Determination of nanoparticle uptake, distribution and characterization

Apr 9, 2019 - Sandra Wagener , Harald Jungnickel , Nils Dommershausen , Thomas Fischer , Peter Laux , and Andreas Luch. Environ. Sci. Technol. , Just ...
1 downloads 0 Views 782KB Size
Subscriber access provided by UNIV OF LOUISIANA

Ecotoxicology and Human Environmental Health

Determination of nanoparticle uptake, distribution and characterization in plant root tissue after realistic long term exposure to sewage sludge using information from mass spectrometry Sandra Wagener, Harald Jungnickel, Nils Dommershausen, Thomas Fischer, Peter Laux, and Andreas Luch Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b07222 • Publication Date (Web): 09 Apr 2019 Downloaded from http://pubs.acs.org on April 10, 2019

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 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 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.

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 22

Environmental Science & Technology

Nanomaterials

Dispersion in Sewage Sludge

Distribution

Uptake & Dissolution in planta? ACS Paragon Plus Environment

Environmental Science & Technology

1

Determination of nanoparticle uptake, distribution and characterization in plant root tissue after realistic

2

long term exposure to sewage sludge using information from mass spectrometry

3 4

Sandra Wagener1*, Harald Jungnickel1, Nils Dommershausen1, Thomas Fischer1, Peter Laux1, Andreas Luch1

5

1Department

6

Straße 8-10, D-10589, Berlin, Germany

of Chemical and Product Safety, German Federal Institute for Risk Assessment (BfR), Max-Dohrn-

7 8

*Corresponding author at: Department of Chemical and Product Safety, German Federal Institute for Risk

9

Assessment (BfR), Max-Dohrn-Strasse 8-10, 10589 Berlin, Germany.

10

Phone: +49 30 18412 4164; fax: +49-(0)30-18412-47 41

11

E-mail: [email protected]

12 13

Abstract

14 15

The use of nanoparticles (NPs) in numerous products and their potential accumulation causes major concern for

16

humans and the environment. Until now, the uptake of NPs in plant tissue was mostly shown under greenhouse

17

conditions at high doses and short exposure periods. Here we present results on the uptake of particulate silver

18

(Ag) and cerium dioxide (CeO2) in the tissue of Triticum aestivum, Brassica napus and Hordeum vulgare, after

19

exposure to sewage sludge treated with nano-Ag (NM300K at 1.8 and 7.0 mg/kg sludge per dm soil) and nano-

20

CeO2 (NM212 at 10 and 50 mg/kg sludge per dm soil). All plants were cultivated in a rural area near the German

21

town Schmallenberg according to the common regional crop rotation on outdoor lysimeters. The highest

22

concentration measured was 86.4 mg/kg for Ag (Hordeum vulgare), and 94 mg/kg for Ce (Triticum sativum).

23

Analysis of plant samples revealed the presence of Ag mainly in its ionic form. However, the occurrence of

24

nano- and larger sized particles of Ag and CeO2 was observed as well. Quantitative shares of the particulate

25

fraction of the total element concentration were estimated with up to 22.4% for Ag and with up to 85.1% for

26

CeO2. A high abundance of particle agglomerates in the phloem suggests upward transport of the nanoparticles

27

to other plant parts. A small number of agglomerates in the xylem suggests a downward transport, and

28

subsequent accumulation in the root phloem. Exemplary investigations of Brassica napus root exposed to nano-

29

CeO2 revealed no accumulation of the pristine material in the cell nucleus, however, CePO4 was found. The

30

presence of this substance points to a dissolution of the low soluble CeO2 in planta and subsequent precipitation.

ACS Paragon Plus Environment

Page 2 of 22

Page 3 of 22

Environmental Science & Technology

31

Furthermore, for the first time, mixed NP-salt agglomerates, composed of Ca3PO4+ and K3SO4+ NPs, could be

32

observed within Brassica napus root tissue.

33 34

Introduction

35 36

In the past 15 years, the use of nanoparticles (NP) in daily life and medical products has constantly increased 1-3.

