Real-Time Visualization of Perylene Nanoclusters in Water and Their

Jun 9, 2015 - State Key Laboratory of Environment Simulation, School of Environment, Beijing Normal University, No. 19 Xinjiekouwai Street, Beijing ...
0 downloads 9 Views 1MB Size
Subscriber access provided by NEW YORK UNIV

Article

Real Time Visualization of Perylene Nano-clusters (PNCs) in Water and Their Partitioning to Graphene Surface and Macrophage Cells Xuejun Guo, Xin Jin, Xiaofang Lv, Yingying Pu, and Fan Bai Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.5b01880 • Publication Date (Web): 09 Jun 2015 Downloaded from http://pubs.acs.org on June 17, 2015

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 33

Environmental Science & Technology

1 2 3 4

Real Time Visualization of Perylene Nano-clusters (PNCs) in Water and Their Partitioning to Graphene Surface and Macrophage Cells

5 6

Xuejun Guo†*, Xin Jin†, Xiaofang Lv†, Yingying Pu‡, Fan Bai‡*

7



8

Normal University, No. 19 Xinjiekouwai Street, Beijing 100875, China

9



State Key Laboratory of Environment Simulation, School of Environment, Beijing

Biodynamic Optical Imaging Center, School of Life Science, Peking University, No. 5

10

Yiheyuan Road, 100871, China

11

*Corresponding author

12

Guo, X. Email: [email protected]; Phone: 86-10-5880-7808 Fax: 86-10-5880-7808;

13

Bai, F. Email: [email protected]; Phone: 86-10-6275-6164 Fax: 86-10-6275-6164.

14

Guo X, Jin X, and Bai F contributed equally to this work.

15 16

ABSTRACT Hydrophobic organic chemicals (HOCs) are of special ecotoxicological

17

concern because they can be directly incorporated and bio-concentrated in living

18

organisms. However, the effects of self-clustering of HOCs on their environmental

19

behavior and toxicology have not yet received enough attention. With the use of a

20

recently developed technique, single-molecule fluorescence microscopy, the motion and

21

distribution of perylene nano-clusters (PNCs) formed in water at very low concentration

ACS Paragon Plus Environment

1

Environmental Science & Technology

Page 2 of 33

22

(1 µM) were visualized with high temporal and spatial resolution. The liquid-solid

23

interface process of PNCs adsorbing onto graphene was also recorded. Instead of the

24

traditional view of HOCs adsorption as a single molecule, our study revealed the

25

characteristic of irreversible adsorption of perylene onto the carbonaceous surface in the

26

form of nano-clusters, exhibiting random sequential ‘car-parking’ events. More

27

interestingly, the transport of PNCs across the cell membrane was also captured in

28

real-time, demonstrating that they entered macrophage cells by endocytosis.

29

Supplementing the well-recognized routine of passive diffusion through a membrane

30

lipid bilayer, the uptake of HOCs in the form of nano-clusters by endocytosis is proposed

31

to be an additional but important mechanism for their uptake into living cells. HOCs

32

distributing in the environmental systems in the form of nano-clusters, exemplified by

33

PNCs in this study, may have significant implications for understanding their

34

environmental fate and potential toxicological effects.

35

TOC abstract

ACS Paragon Plus Environment

2

Page 3 of 33

Environmental Science & Technology

36 37

INTRODUCTION

38

Hydrophobic organic chemicals (HOCs) are of special ecotoxicological concern

39

among numerous anthropogenic chemicals due to their capacity to incorporate directly

40

into the tissues of living organisms1-2. Resistant to biodegradation, many HOCs may be

41

bio-concentrated as a result of continuous exposure to a contaminated environment.3

42

Furthermore, HOCs’ accumulation may occur in predators and humans by

43

biomagnification along food chains due to the ingestion of contaminated foodstuffs.4-5

44

The species of a given HOC is extremely important because it determines behaviors

45

such as persistence, transport, transformation and toxicology in the environment.6-7 When

46

addressing the species of HOCs in environment, classic books and previous studies

47

usually place emphasis on the chemical structure of these molecules and their

ACS Paragon Plus Environment

3

Environmental Science & Technology

Page 4 of 33

48

consequential environmental behaviors.8 Scientists generally consider HOCs are

49

uniformly distributed within a solid phase in single molecular form, for instance, a plant

50

cuticle or mammalian adipose tissue.9-11 Consequently, all environmental studies of

51

partitioning, toxicity and modeling are based on the assumption that the partitioning

52

processes of environmental contaminants are uniform behaviors of single molecules,8,12

53

although with no direct evidence supporting such hypothetical prerequisite. The effects of

54

an important physical state, namely the self-clustering (also as self-aggregation or

55

self-coiling) of HOCs, on their environmental behavior and ecotoxicology have not yet

56

received attention, and relatively few papers paying attention to this subject have been

57

published.9,11

58

However, in vivo visualization of the HOCs taken up into living vegetation and other

59

matrixes showed molecular clustering within specific plant cellular structures.10-11 Further

60

studies revealed that HOCs can form nano- and micro-clusters over time in compatible

61

lipid media such as eicosane, dotricontane and even polymeric lipids in leaf cuticle.9,11

62

Actually, aggregation of HOCs to molecular clusters and nano-aggregates is an important

63

phenomenon in many regions, such as in the ambient troposphere, natural waters, and

64

even in the interstellar medium.13-17 Hydrophobic-lipophilic interaction (HLI), is one of

65

the most important and essential driving forces for the formation of these micro- or

66

nano-sized molecular clusters, aggregates, micelles, and vesicles, which play important

67

roles in affecting the chemical reactivity and bioprocesses of organic substrates.18-19

68

Many HOCs (such as polycyclic aromatic hydrocarbons (PAH), heterocyclic compounds,

ACS Paragon Plus Environment

4

Page 5 of 33

Environmental Science & Technology

69

and organometallic compounds) are strongly hydrophobic and will form nano-clusters in

70

water even at very low concentration, given that the partition coefficients of HOCs

71

between themselves and water can reach the order of 109.9, 19-21

72

The exciting development of single-molecule fluorescence microscopy (SMFM)

73

makes in situ visualization of fluorescence chemicals possible. In contrast with ensemble

74

methods, SMFM enables the scientist to examine individual members of a heterogeneous

75

population and discover the information on distributions and time trajectories of

76

observables that would otherwise be hidden. 22-23 By performing SMFM, many biological

77

process, such as the infection of individual viruses, endocytosis of nano-particle, RNA

78

catalysis and folding, and transcription factor dynamics can be detected, sorted, and

79

quantitatively compared among their subpopulations.24-28

80

Employing SMFM, this study used perylene as a chemical probe to investigate HOC

81

self-clustering and partitioning in real time between water, graphene and macrophage

