+ All documents
Home > Documents > Changes in agglomeration of fullerenes during ingestion and excretion in Thamnocephalus platyurus

Changes in agglomeration of fullerenes during ingestion and excretion in Thamnocephalus platyurus

Date post: 14-Nov-2023
Category:
Upload: independent
View: 1 times
Download: 0 times
Share this document with a friend
31
Ac c e p te d P r e p r i n t 1 Environmental Chemistry DOI 10.1002/etc.468 CHANGES IN AGGLOMERATION OF FULLERENES DURING INGESTION AND EXCRETION IN THAMNOCEPHALUS PLATYURUS Manomita Patra , Xin Ma , Carl Isaacson , Dermont Bouchard , Helen Poynton § , James M. Lazorchak § , Kim R. Rogers †* U.S. Environmental Protection Agency, Las Vegas, NV U.S. Environmental Protection Agency, Athens, GA § U.S. Environmental Protection Agency, Cincinnati, OH Running Head: Agglomeration of Ingested Fullerenes Corresponding Author: K.R.Rogers U.S. Environmental Protection Agency 944 E. Harmon Ave. Las Vegas, NV 89119 Tel: (702) 798-2299 [email protected] To whom correspondence should be addressed ([email protected]) © 2011 SETAC Submitted 6 January 2010; Returned for Revision 1 November 2010; Accepted 15 November 2010 Environmental Toxicology and Chemistry
Transcript

AcceptedPrepri n

t

1

Environmental Chemistry DOI 10.1002/etc.468

CHANGES IN AGGLOMERATION OF FULLERENES DURING INGESTION AND

EXCRETION IN THAMNOCEPHALUS PLATYURUS

Manomita Patra†, Xin Ma

‡, Carl Isaacson

‡, Dermont Bouchard

‡,

Helen Poynton§, James M. Lazorchak

§, Kim R. Rogers

†*

†U.S. Environmental Protection Agency, Las Vegas, NV

‡U.S. Environmental Protection Agency, Athens, GA

§U.S. Environmental Protection Agency, Cincinnati, OH

Running Head: Agglomeration of Ingested Fullerenes

Corresponding Author:

K.R.Rogers

U.S. Environmental Protection Agency

944 E. Harmon Ave.

Las Vegas, NV 89119

Tel: (702) 798-2299

[email protected]

To whom correspondence should be addressed ([email protected])

© 2011 SETAC Submitted 6 January 2010; Returned for Revision 1 November 2010; Accepted 15 November 2010

Environmental Toxicology and Chemistry

AcceptedPrepri n

t

2

Abstract: The crustacean Thamnocephalus platyurus was exposed to aqueous suspensions of

fullerenes C60 and C70. Aqueous fullerene suspensions were formed by stirring C60 and C70 as

received from a commercial vendor in deionized water (termed aqu/C60 and aqu/C70) for

approximately 100 d. The Z-average diameters of aqu/C60 and aqu/C70 aggregates as measured

by dynamic light scattering were 517 ± 21 nm and 656 ± 39 nm (mean ± 95% confidence limit),

respectively. Exposure of T. platyurus to fullerene suspensions resulted in the formation of dark

masses in the digestive track visible under a stereo microscope (40 x magnification). Fullerene

ingestion, over one hour exposure was quantitatively determined after extraction and analysis by

HPLC-MS. One hour exposures (at 3 mg/L and 6 mg/L) resulted in aqu/C60 burdens of 2.7 ± 0.4

µg/mg and 6.8 ± 1.5 µg/mg wet weight, respectively. Thin section TEM images of aqu/C60

exposed T. platyurus showed the formation in the gut of fullerene agglomerates (5 to 10 µm),

which were an order of magnitude larger in size than the suspended fullerene agglomerates.

Upon excretion, the observed fullerene agglomerates were in the 10 to 70 µm size range and

settled to the bottom of the incubation wells. In contrast to the control polystyrene microspheres,

which dispersed after depuration, the aqu/C60 agglomerates (greater than two orders of

magnitude larger than the suspended fullerenes) remained agglomerated for up to six months.

When exposed to fullerenes, crustacean T. platyurus shows the potential to influence the

agglomerate size and may facilitate movement of these nanoparticles from the water column into

the sediment.

Key words: Nanomaterials, Fullerenes, C60, C70, Thamnocephalus platyurus.

AcceptedPrepri n

t

3

INTRODUCTION

Nanotechnology is a rapidly growing industrial enterprise encompassing a diverse array

of engineered nanomaterials, developed and applied in fields such as medicine, plastics, energy,

electronics, and cosmetics. The characteristics of the nanomaterials, as well as concerns

involving nanoparticle transport, fate, exposure, dose and potential effects on ecosystems, have

been reviewed in the literature [1]. Because of the potential for nanomaterials to contaminate

various environmental settings (e.g., water soils, etc.) and contribute to ecosystem and human

exposure and risk, the United States Environmental Protection Agency (U.S. EPA) views these

materials as potential and emerging contaminants.

