, 77e89.
References Kahru, A., Dubourguier, H.-C., 2010. From ecotoxicology to nanoecotoxicology.
Toxicology 269, 105e119.
Aruoja, V., Dubourguier, H.-C., Kasemets, K., Kahru, A., 2009. Toxicity of nano- Khan, S.S., Srivatsan, P., Vaishnavi, N., Mukherjee, A., Chandrasekaran, N., 2011.
particles of CuO, ZnO and TiO2 to microalgae Pseudokirchneriella subcapitata. Interaction of silver nanoparticles (SNPs) with bacterial extracellular proteins
Sci. Total Environ. 407, 1461e1468. (ECPs) and its adsorption isotherms and kinetics. J. Hazard. Mater. 192,
ASTM, 1975. Standard Specification for Substitute Ocean Water. Designation D 299e306.
1141e75 American Standard for Testing and Materials. Lee, W.-M., An, Y.-J., 2013. Effects of zinc oxide and titanium dioxide nanoparticles
Barmo, C., Ciacci, C., Canonico, B., Fabbri, R., Cortese, K., Balbi, T., Marcomini, A., on green algae under visible, UVA, and UVB irradiations: No evidence of
Pojana, G., Gallo, G., Canesi, L., 2013. In vivo effects of n-TiO2 on digestive gland enhanced algal toxicity under UV pre-irradiation. Chemosphere 91, 536e544.
48 M. Sendra et al. / Environmental Pollution 227 (2017) 39e48

Li, S., Wallis, L.K., Ma, H., Diamond, S.A., 2014. Phototoxicity of TiO2 nanoparticles to of Photosynthesis. Springer, pp. 49e70.
a freshwater benthic amphipod: are benthic systems at risk? Sci. Total Environ. Sendra, M., S anchez-Quiles, D., Blasco, J., Moreno-Garrido, I., Lubi rez-
an, L.M., Pe
466e467, 800e808. García, S., Tovar-Sa nchez, A., 2017. Effects of TiO2 nanoparticles and sunscreens
Lin, X., Li, J., Ma, S., Liu, G., Yang, K., Tong, M., Lin, D., 2014. Toxicity of TiO2 nano- on coastal marine microalgae: Ultraviolet radiation is key variable for toxicity
particles to Escherichia coli: effects of particle size, crystal phase and water assessment. Environ. Int. 98, 62e68.
chemistry. PLoS One 9 e110247. Shi, H., Magaye, R., Castranova, V., Zhao, J., 2013. Titanium dioxide nanoparticles: a
Liu, Y., Li, S., Chen, Z., Megharaj, M., Naidu, R., 2014. Influence of zero-valent iron review of current toxicological data. Part. Fibre Toxicol. 10, 1e33.
nanoparticles on nitrate removal by Paracoccus sp. Chemosphere 108, 426e432. Soldo, D., Hari, R., Sigg, L., Behra, R., 2005. Tolerance of Oocystis nephrocytioides to
Ma, S., Zhou, K., Yang, K., Lin, D., 2015. Heteroagglomeration of oxide nanoparticles copper: intracellular distribution and extracellular complexation of copper.
with algal cells: effects of particle type, ionic strength and pH. Environ. Sci. Aquat. Toxicol. 71, 307e317.
Technol. 49, 932e939. SPSS, 2005. Linear Mixed Effects Modeling in SPSS: an Introduction to the MIXED
Maeda, S., Sakaguchi, T., 1990. Accumulation and detoxification of toxic metal ele- Procedure. Technical Report LMEMWP-0305.
ments by algae. Introd. Appl. Phycol. 109e136. Stachowski-Haberkorn, S., Je ro^ me, M., Rouxel, J., Khelifi, C., Rince
, M., Burgeot, T.,
Maurer-Jones, M.A., Gunsolus, I.L., Murphy, C.J., Haynes, C.L., 2013. Toxicity of 2013. Multigenerational exposure of the microalga Tetraselmis suecica to diuron
engineered nanoparticles in the environment. Anal. Chem. 85, 3036e3049. leads to spontaneous long-term strain adaptation. Aquat. Toxicol. 140, 380e388.
McCullagh, P., Nelder, J.A., 1989. Generalized Linear Models. CRC press. Taurozzi, J., Hackley, V., Wiesner, M., 2012a. Preparation of a nanoscale TiO2 aqueous
Miao, A.J., Schwehr, K.A., Xu, C., Zhang, S.-J., Luo, Z., Quigg, A., Santschi, P.H., 2009. dispersion for toxicological or environmental testing. NIST Spec. Publ. 1200, 3.
The algal toxicity of silver engineered nanoparticles and detoxification by Taurozzi, J., Hackley, V., Wiesner, M., 2012b. Preparation of nanoscale TiO2 disper-
exopolymeric substances. Environtal Pollut. 157, 3034e3041. sions in biological test media for toxicological assessment. NIST Spec. Publ.
Miller, R.J., Bennett, S., Keller, A.A., Pease, S., Lenihan, H.S., 2012. TiO2 nanoparticles 1200, 4.
are phototoxic to marine phytoplankton. PLoS One 7 e30321. Uchino, T., Tokunaga, H., Ando, M., Utsumi, H., 2002. Quantitative determination of
Miller, R.J., Lenihan, H.S., Muller, E.B., Tseng, N., Hanna, S.K., Keller, A.A., 2010. Im- OH radical generation and its cytotoxicity induced by TiO2-UVA treatment.
pacts of metal oxide nanoparticles on marine phytoplankton. Environ. Sci. Toxicol. Vitro 16.
