Bioaccumulation of CdTe Quantum Dots in a Freshwater Alga

Aug 14, 2013 - The bioaccumulation kinetics of thioglycolic acid stabilized CdTe quantum dots (TGA-CdTe-QDs) in a freshwater alga Ochromonas danica wa...
0 downloads 12 Views 516KB Size
Subscriber access provided by ANDREWS UNIV

Article

Bioaccumulation of CdTe quantum dots in a freshwater alga Ochromonas danica: a kinetics study Ying Wang, Ai-Jun Miao, Jun Luo, Zhongbo Wei, Jun-Jie Zhu, and Liuyan Yang Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/es4017188 • Publication Date (Web): 14 Aug 2013 Downloaded from http://pubs.acs.org on August 18, 2013

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Environmental Science & Technology is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 27

Environmental Science & Technology

Table of Contents Art

Environment

Algal cell

TGA-CdTe-QDs

Plasma membrane

[QDs]cell (a.u./cell)

Diffusion 3e+4

Macropinocytosis

Dissolution Elimination

Accumulation kinetics

2e+4 Uptake 1e+4

Elimination

0 0

50 100 Time (min)

Vacuole

Expulsion Surface modification

150

ACS Paragon Plus Environment

Environmental Science & Technology

1 2 3 4

Bioaccumulation of CdTe quantum dots in a freshwater alga

5

Ochromonas danica: a kinetics study

6

7

8

Ying Wang†, Ai-Jun Miao†*, Jun Luo†, Zhong-Bo Wei†, Jun-Jie Zhu‡, Liu-Yan Yang†

9 10 11



12

Nanjing University, Nanjing, Jiangsu Province, 210046

13



14

Province, 210046

State Key Laboratory of Pollution Control and Resource Reuse, School of the Environment,

School of Chemistry & Chemical Engineering, Nanjing University, Nanjing, Jiangsu

15 16 17 18 19 20 21

*Corresponding author: [email protected] (Email), +86 25 89680255 (Tel.), +86 25

22

89680569 (Fax)

1

ACS Paragon Plus Environment

Page 2 of 27

Page 3 of 27

Environmental Science & Technology

23

ABSTRACT: The bioaccumulation kinetics of thioglycolic acid stabilized CdTe quantum

24

dots (TGA-CdTe-QDs) in a freshwater alga Ochromonas danica was comprehensively

25

investigated. Their photoluminescence (PL) was determined by flow cytometry. Its cellular

26

intensity increased hyperbolically with exposure time suggesting real internalization of

27

TGA-CdTe-QDs. This hypothesis was evidenced by the nanoparticle uptake experiment with

28

heat-killed or cold-treated cells and by their localization in the vacuoles. TGA-CdTe-QD

29

accumulation could further be well simulated by a biokinetic model used previously for

30

conventional pollutants. Moreover, macropinocytosis was the main route for their

31

internalization. As limited by their diffusion from the bulk medium to the cell surface,

32

TGA-CdTe-QD uptake rate increased proportionally with their ambient concentration. Quick

33

elimination in the PL of cellular TGA-CdTe-QDs was also observed. Such diminishment

34

resulted mainly from their surface modification by vacuolar biomolecules, considering that

35

these nanoparticles remained mostly undissolved and their expulsion out of the cells was slow.

36

Despite the significant uptake of TGA-CdTe-QDs, they had no direct acute effects on O.

37

danica. Overall, the above research shed new light on nanoparticle bioaccumulation study

38

and would further improve our understanding about their environmental behavior, effects and

39

fate.

40 41

2

ACS Paragon Plus Environment

Environmental Science & Technology

42



INTRODUCTION

43

Engineered nanoparticles are defined as particles with size in the range of 1-100 nm in at

44

least two dimensions.1 They are widely used in various areas like medicine, electronics,

45

textile, and environmental remediation. With the ever-increasing development of

46

nanotechnology, a large quantity of nanoparticles will be released inevitably and find their

47

way into the aquatic environment yet their potential risks remain unclear. In such ecosystems,

48

algae with cell size in micron scale are often the dominant primary producers and serve as an

49

important food source for organisms at higher trophic levels. Potential nanoparticle impacts

50

on these autotrophs will not only affect the biological community composition of the aquatic

51

ecosystems but also determine the environmental fate of nanoparticles themselves. Therefore,

52

the algal toxicity of nanoparticles has been extensively investigated in recent years.2-5 Both

53

indirect and direct effects were reported. Namely, nanoparticles could either liberate toxicants

54

(e.g., metal ions or reactive oxygen species) in the bulk medium and suppress the algal

55

growth indirectly or enter the cells and then impose toxic effects directly. In the latter case,

56

the nanoparticles’ capability to accumulate in algae (i.e., bioavailability) determines their

57

behavior, effects and fate in the natural environment. However, this property was rarely

58

examined for nanoparticles due to the lack of simple, instantaneous, and non-destructive

59

techniques to quantify their concentration in algae and other aquatic organisms. By contrast,

60

the bioavailability and accumulation kinetics of heavy metals and organic pollutants have

61

been well studied. Various models were established to elucidate how the accumulation

62

processes are influenced by different biological and physicochemical factors.6-8 Nevertheless,

63

it remains unknown whether the methodology used in these studies of conventional pollutants

64

could be applied to nanoparticles.

65

In these perspectives, we investigated the bioaccumulation kinetics of quantum dots

66

(QDs) in a freshwater alga Ochromonas danica. These nanoparticles are extremely small (
2 µm). Their PL inside the

260

vacuoles also intensified with time indicating a continuous uptake of these nanoparticles as

261

the intracellular PL of each particle was substantially weakened which would be discussed

262

below.