Fullerenes as a class of carbon nanomaterials are used in a wide range of commercial

products and are expected to be one of the major contributors in future product development

[2 (http://www.wilsoncenter.org/index.cfm?topic_id=166192&fuseaction=topics.home)]. Many

environmental studies with fullerenes to date have used solvent exchange techniques for creating

fullerene suspensions of C60 in water [3-5]. For these studies, fullerenes were dissolved in an

organic solvent, usually toluene or tetrahydrofuran (THF), the solution was dispersed in water,

and the organic solvent removed by distillation. Another approach to suspending fullerenes in

water is simply to stir them for extended time periods (> 100 days). Both stirring in water and

organic solvent-assisted techniques result in the formation of kinetically stable suspensions [6,7];

however, colloidal fullerene aggregates formed using solvent exchange and those formed in

water-only systems differ with respect to size, shape, charge, polydispersity and also in the toxic

effects they elicit [6,8-10]. The remaining THF in C60 suspensions has been found to contribute

to the toxicity effect for microorganisms such as Bacilus subtilis [11] and to the toxicity of a

mouse macrophage cell line [12]. Extended stirring in water without the aid of organic solvents

AcceptedPrepri n

t

4

was used in the present study to create colloidal suspensions of aqu/C60 and aqu/C70 that

contained fewer preparation artifacts and that are more representative of natural environmental

systems [7,13].

Aquatic environments may be particularly vulnerable to contamination of nanomaterials,

such as fullerenes, that can form colloidal suspensions. Factors that contribute to the formation of

these suspensions include the relative size of fullerene agglomerates as well as their morphology

and surface zeta potentials which are a function of their media environments as well as how they

were prepared in suspension [7,13,14]. In addition, the transport behavior and ecotoxicity of

fullerenes in aqueous environments also significantly depend on the properties of these

agglomerates [15,16]. Other important factors that limit the understanding of potential

ecosystem impacts include uncertainties concerning ingestion, bioavailability and

bioaccumulation of these materials in indicator species such as crustacean zooplankton.

Planktonic invertebrates are particularly important with respect to contamination of the

aquatic environment because they are the bridge in the food chain between pollutants bound to

suspended particulates, algae and fish [17,18]. Organisms such as Daphnia magna,

Thamnocephalus platyurus, and Brachiomus clyciflorus have been widely used as indicator

species for both ecosystems and human health [19]. Microcrustaceans have also been shown to

ingest and accumulate of carbon nanoparticles [20]. Exposure of D. magna to fullerene C60 and

carbon nanotubes have resulted in the presence of dark masses in the gut of these organisms

[17,21]. The use of radiolabelled nanotube agglomerates has also allowed the observation of

uptake and elimination kinetics under specific conditions [22].

A rapid screening assay that is sensitive to a wide range of pollutants in environmental

matrices has been reported using T. Platyurus. This assay is based on the accumulation of inert

AcceptedPrepri n

t

5

polystyrene microspheres which can be observed in these transparent organisms. Compounds

that induce a toxic stress response inhibit the accumulation of the indicator spheres [23,24].

Based on this microsphere accumulation behavior, this organism was chosen for the present

study.

Although C60 fullerenes have been observed to accumulate in the gut cavity of micro-

crustaceans such as D. magna, with the exception of Tervonen et al., changes in agglomeration

prior to and during ingestion and after excretion have not been well documented [20]. The

objective of the present study is to characterize the uptake of aqu/C60 and aqu/C70 fullerenes into

T. platyurus and measure changes in agglomeration that may occur during ingestion and

excretion of these materials. The implications these processes may have on the transport and fate

of these materials are also discussed.

MATERIALS AND METHODS

Particle characterization

Preparation of fullerene suspensions. The C60 (purity 99.9%) and C70 (purity 99.0%)

fullerenes were purchased from MER Corp. (Tucson, AZ, USA). Rapidtoxkits™ ( the source for

T. platyurus cysts) were purchased from Strategic Diagnostics (Newark, Delaware, USA). All

other chemicals were of reagent grade and purchased from Sigma-Aldrich (St. Louis, MO,

USA). For fullerene suspensions used for T. platyurus exposure, 100 mg of C60 or C70 were

added to 400 ml double de-ionized water (DDI) (>18 M Ohms-cm). The suspensions were stirred

using a magnetic stir plate for approximately 100 d to form nanometer-sized fullerene aggregates

that have been shown to be stable with respect to size and charge [7,13]. For particle size

AcceptedPrepri n

t

6

characterization prior to the exposure experiments, the suspensions were allowed to settle by

gravity for 1 h and the suspensions were sampled at 2-cm below the surface.

Mass quantitation. The aqu/C60 and aqu/C70 were quantified using a combination of

previously published methods [13,25,26]. For mass quantification, aqu/C60 and aqu/C70 were

dissolved into a toluene/methanol mixture (8:2 by volume) then quantified directly using a 13

C60

internal standard with liquid chromatography-atmospheric pressure photoionization mass

spectrometry. The fullerenes were separated with a 95% Toluene : 5% Methanol mobile phase

on a Accela liquid chromatograph (Thermo Fisher, West Palm Beach, FL) with a Cosmosil

column (4 x 150 mm) connected inline with a guard column (3 x 4 mm), both from Phenomenex

(Torrance, CA). Detection was provided by a Quantum Ultra Mass Spectrometer operating in

negative mode with a 10 eV krypton atmospheric pressure photoioization lamp. The mass

spectrometer was set to select molecular ions for C60 (m/z 720) C70 (m/z 840) and 20-30% 13

C

enriched C60 (m/z 734) as no fragment ions were detected.