82

cells. The high sensitivity of SMFM made it possible to detect single or

83

low-copy-number of perylene naon-clusters (PNCs), with a time resolution on

84

millisecond scale.29 SMFM can also be used to real-time visualize the partitioning

85

process of PNCs between different phases. Perylene, a five ring PAH, was selected as the

86

probe reagent. Perylene primarily derives from organic matter diagenesis and

87

anthropogenic processes (e.g. incomplete combustion due to cigarette smoke and engine

88

exhaust).30-32 It is considered as an important PAH in estuarine sediments.33 It has been

89

widely distributed in many natural waters (0-520 ng/l), wastewaters (0.03-3.0 µg/l) and

ACS Paragon Plus Environment

5

Environmental Science & Technology

Page 6 of 33

90

sewage sludges (140-6400 µg/kg).34 Moreover, its high quantum yield (0.94) and

91

excitation spectra at visible light region make perylene an ideal probe for fluorescence

92

microscope. Carbonaceous media is considered as the major sink for HOCs in

93

environment.8 Graphene was used as a simplest surrogate of carbonaceous media in

94

interfacial behavior of PNCs between water and condensed carbonaceous phase.

95

Graphene is a single atom thick sheet of carbon, which allows the light to freely pass

96

through.35 In addition, graphene shows excellent performance and potential future in

97

adsorptive removal of many HOCs and their derivatives.36-37 In cell enrichment process,

98

murine macrophage cell line was chosen since macrophage is a “first responder” to all

99

foreign materials and is widely used for studying various aspects of foreign materials

100

uptake and potentially related toxicity.38-40 In summary, the formation of PNCs in water,

101

its characteristic adsorption onto graphene, and the new pathway of uptake by cell, which

102

were to be revealed in this study, may have significant implications for understanding

103

HOCs environmental fate and potential toxicological effects.

104

MATERIALS AND METHODS

105

Chemicals and Reagents. Perylene (Sw =1.59 nM; logKOW=5.82) and the dye DiO were

106

purchased from Sigma. Dulbecco’s modified eagle medium (DMEM), fetal bovine serum

107

(FBS), phosphate buffer solution (PBS), 5% trypsin, and phenol red free DMEM were

108

purchased from Gibco. Graphene was obtained from Jichang Nano Company (China).

109

Highly oriented pyrolytic graphite (HOPG) was purchased from NT-MDT Company

110

(Russia). Ultrapure water was used throughout this work. The stock solution of 0.1 mM

ACS Paragon Plus Environment

6

Page 7 of 33

Environmental Science & Technology

111

perylene was prepared using ethanol and stored at 4 °C. Before use, the perylene solution

112

was prepared by diluting the stock solution 100 times using ultrapure water.

113

Emission spectra of perylene in water solution

114

The fluorescence spectra of perylene solutions (0.1 nM, 1 nM, 0.01µM, 0.1 µM, 1 µM, 5

115

µM) were recorded using a fluorescence spectrophotometer (Hitachi, F-4600). The

116

solutions (1.0 mL) in a 4 mL tube were excited at 405 nm, and the emission spectra at

117

420-700 nm were recorded.

118

Visualization of PNCs in water

119

Fluorescence microscopy. The real-time Brownian movement of perylene clusters in

120

1µM of perylene solution (1% ethanol) was observed using a fluorescence microscope

121

(Zeiss, Axio Observer). Perylene clusters were excited by a 405 nm laser (Coherent OBIS

122

405), and monitored with the filter set (dichroic mirror 405 nm, emission 570-640 nm).

123

The fluorescent emission was collected by an 100× oil-immersion objective with

124

numerical aperture of 1.46 and imaged onto an electron multiplying charge coupled

125

device (EMCCD) camera (Photometrics, Evolve 512). Image series were recorded at 20

126

frames per second using Epi-illumination.

127

To determine the distribution of the molecule number of individual PNCs, we first

128

calculated the fluorescence contribution of one perylene molecule, I0. One microliter of

129

perylene solution was evenly coated onto a 5mm×5mm area of the coverslip. The

130

coverslip was air dried avoid of light, and was observed using the Zeiss microscope.

131

The calculation of the molecular number of an individual PNC was based on the

ACS Paragon Plus Environment

7

Environmental Science & Technology

Page 8 of 33

132

assumption that the fluorescent contribution of each individual perylene molecule in

133

different sized PNCs is identical. The fluorescent contribution of free perylene molecules

134

was neglected here, since free-state molecules only accounted for 0.1 percent of the total

135

perylene molecules. The fluorescence intensity of one perylene molecule in the cluster, I0

136

was calculated from the following equations of (1) and (2):

137

I0 = (

138

N = N ACV

(1)

A × I)÷ N A'

(2)

139

Where N, NA, c, and V present total molecular number of perylene on the coverslip,

140

Avogadro's constant (6.02×1023), the concentration of perylene solution (1µM) and the

141

volume of perylene solution smeared on the coverslip (1 µL); A, A’, and I refer to the area

142

of the smeared place on coverslip (25 mm2), the area of the captured images by

143

microscope (0.64 mm2), and the averaged fluorescence intensity of the captured images,

144

respectively. The average fluorescence intensity of the captured image (80µm×80µm)

145

was obtained by randomly recording 100 different images in the coated area, and then the

146

average value was taken.

147

Then the images of PNCs in 1 µM of perylene solution were captured using the

148

same fluorescence microscope. The fluorescence intensity of each PNC was extracted,

149

and presented as Ii. A control experiment showed that immersing in water or drying in air

150

has not effect on the fluorescence intensity of a PNC. Thereby, the number of molecules

151

(Ni) for one individual PNC (Ii) in water can be calculated using the equation (3).

ACS Paragon Plus Environment

8

Page 9 of 33

Environmental Science & Technology

Ni =

152

Ii I

(3)

0

153

Atomic force microscopy. Atomic force microscopy was used to further characterize

154

the shape and size of PNCs. Perylene solution (1 µM) was added dropwise onto highly

155

oriented pyrolytic graphite (HOPG) substrate. The HOPG surface was loaded with

156

isolated PNCs via incubating for 1.0 h. Bruker MultiMode 8.0 was used to record the

157

images in tapping mode using a J-scanner and Veeco NP cantilevers with a spring

158

constant of 0.06 N·m-1. Images were recorded under aqueous conditions with a silicone

159

rubber O-ring.

160

Visualization of PNCs adsorption onto graphene. Graphene (1 cm×1 cm) was coated

161

onto a glass coverslip using the method from Suk et al.,41 the Raman spectra of which is

162

shown in Fig. S4. The coverslip with graphene was put on the stage of fluorescence

163

microscope (Zeiss, Axio Observer). Subsequently, a drop of 1 µM perylene solution was

164

added onto the top of graphene. Then the adsorption process was recorded by the

165

fluorescence microscope. Different from visualization of PNCs in water, the illumination

166

was switched to total internal reflection (TIRF) by using Laser TIRF 3 of Zeiss, to avoid

167

the interference of floating PNCs in solution. Image series were recorded at 20 frames per

168

second for analyzing individual PNC adsorption behavior.