263

Bioaccumulation of organic pollutants and heavy metals had been extensively studied in

264

the past few decades. Various models were established on the uptake, elimination, trophic

265

and maternal transfer of these conventional pollutants.6,7 However, such research area is still

266

in its infancy for emerging pollutants like nanoparticles. Therefore, a biodynamic model6,8

267

derived from the ecotoxicological studies of conventional pollutants was adopted herein to

268

validate their applicability in nanotoxicology. According to this model, the variation of

269

[QDs]intra with time in the uptake experiment could be illustrated by the equation below,

270

d[QDs]intra = k u × [QDs]med − kem × [QDs]intra dt 9

ACS Paragon Plus Environment

(1)

Page 11 of 27

Environmental Science & Technology

271

where ku (ml/cell/min) is the uptake rate constant (also called clearance rate). [QDs]med

272

(a.u./ml) means the PL-based concentration of TGA-CdTe-QDs in the uptake medium and

273

the nanoparticle uptake rate could thus be expressed as ku×[QDs]med (a.u./cell/min). Further,

274

kem (min-1) represents the elimination rate constant for the intracellular PL of

275

TGA-CdTe-QDs including those expelled out of the cells (kex, min-1) and those quenched

276

inside as a result of surface modification or dissolution. On the other hand, bioaccumulation

277

could reduce [QDs]med, which may partly be compensated by their expulsion afterwards as

278

follows,

279

d[QDs] med = ( − k u × [QDs] med + k ex × [QDs]intra ) × d dt

(2)

280

where d (cells/ml) means cell density. As TGA-CdTe-QD expulsion out of the cells was

281

rather slow (discussed later), its contribution (i.e., kex × [QDs]intra × d ) to the variation of

282

[QDs]med could be neglected. According to Eqs. (1) and (2), [QDs]intra could be derived, 0

283

[QDs]intra =

ku × [QDs]med × (e −ku ×d ×t − e −kem×t ) kem − ku × d

(3)

284

where [QDs]med0 (a.u./ml) signifies the initial PL-based concentration of TGA-CdTe-QDs in

285

the uptake medium after quick surface adsorption. Further,

286 287

[QDs]cell = [QDs]intra + [QDs]ads

(4)

Therefore, 0

288

[QDs]cell =

k u × [QDs]med × (e −ku ×d ×t − e −kem ×t ) + [QDs]ads k em − k u × d

(5)

289

If [QDs]med remained unchanged (< 20%) during the uptake period as was the case in the

290

present study, Eq. (5) can be further simplified, 0

291

[QDs]cell

k × [QDs]med = u × (1 − e −kem ×t ) + [QDs]ads k em

(6)

292

Eq. (6) was then used to fit the hyperbolic correlation between [QDs]cell and uptake duration.

293

The two constants ku and kem ranged from 2.15×10-9 to 4.92×10-9 ml/cell/min and from 0.018

294

to 0.062 min-1 with no consistent trend between the different concentration treatments. Wu et

295

al.29 quantified the clearance rate of Ochromonas sp. for its predation of various filamentous

296

bacteria. It was in the range of 1.15-3.13×10-8 ml/cell/min, which was approximately one 10

ACS Paragon Plus Environment

Environmental Science & Technology

Page 12 of 27

297

order of magnitude higher than what we found here. One explanation is the difference in their

298

prey size (1.1-13.7 µm for bacteria vs. 88.9 nm for TGA-CdTe-QDs) which determines the

299

clearance rate and uptake routes (phagocytosis vs. pinocytosis).30-32 Further, TGA-CdTe-QD

300

uptake rate increased linearly with [QDs]med (Figure 1d), suggesting the internalization

301

processes were way below saturation. Considering their remarkable dissolution, the actual

302

concentration of TGA-CdTe-QDs in different uptake media was 3.20, 1.95, 1.31, 1.03, 0.83,

303

0.65 mg-Cd/l, as measured by GFAAS after ultrafiltration through a 10 kD membrane.

304

According to their hydrodynamic size in DY-V and the number of Cd for each primary

305

particle, TGA-CdTe-QD uptake rate in the unit of aggregates/cell/min was approximately

306

13.3 - 61.2 which was more rapid than the bacteria ingestion rate (0.1-0.5 bacteria/cell/min)

307

of Ochromonas sp..29 Such discrepancy was mainly due to much higher particle concentration

308

in the former case. If we assume that the migration vesicles with diameter approximately 1

309

µm as shown in Figure 2 were filled with TGA-CdTe-QD aggregates, then it could be

310

estimated that 0.015-0.067% of the plasma membrane was invaginated every minute for each

311

single cell. It was much lower than that of macrophages which can ingest 3% of its plasma

312

membrane every minute.33 The ciliate amoebae have an even faster membrane recycling rate

313

than macrophages suggesting their different internalization pathway or the unsuitability of

314

TGA-CdTe-QDs to be taken up by O. danica.

315

Cellular uptake of nutrients or pollutants (including nanoparticles) generally includes

316

three steps, 1) diffusion from the bulk solution to the cell periphery; 2) surface

317

adsorption/binding; 3) internalization.34 Considering our TGA-CdTe-QDs as rigid spheres,

318

their diffusion coefficient (D, cm2/s) at infinite dilution could be calculated using the

319

Stokes-Einstein relationship,

320

D=

k BT 6πηs RH

(7)

321

where kB (cm2 kg/s-2/K-1) is the Boltzmann constant, T (K) and ηs (kg/s/m) signify absolute

322

temperature and solvent viscosity, RH (m) is the hydrodynamic radius of the particle as

323

determined by DLS. The diffusion coefficient of TGA-CdTe-QDs in DY-V was thus 4.82 ×

324

10-8 cm2/s. Then their maximum diffusion to the cell surface Jmax (a.u./cell/min) could be

325

calculated by the equation below,34

326

J max = 60 D × S × ([QDs]med − [QDs]i ) × (

1 1 + ) R δ

11

ACS Paragon Plus Environment

(8)

Page 13 of 27

Environmental Science & Technology

327

where S (cm2/cell) indicates the surface area of each cell, [QDs]i (a.u./ml) is the nanoparticle

328

concentration in the interface between cells and the bulk solution, R (4×10-4 cm) and δ (cm)

329

signify the radius of O. danica and the diffusion layer thickness. Assuming that

330

TGA-CdTe-QD uptake by the cells is rather quick (i.e., [QDs]i = 0 a.u./ml) and δ is much

331

bigger than R,34 then Jmax could be calculated to be in the range of 10.5-3076.1 a.u./cell/min.

