You are on page 1of 13

Aquatic Toxicology 158 (2015) 1–13

Contents lists available at ScienceDirect

Aquatic Toxicology
journal homepage: www.elsevier.com/locate/aquatox

Toxicity of nano-TiO2 on algae and the site of reactive oxygen species


production
Fengmin Li a,∗ , Zhi Liang a , Xiang Zheng b,∗∗ , Wei Zhao a , Miao Wu a , Zhenyu Wang a
a
Key Laboratory of Marine Environmental Science and Ecology, Ministry of Education, College of Environmental Science and Technology,
Ocean University of China, Qingdao 266100, PR China
b
School of Environment and Natural Resources, Renmin University of China, Beijing 100872, China

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

Article history: Given the extensive use of nanomaterials, they may enter aquatic environments and harm the growth of
Received 11 August 2014 algae, which are primary producers in an aquatic ecosystem. Thus, the balance of an aquatic ecosystem
Received in revised form 15 October 2014 may be destroyed. In this study, Karenia brevis and Skeletonema costatum were exposed to nano-TiO2
Accepted 21 October 2014
(anatase, average particle size of 5–10 nm, specific surface area of 210 ± 10 m2 g−1 ) to assess the effects
Available online 3 November 2014
of nano-TiO2 on algae. The findings of transmission electron microscopy-energy dispersive X-ray spec-
troscopy (TEM-EDX) and scanning electron microscopy (SEM) demonstrate aggregation of nano-TiO2 in
Keywords:
the algal suspension. Nano-TiO2 was also found to be inside algal cells. The growth of the two species
Nano
Titanium dioxide
of algae was inhibited under nano-TiO2 exposure. The 72 h EC50 values of nano-TiO2 to K. brevis and S.
Algae costatum were 10.69 and 7.37 mg L−1 , respectively. TEM showed that the cell membrane of K. brevis was
ROS destroyed and its organelles were almost undistinguished under nano-TiO2 exposure. The malondial-
Oxidative stress dehyde (MDA) contents of K. brevis and S. costatum significantly increased compared with those of the
control (p < 0.05). Meanwhile, superoxide dismutase (SOD) and catalase activities (CAT) of K. brevis and S.
costatum changed in different ways. The reactive oxygen species (ROS) levels in both species were signifi-
cantly higher than those of the control (p < 0.05). The site of ROS production and accumulation in K. brevis
and S. costatum under nano-TiO2 exposure was explored with the addition of inhibitors of different elec-
tron transfer chains. This study indicated that nano-TiO2 in algal suspensions inhibited the growth of K.
brevis and S. costatum. This effect was attributed to oxidative stress caused by ROS production inside algal
cells. The levels of anti-oxidative enzymes changed, which destroyed the balance between oxidation and
anti-oxidation. Thus, algae were damaged by ROS accumulation, resulting in lipid oxidation and inhibited
algae growth. The inhibitors of the electron transfer chain showed that the site of ROS production and
accumulation in K. brevis cells was the chloroplast.
© 2014 Elsevier B.V. All rights reserved.

1. Introduction nanotechnologists have expressed concerns about current and


potential future developments of nanotechnology. The uncertain-
Nanomaterials are increasingly being used in industrial pro- ties about the impact of new nanomaterials on human health
duction, as well as scientific, biological, and medical research (Nel and ecotoxicity require our urgent concern (Manier et al., 2013;
et al., 2009; Patra et al., 2012; Shipway et al., 2000; Zhang, 2003). Melegari et al., 2013; Pakrashi et al., 2013; Rodea-Palomares
Nanoscale TiO2 has been widely applied in industrial applications, et al., 2012). Considering the widespread use of nanomaterials,
including the paint, textile, paper, plastic, sunscreen, cosmetic, they may be released in the aquatic environment. The possible
and food industries (Chong et al., 2010; Dastjerdi and Montazer, effects of these chemicals to aquatic life, especially among algae,
2010; Suh et al., 2009; Tong et al., 2012). Given the extensive which are the primary producers in aquatic environments, have
use of nanomaterials, several nongovernmental organizations and become a large concern. Different toxic effects of nanoparticles
on algae have been studied. Nanoparticles inhibit the growth and
photosynthetic abilities of algae, alter enzymatic activity [e.g.,
∗ Corresponding author. Tel.: +86 053266782780.
malondialdehyde (MDA), superoxide dismutase (SOD), catalase
∗∗ Corresponding author. (CAT), and glutathione (GSH)] in algal cells, and cause DNA dam-
E-mail addresses: lifengmin@ouc.edu.cn (F. Li), zhengxiang7825@sina.com age (Aruoja et al., 2009; Chen et al., 2012; Ma et al., 2013; Melegari
(X. Zheng). et al., 2013; Wang et al., 2008; Wong et al., 2010). However, the

http://dx.doi.org/10.1016/j.aquatox.2014.10.014
0166-445X/© 2014 Elsevier B.V. All rights reserved.
2 F. Li et al. / Aquatic Toxicology 158 (2015) 1–13