37

As a consequence, NPs cause major concern as a potential contaminant of the environment, e.g. through waste

38

water, deposition or precipitation.4-6 Furthermore, the uptake of NPs by agricultural plants is also relevant for

39

human exposure.7 If not removed systematically after their use, NPs will enter the environment via numerous

40

sources and may accumulate depending on their specific chemical and physical properties.8-9 In order to assess

41

the potential effects of these small sized materials, numerous investigations concerning exposure scenarios,

42

environmental fate and toxic effects of nanomaterials (NMs) have been conducted.10-13 Beside adverse effects

43

including phytotoxicity, suppression of seed germination1 and oxidative stress leading to programmed cell

44

death,14 other studies report improved seed germination and crop performance.15 While excess reactive oxygen

45

species (ROS) formation and oxidative damage are considered main mechanisms of NP toxicity, some materials

46

may also scavenge ROS.16 In a recent study by Tripathi et al.,17 silicon NPs at a concentration of 10 µM were

47

shown to protect Zea mays seedlings against arsenic (As). While accumulation of As and reactive oxygen species

48

was reduced, the levels of superoxide dismutase and gluthathione reductase were increased. Ion release was

49

confirmed as a major factor of NP toxicity in plants. In a study with Glycine max, soluble ZnO- and Ag-NPs

50

have reduced plant growth, rigidity and root cell viability. Simultaneously, normal seedling growth was observed

51

in presence of the Al2O3-NPs, which has been attributed to their poor solubility12,18-19 Thus, the chemical identity

52

needs to be considered with regard to potential effects of NPs on plants. This is of particular importance for the

53

investigation of the food chain, which remains as one major research need.7, 14, 20

54

So far, there is no evidence for significant human and environmental health effects that can be specifically

55

attributed to the nano-size of materials.21-23 Instead, a number of other properties like surface area, solubility and

56

morphology are being discussed as main determinants of

57

guidelines for ecotoxicity and toxicity testing of chemicals are in principle considered suitable for NPs as well.

58

However, specific adaptations in particular with regard to sample preparation, dosimetry and biokinetics are

59

required.25-26 With regard to a possible accumulation of NP along the food chain, the influence of abiotic stress

60

such as a variation of trace elements27 should be considered as a confounding factor. High salinity (100 mM

NP toxicity.24 In consequence, the established

ACS Paragon Plus Environment

Environmental Science & Technology

61

NaCl) for example, was shown to cause increased concentrations of Ce in roots and leaves of Brassica plants

62

exposed to CeO2-NPs.28

63

Thus, particle accumulation and fate at field conditions are of particular interest. In this context both, NP dose

64

and exposure duration are important issues. Whereas several studies, mostly conducted in growth chambers,

65

exist concerning fate and short time effects of NPs on plant germination,6, 20, 28-30 the long-term fate of these

66

materials in the environment under real conditions has not yet been sufficiently addressed. Depending on their

67

physical and chemical properties, NPs may dissolve in ions from which particles might be formed again during

68

their transport through the environment. This may not only affect their size, but also significantly change their

69

physical-chemical behaviour. In the case of CeO2, parameters like particle surface coating and presence of

70

organic matter were demonstrated to affect the uptake by higher plants.31 Furthermore, chemical alteration of the

71

particle itself, as e.g. the formation of CePO4 upon exposure of Cucumis sativus to CeO2,32 or changes of the

72

plant protein profile, e.g. of Oryza sativa following treatment with Ag-NPs33 have been demonstrated. For a

73

thorough assessment of NP fate within living organisms, methods are needed to allow for the following tasks: 1)

74

detection and characterization of the particles within the organism, including concentration and size distribution,

75

2) localization of the particles on tissue and cellular level 3) characterization of NP and aggregate composition

76

including other substances that may have accumulated during their transport. Based on these data, conclusions

77

on transformation and possible adverse effects could be drawn.