332

Their actual uptake rate (2.3-637.2 a.u./cell/min) was 14.8-33.9% of the maximum diffusion.

333

Therefore, TGA-CdTe-QD diffusion from the bulk solution to the cell surface determined

334

their internalization by O. danica.35 It could further be anticipated that nanoparticle

335

bioaccumulation may be restricted by their physical transport unless their hydrodynamic size

336

is below 10 nm. This hypothesis was supported by the study of Chen et al.36, in which uptake

337

of “colloidal iron” (~ 5 nm) by the marine diatom Thalassiosira pseudonana was investigated.

338

Its internalization was 0.08-0.2% of the maximum diffusive fluxes and thus was not limited

339

by diffusion.

340

Uptake and Elimination Mechanisms. Endocytosis has been well-recognized to be the

341

major mechanism for nanoparticles to enter cells.31 Two main categories, phagocytosis and

342

pinocytosis, are included and the latter is further classified into at least four mechanisms (i.e.,

343

clathrin-mediated,

caveolae-mediated,

clathrin/caveolae-independent

endocytosis, 30,37

and

344

macropinocytosis) depending on their formation of intracellular vesicles.

345

experiment of the present study, the mixotrophic freshwater alga O. danica was growing

346

quite well excluding the possibility that TGA-CdTe-QDs entered the cells by altering their

347

plasma membrane permeability.4 This alga is also well-known for its ability to ingest bacteria

348

and yeasts (i.e., “cell eating”) through phagocytosis,15 which is mainly employed to take up

349

particles larger than 0.5 µm.30 Therefore, this mechanism may not be applicable to the

350

nano-sized particles, which would be corroborated by the following finding. To gain further

351

insight into the specific mechanism for the translocation of TGA-CdTe-QDs into O. danica,

352

six inhibitors were applied. Among them, CCCP and NaN3 are able to impair ATP production

353

by uncoupling oxidative phosphorylation38 or impeding cytochrome oxidation.39 The

354

presence of both inhibitors resulted in a marked decline of TGA-CdTe-QD uptake rate (Table

355

1), indicating their internalization is an energy-consuming process. MβCD is a cyclic

356

heptasaccharide and could deplete cholesterol or modify cholesterol-rich domains (lipid rafts)

357

within the cell membrane.40 It is thus often used as a selective inhibitor of caveolae-mediated

358

uptake.41 But its suppression of macropinocytosis and clathrin-dependent endocytosis was

359

also reported.42,43 By contrast, genistein is a depressor of several tyrosine kinases and could

360

mainly restrain the caveolae- and clathrin-mediated endocytosis.44 Further, dynasore impedes 12

ACS Paragon Plus Environment

In the uptake

Environmental Science & Technology

361

the function of dynamin and this protein is required in various endocytosis processes (e.g.,

362

phagocytosis) but not including macropinocytosis.30 The latter pathway could however be

363

blocked by amiloride due to the reduction of submembraneous pH together with the

364

suppression of Rac1 and Cdc42 signaling.45 As shown in Table 1, a substantial decrease in

365

TGA-CdTe-QD uptake was observed when MβCD or amiloride was applied, which was

366

further aggravated at higher concentrations of these two inhibitors. On the contrary, both

367

genistein and dynasore had no significant effects (p > 0.05). Moreover, the vacuoles

368

encapsulating TGA-CdTe-QDs inside the cells (> 1 µm, Figure 2) were much larger than

369

those involved in various pinocytosis processes (~ 100 nm, Table 1) other than

370

macropinocytosis.31,46 In viewed of all these phenomena, macropinocytosis was thus

371

considered to be the main route for TGA-CdTe-QDs to be taken up by O. danica. Similarly,

372

this mechanism was previously found to be responsible for the uptake of peptide-conjugated

373

QDs by HeLa cancer cells.47 Nevertheless, QDs could still be endocytosed via other routes

374

like clathrin-mediated pinocytosis and sometimes their bioaccumulation may even be

375

accomplished through multiple pathways collectively.48 Further research is required to

376

illuminate how the internalization routes are selected based on the characteristics of

377

nanoparticles or cells themselves.

378

[QDs]cell (a.u./cell) represents the cellular concentration of TGA-CdTe-QDs based on

379

their PL. Therefore, the hyperbolic correlation between [QDs]cell and exposure time in the

380

uptake experiment was speculated to be caused by the quick elimination of TGA-CdTe-QD

381

PL intracellularly as a result of their dissolution, surface modification, or expulsion out of the

382

cells. The intracellular diminishment of TGA-CdTe-QD PL was supported by the finding that

383

[QDs]cell decreased by more than 90% within a few hours after the ambient nanoparticles

384

were removed in the efflux experiment (Figure 3a). Moreover, [QDs]cell went up again once

385

the cells were re-exposed to TGA-CdTe-QDs, excluding the possibility that [QDs]cell leveled

386

off due to the physiological changes of O. danica in the uptake experiment. On the other hand,

387

no PL signal of TGA-CdTe-QDs was found in the efflux medium and the cellular contents of