169

Visualization of PNCs uptake by macrophage cells

170

Cell cultures. The murine macrophage J774A.1 cell line was purchased from Cell

171

Resource Center, Chinese Academy of Medical Sciences & Peking Union Medical

ACS Paragon Plus Environment

9

Environmental Science & Technology

Page 10 of 33

172

College. The cells were cultured in DMEM supplemented with 10% FBS in an incubator

173

at 37 ºC with 5% CO2.

174

Real-time tracking the endocytosis of PNCs by the macrophages. Before the

175

fluorescence imaging, cells grown in the Petri dish with glass coverslip on the bottom

176

were washed with PBS. One milliliter of phenol red free DMEM was added to the dish.

177

The process of endocytosis of PNCs was recorded by the fluorescence microscope (Zeiss,

178

Axio Observer) after 1 mL of phenol red free DMEM containing 1µM perylene was

179

added. Illumination was switched to TIRF using Laser TIRF 3 of Zeiss. Image series

180

were recorded at 1 frame per 30 seconds for 15 minutes using time lapse imaging.

181

Colocalization. Cells grown on the Petri dish with glass coverslip on the bottom

182

were washed with phosphate buffer solution and incubated with 16 µM DiO for 15 min at

183

37°C to label the cell membrane. After that, cells were washed with PBS and incubated in

184

phenol red free DMEM containing 1 µM perylene for 15 min. Then the cells were

185

washed again with PBS. The DiO labeled cells were imaged by laser confocal

186

microscope (Zeiss, LSM 710 NLO & DuoScan System). The PNCs were excited with a

187

405 nm laser, and monitored with the filter set (dichroic mirror 405 nm, em 570-640 nm).

188

DiO was excited with a 488 nm laser, and monitored using another filter set (dichroic

189

mirror 488 nm, em 500-520 nm). The fluorescent emission was collected by an 40×

190

objective.

191

PFA fixed cells. To inhibit metabolism, cells were incubated with 4% PFA for 30 min

192

at 37°C. After that, the treatment was withdrawn and cells were washed using PBS. Cells

ACS Paragon Plus Environment

10

Page 11 of 33

Environmental Science & Technology

193

were then incubated in phenol red free DMEM containing 1 µM perylene for 15 min.

194

Before fluorescence imaging, the cells were washed with PBS. Images were recorded

195

using Axio Observer Zeiss microscope with Epi-illumination (excitation 405 nm;

196

emission 570-640 nm for PNCs, 435-485 nm for single molecular perylene).

197 198

RESULTS AND DISCUSSION

199

Characterization of PNCs in water. In our study, SMFM was used first to visualize the

200

PNCs formed in water by precipitation. When the perylene concentration in water is well

201

above 1.59 nM of its water solubility at 25 ºC, the molecular forces exhibited between

202

conjugate aromatic molecules, such as π-π interaction and hydrophobic-lipophilic

203

interactions, will lead to self-clustering of these molecules and formation of PNCs (Fig.

204

1A). As indicated in Fig. 1B, the PNCs exhibited unique optical properties differing from

205

single-molecular perylene by a new emission peak at 550~650 nm, while the emission

206

peak of single molecular perylene is around 435~470 nm. This new emission peak

207

(550~650 nm) was identified as a characteristic peak for real-time monitoring PNCs in

208

our study. The red-shift of emission maxima of PNCs compared with individual perylene

209

molecule was also observed in previous studies.21, 42

ACS Paragon Plus Environment

11

Environmental Science & Technology

Page 12 of 33

210 211

Figure 1. Characterization of PNCs. (A) Ball-stick model of perylene and the simulated

212

nano-cluster. (B) Fluorescence spectra of perylene solution at different concentrations

213

(0.1 nM, 1 nM,0.01µM, 0.1 µM, 1 µM, 5 µM) excited at 405 nm. The emission peak

214

ranging from 550 nm to 650 nm is the characteristic peak of PNCs. (C) Atomic force

215

microscopy image of PNCs (1 µM) loaded on HOPG surface. Image was recorded under

216

aqueous conditions. The inset is the section of one typical PNC (with white arrow) on

217

HOPG. The scale bar is 1 µm. (D) Distribution histogram of the height and radius of

ACS Paragon Plus Environment

12

Page 13 of 33

Environmental Science & Technology

218

PNCs deposited onto HOPG. (E) A representative Brownian motion trail of PNC (2.0 s)

219

extracted from continuous imaging of PNCs in water (20 frames per second). (F)

220

Distribution histogram of the number of molecules in PNCs (1 µM). The results are

221

calculated from the analysis of single PNC’s fluorescence intensity using equations (1-3).

222 223

The Brownian motion of PNCs in water was real-time recorded by SMFM (Movie 1).

224

The fluorescent intensity for an individual PNC was presented as maxima as it moved

225

within the focusing plane of microscope. When it moved vertically out of the focal plane,

226

the PNC particle in image became fuzzy with decreasing fluorescent intensity, and finally

227

disappeared. The Brownian motion trail of one typical PNC in water was captured, and

228

shown in Fig. 1E. From the Brownian motion, we calculated the mean hydrodynamic

229

radius (RH) of PNCs in 1 µM perlylene solution based on the Stoke-Einstein relation:24, 43

RH =

230

K BT 6πηD

(4)

231

Where KB = Boltzmann constant (1.38×10-23 J/K); T = Temperature (298K), η =

232

viscosity of the solvent (0.8937×10-3 Pa·s). D is a factor of proportionality with direct

233

relationship to the solubility and molecular weight of the molecule. The slope of the

234

following equation gives the value of D:

235 236 237

4 D(ti − t 0 ) = [(R(ti ) − R(t 0 )]2

(5)

Where t0 = initial time, and ti = time after t. R refers to a two-dimensional vector indicating the spatial location of the center of mass of the cluster.

ACS Paragon Plus Environment

13

Environmental Science & Technology

Page 14 of 33

238

The mean hydrodynamic radius (RH) of PNCs in 1 µM perlylene solution was

239

calculated as 42.3 nm from capturing 30 individual clusters. Atomic force microscopy

240

(AFM) images of PNCs at the same perylene concentration deposited onto HOPG (Fig.

241

1C) showed the mean radius of 49 nm, which is in agreement with the average value of

242

RH obtained from calculation. The distribution histogram of the height and radius of

243

PNCs deposited onto HOPG is shown in Fig. 1D. AFM provided further evidence of the

244

disk-like shape of PNCs deposited on the HOPG, with average thickness of 4.03 nm.