78

Several methods for particle characterization have already been described in the literature, such as scanning

79

electron microscopy (SEM),34 transmission electron microscopy (TEM)35 or X-ray Absorption Spectroscopy

80

(XAS)30. However, for the reliable detection of relevant elements, images have to be strongly magnified, leading

81

to either a limitation of the area of investigation or to a time consuming analysis. A method which allows for

82

element-specific detection over a larger sample section is Time-of-Flight Secondary Ion Mass Spectrometry

83

(ToF-SIMS).19,

84

entities together with the generation of three dimensional (3D) information.38 By imaging the elemental

85

constitution, not only information about the precise localisation of particles, but also on accumulation of other

86

substances can be gathered.39-40 A further benefit of this technique is the possibility of an automated statistical

87

agglomerate cluster analysis, allowing for the determination of particle or aggregate size distribution patterns.40

88

This study is a part of the joint project on “Design criteria for sustainable nanomaterials – DENANA” and aims

89

to investigate about the long-term fate of nano-Ag and nano-CeO2 in root tissues of Triticum aestivum, Brassica

90

napus (Ag and CeO2) and Hordeum vulgare (only Ag). The plants were grown on different lysimeters according

91

to their actual growth period and exposed to NPs via sewage sludge41 to mimic a real life situation.42 With nano-

36-37

The major advantages of this technique are the area-wide detection of several chemical

ACS Paragon Plus Environment

Page 4 of 22

Page 5 of 22

Environmental Science & Technology

92

Ag and nano-CeO2, a soluble and a biopersistent NP species relevant for environmental exposure,43 have been

93

chosen to elucidate the influence of solubility on NP fate. While nano-Ag is commonly used for its antimicrobial

94

activity in numerous products such as medicinal applications and fabrics,44-46 nano-CeO2 mainly occurs in

95

catalysts for vehicles,47-48 but is also used for industrial and construction purposes.49-50 Beside a quantification of

96

Ag and Ce, the particulate occurrence of both materials and their systemic distribution are addressed. Single

97

particle inductive coupled mass spectrometry (spICP-MS) and ToF-SIMS are used to quantify and characterize

98

NPs within plant tissues, enabling not only information about size distribution, but also precisely localizing the

99

particle deposition.

100 101

Materials and methods

102 103

Plant samples and nanomaterials

104

Three different plant species, Triticum aestivum (Triticum aestivum ´Tybalt A‘ Saaten Union GmbH, Isernhagen,

105

Germany), Brassica napus (Brassica napus ‘Treffer‘ KWS Saat SE, Einbeck, Germany) and Hordeum vulgare

106

(Hordeum vulgare ‘SY Typee‘ Syngenta, Maintal, Germany, Table 1), were cultivated in outdoor lysimeters (0.9

107

m × 0.9 m × 0.9 m, width x length x depth) in a rural area near the German town of Schmallenberg (North Rhine-

108

Westphalia, 500 m above sea level). The chosen plant species are most relevant for sewage sludge fertilization in

109

Germany.51

110

http://www.refesol.de/english/analysedaten.shtml) and exposed to either nano-Ag (NM300K) or nano-CeO2

111

(NM212) (see Table 1 for exposure concentrations, and supplement Table S1 for particle specification). Ag and

112

CeO2-NPs were supplied from the Joint Research Center (JRC) Nanomaterials Repository and have been

113

characterized within the OECD testing programme. Both NPs have been subject of numerous research

114

projects.52-53 The materials were incorporated into sewage sludge of municipal sewage from the treatment plant

115

of Schmallenberg, according to the German Sewage Sludge Ordinance,51 before application to lysimeters at

116

given concentrations. For the incorporation of the NPs, the sewage sludge was dried, meshed to particles of a

117

size smaller than 2 mm and then stored in a vessel under stable moderate stirring and aeration (2.5 mg O2/L).

118

NMs were then spiked into the sewage. The sludge was spread over the first 20 cm of soil on the lysimeters,

119

immediately after sowing. At harvest, the plants of each exposure group were divided into root, sprout and grain

120

and frozen at -18 °C.