Characterization of fullerenes by dynamic light scattering (DLS). Changes in fullerene

aggregate size were examined by DLS (ZetaSizer Nano ZS) [13]. Six measurements (12 runs

per measurement) were acquired from each sample. Instrument performance was verified using

National Institute of Standards and Technology (NIST)-traceable polystyrene microsphere

standards. The autocorrelation function was analyzed by the cumulant method to obtain the

moments of the aqueous fullerene aggregate size distribution. The fluctuations of scattered light

from particles can be mathematically correlated to the diffusion coefficient. The intensity

average (Z-average) hydrodynamic diameter was calculated from measured diffusivities using

the Stokes-Einstein equation. For surface charge characterization, zeta ( ) potentials of the

suspensions were monitored using a ZetaSizer Nano ZS instrument (Malvern Instruments,

AcceptedPrepri n

t

7

Worcestershire, UK). This instrument uses phase analysis light scattering (PALS) to measure

the electrophoretic mobility of charged particles. The Smoluchowski equation was used to

calculate -potential from electrophoretic mobility. For electron microscopic analysis, the

fullerenes were analyzed from aqueous suspensions. Fullerene suspensions (120 mg/L) were

placed drop-deposited onto a 200 mesh Fomvar-coated copper grids. Images were acquired at

300 kV using a Tecnai transmission electron microscope (TEM) (FEI, Hillboro, OR, USA) with

a 2k Gatan camera [27]. The imaged aggregates were also analyzed by energy dispersive X-ray

spectroscopy for elemental composition using an EDEX detector. Thin section images were

acquired using an FEI/Philips CM-10 TEM at 300 kV.

TEM and chemical analysis of the gut contents. For thin section TEM analysis, after

exposure to aqu/C60, the T. platyurus were fixed in 2% glutaraldehyde then exchanged into 1%

OsO4 in 0.1 M cacodylate buffer pH 7.4. The organisms were then dehydrated by stepwise

ethanol exchange (10, 30, 50, 90, 100% ethanol) then further exchanged from propylene oxide

into bisphenol A-epichlorhydrin (EPON) 828 resin. The resin was then polymerized for 48 hr at

60º C prior to thin sectioning and TEM analysis.

Chemical analysis of gut contents. After lyophilization, 50 of the exposed organisms

were placed in the extraction vessel. A 1-ml volume of toluene was then added to the extraction

vessel, the vessel was vortexed several seconds and then placed horizontally on an orbital shaker

at 200 rpm for 30 min. After removal from the shaker, the sample phases were allowed to

separate for 15 min before approximately 800-µlwas removed from the toluene supernatant layer

for analysis by HPLC-UV. The samples were then placed in a freezer at 25 C for 2 h to freeze

the aqueous layer and thus facilitate decantation of the remaining toluene layer. Following

decantation of the toluene layer, the aqueous layer was allowed to melt and a second 1-ml

AcceptedPrepri n

t

8

volume of toluene was added to the extraction vessel. The aqueous phase was then re-extracted

as described above. A total of three sequential extractions were performed using this technique

and the fullerene mass extracted in each sequential extraction was determined by HPLC [13].

Using this technique, the recovery and 95% confidence interval for the T. Platyurus extraction

was 57.3 ± 12.1%.

Thamnocephalus platyurus assays.

The relative rate of ingestion of colored polystyrene microspheres (5 µm) in T. platyurus

has been used as the basis for a rapid toxicity screening assay [18]. Due to this uptake behavior

as well as the visibility of these microspheres in the gut cavity of T. platyurus, these

microspheres were used as a comparative control for exposure of this organism to aqu/C60

fullerenes (which can be visibly differentiated by optical microscopy).

Exposure conditions. Thamnocephalus platyurus were hatched from cysts following the

Rapidtox™ protocol. The cysts were hydrated with moderately hard reconstituted water (CaSO4,

0.06 g/L; MgSO4, 0.06 g/L; NaHCO3, 0.096 g/L; KCl, 0.004 g/L) and incubated at 25º C under

4000 lux illumination for 30 to 45 h. After hatching, the larval suspensions were transferred into

double-deionized (DDI) water and mixed (1:1) with gently vortexed fullerene suspensions and

incubated for 1 hr prior to addition of the colored indicator microspheres (5 µm red polystyrene)

into the fullerene-organism mixtures. The organisms were exposed to the microspheres for 15

min prior to fixation with Lugol solution (final concentration, 1% KI, 0.5% I2) and final

observation.

Visual quantitation of fullerene uptake. For the fullerene visual uptake quantitation

experiment, T. platyurus were exposed to fullerenes at concentrations of 1.9, 3.7, 7.5, 30, and 60

mg/L for one h prior to fixation with Lugol solution and quantitative observation. Organisms

AcceptedPrepri n

t

9

with greater than 20% of their gut occupied by dark matter were counted and the percent of total

(number) of organisms accumulating fullerenes determined. For the algae accumulation

experiment, algae (Selenastrum capriconutum at 100,000 cells/ml) were added after 1 h exposure

to the fullerenes (60 mg/l). Algae stocks were maintained as previously described [28]. Fifteen

min after addition of algae, the organisms were fixed with Lugol Solution for observation. For all

visible microscopy observations, organisms were observed using a stereo microscope at 40 x

magnification and images recorded using Motic Image Plus 2.0 software.