245

The distribution of molecular number of PNCs can be calculated from equation

246

(1)-(3) built above in the experimental section. Obtained from analysis of at least 100

247

individual PNCs, the distribution histogram of the molecular number of PNCs in 1.0 µM

248

perylene solution is shown in Fig. 1F. The histogram was well fitted with a Gaussian

249

distribution and the mean molecular number per PNC was 1.03×105. After the

250

verification of cluster form of perylene in water, the settleability of PNCs was determined

251

by centrifugation at different rotating speed. Even with 5000 rpm of rotating rate, the

252

fluorescence intensity of the supernatant derived from PNCs only decreased 12%,

253

demonstrating a high stability of PNCs formed in water (Fig. S1).

254

The formation and aggregation state of HOCs in the aquatic environment should be

255

dependent on a multiple of influencing processes and factors, including the equilibrium

256

between clustering and dis-clustering, the chemical structure of different HOCs, the

257

different HOC sources, the broad range of HOC concentrations and solubility, complex

258

water chemistry, water mixing and so on.44

ACS Paragon Plus Environment

14

Page 15 of 33

Environmental Science & Technology

259

Using SMFM, we visualized in real-time and characterized the PNCs formed in

260

1 µM perylene solution, the content of which is well above its water solubility. PNCs in

261

0.1 µM and 0.01 µM (only 6.3 times of its solubility) perylene solution were recorded too,

262

although with significant decrease in counting number of PNCs in SMFM images. This

263

observation is reasonable and predictable, given the water solubility of perylene is at

264

nanomole level (Sw =1.59 nM). The total concentration of perylene can be up to

265

0-2.06 nM in virgin natural water body, 11.9 nM for wastewaters and 25.4 nM from

266

sewage sludges.34 It was reported that the total PAHs concentration in many rivers

267

reached a few or tens of thousands of nanograms per liter, and generally PAHs occurrence

268

was even more prevalent in sediment of these rivers.45-48 Considering the partition

269

coefficients of many HOCs between themselves and water can reach the order of 109, we

270

propose here that most of these HOCs are essentially distributed as homogeneous or

271

heterogeneous forms of micro-aggregates and nano-clusters in these aqueous systems due

272

to hydrophobic-lipophilic and π-π interactions between themselves and each other.9, 19

273

These HOCs clusters in the aqueous environment and their associations with some other

274

carbonaceous and inorganic solid phases are likely to account for a significant fraction of

275

HOC-suspended solids (HOC-SS), which are generally considered as the major species of

276

most HOCs distributed in natural water columns.

277

The occurrence and distribution of these HOCs micro- and nano-clusters in the

278

aqueous systems should have important implications for understanding their

279

environmental fate and their potential toxicological effects. Here, the PNCs formed in

ACS Paragon Plus Environment

15

Environmental Science & Technology

Page 16 of 33

280

water by a simple method of precipitation were visualized and characterized using

281

SMFM, which thereby stimulated us to address the important implications of molecular

282

clustering phenomenon of HOCs in fields of environmental partitioning and related

283

ecotoxicology. The adsorption process of PNCs from water onto graphene, and a specific

284

cell enrichment process, endocytosis uptake of PNCs by macrophage, were presented in

285

the following sections.

286

Adsorption process of PNCs from water onto graphene. After the characterization of

287

molecular clusters was identified, the adsorption behavior of perylene onto the graphene,

288

as presented in Movie 2, was exhibited as the irreversible adsorption (or deposition) of

289

colloids or particles onto a carbonaceous surface instead of the traditional

290

single-molecular adsorption model. Fig. 2 shows the representative images of PNCs

291

sequestered by graphene at different time, which clearly revealed a random sequential

292

adsorption of PNCs onto the substrate. The time course of the PNCs adsorbing onto

293

graphene was like a ‘car-parking’ event, neither with further PNCs migration on the

294

surface nor immediate desorption, indicating a strong sticking-force exhibited between

295

PNCs and graphene substrate. The increase in the number of adsorbed PNCs on graphene

296

and the change of total fluorescence intensity was recorded over a focal area of 80 µm ×

297

80 µm with time-lapse, the typical results of which are shown in Fig. 3A. The adsorption

298

of PNCs was rapid at the first 10 seconds, and then slowed down considerably. The total

299

fluorescence intensity was positively correlated with the number of adsorbed PNCs,

300

exhibiting an obvious stepwise growth curve as captured by SMFM with high temporal

ACS Paragon Plus Environment

16

Page 17 of 33

Environmental Science & Technology

301

resolution. Interestingly, the increase in PNCs counts with time was not a monotonically

302

increasing curve, probably due to a phenomenon of overshooting generally observed in

303

colloid and polymer adsorption.49-50

304 305

Figure 2. Representative fluorescence images of PNCs (1 µM) adsorbed onto graphene

306

surface in solution for 0 s, 5 s, 10 s, 15 s, 20 s, and 25 s. All Scale bars are 10 µm.

307 308

Owing to the high sensitivity of time resolution offered by SMFM, the individual

309

adsorption behavior of PNC can be visualized. Fig. 3B recorded the irreversible and

310

independent adsorption of an individual PNC onto the surface of graphene. We captured

311

2-4 images during the processes of a PNC moving close to the graphene surface,

312

modifying its position during the process until final adsorption. This adsorption process

ACS Paragon Plus Environment

17

Environmental Science & Technology

Page 18 of 33

313

of PNCs on graphene typically took 100-150 ms. Although most of PNCs were adsorbed

314

stably (Fig. 3B), desorption process of a small fraction of PNCs was also captured.

315

Fig 3C recorded one PNC shot the graphene substrate and then escaped immediately. In

316

Fig. 3D, the PNC was recorded respectively with 2 times of steep rise in fluorescence

317

intensity, spaced by gradual or sharp decline of fluorescence intensity at intervals. It is

318

speculated that the steep rise in fluorescence intensity was caused by the overlapped PNC

319

adsorption. The sharp decline of fluorescence intensity for an adsorbed PNC was due to

320

the detachment of smaller PNCs away from the original one. The gradual decline of the

321

original was probably as a result of gradual dissolution or dis-aggregation of PNC from

322

the substrate.

4

80 Count

(B)

100

60

40

40 Count Intensity

20 5

10 Time (s)

15

Intensity(X10 )

60

6

(A)

20 0 20

4

2

0

0

3.0

323

2.4

4

Intensity(X10 )

4

Intensity(X10 )

0.8 0.4 0.0

1000 1500 2000 2500 Time (ms)

(D)

(C) 1.2

500

0

500

1000 1500 2000 2500

1.8 1.2 0.6 0.0

0

Time (ms)

ACS Paragon Plus Environment

2000

4000

6000

Time (ms)

18

Page 19 of 33

Environmental Science & Technology

324

Figure 3. (A) Adsorption kinetics of perylene adsorbed on graphene. The green squares

325

show the number of adsorbed PNCs on graphene surface. The pink circles show the total

326

fluorescence intensity of adsorbed PNCs on graphene surface. Original fluorescence

327

image series were recorded at 20 frames per second. (B-D) Adsorption process of

328

individual PNCs on graphene extracted from fluorescence images. The figures describe

329

the dynamic fluorescence intensity of three individual PNCs deposited on graphene

330

during the imaging process.