The

lysimeters

were

filled

with

reference

121 122

ACS Paragon Plus Environment

soil

01A

(RefeSol;

Environmental Science & Technology

123

Table 1: Overview of investigated plant species and exposure concentrations for Ag and CeO2 NP.

Plant species

Triticum aestivum Brassica napus Hordeum vulgare

124

Page 6 of 22

Triticum aestivum Brassica napus

Seeding

June 2014 October 2014 August 2015

April 2015 August 2015

Harvest

Control C0*

Exposure Group C1*

Exposure to Ag-NP (NM300K) September 0 2014

Exposure Group C2*

1.8

7.0

10

50

August 2015 July 2016 Exposure to CeO2-NP (NM212) 0 August 2015 July 2016

*Concentration is given in mg kg-1 dry matter sludge

125 126

Analysis and measurement techniques

127

The characterization of particle fate and localization within plant tissue was conducted following the scheme

128

described in

129 130

. After harvest, the plants were divided into corn / shell, sprout and roots. The parts were then analyzed using

131

ICP-MS for total Ce and Ag content. Only samples with a significant Ce or Ag amount were further analyzed

132

with spICP-MS and ToF-SIMS.

133 134

Quantification of the total Ag content with ICP-MS

135

Approximately 100 mg of plant tissue was washed and dried in a drying stove for 24 hours. After cooling, the

136

exact weight of the dry samples was determined. 2 ml of nitric acid (HNO3, 69%), 2 ml of hydrogen peroxide

137

(H2O2) and 2.5 ml of water (H2O dest.) were then added to the sample. Samples were subsequently digested in a

138

microwave oven (FA MLS, Ethos.start) following a temperature program suitable for biological material (step 1:

139

heating up to 100° C in 2 min; step 2: heating up to 180° C in 6 min; step 3: heating up to 210° in 4,5 min: step

140

4: holding at 210° C for 10 min). Samples were analyzed for their total Ag and Ce content using ICP-MS (X-

141

Series II, FA Thermo). The limit of detection (LOD) was determined by taking the mean + 3 times the standard

142

deviation of root blanks.

143 144

Single particle analysis using spICP-MS

145

ICP-MS was additionally operated in the single particle (sp) mode, which enables the determination of particle

146

mass, number and size distribution of the particulate fraction. A further feature is the elemental identification, ACS Paragon Plus Environment

Page 7 of 22

Environmental Science & Technology

147

provided for the respective single particles of interest. Details about the calibration and the determination of

148

single particle number, and dwell time can be found in Wagener et al.54

149

Particle extraction and minimization of particle alteration is a crucial step of single particle determination in the

150

plant matrix. All plant tissue samples were prepared according to an enzymatic digestion method published by

151

Dan et al.29 Samples were cut with a ceramic knife and washed twice with MilliQ-water to remove remaining

152

soil. The cleaned pieces were milled under liquid nitrogen in a mortar. Subsequently, 8 ml of citrate buffer (100

153

mM) and 2 ml of Macerozyme R10 (50 mg/ml) were added at room temperature. Samples were incubated at

154

1400 rpm and 37°C in a thermos shaker for 24 h. To determine the exact plant weight, the remaining water

155

content was determined and subtracted. The supernatant was removed and diluted for subsequent spICP-MS

156

analysis.

157

The LOD for spICP-MS is given by the frequency of single particles signals above the background, which is

158

governed by dissolved element species.55 As no reference materials, i.e. plant materials with a known amount of

159

particulate Ag or CeO2 exist, the recovery rate was determined by adding NPs to the root pieces before the

160

enzymatic digestion. Similar to ICP-MS, spICP-MS measures only elements, but particle size and concentration

161

can be calculated for molecules based on particle density and molar mass. As NM212 was used for these

162

experiments and for determining the recovery rate; calculations with spICP-MS were conducted under the

163

assumption that all particles are CeO2. For further interpretation of the results however, it is important to bear in

164

mind that the Ce content measured with spICP-MS may actually be in another form. Similar issues must be

165

considered for Ag. The calculation of pure Ag was used here because of the use of nano-Ag NM300K. Thus the

166

sizes and concentrations presented below only apply to pure Ag particles and do not consider possible formation

167

of larger particles containing other elements.