mechanisms of these toxic effects on aquatic organisms are largely In human and animal cells, the mitochondrion was considered
unknown. to be the main locus of ROS production. During aerobic respira-
The indirect effects of nanoparticles are caused mainly by phys- tion, most electrons transfer to the mitochondrion, which is the
ical restraints or release of toxic ions (e.g., metal nanoparticles) or terminal of the ETC, and combine with oxygen to produce water.
the production of reactive oxygen species (ROS). Physical restraints However, a small part of electrons leak out of the ETC and combine
can cause toxicity because of the addition of nanoparticles, such with oxygen to produce superoxide free radicals. Hydroxyl radicals
as the shading effect. The production of nanoparticles on the sur- and hydrogen peroxide radicals are produced through specific reac-
face of algal cells might reduce light availability for photosynthesis tions. Many studies showed that the ETC in the mitochondrion is
(Navarro et al., 2008; Schwab et al., 2011). Aggregation of nanopar- affected by toxic substances. Li et al. (2010) suggested that 3-(4-
ticles on the surface of algal cells increased the weight of cells (benzo[d]thiazol-2-yl)-1-phenyl-1H-pyrazol-3-yl) phenyl acetate
by two- to three-fold (Huang and Ismat, 2005). Nanoparticles on (DPB-5) has a significant cytotoxic effect on Hep G2 cells and
the surface of cells might block the cell wall and prevent nutri- induces cell apoptosis. In the mitochondrion, the GSH content
ent molecules from entering cells (Chen et al., 2012; Metzler et al., increases, ROS accumulates, and the membrane potential of the
2011). By contrast, Hund-Rinke and Simon (2006) showed that mitochondrion is destroyed. The addition of rotenone, an inhibitor
nanoparticles have no impact on the growth and light availability of the mitochondrial ETC (Reynafarje et al., 1976), significantly
of algal cells. Thus, whether the shading effect is the toxic mecha- decreases the ROS content caused by DPB-5. Therefore, the mito-
nism by which nanoparticles influence algal cells requires further chondrion is one of the main sites of ROS production.
research. The release of toxic ions of nanoparticles, especially metal Lovern et al. (2007) pointed out that certain nanoparticle types
nanoparticles, also causes toxicity toward algal cells (Borm et al., may have impacts on the population and food web dynamics in
2006). Because TiO2 could not release ions in suspension (Wang aquatic systems. Algae are primary producers that provide energy
et al., 2008), the toxicity of releasing ions was excluded. and nutrients for other organisms through the food web. Changes
The toxicity of nanoparticles to cells may also be attributed to that occur to algal populations under nanoparticle exposure can
ROS production. Nanoparticles might produce ROS upon interac- alter the entire ecosystem. In the present study, Karenia brevis and
tion with organisms or agents present in the environment (Navarro Skeletonema costatum were used as target organisms. Karenia brevis
et al., 2008). Nanoparticles, such as TiO2 , produce ROS in aquatic is a kind of red-tide algae. They were found in many continents such
environments because of their photocatalytic properties (Kus et al., as Europe, North America, Asian and Australia (Guiry and Guiry,
2006). The toxicity of nano-TiO2 toward bacteria increases under 2014). Marine diatom Skeletonema costatum can provide bait for
ultraviolet (UV) exposure (Adams et al., 2006). Other nanoparticles, fish and shrimp (Hasle and Syvertsen, 1997). They were often found
such as fullerol, also generate ROS after UV exposure (Badireddy in China coast (Guo et al., 2014). These organisms are important pri-
et al., 2007). When nanoparticles entered algal cells through the cell mary producers which can affect the balance of ocean ecosystem.
membrane, ROS might be induced to generate in algal cells; thus, This research aimed to determine the effects of nano-TiO2 on
the antioxidant (e.g., ascorbate, glutathion, etc.) and antioxidant the growth of two algae species, and observe the changes in ROS
enzymes (such as SOD, CAT, etc.) change following ROS genera- levels and antioxidant enzymes in algal cells under exposure to
tion to maintain the pro-oxidation/anti-oxidation balance. Patra nano-TiO2 suspension. Moreover, the mechanisms underlying the
et al. (2012) studied the antifungal effects of zinc oxide nanoparti- inhibition of algae growth by nano-TiO2 were investigated. The site
cles, and discovered that these effects were due to ROS generation. of ROS production and accumulation in algal cells under nano-TiO2
The activities of SOD and CAT increased to eliminate excess ROS. exposure was also explored.
Lipid peroxidation is an important process involved in the toxic
mechanism of nanoparticles. Chen et al. (2012) suggested that lipid 2. Materials and methods
peroxidation is induced under nano-TiO2 exposure, and found that
the MDA contents initially increased and then decreased after 8 h 2.1. Algae cultures
of exposure. Chen associated these results to ROS production or
the shading effect, but these findings have not been confirmed. K. brevis and S. costatum were purchased from the Institute of
Therefore, the present study aimed to determine the site of ROS Oceanography, Chinese Academy of Sciences. K. brevis and S. costa-
production. tum were maintained in F/2 medium (Guillard and Ryther, 1962) in
Three main types of ROS producers exist in plants, namely, 5 L flasks. Cultures were pre-cultured to logarithmic phase before
electron transport chains (ETCs) in chloroplasts and mitochon- exposure at 22 ± 1 ◦ C, with cool-white fluorescent light at a photon
dria, some peroxidases and oxidases (e.g., NADPH oxidase, NADH intensity of 12,000 lux and a 14:10 h light dark photocycle. Flasks
oxidase, xanthine oxidase, lipoxygenase, glycolate oxidase, amine were shaken by hand twice each day to obtain a homogenous cell
oxidase, etc.), and photosensitizers, such as chlorophyll molecules distribution.
(Dat et al., 2000; Edreva, 2005; Song et al., 2012). Chloroplast is
one of the main source and site of ROS production (Alberts and
Lewis, 2002; Gill and Tuteja, 2010). If light energy is not com- 2.2. Nano-TiO2
pletely consumed, O2 , which has been converted into an electron,
will be a selective receptor of excess light energy that is con- Nano-TiO2 (anatase) was purchased from Wanjingxin Company
verted to superoxide free radicals. Many studies showed that plant (Hangzhou, China). It had an average particle size of 5–10 nm and
cells can produce ROS when exposed to toxic substances. Forti and specific surface area of 210 ± 10 m2 g−1 .
Gerola (1977) found inhibited peroxidase activity and photosyn-
thesis rate, as well as low H2 O2 content, when chloroplasts are 2.2.1. Particle size and morphology of nano-TiO2
isolated from spinach exposed to azide and cyanide. Tripathy and Nano-TiO2 was diluted with distilled water to a final concentra-
Mohanty (1980) declared that Zn2+ ions affect the ETC in chloro- tion of 20 mg L−1 . After 30 min of ultrasonic treatment, nano-TiO2
plasts. Photosystem II (PS II) is inhibited under low concentrations was dipped by a copper screen and air-dried under sunlight for 2 h.
of Zn2+ exposure, whereas photosystem I (PS I) is inhibited under The particle size and crystal parameter of nano-TiO2 were detected
high concentrations of Zn2+ exposure. Therefore, metal nanopar- by transmission electron microscopy energy-dispersive X-ray spec-
ticles that can release metal ions may also affect the ETC in algal troscopy (TEM-EDX). After K. brevis cultures were exposed to
chloroplasts. nano-TiO2 (0, 10, and 100 mg L−1 ) for 24 h under similar culture
F. Li et al. / Aquatic Toxicology 158 (2015) 1–13 3

parameters, nano-TiO2 morphology was observed by scanning 100 mmol L−1 EDTA). Nano-TiO2 suspension was added, and the
electron microscopy (SEM). final concentration of nano-TiO2 was 10 mg L−1 . The chlorophyll
suspension without nano-TiO2 was used as the control. Groups
2.2.2. Preparation of nano-TiO2 suspension were prepared in triplicate. Given that O2 •− is too active to detect
Nano-TiO2 was suspended in F/2 medium to obtain stock solu- and • OH is the product of the secondary reaction, the • OH content
tions (1000 mg L−1 (12.5 mol L−1 )) followed by 30 min of sonication, was measured in the isolated chloroplast after 24 h of exposure.
and shaken for 2 min. Prior to the exposure experiment, the stock
solutions (1000 mg L−1 ) of nano-TiO2 were diluted to four con- 2.5. Test methods
centrations, namely, 125, 250, 500, and 750 mg L−1 (1562.5, 3125,
6250, 9375 mmol L−1 ). The solutions were disinfected with UV light 2.5.1. Cell number and EC50
before further analyses. The cell number (cell mL−1 ) was counted using a hemocytome-
ter under a light microscope (Nikon, YS100, Japan) every 24 h. The
2.3. Effects of nano-TiO2 on algae growth EC50 value was calculated using SPSS version 19.