For the experiments involving imaging of gut content by thin section TEM, T. platyurus

were exposed to aqu/C60 at 69 mg/L for 1 h prior to separation, preservation and analysis. For

chemical analysis of gut contents, the organisms were washed 3 times with DDI water by

centrifugation (3000xg, 3 min). After centrifugation, the supernatant was removed each time and

the pelleted organisms re-suspended in DDI water using a pasture pipette. For the sample blank,

aqu/C60 was added after the control organisms were fixed to account for the non-ingested

particles that were adhered to the surface of the organisms or carried over in the washes. The

pelleted organisms were then placed onto a pre-weighed weigh boat and the excess liquid

removed using the corner of a small paper towel. The organisms were then separated into groups,

frozen, lyophilized.

RESULTS AND DISCUSSION

Aqu/C60 and Aqu/C70 fullerene preparations.

The Z-average diameters for the aqu/C60 and aqu/C70 aggregates prior to feeding showed

size distribution averages of 517 ± 21 nm and 656 ± 39 nm (mean ± 95% confidence limits),

AcceptedPrepri n

t

10

respectively. Similar to the results by DLS, analysis of the vortexed aqu/C60 suspensions by TEM

showed aggregates that ranged from about 200 to 500 nm in length (Fig. 1A, B). Zeta ( )

potentials for the aqu/C60 and aqu/C70 suspensions were -27.5 ± 0.7 mV and -47.5 ± 5.4 mV,

respectively. Generally, colloidal suspensions of particles with absolute zeta potentials greater

than 30 mV will be stable and will not further aggregate or precipitate as long as background

solution parameters remain constant [29]. In the present study, the background solution ionic

environment did not change as all experiments were conducted in DDI. While the aqu/C70 zeta

potential was significantly more negative than -30 mV, the aqu/C60 zeta potential was

significantly higher, approaching the rule of thumb (-30mV) value.

Exposure of T. platyurus to fullerene suspensions.

The experimental design primarily involved exposure of T. platyurus to aqueous

suspensions of fullerenes (aqu/C60 and aqu/C70) and combinations of fullerenes with

microspheres or algae. In the latter case, the organisms were exposed to fullerenes prior to the

microspheres or algae to determine whether or not the accumulation of the smaller fullerene

agglomerates could exclude the larger indicator beads or algae (typical food supply). The relative

size distribution of the fullerene agglomerates were monitored prior to, during ingestion and after

excretion.

Because T. platyurus are semi-transparent organisms, exposure and accumulation of the

red indicator beads could be visualized as a red opaque outline of the gut cavity (Fig. 2A). When

T. platyurus were exposed to aqu/C60 or aqu/C70 fullerenes prior to exposure to the indicator

beads, they accumulated the fullerenes as dark masses in their digestive tract (Fig. 2B). The

fullerenes appeared to compete with the indicator microspheres with respect to accumulation in

AcceptedPrepri n

t

11

the gut. Although some of the organisms’ digestive tracks contained only red microspheres or

fullerenes, many of them showed segments of each (Fig. 2B). After exposure of T. platyurus to

aqu/C60 or aqu/C70 with no addition of indicator microspheres, the dark masses visible in the

digestive tract also appeared to be segmented (with intervening light areas) indicative of the

formation of excretion pellets (Fig. 2C).

The observed percentage of organisms accumulating the dark masses (in 20% of the gut

tract length) was concentration dependent for both aqu/C60 and aqu/C70 fullerenes (Fig. 3). After

a 1 h exposure, about 50% of the organisms accumulated observable fullerene masses in their gut

at an initial aqu/C60 concentration of 7.5 mg/L. At an initial concentration of 30 mg/L, about

80% of the organisms had accumulated these dark masses (Fig. 3). Visual observation did not

reveal whether or not the accumulated fullerenes were coating the inside of the digestive track or

filling the lumen. However, these observations may suggest that the fullerenes coalesced in

specific areas and forming the observed banding pattern.

When the T. platyurus were exposed to fullerenes followed by unicellular algae, the

ingested algae could also be observed as green bands between the dark masses of aqu/C60 or

aqu/C70 fullerenes, similar to the results observed with the red indicator spheres. Contrast

between the green and black bands could not be observed well in the photographs and these data

are not shown.

Quantitative Assessment of Uptake

To further confirm the uptake of fullerenes, the accumulation of aqu/C60 in T. platyurus

was quantified by solvent extraction and analysis by HPLC-MS. The fullerene agglomerate

burden was measured for two initial fullerene concentrations (3 mg/L and 6 mg/L) over a 60 min

time course (Fig 4). The uptake of aqu/C60 between 0 and 10 min was rapid for both initial

AcceptedPrepri n

t

12

exposure concentrations. However, after the initial uptake period, the C60 concentration per unit

mass of organism did not change significantly over the remaining 60 min measurement period.