388

Cd or Te decreased by no more than 10% after 4.5 h (Figure 3b). It implies that

389

TGA-CdTe-QD expulsion out of the cells either in the form of nanoparticles or metal ions

390

was relatively slow and intracellular PL quenching was the main cause for the substantial

391

diminishment of [QDs]cell. The exocytosis/expulsion of single-walled carbon nanotubes and

392

gold nanoparticles by the fibroblast NIH-3T3 cells were examined with the efflux rate

393

constant (i.e., kex) in the range of 10-4 to 10-3 min-1.49 In other words, approximately

394

2.4-21.3% of the intracellular nanoparticles were expelled out during a 4.5-h period, 13

ACS Paragon Plus Environment

Page 14 of 27

Page 15 of 27

Environmental Science & Technology

395

comparable to what was estimated based on the variation of cellular metal contents in the

396

efflux experiment of the present study. The decrease of [QDs]cell with time was then

397

simulated by the first order kinetics below,

398

[QDs]cell = [QDs]cell × e - kem ×t

0

(9)

399

where [QDs]cell0 (a.u./cell) is the cellular concentration of TGA-CdTe-QDs at the beginning

400

of the efflux experiment. The elimination rate constant kem thus obtained was 0.057 min-1 as

401

was similar to its counterpart (0.062 min-1) in the uptake experiment. Then was the

402

intracellular PL quenching because of TGA-CdTe-QD surface modification or their

403

dissolution in the acidic vacuoles?31,50 To answer this question, the PL of TGA-CdTe-QDs in

404

the cell homogenate of O. danica as a simulation of the intracellular environment was

405

monitored. It diminished by more than 90% within 10 min while the wavelength of their PL

406

maximum remained unchanged as in contrast to what was found in Domingos et al.51. In this

407

situation, the characteristic XRD peaks of CdTe still existed with similar height to the control

408

treatment (Figure 3c) where the same proportion of TGA-CdTe-QDs and cell homogenate

409

were freeze-dried before being mixed with each other (i.e., possible dissolution of the

410

nanoparticles was avoided). Additionally, our preliminary experiment illustrated that the PL

411

of TGA-CdTe-QDs remained constant with the ambient pH in the range of 4-8, excluding the

412

potential effects of vacuolar acidic environment on PL quenching. Therefore, it is mainly

413

TGA-CdTe-QD surface modification due to the existence of biomolecules (e.g., amino acids,

414

proteins, carbohydrates and lipids etc.) in the vacuoles which eliminated their PL abruptly.

415

Likewise, nanoparticles in biological fluids tend to be coated by proteins which was also

416

named as protein corona.52 Furthermore, an irreversible quenching of QD PL by DNA,

417

nucleotides, amino acids was reported,53 which may result from the redox reaction between

418

the QDs and biomolecules or the binding of the quenchers to the nanoparticle surface via

419

electrostatic attraction, hydrogen bonding and so on.

420

Toxicity of TGA-CdTe-QDs. Metallic nanoparticles may have either direct or indirect

421

adverse effects on algae depending on their ability to enter cells.2,16 In the literature, most of

422

the algal toxicity of nanoparticles was attributed to their indirect effects with relatively few

423

observations of direct ones.4,5,16 Silver nanoparticles could get into the cells of the same algal

424

species as what was used herein and were concentrated also in vacuoles.16 Further, these

425

nanoparticles were noxious even when the ambient free silver ion concentration was orders of

426

magnitude lower than the non-observed effect level, suggesting that silver nanoparticles

427

inside the cells did have some direct effects. Similarly, natural organic matter was able to 14

ACS Paragon Plus Environment

Environmental Science & Technology

428

induce the uptake of CuO nanoparticles with size smaller than 5 nm by a prokaryotic alga

429

Microcystis aeruginosa and thus increase their nanotoxicity.54 In the present study, a large

430

quantity of TGA-CdTe-QDs was internalized by O. danica. However, their adverse effects

431

were comparable to what was found in the < 10 kD ultrafiltrate of these nanoparticles with

432

similar [Cd]dis (Figure 4). Accordingly, the relative change of µ with or without the presence

433

of TGA-CdTe-QDs could be fitted to the same dose-response curve. Although Te was also

434

present in the ultrafiltrate, its concentration was not high enough to be toxic based on the

435

inhibition results of TeO32- (SI, Figure S5), the most noxious form of Te.55 Therefore, our

436

findings revealed that the intracellular TGA-CdTe-QDs had negligibly direct acute effects on

437

O. danica and their toxicity was mainly caused by Cd ion liberation into the bulk medium.

438

Possible explanations include the disposition of the nanoparticles in relatively inert

439

compartments (e.g., vacuoles). Surface modification by intracellular biomolecules may also

440

alleviate their toxicity.56 Kirchner et al.57 found that the Cd released from intracellular QDs

441

were more toxic than those dissolved outside. It implies that most of TGA-CdTe-QDs

442

remained undissolved in O. danica further supporting our hypothesis that intracellular PL

443

quenching was mainly brought about by nanoparticle modification on the surface instead of

444

dissolution.

445

Overall, our study manifests that TGA-CdTe-QDs could enter the freshwater alga O.

446

danica directly through macropinocytosis. Once they were in the cells, these nanoparticles

447

were concentrated in vacuoles with limited exocytosis/expulsion and dissolution, resulting in

448

negligible toxicity to the alga. Moreover, the PL of intracellular TGA-CdTe-QDs got

449

eliminated rather quickly as a result of their surface modification. Such instability of PL

450

inside the cells cannot be disregarded when cell-targeting QDs with new functions are to be

451

synthesized in the future. More importantly, a biodynamic model previously used for

452

conventional pollutants was for the first time systematically applied herein to illustrate

453

nanoparticle bioaccumulation kinetics in aquatic organisms, thanks to the PL of QDs. The

454

methodology thus established might be extended to other environmental studies using

455

different organisms in various environments as long as the objective nanoparticles could be

456

analyzed rather simply, instantaneously, and non-destructively (e.g., fluorescence or isotope

457

labeling).