2.3.1. Shading effect 2.5.2. Chl-a content


A 500 mL flask containing 250 mL of algae culture was mixed After 48 h of exposure, Chl-a of test algae was extracted in 90%
in a 2 L flask filled with 50 mg L−1 nano-TiO2 suspension, and the ethanol in the dark, followed by 5 min of centrifugation at 3600 rpm
control group was filled with distilled water. The water level out- for K. brevis and 6 min of centrifugation at 4200 rpm for S. costatum.
side the 500 mL flask should be higher than that inside the flask. Chl-a was diluted to 10 mL with 90% ethanol. The Chl-a content
All samples were prepared in triplicate. The cell number (cell ml−1 ) was measured at 750, 663, 645, and 630 nm with a UV–vis scanning
was counted every 24 h. spectrophotometer (UV-2550, SHIMADZU, Japan), and 90% ethanol
was set as the control.
2.3.2. Growth of algae exposed to nano-TiO2 suspension
Algae cultures (pre-cultured in 5 L flask) in logarithmic phase 2.5.3. Preparation of cell homogenate
with cell number more than 104 cells mL−1 were diluted with K. brevis cells were collected by centrifugation at 3381 rpm and
F/2 medium. Approximately 240 mL of algae was transferred into 4 ◦ C for 5 min, whereas S. costatum cells were collected by centrifu-
15,500 mL flasks equally. Approximately 10 mL of these four dilu- gation at 4200 rpm and 4 ◦ C for 6 min. Cells were homogenized in
tions (125, 250, 500, and 750 mg L−1 ) of nano-TiO2 were added into 4 mL of PBS solution (pH 7.8, 4 ◦ C) by an ultrasonic cell pulverizer
these flasks. The algae cultures, as well as the control without nano- (KS-500F, Kesheng, Ningbo, China) in an ice bath for 5 min. The
TiO2 , were treated under identical conditions. These five groups homogenate was used to measure the MDA content, or transferred
were prepared in triplicate. The final concentrations of nano-TiO2 using 3 mL of PBS into a 10 mL plastic centrifugal tube for centrifu-
in the algae cultures were 0, 5, 10, 20, and 30 mg L−1 . The cell den- gation at 4428 rpm and 4 ◦ C for 10 min. The supernatant was used
sity was counted every 24 h, and the growth curve and 72 h EC50 for enzyme, ROS, and free radical assays.
were analyzed. The chlorophyll a (Chl-a) content was measured
after 48 h of exposure. The nano-TiO2 distributions and ultrastruc- 2.5.4. Lipid peroxidation assays
ture of algal cells in 0 and 20 mg L−1 nano-TiO2 suspension were The MDA content and SOD and CAT activities were detected
detected by TEM after 24 h of exposure. using a reagent kit (Nanjing Jiancheng Biotechnology Institute,
China) according to the manufacturer’s instructions (Deng et al.,
2.4. Lipid peroxidation and site of ROS production 2009; Han et al., 2004). The MDA content was determined based
on the modified NBT method (Rocchetta et al., 2006). SOD activ-
2.4.1. Lipid peroxidation ity was determined based on the WST-1 method. CAT activity was
The MDA concentration, antioxidant enzyme (SOD and CAT) determined based on the method of ammonium molybdate spec-
activities, and contents of three free radicals (• OH, O2 •− , and H2 O2 ) trophotometry.
in algal cells and suspensions were analyzed after 48 h of exposure.
2.5.5. ROS assays
2.4.2. Site of ROS production The cell homogenate was obtained using the method described

Algae cultures were cultured as described in 2.1, and transferred in Section 2.5.3. The levels of three free radicals (i.e., • OH, O2 − ,
into 500 mL flasks. Firstly three inhibitors, namely, rotenone (CAS and H2 O2 ) and ROS were detected using a reagent kit (Nanjing
330-54-1, analytical grade, Sigma–Aldrich, USA), diuron (CAS 83- Jiancheng Biotechnology Institute, China) according to the man-

79-4, analytical grade, Sigma–Aldrich, USA), and dicoumarol (CAS ufacturer’s instructions. The free radicals • OH, O2 − , and H2 O2
66-76-2, analytical grade, Sigma–Aldrich, USA), were added into were detected by a UV–vis spectrophotometer (UV-2550, SHI-

each flask. 30 min later 20 mg L−1 nano-TiO2 was added into each MADZU, Japan), • OH was detected at 550 nm, O2 − was detected
flask with addition of inhibitors. Then algae cultures were exposed at 550 nm, and H2 O2 was detected at 405 nm. ROS levels were
to 0 and 20 mg L−1 nano-TiO2 suspensions, inhibitors, or a mixture determined using a 2,7-dichlorofuorescin diacetate probe. ROS lev-
of the inhibitor and nano-TiO2 for 24 h under similar culture param- els were detected by a fluorescence spectrophotometer (F4600,
eters. The final volume of algae cultures in the flasks was 250 mL, Hitachi, Japan) at the emission and excitation wavelengths of 525
and the groups were prepared in triplicate. The final concentrations and 488 nm, respectively.
of rotenone, diuron, and dicoumarol were 1.25, 5, and 5 ␮mol L−1
(the concentrations at which these three inhibitors had no effects 2.5.6. TEM-EDX and SEM analysis
on algae), respectively. Specific method of operation was shown in After 24 h of exposure, algal cells were collected by centrifu-
Table 1. The ROS levels were then detected. gation (3800 rpm, 10 min), and the samples were fixed with 2.5%
glutaraldehyde solution in 4 ◦ C for 2 h. The samples were then
2.4.3. Isolated chloroplast washed with 0.1 M PBS (pH 7.8) by centrifugation (3800 rpm,
Algae cultures were cultured for 48 h as mentioned above. 10 min) three times. Algal cells were fixed with 1% osmium tetrox-
Chlorophyll was extracted from algal cells by centrifugation, ide for 1 h in 4 ◦ C, and 0.1 M PBS (pH 7.8) was added to wash the cells
and suspended in buffer solution (pH 8, 150 mmol L−1 NaCl, by centrifugation (3800 rpm, 10 min) three times. Samples were
4 F. Li et al. / Aquatic Toxicology 158 (2015) 1–13

Table 1
A list of steps on how to find out which ETC was the site of ROS production with addition of ETC inhibitors. A: Control; B: rotenone; C: diuron; D: dicoumarol; E: nano-TiO2 ;
F: nano-TiO2 + rotenone; G: nano-TiO2 + diuron; H: nano-TiO2 + dicoumarol.

A B C D E F G H

Algae culture (mL) 250 245 245 245 245 240 240 240
62.6 ␮mol L−1 Rotenone (mL) / 5 / / / 5 / /
250 ␮mol L−1 Diuron (mL) / / 5 / / / 5 /
250 ␮mol L−1 Dicoumarol (mL) / / / 5 / / / 5
Algae culture was treated with inhibitors for 30mins
1000 mg L−1 TiO2 (mL) / / / / 5 5 5 5

dehydrated successively through a series of acetone solutions of


30%, 50%, 70%, 90%, and 100% for 10 min. The samples were embed-
ded in Epon 812 resin, and fixed in gradient temperatures of 37 ◦ C,
45 ◦ C, and 65 ◦ C for 24 h. Sample slices were prepared by a micro-
tome (Ultracut E), and stained with uranyl acetate and lead citrate.
The ultrastructures of algal cells were observed by TME-EDX (JEOL,
JEM-1200EX, Japan). The samples observed by SEM were treated
similar to the procedure mentioned above, except acetone solution
was used instead of alcohol solution. SEM (KYKY-2800B) was used
to observe samples.

2.6. Statistical analysis

The experimental data were analyzed by one-way ANOVA using


SPSS version 19. Tukey’s multiple comparison at a probability of
0.05 was also used. Statistical significance was set at p < 0.05.

3. Results

3.1. Nano-TiO2

3.1.1. Particle size of nano-TiO2 in suspension


The particle size of nano-TiO2 diluted with distilled water was
observed by TEM (Fig. 1A). The mean particle size of nano-TiO2 was
calculated to be less than 40 nm. Fig. 1B shows particle aggregation
detected by TEM-EDX. Comparison with the standard lattice index
of TiO2 showed that the aggregates were TiO2 .

3.1.2. Morphology of nano-TiO2 in algae cultures


After nano-TiO2 suspensions were added into algae cultures,
nano-TiO2 morphology in the suspensions was observed by SEM.
Fig. 2 shows nano-TiO2 aggregates on the surface of algal cells in
algae culture. The particle size of nano-TiO2 in suspensions was
bigger than that of dry powder. Algal cells were almost absolutely
covered by 100 mg L−1 nano-TiO2 . In addition, the aggregates on
the cell surface increased with increasing nano-TiO2 concentra-
tion. Nanoparticles have been used as an efficient sorbent in many
fields because of their high specific surface area. The high specific Fig. 1. Particle size of nano-TiO2 in suspension. Particle size (A) of nano-TiO2 and
surface area and strong adsorption ability of nano-TiO2 may cause its crystal parameter (B) was detected by TEM-EDX.
enwrapping of algal cells.