331 332

Adsorption to the condensed phase is affects the transport, degradation,

333

bioavailability, fate and ecological risk presented by HOCs in the environment. Most, if

334

not all the previous investigations of HOCs adsorption onto the solid phases were

335

performed in an almost fixed condition: the concentration range of HOCs in the aqueous

336

phase was below the solubility of that compound in water.51 Under this restricted

337

paradigm, the adsorption process of HOCs is artificially simplified to a uniform behavior

338

of dissolved single molecule partitioning between the aqueous and the solid phase,

339

without considering the effect of self-clustering of HOCs. Obviously, this simplified

340

experimental frame does not reflect the truth because the discharged level of HOCs into

341

the accepting water body is very often beyond this dissolubility limit.44 As verified in our

342

study, nano-cluster forms of perylene could exist in aqueous solution even at 0.01 µM.

343

With high spatial and temporal resolution, SMFM recorded here the liquid-solid interface

344

process of PNCs with the characteristics of colloids adsorption (deposition) and

ACS Paragon Plus Environment

19

Environmental Science & Technology

Page 20 of 33

345

desorption (detachment), but not the uniform behavior of single molecules as generally

346

assumed.8, 12 As a consequence, taking into account HOC clustering in water and its effect

347

on adsorption will provide a more precise framework for building models that quantify

348

the dynamic behavior of organic chemicals in environmental systems.

349

Observation of PNCs formed in water and their random and irreversible deposition

350

onto the carbonaceous surface has important implications in predicting HOC behaviors in

351

the aqueous environment. Clustering of HOCs in aquatic environment would lead to the

352

increase in the discreteness and heterogeneity of HOCs distributed in the liquid-solid

353

interface. Secondly, traditional view of single molecular HOC adsorption/desorption

354

(A/D) onto the sediments cannot sufficiently explain the impact of hydrodynamic force

355

and perturbation on HOCs release from the sediments into the water column.52-54 The

356

colloidal property of HOCs clusters depositing onto the solid phase might partly explain

357

the significant release of HOCs from sediments by the external forces, such as waves,

358

tides, currents, dredging, trawling and ship traffic. Furthermore, A/D hysteresis is

359

considered as a general phenomenon for HOCs adsorbed onto the solid phases.55-56

360

Besides the previously well-proposed mechanisms,57-59 perylene self-clustering and the

361

resulting irreversibility of PNCs adsorbed onto graphene shown here might present a new

362

but important interpretation to A/D hysteresis.

363 364

PNCs transport through cell membrane. It is generally considered that HOCs pass

365

through the lipid bilayer by passive diffusion.51 However, the formation and stability of

ACS Paragon Plus Environment

20

Page 21 of 33

Environmental Science & Technology

366

PNCs in water even at very low concentration intrigued us to investigate if perylene as a

367

typical HOC can be transported into the cell by endocytosis. Laser scanning confocal

368

microscopy was used to image whether the PNCs were incorporated into the macrophage

369

(Fig. 4A). The cells were first labeled with DiO, a fluorescent lipophilic dye that is

370

commonly used to image cell membranes. The image confirmed that PNCs were located

371

in the intracellular space.

372 373

Figure 4. (A) Merged image of PNCs (red) and DiO (green). Images on right and bottom

374

represent the intrinsic 3D sectioning of the image. The scale bar is 10 µm. (B) Merged

ACS Paragon Plus Environment

21

Environmental Science & Technology

Page 22 of 33

375

images of the uptake and transport of PNCs (red) in J774A.1 cell. The trails of PNCs are

376

marked by blue lines. (C-D) Merged image of PNCs (red) and single molecular perylene

377

(blue) in living cells (C) and 4% PFA fixed cells (D). (C) and (D) are taken under the

378

same imaging conditions. Scales bars are 10 µm. Obvious signals of PNCs could be seen

379

in the living cells (C) but not in the fixed cells (D).

380 381

Using SMFM, we also investigated the transport of PNCs in cells by tracking single

382

PNCs in living cells. Total internal reflection (TIRF) illumination was used here to avoid

383

the interference from PNCs floating in solution or adsorbed onto the cell membrane.

384

Movie 3 shows the processes of PNCs entering and transferring in the macrophage during

385

15 min of real time visualization. Fig. 4 (B) traced the trajectory of an individual PNC

386

moving from the apical domain to the midst of cell.

387

If the incorporation of perylene into the macrophage was primarily due to

388

endocytosis, it should be significantly disrupted by metabolic inhibitor paraformaldehyde

389

(PFA), since endocytosis uptake of macromolecules or particles is an energy dependent

390

process. To confirm this assumption, the uptake of PNCs by the macrophage was studied

391

with the pretreatment of 4% PFA or not. Since PNCs is characterized by the new red shift

392

emission peak, PNCs and single molecular perylene can be captured separately. Fig.

393

4(C–D) shows the fluorescence image of PNCs (em 570-640 nm) and single molecular

394

perylene (em 435-485 nm) in the living (Fig. 4C) and fixed (Fig. 4D) cells incubated with

395

1 µM of perylene for 15 min. As expected, intracellular accumulation of PNCs in 4% PFA

ACS Paragon Plus Environment

22

Page 23 of 33

Environmental Science & Technology

396

pretreated cells was completely blocked compared with the living cell. The accumulation

397

of single-molecular perylene was also dramatically reduced for the PFA-pretreated cells

398

(Fig. S2-3). Note here that the significant reduction in uptake of single-molecular

399

perylene by the PFA-pretreated macrophage should not be explained by the blocking of

400

spontaneous diffusion of small molecule through lipid bilayer because PFA does not

401

affect on this. This was perhaps due to the obstruction of endocytosis uptake of PNCs.

402

These were to be dis-aggregated to single-molecular perylene with the involvement of

403

various intracellular components (i.e., lipoproteins and nonpolar protein domains), and

404

subsequently partitioned into different subcellular compartments.