458



ACKNOWLEDGEMENTS

459

We thank two anonymous reviewers for their constructive suggestions on this paper. The

460

financial supports offered by the National Natural Science Foundation of China (41001338, 15

ACS Paragon Plus Environment

Page 16 of 27

Page 17 of 27

Environmental Science & Technology

461

41271486, and 21237001) and the Natural Science Foundation of Jiangsu Province

462

(BK2010371) to A. J. Miao have made this work possible.

463



SUPPORTING INFORMATION

464

Detailed procedures about nanoparticle synthesis and characterization, the PL spectra of

465

TGA-CdTe-QDs and O. danica, physicochemical characterization of TGA-CdTe-QDs,

466

nanoparticle PL stability in DY-V and verification of flow cytometer’s ability to properly

467

detect TGA-CdTe-QDs in the medium, discrimination of extracellular nanoparticles from O.

468

danica and quantification of nanoparticle bioaccumulation by flow cytometry, as well as the

469

toxicity of TeO32- are included as Supporting Information. This information is available free

470

of charge via the Internet at http://pubs.acs.org.

471 472

16

ACS Paragon Plus Environment

Environmental Science & Technology

473



474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516

(1) Klaine, S. J.; Alvarez, P. J. J.; Batley, G. E.; Fernandes, T. F.; Handy, R. D.; Lyon, D. Y.; Mahendra, S.; McLaughlin, M. J.; Lead, J. R. Nanomaterials in the environment: Behavior, fate, bioavailability, and effects. Environ. Toxicol. Chem. 2008, 27, (9), 1825-1851. (2) Miao, A. J.; Schwehr, K. A.; Xu, C.; Zhang, S. J.; Luo, Z.; Quigg, A.; Santschi, P. H. The algal toxicity of silver engineered nanoparticles and detoxification by exopolymeric substances. Environ. Pollut. 2009, 157, (11), 3034-3041. (3) Miao, A. J.; Zhang, X. Y.; Luo, Z.; Chen, C. S.; Chin, W. C.; Santschi, P. H.; Quigg, A. Zinc oxide engineered nanoparticles dissolution and toxicity to marine phytoplankton. Environ. Toxicol. Chem. 2010, 29, (12), 2814-2822. (4) Navarro, E.; Baun, A.; Behra, R.; Hartmann, N. B.; Filser, J.; Miao, A. J.; Quigg, A.; Santschi, P. H.; Sigg, L. Environmental behavior and ecotoxicity of engineered nanoparticles to algae, plants, and fungi. Ecotoxicology 2008, 17, (5), 372-386. (5) Hartmann, N. B.; Von der Kammer, F.; Hofmann, T.; Baalousha, M.; Ottofuelling, S.; Baun, A. Algal testing of titanium dioxide nanoparticles-Testing considerations, inhibitory effects and modification of cadmium bioavailability. Toxicology 2010, 269, (2-3), 190-197. (6) Fan, C. W.; Reinfelder, J. R. Phenanthrene accumulation kinetics in marine diatoms. Environ. Sci. Technol. 2003, 37, (15), 3405-3412. (7) Tsui, M. T. K.; Wang, W. X. Biokinetics and tolerance development of toxic metals in Daphnia magna. Environ. Toxicol. Chem. 2007, 26, (5), 1023-1032. (8) Wang, W. X.; Fisher, N. S. Delineating metal accumulation pathways for marine invertebrates. Sci. Total Environ. 1999, 238, 459-472. (9) Alivisatos, A. P. Semiconductor clusters, nanocrystals, and quantum dots. Science 1996, 271, (5251), 933-937. (10) Anas, A.; Okuda, T.; Kawashima, N.; Nakayama, K.; Itoh, T.; Ishikawa, M.; Biju, V. Clathrin-mediated endocytosis of quantum dot-peptide conjugates in living cells. ACS Nano 2009, 3, (8), 2419-2429. (11) Zhao, F.; Zhao, Y.; Liu, Y.; Chang, X.; Chen, C.; Zhao, Y. Cellular uptake, intracellular trafficking, and cytotoxicity of nanomaterials. Small 2011, 7, (10), 1322-1337. (12) Lin, S.; Bhattacharya, P.; Rajapakse, N. C.; Brune, D. E.; Ke, P. C. Effects of quantum dots adsorption on algal photosynthesis. J. Phys. Chem. C 2009, 113, (25), 10962-10966. (13) Domingos, R. F.; Simon, D. F.; Hauser, C.; Wilkinson, K. J. Bioaccumulation and effects of CdTe/CdS quantum dots on Chlamydomonas reinhardtii - nanoparticles or the free ions? Environ. Sci. Technol. 2011, 45, (18), 7664-9. (14) Miralles, P.; Church, T. L.; Harris, A. T. Toxicity, uptake, and translocation of engineered nanomaterials in vascular plants. Environ. Sci. Technol. 2012, 46, (17), 9224-9239. (15) Aaronson, S. The biology and ultrastructure of phagotrophy in Ochromonas danica (Chrysophyceae : Chrysomonadida). J. Gen. Microbiol. 1974, 83, 21-29. (16) Miao, A. J.; Luo, Z.; Chen, C. S.; Chin, W. C.; Santschi, P. H.; Quigg, A. Intracellular uptake: A possible mechanism for silver engineered nanoparticle toxicity to a freshwater alga Ochromonas danica. Plos One 2010, 5, (12). (17) Zhang, H.; Wang, L.; Xiong, H.; Hu, L.; Yang, B.; Li, W. Hydrothermal synthesis for high-quality CdTe nanocrystals. Adv. Mater. 2003, 15, (20), 1712-1715.