3.3. Effects of nano-TiO2 on algae growth


3.2. Shading effect
3.3.1. Inhibition of nano-TiO2 on algae growth
Aggregation of nano-TiO2 on the surface of algal cells in algae The inhibition of nano-TiO2 on algae growth is shown in Fig. 4.
cultures may decrease the photosynthetic abilities of algal cells and Fig. 4A shows the effects of nano-TiO2 on the growth of K. brevis.
inhibit growth. The shading effect may be a mechanism of nano- The inhibitory rate increased with increasing nano-TiO2 concen-
TiO2 toxicity. Fig. 3 shows that the cell numbers of K. brevis and tration, but stopped increasing with time. After 24 h of exposure,
S. costatum in the treatment groups changed, but were not signifi- the inhibitory rates of each group were 32%, 63%, 74%, and 84%. The
cantly different from those of the control. Thus, the shading effect results show that K. brevis was inhibited by nano-TiO2 , and the 72 h
was not a mechanism of nano-TiO2 toxicity in algal cells. If nano- EC50 value was 10.69 mg L−1 .
TiO2 can enter algal cells, we then aimed to determine whether it Fig. 4B shows the effects of nano-TiO2 on the growth of S.
can inhibit algae growth. costatum. The inhibitory rate increased with increasing nano-TiO2
F. Li et al. / Aquatic Toxicology 158 (2015) 1–13 5

Fig. 2. Morphology of nano-TiO2 in algae cultures. Morphology of nano-TiO2 was observed by SEM in K. brevis cultures. (A): K. brevis treated by 0 mg L−1 nano-TiO2 ; (B): K.
brevis treated by 10 mg L−1 nano-TiO2 ; (C): K. brevis treated by 100 mg L−1 nano-TiO2 .

Fig. 3. Growth of algae cells with shading effects. (A): K. brevis; (B): S. costatum.
Fig. 4. Inhibition rates of different concentrations of nano-TiO2 on the growth of
Values represent mean ± S.D. n = 3.
algae. (A): K. brevis; (B): S. costatum. Values represent mean ± S.D. n = 3.
6 F. Li et al. / Aquatic Toxicology 158 (2015) 1–13

Table 2
The analysis of element content. A: arrows 1 in Fig. 6B; B: arrows 2 in Fig. 6B.

A
Element Weight percent (%) Atomic percent (%)

CK 28.58 67.51
Ti K 4.18 2.48
Cu K 67.23 30.02

Element Weight percent (%) Atomic percent (%)

CK 26.36 59.88
OK 6.28 10.71
Ti K 4.18 2.48
Cu K 63.93 27.45

the standard lattice index of TiO2 showed that these particles were
TiO2 . The element content of the particles marked by arrows in
Fig. 6B was also analyzed. As shown in Table 2, the weight percent
of Ti outside the cells was 4.18, which was higher than that inside
the cells. Given that the carrier of slices of algal cells was a cop-
per network, the weight percent of Cu was the highest based on
elemental analysis.

3.3.3. Ultrastructure of algal cells


The ultrastructures of algal cells were detected by TEM. S. costa-
tum is a diatom with a thick cell wall, so operation in the experiment
for algal cell immobilization failed. The ultrastructure of K. bre-
vis cells was shown in Fig. 7A. A complete cell of K. brevis with
organelles was observed, and the chloroplast exhibited a clear and
complicated architecture with numerous thylakoids (Fig. 7A and
C). Under exposure to nano-TiO2 , the stroma lamellae of chloro-
plast showed a misty appearance (Fig. 7D). Fig. 7B showed that the
treated cells had damaged cell walls.

Fig. 5. Chl-a of algae cells exposed to different concentrations of nano-TiO2 . (A):


K. brevis; (B): S. costatum. Values represent mean ± S.D. Different letters represent 3.4. Lipid peroxidation and ROS production
significant differences between the treatment means (p < 0.05, Tukey), n = 3.

3.4.1. Lipid peroxidation assays


concentration, and continued to increase with time. For instance, The membranes of algal cells were damaged based on the
the inhibitory rate of the 5 mg L−1 nano-TiO2 group was 27.1%. ultrastructure of algal cells, as described in 3.3.4. Thus, nano-TiO2
After 24 h of culture, it increased to 61% at the seventh day. The toxicity may be related to lipid peroxidation. SOD and CAT activities
other groups showed similar trends. The results show that S. costa- and MDA content in algal cells were detected to determine whether
tum was inhibited by nano-TiO2 , and the 72 h EC50 value was nano-TiO2 causes lipid peroxidation. As shown in Fig. 8, SOD
7.37 mg L−1 . and CAT activities and MDA content of algal cells changed under
The chlorophyll content is an indicator of the growth situation of nano-TiO2 exposure. As the nano-TiO2 concentration exceeded
algal cells. The Chl-a content of algal cells exposed to different nano- 10 mg L−1 , SOD activity decreased significantly (p < 0.05, Fig. 8A. By
TiO2 concentrations is shown in Fig. 5. The Chl-a content of K. brevis contrast, SOD activity in S. costatum cells significantly increased
cells significantly decreased under treatment with 10 mg L−1 nano- with increasing nano-TiO2 concentration (p < 0.05, Fig. 8B). CAT
TiO2 (p < 0.05), as shown in Fig. 5A. The Chl-a contents of the control activity in K. brevis cells increased significantly with increasing
and 30 mg L−1 groups were 448 and 209 mg m−3 , respectively. The nano-TiO2 concentration (p < 0.05), whereas the opposite trend was
Chl-a content of the 30 mg L−1 group decreased by 53.3% compared observed in S. costatum (Fig. 8C and D). In K. brevis cells, CAT activity
with that of the control. Fig. 5B indicates that the Chl-a content of was not inhibited under nano-TiO2 exposure, but promoted by the
S. costatum cells significantly decreased with increasing nano-TiO2 increase in hydrogen peroxide radicals. SOD activity was inhibited
concentration (p < 0.05). The Chl-a content of the 30 mg L−1 group under nano-TiO2 exposure. In S. costatum cells, the changes in SOD
was 55 m−3 , and decreased by 61.5% compared with that of the and CAT were consistent with the changes in hydrogen peroxide
control. radicals. SOD activity increased as • OH decreased, and CAT activity
decreased as • OH decreased. In K. brevis cells, the MDA contents of
3.3.2. Distribution of nano-TiO2 in algae the treatment groups were significantly different from those of the
As shown in Fig. 6A, the cell walls of K. brevis were covered control (p < 0.05, Fig. 8E), and showed an increasing trend. The MDA
by particles and broken down where the particles gathered. The content of S. costatum significantly increased with increasing nano-
arrows in Fig. 6B show that the particles were found not only out- TiO2 concentration (p < 0.05, Fig. 8F). The MDA content of K. brevis
side algal cells but also in the chloroplast inside algal cells. These and S. costatum under 30 mg L−1 nano-TiO2 exposure was 2.4- and
particles were analyzed by TEM-EDX (Fig. 6C). Comparison with 2.7-fold higher than that of the control, respectively.
F. Li et al. / Aquatic Toxicology 158 (2015) 1–13 7

Fig. 6. Distribution of nano-TiO2 in algae. Arrows in Fig. 6B showed the site of nano-TiO2 entered into cells. (A, B): K. brevis treated by 20 mg L−1 nano-TiO2 ; (C): Crystal
parameter of particles detected by TEM-EDX.

Fig. 7. Ultrastructure of K. brevis cells treated by nano-TiO2 . (A, C): control cells; (B, D): cells treated by 20 mg L−1 nano-TiO2 . CW, cell wall; CHL, chloroplast.
8 F. Li et al. / Aquatic Toxicology 158 (2015) 1–13

Fig. 8. SOD, CAT activities and MDA content of algae cells. Algae were exposed to nano-TiO2 for 48 h. (A, C, E): K. brevis; (B, D, F): S. costatum. Values represent mean ± S.D.
Different letters represent significant differences between the treatment means (p < 0.05, Tukey), n = 3.