168 169

Characterization of particle fate using ToF-SIMS

170

A ToF-SIMS5 (ION-TOF GmbH, Münster, Germany) instrument was used for the identification of the chemical

171

composition and distribution patterns of particles within plant tissue sections. ToF-SIMS uses a focused ion

172

beam to sputter secondary ions arising from the sample surface. An image of the chemical composition of the

173

sample can then be established. The instrument can be used in dual ion beam mode, where multiple surface

174

layers are analysed and three-dimensional chemical information can be acquired. The lateral resolution of this

175

method is ca. 80 nm.

176

Thin 70 µm layers of root tissue were prepared using a cryo microtome. Small pieces of roots were embedded in

177

resin medium and cut at a temperature of -15°C (object holder) and -25°C (cutting chamber). In this study, a

ACS Paragon Plus Environment

Environmental Science & Technology

178

lateral resolution below 100 nm 56 and a depth resolution of approximately 10 nm per layer was achieved.57 Ion

179

images and depth profiles were performed using a 30 keV nano-bismuth primary ion beam source and a 30kV

180

argon cluster ion beam. 3D depth profiles were acquired in the dual beam analysing mode using a 30 keV nano-

181

bismuth (Bi)x(y+)-cluster primary ion beam source with a BiMn emitter as analysing ion beam and a a 20 keV

182

argon gas cluster ion source both mounted at 45° with respect to the sample surface as sputter ion beam and an

183

electron flood gun for charge compensation. The methodology has been demonstrated successfully for the

184

reconstruction of 3D distribution patterns of NPs from living cell systems.19

185 186

Results and Discussion

187 188

Total concentration of Ag and Ce in planta

189

The LOD for Ag was determined as 0.3 ppb, 0.1 ppb and 0.8 ppb for Triticum aestivum, Brassica napus and

190

Hordeum sativum, corresponding to 0.04, 0.01 and 0.1 mg kg-1 plant tissue, respectively when using a dried plant

191

sample of 75 mg. Correspondingly, the LOD for Ce in Triticum aestivum and Brassica napus were 127 ppb and

192

40 ppb or 21 mg kg-1 and 8 mg kg-1 plant tissue, respectively. The increased LOD for Ce arises from the natural

193

occurrence of Ce in soil. The recovery rate for the digestion of Ag was determined with 97% for a concentration

194

range of 20 ppb and with 123% for a concentration range of 2 ppb. Due to the increased Ce levels in the roots, no

195

reproducible recovery rate for Ce was achieved. These circumstances must be considered in the interpretation of

196

the results.

197

Analysis with ICP-MS revealed that plant roots absorb Ag and Ce in significant amounts after long-term

198

exposure to nano-Ag and CeO2 containing soils (Table 2). In contrast, in the sprouts, corn and shells the

199

concentrations for both, Ce and Ag, were below the LOD. For Ag, concentrations between 1.3 and 86.4 mg kg-1

200

were observed in roots. For Ce, quantification was shown to be more difficult due to the large background levels

201

in root tissue arising from naturally occurring Ce in the soil. However, when applying a larger amount of CeO2

202

to the lysimeters (C2), a significant accumulation was revealed. The Ce content varied between below LOD and

203

94 mg kg-1. The concentrations of Ag measured in Hordeum vulgare are significantly higher compared to the

204

other plant species. The high standard deviation (SD) of the highest detected concentrations of Ag in Hordeum

205

vulgare and of Ce in Triticum sativum demonstrate the inhomogeneous distribution of these elements in the

206

roots. However, also an inhomogenous distribution of the NPs within the lysimeter soil after sludge application

207

may lead to high variation of both elements.