405

Passive diffusion through the lipid bilayer is usually accepted as the primary route

406

for HOCs incorporation into organisms.60-61 This is certainly a reasonable assumption

407

when the concentration of a specific HOC is less than its solubility in that liquid matrix,

408

because HOC in single molecular form is compatible with the lipid bilayer and can

409

diffuse freely through the cell membrane. However, in the blood plasma, most HOCs (i.e.,

410

Benzo(a)pyrene

411

lipoprotein-associated physiological form, the cellular uptake of BaP and other HOCs

412

was demonstrated to be a simple and spontaneous diffusion through the cell

413

membrane..53,A wide variety of organism cells, such as epithelial cells of the skin,

414

digestive tract and lungs, are exposed to the aquatic environment in the absence of

415

lipoproteins and some such dynamic carriers.60 These cells act as basic physical barriers,

416

through which only various contaminants can enter into the body from an outside aquatic

(BaP))

are

noncovalently

associated

ACS Paragon Plus Environment

with

lipoproteins.

In

23

Environmental Science & Technology

Page 24 of 33

417

environment. Additionally, many single-celled eukaryotic species, such as algae and

418

protozoa, live in water environment.62 As shown in this study, PNCs were stably formed

419

in water and accessible to cell via endocytosis uptake even at very low incubated doses.

420

Therefore, the endocytosis uptake of cluster forms is proposed here to be an additional

421

but important mechanism for HOCs bio-concentration into living cells. Our speculation

422

on endocytosis uptake of HOCs was supported by recent research from Fayeulle et al.

423

showing F. solani used an ‘endocytosis-like’ process for BaP uptake dependent on

424

energy.63 This is contrary to the general consensus of BaP uptake by diffusion.

425

The endocytosis uptake of HOCs in cluster form, as PNCs observed here, would

426

enhance the quantity of pollutant incorporated into cells even though these compounds

427

have very low water solubility in aquatic environments. This process may permit faster

428

incorporation of HOCs into cells and may further influence their intracellular localization

429

and toxicological properties. Further investigations are required to address the dynamics

430

and mechanisms of intracellular distribution, transport, dis-aggregation and partitioning

431

of these clusters between different cell compartments.

432

Supporting Information Available

433

Fluorescence intensity of perylene in supernatant after centrifugation with different

434

rotating speed, supplemented images of PNCs (red) and single molecular perylene (blue)

435

in living and fixed cells, comparison of the fluorescence intensity from single molecular

436

perylene in living and fixed cells, Raman spectra of graphene on coverslip, the movie of

437

Browain motion of PNCs, the movie of PNCs adsorbing onto graphene, the movie of

ACS Paragon Plus Environment

24

Page 25 of 33

Environmental Science & Technology

438

uptake and transport of PNCs in J774A.1 cell. These materials are available free of

439

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

440

Acknowledgments

441

This work was supported by National Natural Science Foundation of China (NSFC)

442

(Grant Nos. 41371440) and the National Key Basic Research Program of China

443

(2013CB430406). FB also acknowledge the financial support from the Recruitment

444

Program of Global Youth Experts.

445 446

REFERENCES

447

(1) Minh, T. B.; Watanabe, M.; Tanabe, S.; Yamada, T.; Hata, J.; Watanabe, S. Occurrence

448

of tris(4-chlorophenyl)methane, tris(4-chlorophenyl)methanol, and some other persistent

449

organochlorines in japanese human adipose tissue. Environ. Health Perspect. 2000, 108

450

(7), 599-603.

451

(2) Hendriks, A. J.; Van der Linde, A.; Cornelissen, G.; Sijm, D. The power of size. 1. rate

452

constants and equilibrium ratios for accumulation of organic substances related to

453

octanol-water partition ratio and species weight. Environ. Toxicol. Chem. 2001, 20 (7),

454

1399-1420.

455

(3) Streit, B. Bioaccumulation Processes in Ecosystems. Experientia. 1992, 48 (10),

456

955-970.

457

(4) Kelly, B. C.; Ikonomou, M. G.; Blair, J. D.; Morin, A. E.; Gobas, F. A. P. C. Food

458

web–specific biomagnification of persistent organic pollutants. Science. 2007, 317 (5835),

ACS Paragon Plus Environment

25

Environmental Science & Technology

Page 26 of 33

459

236-239.

460

(5) Burreau, S.; Zebuhr, Y.; Broman, D.; Ishaq, R. Biomagnification of PBDEs and PCBs

461

in food webs from the baltic sea and the northern atlantic ocean. Sci. Total Environ. 2006,

462

366 (2-3), 659-672.

463

(6) Lim, S.J.; Fox, P. Estimating the persistence of organic contaminants in indirect

464

potable reuse systems using quantitative structure activity relationship (QSAR). Sci. Total

465

Environ. 2012, 433, 1-7.

466

(7) Altenburger, R.; Nendza, M.; Schuurmann, G. Mixture toxicity and its modeling by

467

quantitative structure-activity relationships. Environ. Toxicol. Chem. 2003, 22 (9),

468

1900-1915.

469

(8) Schwarzenbach, R. P.; Gschwend, P. M.; Imboden, D. M. Environmental Organic

470

Chemistry. Wiley- Interscience: New York, 2002.

471

(9) Wild, E.; Cabrerizo, A.; Dachs, J.; Jones, K. C. Clustering of nonpolar organic

472

compounds in lipid media: evidence and implications. J. Phys. Chem. A. 2008, 112 (46),

473

11699-11703.

474

(10) Wild, E.; Dent, J.; Thomas, G. O.; Jones, K. C. Visualizing the air-to-leaf transfer

475

and within-leaf movement and distribution of phenanthrene: Further studies utilizing

476

two-photon excitation microscopy. Environ. Sci. Technol. 2006, 40, (3), 907-916.

477

(11) Li, Q. Q.; Chen, B. L. Organic Pollutant Clustered in the Plant Cuticular Membranes:

478

Visualizing the Distribution of Phenanthrene in Leaf Cuticle Using Two-Photon Confocal

479

Scanning Laser Microscopy. Environ. Sci. Technol. 2014, 48, (9), 4774-4781.

ACS Paragon Plus Environment

26

Page 27 of 33

Environmental Science & Technology

480

(12) Mackay, D. Multimedia Enironmental Models: The Fugacity Approach, 3rd ed.

481

Lewis Publishers: London, 2001.

482

(13) Kato, T.; Kutsuna, T.; Yabuuchi, K.; Mizoshita, N. Anisotropic self-aggregation of an

483

anthracene derivative: formation of liquid-crystalline physical gels in oriented states.

484

Langmuir. 2002, 18 (18), 7086-7088.

485

(14) Cai, W.; Wang, G.; Xu, Y.; Jiang, X.; Li, Z. Vesicles and organogels from foldamers:

486

a solvent-modulated self-assembling process. J. Am. Chem. Soc. 2008, 130 (22),

487

6936-6937.

488

(15) Qian, M. C.; Reber, A. C.; Ugrinov, A.; Chaki, N. K.; Mandal, S.; Saavedra, H. M.;

489

Khanna, S. N.; Sen, A.; Weiss, P. S. Cluster-assembled materials: toward nanomaterials

490

with precise control over properties. ACS Nano. 2010, 4 (1), 235-240.