REFERENCES

17

ACS Paragon Plus Environment

Page 18 of 27

Page 19 of 27

517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561

Environmental Science & Technology

(18) Domingos, R. F.; Baalousha, M. A.; Ju-Nam, Y.; Reid, M. M.; Tufenkji, N.; Lead, J. R.; Leppard, G. G.; Wilkinson, K. J. Characterizing manufactured nanoparticles in the environment: Multimethod determination of particle sizes. Environ. Sci. Technol. 2009, 43, (19), 7277-7284. (19) Miao, A. J.; Wang, W. X. Fulfilling iron requirements of a coastal diatom under different temperatures and irradiances. Limnol. Oceanogr. 2006, 51, (2), 925-935. (20) Al-Hajaj, N. A.; Moquin, A.; Neibert, K. D.; Soliman, G. M.; Winnik, F. o. M.; Maysinger, D. Short ligands affect modes of QD uptake and elimination in human cells. ACS Nano 2011, 5, (6), 4909-4918. (21) Yu, W. W.; Qu, L.; Guo, W.; Peng, X. Experimental determination of the extinction coefficient of CdTe, CdSe, and CdS nanocrystals. Chem. Mater. 2003, 15, (14), 2854-2860. (22) Derfus, A. M.; Chan, W. C. W.; Bhatia, S. N. Probing the cytotoxicity of semiconductor quantum dots. Nano Lett. 2003, 4, (1), 11-18. (23) Schindler, P.; Althaus, H.; Hofer, F.; Minder, W. Löslichkeitsprodukte von Metalloxiden und -hydroxiden. 10. Mitteilung. Löslichkeitsprodukte von Zinkoxid, Kupferhydroxid und Kupferoxid in Abhängigkeit von Teilchengrösse und molarer Oberfläche. Ein Beitrag zur Thermodynamik von Grenzflächen fest-flüssig. Helvetica Chimica Acta 1965, 48, (5), 1204-1215. (24) Xu, M.; Deng, G.; Liu, S.; Chen, S.; Cui, D.; Yang, L.; Wang, Q. Free cadmium ions released from CdTe-based nanoparticles and their cytotoxicity on Phaeodactylum tricornutum. Metallomics 2010, 2, (7), 469-73. (25) Zanella, M.; Abbasi, A. Z.; Schaper, A. K.; Parak, W. J. Discontinuous growth of II−VI semiconductor nanocrystals from different materials. J. Phys. Chem. C 2010, 114, (14), 6205-6215. (26) Weng, J.; Song, X.; Li, L.; Qian, H.; Chen, K.; Xu, X.; Cao, C.; Ren, J. Highly luminescent CdTe quantum dots prepared in aqueous phase as an alternative fluorescent probe for cell imaging. Talanta 2006, 70, (2), 397-402. (27) Chang, E.; Thekkek, N.; Yu, W. W.; Colvin, V. L.; Drezek, R. Evaluation of quantum dot cytotoxicity based on intracellular uptake. Small 2006, 2, (12), 1412-7. (28) Lesniak, A.; Salvati, A.; Santos-Martinez, M. J.; Radomski, M. W.; Dawson, K. A.; Aberg, C. Nanoparticle adhesion to the cell membrane and its effect on nanoparticle uptake efficiency. J. Am. Chem. Soc. 2013, 135, (4), 1438-44. (29) Wu, Q. L.; Boenigk, J.; Hahn, M. W. Successful predation of filamentous bacteria by a nanoflagellate challenges current models of flagellate bacterivory. Appl. Environ. Microbiol. 2004, 70, (1), 332-339. (30) Iversen, T. G.; Skotland, T.; Sandvig, K. Endocytosis and intracellular transport of nanoparticles: Present knowledge and need for future studies. Nano Today 2011, 6, (2), 176-185. (31) Hild, W. A.; Breunig, M.; Goepferich, A. Quantum dots - Nano-sized probes for the exploration of cellular and intracellular targeting. Eur. J. Pharm. Biopharm. 2008, 68, (2), 153-168. (32) Schmidtke, A.; Bell, E. M.; Weithoff, G. Potential grazing impact of the mixotrophic flagellate Ochromonas sp (Chrysophyceae) on bacteria in an extremely acidic lake. J. Plankton Res. 2006, 28, (11), 991-1001. (33) Alberts, B.; Johnson, A.; Lewis, J.; Raff, M.; Roberts, K.; Walter, P. Molecular Biology of 18