3.4.2. ROS in Fig. 9. The contents of the three free radicals demonstrated
To determine whether ROS production causes lipid peroxida- similar changes in K. brevis and S. costatum. When the nano-TiO2
tion in algal cells, the contents of three free radicals in algal cells concentration was 30 mg L−1 , the contents of • OH and H2 O2 were
under nano-TiO2 exposure were analyzed and the results are shown significantly different from those of the other groups and increased
F. Li et al. / Aquatic Toxicology 158 (2015) 1–13 9

Fig. 9. Contents of three free radicals of algae cells. Algae were exposed to nano-TiO2 for 48 h. (A, C, E): K. brevis; (B, D, F): S. costatum. Values represent mean ± S.D. Different
letters represent significant differences between the treatment means (p < 0.05, Tukey), n = 3.


compared with that of the control. The O2 − contents showed no 3.5. Site of ROS production in algal cells under nano-TiO2
significant differences between the treatment and control groups, exposure

possibly because O2 − was too active to be detected before its
transformation. No ROS were detected in the nano-TiO2 suspen- 3.5.1. Effects of inhibitors of ETC on ROS level
sion. Therefore, ROS were produced and accumulated in algal To determine the site of ROS production in algal cells, the effects
cells. of inhibitors rotenone, diuron, and dicoumarol of ETC on ROS levels
10 F. Li et al. / Aquatic Toxicology 158 (2015) 1–13

Fig. 10. Effects of inhibitors on ROS level induced by nano-TiO2 . Cells were treated with 20 mg L−1 nano-TiO2 for 12 h after pretreatment with rotenone (1.25 ␮M L−1 ),
diuron (5 ␮M L−1 ), and dicoumarol (5 ␮M L−1 ) for 30 min. A: K. brevis; B: S. costatum. (A): Control; (B): rotenone; (C): diuron; (D): dicoumarol; (E): nano-TiO2 ; (F): nano-
TiO2 + rotenone; (G): nano-TiO2 + diuron; (H): nano-TiO2 + dicoumarol. Values represent mean ± SD. Different letters represent significant differences between the treatment
groups (p < 0.05, Tukey, n = 3).

induced by nano-TiO2 are shown in Fig. 10. Cells were treated with
nano-TiO2 for 12 h (ROS level peaked in the preliminary experi-
ment) after pretreatment with rotenone, diuron, and dicoumarol
for 30 min. The fluorescence degree indicates the ROS content.
Fig. 10A shows that the ROS content in K. brevis cells significantly
increased under exposure to one of the three inhibitors or nano-
TiO2 as the control (p < 0.05). Nano-TiO2 induced more ROS than the
other three inhibitors. Groups of rotenone or dicoumarol followed
by addition with nano-TiO2 both significantly increased compared
with groups in which only one of them was added (p < 0.05). In
addition, the ROS content in the diuron with nano-TiO2 group sig-
nificantly decreased compared with that in the diuron only and
nano-TiO2 only groups (p < 0.05). Addition of diuron inhibited the
transfer of electrons in the ETC of the chloroplast, and the ROS con-
tent decreased. Thus, we considered the chloroplast to be the site
of ROS production induced by nano-TiO2 in K. brevis. In S. costa-
tum, the ROS content in the rotenone, diuron, or nano-TiO2 only
groups slightly increased compared with that of the control, except
in the dicoumarol group (Fig. 10B). This result could be due to the
high dicoumarol toxicity in S. costatum cells. Nano-TiO2 induced
Fig. 11. Effects of nano-TiO2 on hydroxyl radical production of isolated chloro-
more ROS than the other three inhibitors. Groups of one of the plast. The concentration of nano-TiO2 was 10 mg L−1 . Values represent mean ± SD
three inhibitors with the addition of nano-TiO2 slightly increased (*p < 0.05, n = 3).
compared with groups containing only one inhibitor. No significant
differences were observed among the groups. Thus, the site of ROS
production induced by nano-TiO2 in S. costatum remains unknown. nano-TiO2 on Chlamydomonas reinhardtii was 10 mg L−1 . Algal toxi-
cities of different nanoparticles were based on the types of algae and
3.5.2. Effects of nano-TiO2 on isolated chloroplast nanoparticles. The 72 h EC50 value of CuO nanoparticles on C. rein-
To prove that the chloroplast was the site of ROS production hardtii was 150.45 mg L−1 , and the NOEC ≤ 100 mg L−1 (Melegari
induced by nano-TiO2 in K. brevis cells, the • OH content of the iso- et al., 2013). Our results provide data for assessing the algal toxic-
lated chloroplast extracted from K. brevis cells was determined. ity of nano-TiO2 on marine algae. Accurate quantification of Chl-a
The • OH content induced by nano-TiO2 significantly increased by is an important step in estimating phytoplankton biomass in both
1.17-fold compared with that of the control (Fig. 11). These results marine and freshwater environments (Simon and Helliwell, 1998).
prove that the chloroplast was the site of ROS production induced The Chl-a content can be an indicator of the growth situation of
by nano-TiO2 in K. brevis cells. algal cells. In the current study, the Chl-a contents of both K. brevis
and S. costatum sharply decreased, possibly weakening their pho-
4. Discussion tosynthetic abilities. Thus, the growth of K. brevis and S. costatum
was inhibited by nano-TiO2 .
4.1. Inhibition of nano-TiO2 on algae growth
4.2. Lipid peroxidation as one of the mechanisms
To investigate the toxicity of nano-TiO2 on the growth of K.
brevis and S. costatum, several treatments were conducted. Both Some studies have reported the mechanisms by which nanopar-
K. brevis and S. costatum were inhibited, and their 72 h EC50 ticles inhibit the growth of algae. Although the shading effect is
values were 10.69 and 7.37 mg L−1 , respectively (Fig. 4). Simi- one of the mechanisms, few have been proven (Chen et al., 2012;
larly, Huang and Ismat (2005) reported that the EC50 value of Sadiq et al., 2011). Wang et al. (2008) showed that nanoparticles
F. Li et al. / Aquatic Toxicology 158 (2015) 1–13 11