One hour exposures (at 3mg/L and 6 mg/L) resulted in aqu/C60 burdens of 2.7 ± 0.4

µg/mg and 6.8 ± 1.5 µg/mg wet weight, respectively. The amount of aqu/C60 that bound to the

organisms after they had been fixed with Lugol solution was about 15% of the total uptake. A

comparison of the amount of aqu/C60 per organism for the two exposure levels after 60 min

differed by a factor of 2.5. This level of increase for the quantitative extraction and measurement

of accumulated aqu/C60 averaged for a group of 50 organisms was similar to results shown in ` 3

where an increase in exposure level from 3.7 mg/L to 7.5 mg/L resulted in a 2.7 times increase in

the percentage of organisms observed to accumulate dark masses (in 20%) of gut length. This

result suggests that the fullerene uptake behavior for individual organisms was similar. The

herein reported uptake values for aqu/C60 into T. platyurus were similar to the value of 4.5

µg/mg wet weight reported for the uptake of C60 into D. magna after a 24 h exposure at a

fullerene concentration of 2 mg/l [20].

Structure and composition of dark masses in the gut.

Electron micrographs of the transverse cross-sections of T. platyurus exposed to aqu/C60 and

polystyrene beads show the microvilli-lined lumen of the mid-gut containing aqu/C60

agglomerates (Fig. 5 A, B). Compared to the aqu/C60 agglomerates used for the exposure

experiments (200-500 nm by TEM), these aqu/C60 agglomerates appeared to be significantly

larger (5 to 10 µm) and composite in structure. These larger clusters observed in the gut were

presumably assembled from smaller agglomerates that were ingested by the organisms. The

micrograph in Figure 5B shows one of these agglomerates (a) next to one of the polystyrene

AcceptedPrepri n

t

13

microspheres (b). Part of this agglomerate appears to have coalesced into a larger structure. The

dense fullerene agglomerates did not appear to cross the gut lining in any of the numerous thin

section electron micrographs that we analyzed (not shown).

Exposure of T. platyurus to aqu/C60 for several hours resulted in the formation of pellets

which were excreted and settled to the bottom of the incubation wells. Analysis of this material

by light microscopy showed a population of agglomerates ranging in size from about 10 µm to

70 µm (Fig. 6A). Observation of the agglomerates from the excreted material by TEM also

showed the presence of micrometer sized agglomerates of aqu/C60 (Fig. 6B). These larger

aqu/C60 agglomerates from the excretion pellets appeared to be composed of smaller

agglomerates in the 200 to 500 nm size range that had coalesced into larger structures. The

excreted C60 agglomerates were stored in DDI water at room temperature and did not appear to

disperse even after 6 months. When the indicator microspheres were added in the absence of

fullerenes, they accumulated in the gut as evidenced by the red color in the gut track (see Fig.

2A). After exposure of the indicator spheres for several hours, there did not appear to be any

aggregates of these particles. Although unmodified polystyrene microspheres (5 µm) will settle

from suspension after several days, they tend to remain suspended and monodispersed for several

hours. Differences in the post-excretion dispersion behavior of aqu/C60 and polystyrene

microspheres may be due, in part, to their surface charge. It has been suggested that, in general,

particles with an absolute zeta potential of greater than 30 mV will form stable suspensions due

to their strong electrostatic repulsion [7]. Unmodified polystyrene microspheres have a zeta

potential of -70.8 mV [30], whereas the herein reported zeta potential for the aqu/C60 is -27.5

mV. Although the aqu/C60 used as the source material for the current exposure measurements

remained in suspension (without stirring) for six months after the experiment, it may be the case

AcceptedPrepri n

t

14

that the forced particle proximity resulting from the confined space of the gut resulted in the

observed coalescing of the aqu/C60 particles which showed a zeta potential less than typically

indicated for stable suspensions.

Table 1 shows size distributions visible microscopy and electron microscopy as well as

hydrodynamic size distributions by DLS before and after ingestion by T. platyurus. The relative

differences in size between the DLS and TEM measurements for the aqu/C60 and aqu/C70

suspensions prior to ingestion are likely to be due to inherent differences in these techniques.

More specifically, DLS measures a signal intensity-weighted hydrodynamic diameter

distribution, where as TEM measures a number-weighted population size distribution. Because

the larger particles tend to dominate the population size distribution measured by DLS, we

believe that the measurements by these two techniques are consistent. Negative potentials are

reported for the aqu/C60 and aqu/C70 and similar values have been previously observed for these

aqueous fullerene suspensions [7].

Potential ecological implications

Although aquatic crustacean filter-feeders can feed on a range of biological particulate

materials, they typically ingest unicellular algae in the micrometer size range with a preferred

size for food particles of >1 µm. In studies using micro spheres, D. cucullae have been shown to

accumulate polystyrene beads in the 5 to 35 µm size range [31]. Fairy shrimp (anostracans) have

been grown in the laboratory on diets ranging from bacteria to baker’s yeast (1 to 10 µm) [32].

For each of these examples, the sizes of food sources are typically greater than 1 µm. The

observation that zooplankton such as D. magna can accumulate relatively small polystyrene

spheres (50 nm) which were expected to pass through the filter combs has lead researchers to

propose mechanisms of accumulation of very fine particles based on particle adhesion as well as

AcceptedPrepri n

t

15

physical filtration [33]. We observed that in addition to the 5 µm indicator spheres, T. platyurus

accumulated the smaller C60 agglomerates suggesting that the fullerenes may also have, to some

extent, adhered to the organism’s filtering apparatus through particle adhesion mechanisms.