491

(16) Raj, A.; Sander, M.; Janardhanan, V.; Kraft, M. A study on the coagulation of

492

polycyclic aromatic hydrocarbon clusters to determine their collision efficiency. Combust.

493

Flame. 2010, 157 (3), 523–534.

494

(17) Rapacioli, M.; Calvo, F.; Joblin, C.; Parneix, P.; Toublanc, D.; Spiegelman, F.

495

Formation and destruction of polycyclic aromatic hydrocarbon clusters in the interstellar

496

medium. Astron. Astrophys. 2006, 460 (2), 519–531.

497

(18) Blyth, C. A.; Knowles, J. R. Kinetic consequences of intermolecular attraction.2.

498

hydrolysis of a series of fatty acid para-nitrophenyl esters catalyzed by a series of n

499

alkylimidazoles-very simple esterase model. J. Am. Chem. Soc. 1971, 93 (12),

500

3021-3027.

ACS Paragon Plus Environment

27

Environmental Science & Technology

Page 28 of 33

501

(19) Jiang, X. Hydrophobic-Lipophilic Interactions. Aggregation and self-coiling of

502

organic molecules. Acc. Chem. Res. 1988, 21 (10), 362-367.

503

(20) Shi, J.; Zhang, Y.; Xia, L.; Jiang, X. Do flat molecules without a side-chain readily

504

form aggregates? A fluorescence study of pyrene eximer and exciplex formation on

505

EtOH-H2O and MeOH-H2O solvent systems. Chin. Chem. Lett. 1990, 2 (1), 149-152.

506

(21) Tachikawa, T.; Chung, H. R.; Masuhara, A.; Kasai, H.; Oikawa, H.; Nakanishi, H.;

507

Fujitsuka, M.; Majima, T. In situ and ex situ observations of the growth dynamics of

508

single perylene nanocrystals in water. J. Am. Chem. Soc. 2006, 128 (50), 15944-15945.

509

(22) Weiss, S. Fluorescence spectroscopy of single biomolecules. Science 1999, 283,

510

(5408), 1676-1683.

511

(23) Cornish, P. V.; Ha, T. A survey of single-molecule techniques in chemical biology.

512

Acs Chem. Biol. 2007, 2, (1), 53-61.

513

(24) Martens, T. F.; Remaut, K.; Demeester, J.; De Smedt, S. C.; Braeckmans, K.

514

Intracellular delivery of nanomaterials: How to catch endosomal escape in the act. Nano

515

Today 2014, 9, (3), 344-364.

516

(25) Elf, J.; Li, G. W.; Xie, X. S. Probing transcription factor dynamics at the

517

single-molecule level in a living cell. Science 2007, 316, (5828), 1191-1194.

518

(26) Zhuang, X. W.; Bartley, L. E.; Babcock, H. P.; Russell, R.; Ha, T. J.; Herschlag, D.;

519

Chu, S. A single-molecule study of RNA catalysis and folding. Science 2000, 288, (5473),

520

2048-2053.

521

(27) Lakadamyali, M.; Rust, M. J.; Babcock, H. P.; Zhuang, X. W. Visualizing infection

ACS Paragon Plus Environment

28

Page 29 of 33

Environmental Science & Technology

522

of individual influenza viruses. Proc. Natl. Acad. Sci. U.S.A.

523

9280-9285.

524

(28) Lew, M. D.; Moerner, W. E. Azimuthal polarization filtering for accurate, precise,

525

and robust single-molecule localization microscopy. Nano Lett. 2014, 14, (11),

526

6407-6413.

527

(29) Waters, J. C. Accuracy and precision in quantitative fluorescence microscopy. J. Cell

528

Biol. 2009, 185 (7), 1135-1148.

529

(30) Slater, G. F.; Benson, A. A.; Marvin, C.; Muir, D. PAH Fluxes to Siskiwit Revisted:

530

Trends in Fluxes and Sources of Pyrogenic PAH and Perylene Constrained via

531

Radiocarbon Analysis. Environ. Sci. Technol. 2013, 47, (10), 5066-5073.

532

(31) Booij, K.; Arifin, Z.; Purbonegoro, T. Perylene dominates the organic contaminant

533

profile in the Berau delta, East Kalimantan, Indonesia. Mar. Pollut. Bull. 2012, 64, (5),

534

1049-1054.

535

(32) Cunha, A.; Almeida, A.; Re, A.; Martins, A.; Alcantara, F. Perylene toxicity in the

536

estuarine environment of Ria de Aveiro (Portugal). Ecotoxicology 2006, 15, (2), 171-185.

537

(33) Budzinski, H.; Jones, I.; Bellocq, J.; Pierard, C.; Garrigues, P. Evaluation of

538

sediment contamination by polycyclic aromatic hydrocarbons in the Gironde estuary. Mar.

539

Chem. 1997, 58, (1-2), 85-97.

540

(34) Monograph on the Evaluation of the Carcinogenic Risk of Chemicals to Humans.

541

Polynuclear Aromatic Compounds. Part I: Chemical Environmental and Experimental

542

Data. Vol. 32: 412; International Agency for Research on Cancer: France, 1983;

ACS Paragon Plus Environment

2003, 100, (16),

29

Environmental Science & Technology

Page 30 of 33

543

http://monographs.iarc.fr/ENG/Monographs/vol1-42/mono32.pdf.

544

(35) Novoselov, K. S.; Jiang, D.; Schedin, F.; Booth, T. J.; Khotkevich, V.V.; Morozov, S.

545

V.; Geim, A. K. Two-dimensional atomic crystals. Proc. Natl. Acad. Sci. U.S.A. 2005, 102

546

(30), 10451-10453.

547

(36) Ji, L.; Chen, W.; Xu, Z.; Zheng, S.; Zhu, D. Graphene nanosheets and graphite oxide

548

as promising adsorbents for removal of organic contaminants from aqueous solution. J.

549

Environ. Qual. 2013, 42 (1), 191-198.

550

(37) Wang, J.; Chen, Z.; Chen, B. Adsorption of polycyclic aromatic hydrocarbons by

551

graphene and graphene oxide nanosheets. Environ. Sci. Technol. 2014, 48 (9), 4817-4825.

552

(38) Yen, H. J.; Hsu, S. H.; Tsai, C. L., Cytotoxicity and Immunological Response of

553

Gold and Silver Nanoparticles of Different Sizes. Small 2009, 5, (13), 1553-1561.

554

(39) Yue, H.; Wei, W.; Yue, Z. G.; Lv, P. P.; Wang, L. Y.; Ma, G. H.; Su, Z. G., Particle size

555

affects the cellular response in macrophages. Eur. J. Pharm. Sci. 2010, 41, (5), 650-657.