aggregate in seawater, thereby reducing specific surface area and


active sites of the surface. A similar phenomenon was observed
by SEM. Nano-TiO2 aggregated on the surface of algal cells after
preparation in solution, possibly because of the large specific sur-
face and high adsorption capacity of nano-TiO2 . These aggregations
may prevent chlorophyll from absorbing light and inhibit algae
growth (Schwab et al., 2011). To prove that nano-TiO2 could inhibit
algae growth by preventing light absorption of algae, the shading
effect was analyzed in our study. At 50 mg L−1 nano-TiO2 , algae
growth in the treatment groups showed no significant differences
from that of the control. Therefore, a nano-TiO2 concentration of
<50 mg L−1 could not inhibit the growth of algal cells via the shading
effect. A similar result showed that algae growth and light uti-
lization did not change in the presence of nano-TiO2 (Hund-Rinke
and Simon, 2006). Ji et al. (2011) assessed algal toxicities of nano-
ZnO and nano-TiO2 in the presence and absence of light. Their
results showed that toxicities of groups with nanoparticles in the
dark were higher than those of the groups without nanoparticles Fig. 12. Site of diuron inhibited electron transport in the chloroplast ETC. Diuron
inhibited the transfer of electrons from QA to QB .
in the dark. The shading effect was not the main factor inhibiting
algae growth. Nano-TiO2 entered the algal cells and adhered to the
chloroplast (Fig. 6). In theory, nanoparticles with particle sizes less relationship. The ROS contents in Chlorella sp. cells exposed to ROS
than cell wall apertures could go through the cell wall of algal cells levels in the cells increased and were not eliminated under oxi-
(Navarro et al., 2008). Wang et al. (2011) found that nanoparti- dant stress caused by nanoparticles, such as nano-ZnO, thereby
cles enter algal cells by entering the apertures in the cell wall, as increasing lipid peroxidation (Lin and Xing, 2008). In the present
well as endocytosis of the cell membrane. By contrast, our current study, the contents of the three free radicals increased in both K.
research indicated that nano-TiO2 might enter the cell by damag- brevis and S. costatum cells. Moreover, no ROS were detected in
ing the cell wall where nanoparticles were aggregated (Fig. 7B). In the nano-TiO2 suspensions. ROS were produced in algal cells. Thus,
algae suspensions with nano-TiO2 , analysis of the crystal parame- the balance in the oxidation-reduction system in algal cells was
ters of particles both outside and inside algal cells proved that the destroyed. Intracellular ROS production increased oxidative stress,
particles were TiO2 , as detected by TEM-EDX. Given that nano-TiO2 which could explain the toxic effects of nano-TiO2 on K. brevis and
entered algal cells, algae growth was inhibited by nano-TiO2 itself. S. costatum. Therefore, lipid peroxidation caused by oxidative stress
Based on the ultrastructure of K. brevis (Fig. 7), algal cells were dam- might be one of the mechanisms of nano-TiO2 in algal cells.
aged. Less organelles and more crystal particles were observed. Our
results concur with many studies that nanoparticles can enter cells. 4.3. Chloroplast as the main site of ROS accumulation
Chen et al. (2012) claimed that nano-TiO2 was located inside algal
cells. They considered that nano-TiO2 inside algal cells might be Rotenone, diuron, and dicoumarol are inhibitors of
more toxic to the cells. Navarro et al. (2008) reported that pores the mitochondrial ETC, chloroplast ETC, and membrane
in the cell wall might be induced by nanoparticles to grow; thus, oxidation–reduction ETC, respectively (Choi et al., 2002; Li
nanoparticles could pass the cell wall through the new pores. By et al., 2010; Yu et al., 2007). Rotenone and diuron have been
contrast, our results show that nano-TiO2 might enter the cells by used as ETC inhibitors in some studies. Complexes I and III of
damaging the cell wall. TEM images (Fig. 7) also provided direct ETC in the mitochondria are the major sites for ROS production.
evidence that the cell structures were damaged by nanoparticles Rotenone inhibits complex I and increases ROS generation in
inside the cells. submitochondrial particles (Li et al., 2003). The effects of diuron
Lipid peroxidation induced by ROS production is one of the on the chloroplast ETC are shown in Fig. 12. Diuron can bind with
main mechanisms of algal toxicity of nanoparticles (Ma et al., 2013; the QB protein and inhibit PQ reduction (Laisk and Oja, 2013); thus,
Melegari et al., 2013; Rodea-Palomares et al., 2012). Antioxidant the action site of diuron is between QB and QA (Choi et al., 2002).
enzymes, SOD and CAT activities, and MDA as the end product In the present study, ROS production significantly increased under
of lipid peroxidation were detected to estimate the level of lipid exposure of nano-TiO2 with addition of dicoumarol/rotenone
peroxidation (Han et al., 2004; Qiu et al., 2013). In this study, the compared with that under exposure of nano-TiO2 alone. ROS
MDA contents in algal cells significantly increased with increasing production significantly decreased under exposure of nano-TiO2
nano-TiO2 concentrations, indicating that nano-TiO2 could impose with addition of diuron compared with that under exposure of
oxidative stress and cause lipid peroxidation in algal cells (Lin et al., nano-TiO2 alone (Fig. 10A). This finding indicates that diuron could
2012). Activities of SOD and CAT in K. brevis and S. costatum cells remove excess ROS produced by nano-TiO2 , or ROS produced by
changed, exhibiting various trends. In K. brevis cells, CAT activity nano-TiO2 were inhibited. Given that diuron was the inhibitor
increased and SOD activity decreased with increasing nano-TiO2 of the chloroplast ETC, ROS produced by TiO2 were found on the
concentration. The changes in SOD and CAT activities in S. costa- chloroplast ETC. We concluded that the action site of nano-TiO2
tum cells were different from those in K. brevis. CAT and SOD can on K. brevis cells was the chloroplast. Similar results were also
both increase to remove excess ROS (Hu et al., 2014). In contrast to reported. Li et al. (2010) suggested that rotenone can significantly
the current study, increased H2 O2 inhibited SOD activity in K. bre- decrease the ROS content caused by DPB-5. Similarly, a previous
vis cells and further increased H2 O2 (San Mateo et al., 1998). In S. study showed that H2 O2 production induced by antimycin A can
costatum cells, SOD activity was less by four orders of magnitudes, be prevented with the addition of rotenone in the mitochondria
and CAT activity was less by one order of magnitude than those in with complex I substrates (Chen et al., 2003).
K. brevis cells. As SOD and CAT activities were inhibited, the • OH The chloroplast is the site of the affected toxicants, such as
content increased by seven orders of magnitudes. selenite, which causes sparse substrates, dissolution of proteins,
Although lipid peroxidation is considered to be caused by and changes in the thylakoid system (Morlon et al., 2005). A
intracellular ROS production, few studies have confirmed this previous study showed that nano-TiO2 with low concentration
12 F. Li et al. / Aquatic Toxicology 158 (2015) 1–13

Fig. 13. The toxicity mechanism of nano-TiO2 in K. brevis.