The basis of the T. platyurus feeding inhibition assay involves the organism’s sensing the

presence of toxins and subsequently slowing or stopping its feeding behavior [34]. Nevertheless,

T. platyurus appeared to ingest aqu/C60 to a similar extent as the indicator microspheres (see Fig

2B) suggesting that the fullerene agglomerates were mistaken as food [18]. Our observations are

consistent with reports that D. magna accumulates C60 agglomerates which are not acutely toxic

even at relatively high concentrations (5 mg/L) [17]. It has further been suggested, however, that

even though ingestion of nanoparticles may not be acutely toxic to aquatic filter feeders, the

depletion of energy reserves (due to nanoparticle competition) during periods of low food

supply, may be detrimental to these populations [35].

Filter-feeding organisms such as T. platyurus and D. magna play important roles in the

ecosystem due to their rapid filtration rate and have been shown to have significant impacts on

water turbidity and algal composition [36]. Filella et al. recently illustrated the potential for

Daphnia sp. to impact the size distribution of inorganic colloids, in fresh water lakes [37]. The

tendency for D. magna to filter and excrete carbon-based nanomaterials in an agglomerated state,

potentially change the size distribution and fate of these nanoparticles in the environment has

been suggested by Tervonen et al. [20]. We suggest that anostracans such as T. platyurus may

also show this potential to alter the size distribution of fullerene agglomerates.

Crustacean filter feeders have been shown to accumulate smaller fullerene agglomerates

from relatively low concentration suspensions. For example, D. magna have been shown to

ingest and accumulate agglomerates of C60 at sizes below 0.45 µm [17]. D. magna have also

AcceptedPrepri n

t

16

been shown to ingest and accumulate carbon nanotubes with an average length of < 200 nm [38].

These authors also reported that the nanotubes aggregated in the gut and were not efficiently

eliminated without continuous exposure to dispersed nanotubes or when additional algae were

available in the media. In another experiment reported by Roberts et al. [21], lipid-coated

nanotubes were ingested by D. magna and after accumulation in the gut, the nanotubes were

reported to have been excreted by the organisms as aggregated material that was stripped of the

original lipid coating that allowed the nanotubes to remain dispersed in suspension [38].

CONCLUSION

When T. platyurus are exposed to polystyrene indicator spheres, the spheres are initially

concentrated in the gut then disperse after passing through the organism. However, the

aqu/fullerenes form larger agglomerates during the process of ingestion and excretion. The

aqu/C60 agglomerates used for exposure were in the 200 nm to 500 nm size range. These

agglomerates remained in suspension prior to ingestion by T. platyurus. In the mid-gut, these

agglomerates appeared to coalesce into larger agglomerates that were in the 5 to 10 µm size

range. Upon excretion, the pellets appeared (by light microscopy) to form agglomerates in the 10

to 70 µm size range. These agglomerates were large enough to settle to the bottom of the flask

and remained stable for up to six months without stirring.

The present study illustrates that anostacans such as T. platyurus show the potential to

influence aqu/C60 fullerene agglomerate size. As a result, these organisms may facilitate the

movement of fullerene agglomerates from the water column and into the sediment potentially

shifting exposure to benthic organisms in the sediment.

AcceptedPrepri n

t

17

Acknowledgement - We thank Dr. Marisol Sepulveda and Debby Sherman (Purdue University).

Manomita Patra gratefully acknowledges a National Research Council Research Associateship

Award at the National Exposure Research Laboratory, Human Exposure and Atmospheric

Sciences Division, Las Vegas, Nevada.

The United States Environmental Protection Agency (EPA), through its Office of

Research and Development (ORD), has funded and managed the research described here. It has

been subjected to the Agency’s administrative review and has been approved for publication.

Mention of trade names or commercial products does not constitute endorsement or

recommendation for use.

AcceptedPrepri n

t

18

REFERNCES

1. Klaine SJ, Alvarez PJJ, Batley GE, Fernandes TF, Handy RD, Lyon DY, Mahendra S,

Lead JR. 2008. Nanomaterials in the environment: Behavior, fate, bioavailability, and

effects. Environ Toxicol Chem 27: 1825-1851.

2. Woodrow Wilson International Center for Scholars. 2009. Project on Emerging

Nanotechnologies. Washington, DC, USA.

3. Fortner JD, Lyon DY, Sayes CM, Boyd AM, Falkner JC, Hotze EM, Alemany LB, Tao

YJ, Guo W, Ausman KD, Colvin VL, Hughes JB. 2005. C-60 in water: Nanocrystal

formation and microbial response. Environ Sci Technol 39: 4307-4316.

4. Deguchi S, Alargova RG, Tsujii K. 2001. Stable dispersions of fullerenes, C60 and C70, in

water: Preparation and characterization. Langmuir 17: 6013-6017.

5. Li Q, Xie B, Hwang YS, Xu YJ. 2009. Kinetics of C60 fullerene dispersion in

water enhanced by natural organic matter and sunlight. Environ Sci Technol 43: 3574–

3579.

6. Brant J, Lecoanet H, Hotze M, Wiesner M. 2005. Comparison of electrokinetic properties

of colloidal fullerenes (n-C60) formed using two procedures. Environ Sci Technol 39:

6343-6351.

7. Ma X, Bouchard D. 2009. Formation of aqueous suspensions of fullerenes. Environ Sci

Technol 43: 330-336.