556

(40) Wang, H. Y.; Wu, L. X.; Reinhard, B. M., Scavenger Receptor Mediated Endocytosis

557

of Silver Nanoparticles into J774A.1 Macrophages Is Heterogeneous. Acs Nano 2012, 6,

558

(8), 7122-7132.

559

(41) Suk, J. W.; Kitt, A.; Magnuson, C. W.; Hao, Y. F.; Ahmed, S.; An, J.H.; Swan, A. K.;

560

Goldberg, B.B.; Ruoff, R.S. Transfer of CVD-grown monolayer graphene onto arbitrary

561

substrates. ACS Nano. 2011, 5 (9), 6916-6924.

562

(42) Weinberger, R.; Love, L. J. C. Luminescence properties of polycyclic aromatic

563

hydrocarbons in colloidal or microcrystalline suspensions. Spectrochim. Acta, Part A.

ACS Paragon Plus Environment

30

Page 31 of 33

Environmental Science & Technology

564

1984, 40 (1), 49-55.

565

(43) Vestergaard, M.; Hamada, T.; Saito, M.; Yajima, Y.; Kudou, M.; Tamiya, E.; Takagi,

566

M. Detection of alzheimer's amyloid beta aggregation by capturing molecular trails of

567

individual assemblies. Biochem. Biophys. Res. Commun. 2008, 377 (2), 725-728.

568

(44) Neff, J. M. Polycyclic Aromatic Hydrocarbons in the Aquatic Environment: Sources,

569

Fates and Biological Effects. Applied Science Publishers: London, 1979.

570

(45) Doong, R. A.; Lin, Y. T., Characterization and distribution of polycyclic aromatic

571

hydrocarbon contaminations in surface sediment and water from Gao-ping River, Taiwan.

572

Water Res. 2004, 38, (7), 1733-1744.

573

(46) Zhang, Z. L.; Huang, J.; Yu, G.; Hong, H. S., Occurrence of PAHs, PCBs and

574

organochlorine pesticides in the Tonghui River of Beijing, China. Environ. Pollut. 2004,

575

130, (2), 249-261.

576

(47) Simpson, C. D.; Mosi, A. A.; Cullen, W. R.; Reimer, K. J., Composition and

577

distribution of polycyclic aromatic hydrocarbon contamination in surficial marine

578

sediments from Kitimat Harbor, Canada. Sci. Total Environ. 1996, 181, (3), 265-278.

579

(48) Pereira, W. E.; Hostettler, F. D.; Rapp, J. B., Distributions and fate of chlorinated

580

pesticides, biomarkers and polycyclic aromatic hydrocarbons in sediments along a

581

contamination gradient from a point-source in San Francisco Bay, California. Mar.

582

Environ. Res. 1996, 41, (3), 299-314.

583

(49) Filippov, L. K. F. Overshoots of adsorption kinetics. J. Colloid Interface Sci. 1996,

584

178 (2), 571-580.

ACS Paragon Plus Environment

31

Environmental Science & Technology

Page 32 of 33

585

(50) Ohshima, H.; Sato, H.; Matsubara, H.; Hyono, A.; Okubo, M. A theory of adsorption

586

kinetics with time delay and its application to overshoot and oscillation in the surface

587

tension of gelatin solution. Colloid Polym. Sci. 2004, 282 (10), 1174-1178.

588

(51) Xia, G. S.; Ball, W. P. Adsorption-partitioning uptake of nine low-polarity organic

589

chemicals on a natural sorbent. Environ. Sci. Technol. 1999, 33 (2), 262-269.

590

(52) Chiou, C. T.; Peters, L. J.; Freed, V. H. Physical Concept of Soil-Water Equilibria for

591

Non-Ionic Organic-Compounds. Science 1979, 206, (4420), 831-832.

592

(53) Ciarelli, S.; van Straalen, N. M.; Klap, V. A.; van Wezel, A. P. Effects of sediment

593

bioturbation by the estuarine amphipod Corophium volutator on fluoranthene

594

resuspension and transfer into the mussel (Mytilus edulis). Environ. Toxicol. Chem. 1999,

595

18, (2), 318-328.

596

(54) Wang, L. L.; Shen, Z. Y.; Wang, H. Y.; Niu, J. F.; Lian, G. X.; Yang, Z. F.

597

Distribution characteristics of phenanthrene in the water, suspended particles and

598

sediments from Yangtze River under hydrodynamic conditions. J. Hazard. Mater. 2009,

599

165, (1-3), 441-446.

600

(55) Pavlostathis, S. G.; Jaglal, K. Desorptive Behavior of Trichloroethylene in

601

Contaminated Soil. Environ. Sci. Technol. 1991, 25, (2), 274-279.

602

(56) Braida, W. J.; Pignatello, J. J.; Lu, Y. F.; Ravikovitch, P. I.; Neimark, A. V.; Xing, B.

603

S. Sorption hysteresis of benzene in charcoal particles. Environ. Sci. Technol. 2003, 37,

604

(2), 409-417.

605

(57) Wu, W. L.; Sun, H. W. Sorption-desorption hysteresis of phenanthrene - Effect of

ACS Paragon Plus Environment

32

Page 33 of 33

Environmental Science & Technology

606

nanopores, solute concentration, and salinity. Chemosphere 2010, 81, (7), 961-967.

607

(58) Carroll, K. M.; Harkness, M. R.; Bracco, A. A.; Balcarcel, R. R. Application of a

608

Permeant Polymer Diffusional Model to the Desorption of Polychlorinated-Biphenyls

609

from Hudson River Sediments. Environ. Sci. Technol. 1994, 28, (2), 253-258.

610

(59) Burgos, W. D.; Novak, J. T.; Berry, D. F. Reversible sorption and irreversible binding

611

of naphthalene and alpha-naphthol to soil: Elucidation of processes. Environ. Sci. Technol.

612

1996, 30, (4), 1205-1211.

613

(60) Klaassen, C.D. Casarett and Doull’s Toxicology: The Basic Science of Poisons.

614

McGraw-Hill: New York, 2008.

615

(61) Plant, A. L.; Knapp, R. D.; Smith, L. C. Mechanism and rate of permeation of cells

616

by polycyclic aromatic hydrocarbons. J. Biol. Chem. 1987, 262 (6), 2514.

617

(62) Corpe, W. A.; Jensen, T. E. The diversity of bacteria, eukaryotic cells and viruses in

618

an oligotrophic lake. Appl. Microbiol. Biotechnol. 1996, 46 (5-6), 622-630.

619

(63) Fayeulle, A.; Veignie, E.; Slomianny, C.; Dewailly, E.; Munch, J.; Rafin, C.

620

Energy-dependent uptake of benzo(a)pyrene and its cytoskeleton-dependent intracellular

621

transport by the telluric fungus Fusarium Solani. Environ. Sci. Pollut. Res. 2014, 21 (5),

622

3515-3523.

ACS Paragon Plus Environment

33