ACS Paragon Plus Environment

Environmental Science & Technology

562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606

the Cell. Garland Science: New York, 2002. (34) Buffle, J.; Wilkinson, K. J.; van Leeuwen, H. P. Chemodynamics and bioavailability in natural waters. Environ. Sci. Technol. 2009, 43, (19), 7170-7174. (35) Hudson, R. J. M.; Morel, F. M. M. Iron transport in marine phytoplankton - Kinetics of cellular and medium coordination reactions. Limnol. Oceanogr. 1990, 35, (5), 1002-1020. (36) Chen, M.; Dei, R. C. H.; Wang, W. X.; Guo, L. D. Marine diatom uptake of iron bound with natural colloids of different origins. Mar. Chem. 2003, 81, (3-4), 177-189. (37) Mayor, S.; Pagano, R. E. Pathways of clathrin-independent endocytosis. Nat. Rev. Mol. Cell Biol. 2007, 8, (8), 603-612. (38) Hassler, C. S.; Wilkinson, K. J. Failure of the biotic ligand and free-ion activity models to explain zinc bioaccumulation by Chlorella kesslerii. Environ. Toxicol. Chem. 2003, 22, (3), 620-6. (39) Keilin, D. The action of sodium azide on cellular respiration and on some catalytic oxidation reactions. P. Roy. Soc. Lond. B Biol. 1936, 121, (822), 165-173. (40) Gratton, S. E.; Ropp, P. A.; Pohlhaus, P. D.; Luft, J. C.; Madden, V. J.; Napier, M. E.; DeSimone, J. M. The effect of particle design on cellular internalization pathways. Proc. Natl. Acad. Sci. U. S. A. 2008, 105, (33), 11613-8. (41)Foerg, C.; Ziegler, U.; Fernandez-Carneado, J.; Giralt, E.; Rennert, R.; Beck-Sickinger, A. G.; Merkle, H. P. Decoding the entry of two novel cell-penetrating peptides in HeLa cells: lipid raft-mediated endocytosis and endosomal escape. Biochemistry 2005, 44, (1), 72-81. (42) van Kerkhof, P.; Sachse, M.; Klumperman, J.; Strous, G. J. Growth hormone receptor ubiquitination coincides with recruitment to clathrin-coated membrane domains. J. Biol. Chem. 2001, 276, (6), 3778-3784. (43) Rodal, S. K.; Skretting, G.; Garred, O.; Vilhardt, F.; van Deurs, B.; Sandvig, K. Extraction of cholesterol with methyl-beta-cyclodextrin perturbs formation of clathrin-coated endocytic vesicles. Mol. Biol. Cell 1999, 10, (4), 961-974. (44) Pelkmans, L.; Puntener, D.; Helenius, A. Local actin polymerization and dynamin recruitment in SV40-induced internalization of caveolae. Science 2002, 296, (5567), 535-539. (45) Koivusalo, M.; Welch, C.; Hayashi, H.; Scott, C. C.; Kim, M.; Alexander, T.; Touret, N.; Hahn, K. M.; Grinstein, S. Amiloride inhibits macropinocytosis by lowering submembranous pH and preventing Rac1 and Cdc42 signaling. J. Cell Biol. 2010, 188, (4), 547-563. (46) Conner, S. D.; Schmid, S. L. Regulated portals of entry into the cell. Nature 2003, 422, (6927), 37-44. (47) Ruan, G.; Agrawal, A.; Marcus, A. I.; Nie, S. Imaging and tracking of tat peptide-conjugated quantum dots in living cells: new insights into nanoparticle uptake, intracellular transport, and vesicle shedding. J. Am. Chem. Soc. 2007, 129, (47), 14759-14766. (48) Zhang, L. W.; Baeumer, W.; Monteiro-Riviere, N. A. Cellular uptake mechanisms and toxicity of quantum dots in dendritic cells. Nanomedicine 2011, 6, (5), 777-791. (49) Jin, H.; Heller, D. A.; Sharma, R.; Strano, M. S. Size-dependent cellular uptake and expulsion of single-walled carbon nanotubes: Single particle tracking and a generic uptake model for nanoparticles. ACS Nano 2009, 3, (1), 149-158. (50) Mathieu, Y.; Guern, J.; Kurkdjian, A.; Manigault, P.; Manigault, J.; Zielinska, T.; Gillet, B.; Beloeil, J. C.; Lallemand, J. Y. Regulation of vacuolar pH of plant-cells .1. Isolation and properties of vacuoles suitable for P-31 NMR-studies. Plant Physiol. 1989, 89, (1), 19-26. 19

ACS Paragon Plus Environment

Page 20 of 27

Page 21 of 27

607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629

Environmental Science & Technology

(51) Domingos, R. F.; Franco, C.; Pinheiro, J. P. Stability of core/shell quantum dots—role of pH and small organic ligands. Environ. Sci. Pollut. Res. 2013, 20, (7), 4872-4880. (52) Juganson, K.; Mortimer, M.; Ivask, A.; Kasemets, K.; Kahru, A. Extracellular conversion of silver ions into silver nanoparticles by protozoan Tetrahymena thermophila. Environ. Sci.: Processes Impacts 2013, 15, (1), 244-250. (53) Siegberg, D.; Herten, D. P. Fluorescence quenching of quantum dots by DNA nucleotides and amino acids. Aust. J. Chem. 2011, 64, (5), 512-516. (54) Wang, Z. Y.; Li, J.; Zhao, J.; Xing, B. S. Toxicity and internalization of CuO nanoparticles to prokaryotic alga Microcystis aeruginosa as affected by dissolved organic matter. Environ. Sci. Technol. 2011, 45, (14), 6032-6040. (55) Ghasemi, E.; Najafi, N. M.; Seidi, S.; Raofie, F.; Ghassempour, A. Speciation and determination of trace inorganic tellurium in environmental samples by electrodeposition-electrothermal atomic absorption spectroscopy. J. Anal. At. Spectrom. 2009, 24, (10), 1446-1451. (56) Clift, M. J. D.; Rothen-Rutishauser, B.; Brown, D. M.; Duffin, R.; Donaldson, K.; Proudfoot, L.; Guy, K.; Stone, V. The impact of different nanoparticle surface chemistry and size on uptake and toxicity in a murine macrophage cell line. Toxicol. Appl. Pharmacol. 2008, 232, (3), 418-427. (57) Kirchner, C.; Liedl, T.; Kudera, S.; Pellegrino, T.; Javier, A. M.; Gaub, H. E.; Stolzle, S.; Fertig, N.; Parak, W. J. Cytotoxicity of colloidal CdSe and CdSe/ZnS nanoparticles. Nano Lett. 2005, 5, (2), 331-338.

20

ACS Paragon Plus Environment

Environmental Science & Technology

630 631

Page 22 of 27

Table 1. Suppression of TGA-CdTe-QD uptake in the presence of different concentrations of inhibitors. Pinocytosis Inhibitors

NaN3 CCCP MβCD Amiloride Genistein Dynasore 632 633 634 635 636

Conc.