8. Duncan LK, Jinschek JR, Vikesland PJ. 2008. C60 Colloid Formation in Aqueous

Systems: Effects of Preparation Method on Size, Structure, and Surface Charge. Environ

Sci Technol 42: 173-178.

AcceptedPrepri n

t

19

9. Brant JA, Labille J, Bottero JY, Wiesner MR. 2006. Characterizing the Impact of

Preparation Method on Fullerene Cluster Structure and Chemistry. Langmuir 22: 3878-

3885.

10. Zhu S, Oberdorster E, Haasch ML. 2006. Toxicity of an engineered nanoparticle

(fullerene, C60) in two aquatic species, Daphnia and fathead minnow. Marine Environ

Res 62: S5–S9.

11. Lyon DY, Adams LK, Falkner JC, Alvarez PJJ. 2006. Antibacterial activity of fullerene

water suspensions: effects of preparation method and particle size. Environ Sci Technol

40: 4360-43-66.

12. Kovochich M, Espinasse B, Auffan M, Hotze EM, Wessel L, Xia T, Nel AE,

Wiesner MR. 2009. Comparative toxicity of C60 aggregates toward mammalian cells:

role of tetrahydrofuran (THF) decomposition. Environ Sci Technol 43: 6378–6384.

13. Bouchard D, Ma X. 2008. Extraction and high performance liquid chromatographic

analysis of C60, C70 and [6,6]-phenyl C61-butyric acid methyl ester in synthetic and

natural waters. J Chromatog A 120: 153-159.

14. Xie B, Xu Z, Guo W, Li Q. 2008. Impact of natural organic matter on the physiochemical

properties of aqueous C60 nanoparticles. Environ Sci Technol 42: 2853-2859.

15. Lin D, Xing B. 2008. Tannic acid adsorption and its role for stabilizing carbon nanotube

suspensions. Environ Sci Technol 42: 5917-5923.

16. Espinasse B, Hotze EM, Wiesner MR. 2007. Transport and retention of colloidal

aggregates of C60 in porous media: Effects of organic macromolecules, ionic

composition, and preparation method. Environ Sci Technol 41: 7396-7402.

AcceptedPrepri n

t

20

17. Baun A, Hartman NB, Griegger K, Kisk KO. 2008. Ecotoxicity of engineered

nanoparticles to aquatic invertebrates: a brief review and recommendations for future

toxicity testing. Ecotoxicol 17: 387-395.

18. Persoone G, Janssen C, Coen WD. 1994. Cyst-based toxicity tests X: comparison of the

sensitivity of the acute Daphnia magna test and two crustacean microbiotests for

chemicals and wastes. Chemosphere 29: 2701-2710.

19. Rojickova-Padrtova R, Marsalek B, Holoubek I. 1998. Evaluation of alternative and

standard toxicity assays for screening of environmental samples: selection of an optimal

test battery. Chemosphere 37: 495-507.

20. Tervonen K, Waissi G, Petersen EJ, Akkanen J, Kukkonen JVK. 2010. Analysis of

fullerene C60 and kinetic measurements for its accumulation and depuration in Daphnia

magna. Environ Toxicol Chem 29: 1072-1078.

21. Roberts A, Mount AS, Seda B, Southern J, Aino R, Lin S, Ke PC, Rao AM, Klaine SJ.

2007. In vivo biomodification of lipid-coated carbon nanotubes by Daphnia magna.

Environ Sci Technol 41: 3025-3029.

22. Petersen EJ, Huang Q, Weber WJ Jr. 2008. Ecological uptake and depuration of carbon

nanotubes by Lumbriculus variegates. Environ Health Persp 116: 496-500.

23. Torokne A, Vasdinnyei B, Asztalos BM. 2007. A rapid microbiotest for the detection of

cyanobacterial toxins. Environ Toxicol 22: 64-68.

24. Marsalek, B, Blaha L. 2004. Comparison of 17 biotests for detection of cyanobacterial

toxicity. Environ Toxicol 19: 310-317.

AcceptedPrepri n

t

21

25. Kawano S, Murata H, Mikami H, Mukaibatake K, Waki H. 2006. Method optimization

for analysis of fullerenes by liquid chromatography/atmospheric pressure photoionization

mass spectrometry. Rapid Commun Mass Spectrom 20: 2783-2785.

26. Isaacson CW, Usenko CY, Tanguay RL, Field JA. 2007. Quantitation of fullerenes by

LC-ESI/MS and its application to in-vivo toxicity assays. Anal Chem 79: 9091-9097.

27. Isaacson CW, Bouchard D. 2010. Asymmetic flow field flow fractionation of aqueous

C60 nanoparticles with size determination by dynamic light scattering and quantitation

by liquid chromatography atmospheric pressure photo-ionization mass spectrometry. J

Chromatogr A 1217: 1506-1512.

28. Weber CI. 1993. Methods for measuring the acute toxicity of effluent and receiving

waters to freshwater and marine organisms. U.S. Environmental Protection Agency.

EPA/600/4-90/027F. Washington, DC.

29. Elimelech M, Jia X, Gregory J, Williams R. 1998. Particle Deposition & Aggregation:

Measurement, Modeling and Simulation. Butterworth-Heinemann, Woburn, WA, USA.

30. Tabata, Y, Ikada Y. 1988. Effect of the size and surface charge of polymer microspheres

on their phagocytosis by macrophage. Biomaterials 9: 356-362.