under UV irradiation increases the electron transport rate, energy and indistinct organelles. Finally, algae growth was inhibited. The
conversion rate, and oxygen release of isolated chloroplast site of ROS production induced by nano-TiO2 in K. brevis was the
extracted from spinach (Hong et al., 2005). Nano-TiO2 may bind chloroplast.
with PS II, and no effect has been found on the structure of PS II. The chloroplast was the site of ROS production induced by nano-
These findings were consistent with the results in the current study. TiO2 . Results from the current study cannot determine the specific
To prove that the chloroplast was the main site of ROS pro- site of the chloroplast ETC where ROS were produced.
duction in K. brevis cells, we studied the effect of nano-TiO2 on
the isolated chloroplast extracted from K. brevis cells. Increased
H2 O2 contents indicated that the isolated chloroplast produced Acknowledgements
ROS under nano-TiO2 exposure. Therefore, the chloroplast was a
site of ROS production in K. brevis cells induced by nano-TiO2 . TEM We appreciate the financial support from the National Natu-
analysis (Fig. 7) also showed that the stroma lamellae of the chloro- ral Science Foundation of China (51378480), National Key Basic
plast demonstrated a misty appearance under nano-TiO2 exposure, Research Program of China (2015CB453301), NSFC-Shandong Joint
which explained that the chloroplast was damaged by ROS produc- Fund (U1406403), and Public Science and Technology Research
tion. Funds Projects of Ocean (201305003).
Our study is the first to report on the action site of nanoparticles
on algal cells. We proved that the site of ROS induced by nano-TiO2
in K. brevis cells was the chloroplast. The site of ROS production References
induced by nano-TiO2 in S. costatum remains unknown despite the
Adams, L.K., Lyon, D.Y., Alvarez, P.J.J., 2006. Comparative eco-toxicity of nanoscale
addition of the three ETC inhibitors. However, lipid peroxidation TiO2 , SiO2 , and ZnO water suspensions. Water Res. 40 (19), 3527–3532.
was detected in S. costatum cells under nano-TiO2 exposure. The Alberts, B.J.A., Lewis, J., 2002. Molecular Biology of the Cell, 4th ed. Garland Publishing
toxicity mechanism of nano-TiO2 on K. brevis is shown in Fig. 13. Inc., New York.
Aruoja, V., Dubourguier, H.-C., Kasemets, K., Kahru, A., 2009. Toxicity of nanoparticles
First, nano-TiO2 aggregated on the surface of algal cells.
of CuO, ZnO and TiO2 to microalgae Pseudokirchneriella subcapitata. Sci. Total
Nanoparticles then entered the cell wall by passing through the Environ. 407 (4), 1461–1468.
pores or damaging the cell wall. Endocytosis of the cell mem- Badireddy, A.R., Hotze, E.M., Chellam, S., Alvarez, P., Wiesner, M.R., 2007. Inactivation
brane allowed entry of the nanoparticles into the cells, and some of Bacteriophages via photosensitization of fullerol nanoparticles. Environ. Sci.
Technol. 41 (18), 6627–6632.
nanoparticles attached to the chloroplast. ROS were produced on Borm, P., Klaessig, F.C., Landry, T.D., Moudgil, B., Pauluhn, J., Thomas, K., Trottier, R.,
the ETC of the chloroplast upon nano-TiO2 exposure in cells. Excess Wood, S., 2006. Research strategies for safety evaluation of nanomaterials, Part
ROS were produced, and anti-oxidative enzymes were activated to V: role of dissolution in biological fate and effects of nanoscale particles. Toxicol.
Sci. 90 (1), 23–32.
maintain the balance between pro-oxidation/anti-oxidation. How- Chen, Q., Vazquez, E.J., Moghaddas, S., Hoppel, C.L., Lesnefsky, E.J., 2003. Production
ever, this balance was destroyed by an increase in ROS production of reactive oxygen species by mitochondria – central role of complex III. J. Biol.
because enzymes could not eliminate excess ROS. ROS accumulated Chem. 278 (38), 36027–36031.
Chen, L., Zhou, L., Liu, Y., Deng, S., Wu, H., Wang, G., 2012. Toxicological effects of
in algal cells, thereby causing lipid peroxidation. The ultrastructure nanometer titanium dioxide (nano-TiO2 ) on Chlamydomonas reinhardtii. Ecotox-
of cells was damaged, as evidenced by the incomplete membrane icol. Environ. Saf. 84, 155–162.
F. Li et al. / Aquatic Toxicology 158 (2015) 1–13 13