Technol. 44, 7329e7334. Vance, M.E., Kuiken, T., Vejerano, E.P., McGinnis, S.P., Hochella Jr., M.F., Rejeski, D.,
Montiel-D avalos, A., Ventura-Gallegos, J.L., Alfaro-Moreno, E., Soria-Castro, E., Hull, M.S., 2015. Nanotechnology in the real world: redeveloping the nano-
García-Latorre, E., Caban ~ as-Moreno, J.G., Ramos-Godinez, M.d.P., Lo pez- material consumer products inventory. Beilstein J. Nanotechnol. 6, 1769e1780.
Marure, R., 2012. TiO2 nanoparticles induce dysfunction and activation of hu- Velzeboer, I., Hendriks, A.J., Ragas, A.M., Van de Meent, D., 2008. Nanomaterials in
man endothelial cells. Chem. Res. Toxicol. 25, 920e930. the environment aquatic ecotoxicity tests of some nanomaterials. Environ.
Navarro, E., Baun, A., Behra, R., Hartmann, N.B., Filser, J., Miao, A.-J., Quigg, A., Toxicol. Chem. 27, 1942e1947.
Santschi, P.H., Sigg, L., 2008. Environmental behavior and ecotoxicity of engi- Wallis, L.K., Diamond, S.A., Ma, H., Hoff, D.J., Al-Abed, S.R., Li, S., 2014. Chronic TiO2
neered nanoparticles to algae, plants, and fungi. Ecotoxicology 17, 372e386. nanoparticle exposure to a benthic organism, Hyalella azteca: impact of solar
Nielsen, H.D., Berry, L.S., Stone, V., Burridge, T.R., Fernandes, T.F., 2008. Interactions UV radiation and material surface coatings on toxicity. Sci. Total Environ. 499,
between carbon black nanoparticles and the brown algae Fucus serratus: inhi- 356e362.
bition of fertilization and zygotic development. Nanotoxicology 2, 88e97. Warheit, D.B., Hoke, R.A., Finlay, C., Donner, E.M., Reed, K.L., Sayes, C.M., 2007.
OECD, 1994. OECD Guidelines for the Testing of Chemicals. Organization for Development of a base set of toxicity tests using ultrafine TiO2 particles as a
Economic. component of nanoparticle risk management. Toxicol. Lett. 171.
Piccapietra, F., Allue , C.G., Sigg, L., Behra, R., 2012. Intracellular silver accumulation Wiesner, M.R., Lowry, G.V., Alvarez, P., Dionysiou, D., Biswas, P., 2006. Assessing the
in Chlamydomonas reinhardtii upon exposure to carbonate coated silver nano- risks of manufactured nanomaterials. Environ. Sci. Technol. 40, 4336e4345.
particles and silver nitrate. Environ. Sci. Technol. 46, 7390e7397. Xiao, X., Han, Z.-y., Chen, Y.-x., Liang, X.-q., Li, H., Qian, Y.-c., 2011. Optimization of
Ratnasekhar, C., Sonane, M., Satish, A., Mudiam, M.K.R., 2015. Metabolomics reveals FDAePI method using flow cytometry to measure metabolic activity of the
the perturbations in the metabolome of Caenorhabditis elegans exposed to ti- cyanobacteria, Microcystis aeruginosa. Phys. Chem. Earth 36, 424e429. Parts A/
tanium dioxide nanoparticles. Nanotoxicology 9, 994e1004. B/C.
S
anchez-Quiles, D., Tovar-Sa nchez, A., Horstkotte, B., 2013. Titanium determination Zhao, J., Wang, Z., White, J.C., Xing, B., 2014. Graphene in the aquatic environment:
by multisyringe flow injection analysis system and a liquid waveguide capillary adsorption, dispersion, toxicity and transformation. Environ. Sci. Technol. 48,
cell in solid and liquid environmental samples. Mar. Pollut. Bull. 76, 89e94. 9995e10009.
Sayes, C.M., Wahi, R., Kurian, P.A., Liu, Y., West, J.L., Ausman, K.D., Warheit, D.B., Zhou, G.-J., Peng, F.-Q., Zhang, L.-J., Ying, G.-G., 2012. Biosorption of zinc and copper
Colvin, V.L., 2006. Correlating nanoscale titania structure with toxicity: a from aqueous solutions by two freshwater green microalgae Chlorella pyr-
cytotoxicity and inflammatory response study with human dermal fibroblasts enoidosa and Scenedesmus obliquus. Environ. Sci. Pollut. Res. 19, 2918e2929.
and human lung epithelial cells. Toxicol. Sci. 92. Zhu, X., Zhou, J., Cai, Z., 2011. TiO2 nanoparticles in the marine environment: impact
Schreiber, U., Bilger, W., Neubauer, C., 1995. Chlorophyll Fluorescence as a Nonin- on the toxicity of tributyltin to abalone (Haliotis diversicolor supertexta) em-
trusive Indicator for Rapid Assessment of in Vivo Photosynthesis, Ecophysiology bryos. Environ. Sci. Technol. 45, 3753e3758.

You might also like