31. Bern L. 1990. Postcapture particle size selection by Daphnia cuculata (cladocera).

Limnol Oceanogr 35: 923-926.

32. Maeda-Martinez AM, Obregon-Barboza H, Dumont HJ. 1995. Laboratory culture of fairy

shrimps using baker’s yeast as a basic food in a flow through system. Hydrobiologia 29:

141-157.

33. Gerritsen J, Porter KG. 1982. The role of surface chemistry in filter feeding by

zooplankton. Science 216: 1225-1227.

AcceptedPrepri n

t

22

34. Torokne A, Vasdinnyei R, Asztalos BM. 2007. A rapid microbiotest for the detection of

cyanobacterial toxins. Environ Toxicol Chem 22: 64-68.

35. Rosenkranz P, Chaudhry Q, Stone V, Fernandes TF. 2009. A comparison of nanoparticle

and fine particle uptake by Daphnia magna. Environ Toxicol Chem 28: 2142-2149.

36. Gliwicz MZ. 1986. Suspended clay concentration controlled by filter-feeding

zooplankton in a tropical reservoir. Nature 323: 330-332.

37. Filella M, Rellstab C, Chanudet V, Spaak P. 2008. Effect of the filter feeder on particle

size distribution on organic colloids I freshwater. Water Res 42: 1919-1924.

38. Petersen EJ, Akkanen J, Kukkonen JVK, Weber J Jr. 2009. Biological uptake and

depuration of carbon nanotubes by Daphnia magna. Environ Sci Technol 43: 2969-2975.

AcceptedPrepri n

t

23

Figure Legends

Figure 1. Transmission electron micrographs of aqu/C60 fullerene agglomerates prior to

ingestion by T. platyurus. Selected images were representative of observations. (A) Larger

agglomerate; (B) Smaller agglomerates, (bars = 200 nm).

Figure 2. Accumulation of aqu/C70 fullerene (observed as dark gray) and red (observed as light

gray) polystyrene indicator beads by T. platyurus. (A) Control organisms, 30 to 45 h nauplii

exposed to indicator beads only (15 min); (B) Nauplii exposed to aqu/C70 (60 mg/L) for 60 min

prior to addition of indicator beads; (C) Nauplii exposed to only aqu/C70, 60 mg/ml, 60 min.

Figure 3. Visual observation of uptake of aqu/C60 and aqu/C70 fullerenes by T. platyurus. Bars

represent percentage of organisms showing greater than 20% gut length occupied by

accumulated fullerenes. The organisms (20 to 40) were scored and averaged in 2 trials. Exposure

conditions were the same as for Fig. 2C.

Figure 4. Quantitative assessment of uptake of aqu/C60. Initial concentration of aqu/C60 were 3

mg/L ( ) and 6 mg/L ( ). Other exposure conditions were the same as for Fig. 2C. Extraction

and quantitation described in Methods.

Figure 5. Transmission electron micrographs of transverse thin cross-sections of T. platyurus

after exposure to aqu/C60 (60 min., 69 mg/ml). (A) Arrows indicate aqu/C60 agglomerates (bar =

2 µm); (B) (a) aqu/C60 agglomerate, (b) polystyrene microsphere (bar =1 µm).

AcceptedPrepri n

t

24

Figure 6: Light microscopy and TEM images from T. platyurus excreted material after exposure

to aqu/C60. (A) Light microscopy image of excretion pellets after 30 d in water (white circle = 70

µm). (B) TEM image of one of the large agglomerates from the same sample as (A) (bar =

0.5µm).

Figure 1(A).

AcceptedPrepri n

t

25

Figure 1(B).

AcceptedPrepri n

t

26

Figure 2(A).

Figure 2(B).

Figure 2(C).

AcceptedPrepri n

t

27

Figure 3.

0 0.9 1.9 3.7 7.5 15 30 60

0

10

20

30

40

50

60

70

80

% O

rgan

ism

s A

ccum

ula

tin

g D

ark

Masses

Fullerene Concentration (mg/L)

C-60

C-70

Figure 4.

0 10 20 30 40 50 60

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

5.5

6.0

6.5

7.0

7.5

8.0

8.5

3 mg/L

6 mg/L

C6

0 M

ass (

g/m

g)

Time (min)

AcceptedPrepri n

t

28

Figure 5(A).

Figure 5(B).

AcceptedPrepri n

t

29

Figure 6(A).

AcceptedPrepri n

t

30

Figure 6(B).

AcceptedPrepri n

t

31

Table 1. Changes in fullerenes after ingestion, accumulation and depuration by T. platyurus.

Sample Agglomerate size

by Microscopy

Z average size

(nm) (mean ± 95%

confidence limit)

Polydispersivity

index (PDI)

(mean ± 95%

confidence limit)

Z potential (mV)

(mean ± 95%

confidence limit)

Aqu/C60 prior to feeding 200-500 nm 517 ± 20 0.38 ± 0.01 -27.5 ± 0.7

Aqu/C60 mid-gut thin

section TEM

5-10 µm ND ND ND

Aqu/C60 after excretion 10-70 µm* ND ND ND

C70 prior to feeding 200-500 nm 656 ± 39 0.58 ± 0.04 -47.5 ± 5.4

*200 – 500 nm sub-structure

ND = not determined


Recommended