Choi, S.M., Jeong, S.W., Jeong, W.J., Kwon, S.Y., Chow, W.S., Park, Y.I., 2002. Chloroplast Melegari, S.P., Perreault, F., Costa, R.H., Popovic, R., Matias, W.G., 2013. Evaluation of
Cu/Zn-superoxide dismutase is a highly sensitive site in cucumber leaves chilled toxicity and oxidative stress induced by copper oxide nanoparticles in the green
in the light. Planta 216 (2), 315–324. alga Chlamydomonas reinhardtii. Aquat. Toxicol. 142–143, 431–440.
Chong, M.N., Jin, B., Chow, C.W.K., Saint, C., 2010. Recent developments in photocat- Metzler, D.M., Li, M., Erdem, A., Huang, C.P., 2011. Responses of algae to photocat-
alytic water treatment technology: a review. Water Res. 44 (10), 2997–3027. alytic nano-TiO2 particles with an emphasis on the effect of particle size. Chem.
Dastjerdi, R., Montazer, M., 2010. A review on the application of inorganic nano- Eng. J. 170 (2–3), 538–546.
structured materials in the modification of textiles: focus on anti-microbial Morlon, H., Fortin, C., Floriani, M., Adam, C., Garnier-Laplace, J., Boudou, A.B., 2005.
properties. Colloids Surf. B: Biointerfaces 79 (1), 5–18. Toxicity of selenite in the unicellular green alga Chlamydomonas reinhardtii:
Dat, J., Vandenabeele, S., Vranova, E., Van Montagu, M., Inze, D., Van Breusegem, F., comparison between effects at the population and sub-cellular level. Aquat.
2000. Dual action of the active oxygen species during plant stress responses. Toxicol. 73 (1), 65–78.
Cell. Mol. Life Sci. 57 (5), 779–795. Navarro, E., Baun, A., Behra, R., Hartmann, N.B., Filser, J., Miao, A.-J., Quigg,
Deng, X., Wu, F., Liu, Z., Luo, M., Li, L., Ni, Q., Jiao, Z., Wu, M., Liu, Y., 2009. The splenic A., Santschi, P.H., Sigg, L., 2008. Environmental behavior and ecotoxicity of
toxicity of water soluble multi-walled carbon nanotubes in mice. Carbon 47 (6), engineered nanoparticles to algae, plants, and fungi. Ecotoxicology 17 (5),
1421–1428. 372–386.
Edreva, A., 2005. Generation and scavenging of reactive oxygen species in chloro- Nel, A.E., Maedler, L., Velegol, D., Xia, T., Hoek, E.M.V., Somasundaran, P., Klaes-
plasts: a submolecular approach. Agric. Ecosyst. Environ. 106 (2–3), 119–133. sig, F., Castranova, V., Thompson, M., 2009. Understanding biophysicochemical
Forti, G., Gerola, P., 1977. Inhibition of photosynthesis by azide and cyanide and the interactions at the nano–bio interface. Nat. Mater. 8 (7), 543–557.
role of oxygen in photosynthesis. Plant Physiol. 59 (5), 859–862. Pakrashi, S., Dalai, S., Prathna, T.C., Trivedi, S., Myneni, R., Raichur, A.M.,
Gill, S.S., Tuteja, N., 2010. Reactive oxygen species and antioxidant machinery in Chandrasekaran, N., Mukherjee, A., 2013. Cytotoxicity of aluminium oxide
abiotic stress tolerance in crop plants. Plant Physiol. Biochem. 48 (12), 909–930. nanoparticles towards fresh water algal isolate at low exposure concentrations.
Guillard, R.R., Ryther, J.H., 1962. Studies of marine planktonic diatoms. I. Cyclotella Aquat. Toxicol. 132, 34–45.
nana Hustedt, and Detonula confervacea (cleve) Gran. Can. J. Microbiol. 8, Patra, P., Mitra, S., Debnath, N., Goswami, A., 2012. Biochemical-, biophysical-,
229–239. and microarray-based antifungal evaluation of the buffer-mediated synthe-
Guiry, M.D., Guiry, G.M., 2014. AlgaeBase. World-wide Electronic Publication. sized nano zinc oxide: an in vivo and in vitro toxicity study. Langmuir 28 (49),
National University of Ireland, Galway, http://www.algaebase.org; searched on 16966–16978.
(2014). Qiu, J., Ma, F., Fan, H., Li, A., 2013. Effects of feeding Alexandrium tamarense, a para-
Guo, P.Y., Shen, H.T., Wang, J.H., 2014. Species diversity, community structure and lytic shellfish toxin producer, on antioxidant enzymes in scallops (Patinopecten
distribution of phytoplankton in the Changjiang estuary during dry and flood yessoensis) and mussels (Mytilus galloprovincialis). Aquaculture 396–399,
periods. J. Mar. Biol. Assoc. U.K. 94 (3), 459–472. 76–81.
Han, Y.-T., Han, Z.-W., Yu, G.-Y., Wang, Y.-J., Cui, R.-Y., Wang, C.-B., 2004. Inhibitory Reynafarje, B., Brand, M.D., Lehninger, A.L., 1976. Evaluation of the H+/site ratio of
effect of polypeptide from Chlamys farreri on ultraviolet A-induced oxida- mitochondrial electron transport from rate measurements. J. Biol. Chem. 251
tive damage on human skin fibroblasts in vitro. Pharmacol. Res. 49 (3), (23), 7442–7451.
265–274. Rocchetta, I., Mazzuca, M., Conforti, V., Ruiz, L., Balzaretti, V., de Molina, M.D.R.,
Hasle, G.R., Syvertsen, E.E., 1997. Marine Diatoms. Academic Press, Inc., 1250 Sixth 2006. Effect of chromium on the fatty acid composition of two strains of Euglena
Ave., San Diego, California 92101, USA/14 Belgrave Square, 24-28 Oval Road, gracilis. Environ. Pollut. 141 (2), 353–358.
London NW1 70X, England, UK. Rodea-Palomares, I., Gonzalo, S., Santiago-Morales, J., Leganes, F., Garcia-Calvo, E.,
Hong, F.H., Zhou, J., Liu, C., Yang, F., Wu, C., Zheng, L., Yang, P., 2005. Effect of nano- Rosal, R., Fernandez-Pinas, F., 2012. An insight into the mechanisms of nanoce-
TiO2 on photochemical reaction of chloroplasts of spinach. Biol. Trace Elem. Res. ria toxicity in aquatic photosynthetic organisms. Aquat. Toxicol. 122–123,
105 (1–3), 269–279. 133–143.
Hu, C., Liu, X., Li, X., Zhao, Y., 2014. Evaluation of growth and biochemical indicators Sadiq, I.M., Dalai, S., Chandrasekaran, N., Mukherjee, A., 2011. Ecotoxicity study of
of Salvinia natans exposed to zinc oxide nanoparticles and zinc accumulation in titania (TiO(2)) NPs on two microalgae species: Scenedesmus sp. and Chlorella
plants. Environ. Sci. Pollut. Res. Int. 21 (1), 732–739. sp. Ecotoxicol. Environ. Saf. 74 (5), 1180–1187.
Huang, C.P., Cha, D.K., Ismat, S.S., 2005. Progress Report: Short-Term Chronic Toxicity San Mateo, L.R., Toffer, K.L., Kawula, T.H., 1998. The sodA gene of Haemophilus ducreyi
of Photocatalytic Nanoparticles to Bacteria, Algae, and Zooplankton. University encodes a hydrogen peroxide-inhibitable superoxide dismutase. Gene 207 (2),
of Delaware. 251–257.
Hund-Rinke, K., Simon, M., 2006. Ecotoxic effect of photocatalytic active nanoparti- Schwab, F., Bucheli, T.D., Lukhele, L.P., Magrez, A., Nowack, B., Sigg, L., Knauer, K.,
cles TiO2 on algae and daphnids. Environ. Sci. Pollut. Res. 13 (4), 225–232. 2011. Are carbon nanotube effects on green algae caused by shading and agglom-
Ji, J., Long, Z., Lin, D., 2011. Toxicity of oxide nanoparticles to the green algae Chlorella eration? Environ. Sci. Technol. 45 (14), 6136–6144.
sp. Chem. Eng. J. 170 (2–3), 525–530. Shipway, A.N., Katz, E., Willner, I., 2000. Nanoparticle arrays on surfaces for elec-
Kus, M., Gernjak, W., Fernandez Ibanez, P., Malato Rodriguez, S., Blanco Galvez, J., tronic, optical, and sensor applications. Chemphyschem 1 (1), 18–52.
Icli, S., 2006. A comparative study of supported TiO2 as photocatalyst in water Simon, D., Helliwell, S., 1998. Extraction and quantification of chlorophyll A from
decontamination at solar pilot plant scale. J. Solar Energy Eng. – Trans. Asme 128 freshwater green algae. Water Res. 32 (7), 2220–2223.
(3), 331–337. Song, G., Gao, Y., Wu, H., Hou, W., Zhang, C., Ma, H., 2012. Physiological effect of
Laisk, A., Oja, V., 2013. Thermal phase and excitonic connectivity in fluorescence anatase TiO2 nanoparticles on Lemna minor. Environ. Toxicol. Chem. 31 (9),
induction. Photosynth. Res. 117 (1–3), 431–448. 2147–2152.
Li, N.Y., Ragheb, K., Lawler, G., Sturgist, J., Rajwa, B., Melendez, J.A., Robinson, J.P., Suh, W.H., Suslick, K.S., Stucky, G.D., Suh, Y.-H., 2009. Nanotechnology, nanotoxicol-
2003. Mitochondrial complex I inhibitor rotenone induces apoptosis through ogy, and neuroscience. Prog. Neurobiol. 87 (3), 133–170.
enhancing mitochondrial reactive oxygen species production. J. Biol. Chem. 278 Tong, H., Ouyang, S., Bi, Y., Umezawa, N., Oshikiri, M., Ye, J., 2012. Nano-
(10), 8516–8525. photocatalytic materials: possibilities and challenges. Adv. Mater. 24 (2),
Li, J., Xu, Z., Tan, M., Su, W., Gong, X.-g., 2010. 3-(4-(Benzo d thiazol-2-yl)-1-phenyl- 229–251.
1H-pyrazol-3-yl) phenyl acetate induced Hep G2 cell apoptosis through a ROS- Tripathy, B.C., Mohanty, P., 1980. Zinc-inhibited electron transport of photosynthesis
mediated pathway. Chem. Biol. Interact. 183 (3), 341–348. in isolated barley chloroplasts. Plant Physiol. 66 (6), 1174–1178.
Lin, D., Xing, B., 2008. Root uptake and phytotoxicity of ZnO nanoparticles. Environ. Wang, J., Zhang, X., Chen, Y., Sommerfeld, M., Hu, Q., 2008. Toxicity assessment of
Sci. Technol. 42 (15), 5580–5585. manufactured nanomaterials using the unicellular green alga Chlamydomonas
Lin, D., Ji, J., Long, Z., Yang, K., Wu, F., 2012. The influence of dissolved and surface- reinhardtii. Chemosphere 73 (7), 1121–1128.
bound humic acid on the toxicity of TiO2 nanoparticles to Chlorella sp. Water Wang, Z., Li, J., Zhao, J., Xing, B., 2011. Toxicity and internalization of CuO nanopar-
Res. 46 (14), 4477–4487. ticles to prokaryotic alga Microcystis aeruginosa as affected by dissolved organic
Lovern, S.B., Strickler, J.R., Klaper, R., 2007. Behavioral and physiological changes matter. Environ. Sci. Technol. 45 (14), 6032–6040.
in Daphnia magna when exposed to nanoparticle suspensions (titanium Wong, S.W.Y., Leung, P.T.Y., Djurisic, A.B., Leung, K.M.Y., 2010. Toxicities of nano zinc
dioxide, nano-C-60, and C(60)HxC(70)Hx). Environ. Sci. Technol. 41 (12), oxide to five marine organisms: influences of aggregate size and ion solubility.
4465–4470. Anal. Bioanal. Chem. 396 (2), 609–618.
Ma, H., Williams, P.L., Diamond, S.A., 2013. Ecotoxicity of manufactured ZnO Yu, Y., Kong, F., Wang, M., Qian, L., Shi, X., 2007. Determination of short-term copper
nanoparticles – a review. Environ. Pollut. 172, 76–85. toxicity in a multispecies microalgal population using flow cytometry. Ecotoxi-
Manier, N., Bado-Nilles, A., Delalain, P., Aguerre-Chariol, O., Pandard, P., 2013. col. Environ. Saf. 66 (1), 49–56.
Ecotoxicity of non-aged and aged CeO2 nanomaterials towards freshwater Zhang, W.X., 2003. Nanoscale iron particles for environmental remediation: an
microalgae. Environ. Pollut. 180, 63–70. overview. J. Nanopart. Res. 5 (3–4), 323–332.

You might also like