Inhibition (%)

10 mM

19.7*

20 mM

**

44.1

10 µM

5.45

20 µM

36.3*

50 mM

35.7*

100 mM

47.1**

0.4 mM

28.7*

0.6 mM

37.1**

0.2 mM

-26.3

0.4 mM

-6.5

1 µM

5.1

4 µM

8.8

Phagocytosis (> 1 µm)a

Macropinocytosis (> 0.5 µm)

Clathrin-mediated (~120 nm)

Caveolae-mediated (~50 – 80 nm)

Clathrin/Caveolae-independent (~90 nm)

+b

+

+

+

+

+

+

+

+

+

+

+

+

-c

-

-

-

+

-

a. b. c. * .

The size of the vesicles involved in this endocytosis route. This endocytosis route could be suppressed by the inhibitor of the same row and significant uptake inhibition was observed. This endocytosis route could be suppressed by the inhibitor of the same row but no significant uptake inhibition was observed. p < 0.05. ** . p < 0.01.

21

ACS Paragon Plus Environment

Page 23 of 27

Environmental Science & Technology

637

Figure Legends

638

Figure 1. Variation of cellular TGA-CdTe-QDs ([QDs]cell, a.u./cell) with exposure time (a, b)

639

in the presence of 1.3, 1.5, 1.9, 2.3, 3.1, 4.7 mg-Cd/l nanoparticles and (c) in the uptake

640

experiment with heat-killed (heat, star) or cold-treated (cold-0, square; cold-15, triangle) cells

641

as compared to the normal uptake (circle). The arrow indicates transferring cells from 25 oC

642

to 4 oC after 15-min normal uptake. Ambient TGA-CdTe-QD concentration was kept at 4.7

643

mg-Cd/l throughout this experiment. Solid lines in (a)-(c) represent the simulated increase of

644

[QDs]cell with exposure time; (d) Linear correlation between TGA-CdTe-QD uptake rate and

645

their ambient concentration in the medium. Data are mean ± standard deviation (n = 2).

646

Figure 2. The confocal laser scanning microscopy images of O. danica as obtained via

647

different channels (CH1: differential interference contrast; CH2: 405 nm laser, filter bandpass

648

= 656-702 nm; CH3: 405 nm laser, filter bandpass = 565-610 nm; Merge: combinative

649

images from CH1 to 3) or under the z-scanning mode at a 1.24 µm spacing (30 min only)

650

after exposed to 4.7 mg-Cd/l TGA-CdTe-QDs for 0, 5, 30, and 60 min. Scale bars are 5 µm.

651

Figure 3. (a) Decrease of cellular TGA-CdTe-QDs ([QDs]cell, a.u./cell) with depuration time

652

in a 4.5-h efflux experiment. The arrow indicates a subset of the cells was exposed to 4.7

653

mg-Cd/l TGA-CdTe-QDs again after 30 min in this experiment. Solid lines represent the

654

simulated curves for the elimination of [QDs]cell and the reuptake of TGA-CdTe-QDs; (b) The

655

relative amount of Cd (circle) and Te (triangle) retained as compared to their initial cellular

656

content in the efflux experiment. Solid lines represent the relative changes of cellular metal

657

content with depuration time as simulated by the first-order kinetics; (c) XRD patterns (top)

658

for the freeze-dried mixture of TGA-CdTe-QDs and cell homogenate as well as (bottom) for

659

the control treatment where the nanoparticle solution and cell homogenate were dried

660

separately before being mixed with each other. The cell homogenate was prepared by

661

suspending the algal cells in 0.9% w/w sodium chloride before breaking them through

662

sonication. Sodium chloride also served as an internal standard for the quantification of CdTe.

663

Data are mean ± standard deviation (n = 2).

664

Figure 4. Relative changes of the cell specific growth rate µ at different dissolved

665

concentration of Cd ([Cd]dis) for Ochromonas danica exposed to different concentrations of

666

TGA-CdTe-QDs (1.4, 2.9, 5.9, 12, 32, and 62 mg-Cd/l) or their corresponding < 10 kD

667

ultrafiltrates for 24 h as compared to the control treatment. Solid line is the simulated

668

dose-response curve by the Logistic model. Data are mean ± standard deviation (n = 2). 22

ACS Paragon Plus Environment

Environmental Science & Technology

2e+3

c 2e+4

a e

normal 1.9

[QDs]cell (a.u./cell)

1e+3

1e+4

cold-15 1.5

cold-0 1.3

0 0

20

40

2e+4

heat

60 0 Time (min)

20

40

60 d

b

0

600

4.7

1e+4

300

3.1 2.3

0 0 670 671 672

20 40 Time (min)

60

0

1e+11 2e+11 [QDs]med (a.u./ml)

23

ACS Paragon Plus Environment

[QDs]cell (a.u./cell)

Figure 1

0 3e+11

Uptake rate (a.u./cell/min)

669

Page 24 of 27

Page 25 of 27

673

Environmental Science & Technology

Figure 2 CH 2

CH 3

Z-scanning

60 min

30 min

5 min

0 min

CH 1

674 675 676

24

ACS Paragon Plus Environment

Merge

Environmental Science & Technology

Figure 3 [QDs]cell (a.u./cell)

677

a

3.0e+4

reuptake

1.5e+4 elimination

0.0 Metal retained (%)

150

b Cd

100

Te

50 0 0

100

200

300

Time (min)

678

Counts (a.u.)

1000

c CdTe

500

NaCl

0 10

30 50 2θ (degrees)

679 680 681 682

25

ACS Paragon Plus Environment

70

Page 26 of 27

Page 27 of 27

Figure 4

Relative change of µ

683

Environmental Science & Technology

1.0

0.5 < 10 kD QDs

0.0 0

684

1 2 [Cd]dis (mg/l)

3

685 686 687 688 689 690

26

ACS Paragon Plus Environment