208

ACS Paragon Plus Environment

Page 8 of 22

Page 9 of 22

Environmental Science & Technology

209 210 211

Concentrations of particulate Ce and Ag in plant roots measured with spICP-MS and ToF-SIMS

212

Ag and Ce are not only present in the plant root tissue in ionic form, but also as particulate matter (Figure 1,

213

Figure S2, Table 2). The recovery rate of total Ag was 81.6% (±10%) with a particulate fraction of 12.3% (±2

214

%), and for total CeO2 105.9 (±29%) with a particulate fraction of 70.7% (± 25%). A dissolution rate of ~10%

215

was already determined for NM300K in pure water. The fact that in the recovery experiments the NMs are added

216

to the surface of the roots instead of being incorporated within the matrix may lead to an overestimation of the

217

recovery and therefore to an underestimation of the real particulate matter. Agglomeration may also contribute to

218

the observed diameter for Ag-NP in plant extracts of 68 nm (± 1.1 nm) compared to a primary particle size of 15

219

nm for NM300K. Similarly, for CeO2-NP, the median particle diameter in plant extracts was 89 nm (± 0.7 nm),

220

compared to a primary particle size of 28 nm for the pristine material NM212. Corresponding analysis of both

221

particles in pure water showed primary particle sizes approximating the sizes of the pristine materials. Analysis

222

of roots revealed for plants exposed to Ag a significantly lower share of particulates compared to ions (Table 2).

223 224

Table 2: Total and particulate Ce1 and Ag concentration in mg kg-1 root tissue after exposure to sewage sludge tissue ± standard deviation

Plant species

Analyte

Total C1

Total C2

Particulate C0

Particulate C2

%Particulate C2

Triticum aestivum

Ag

2.3 ± 0.12

9.1 ± 4.13

not determined

not determined

not determined

Brassica napus

Ag

1.3 ± 1.94

7.6 ± 2.35

< LOD

1.7 ± 0.2

22.4

Hordeum vulgare

Ag

12.7 ± 10.1

86.4 ± 45.1

< LOD

0.6 ± 0.1

0.7

Triticum aestivum

Ce/CeO2

24.0 ± 7.0

94 ± 77.6

13.7 ± 5.8

80 ± 6.2

85.1

Brassica napus

Ce/CeO2

< LOD

8.9 ± 1.3

3.5 ± 0.9

3.7 ± 1.3

41.6

225 226 227

1

228

Ag NP were detected in exposed plants of Brassica napus (1.7 mg kg-1) and to a smaller amount in Hordeum

229

vulgare (0.6 mg kg-1) by spICP-MS. The detection of both Ag isotopes, 106.9 u and 108.9 u, by ToF-SIMS

230

measurements of Brassica napus roots confirms the occurrence of particulate Ag (Figure 2, Figure S3). In

231

Triticum aestivum, which was not measured with spICP-MS, ToF-SIMS measurements also verify an

232

accumulation of Ag NPs inside the root tissue (Figure 3). No Ag-NP were detected in control plants with spICP-

233

MS and ToF-SIMS. The low concentration of Ag-NPs measured in Hordeum vulgare is noticeable when

: Total concentrations are given for Ce, particulate concentrations for CeO2. Concentrations revealed for plant samples from the same lysimeter published in Schlich et al.41 were as follows: 2: 2.6 mg kg-1, 3: 10.9 mg kg-1, 4: 3.4 mg kg-1, 5: 8.4 mg kg-1.

ACS Paragon Plus Environment

Environmental Science & Technology

234

compared with the high total concentration of the element. Our results achieved with spICP-MS point to a

235

significant uptake of ionic Ag or to a partial dissolution of Ag particles during or following their uptake. In

236

Triticum aestivum, which was not measured with spICP-MS, the detection of both Ag isotopes, 106.9 u and

237

108.9 u, by ToF-SIMS (Figure 3) confirms an accumulation of Ag-NPs inside the root tissue. For plants exposed

238

to CeO2, the precise quantification for particulate matter, as already described for the measurement of the total

239

content, was more difficult due to the large natural background. The concentration of particulate Ce varied

240

between 3.5 to 13.7 mg kg-1 for the control samples (C0) and between 3.7 and 80 mg kg-1 for the exposed

241

samples (Table 2). The detection of Ce containing particles in the control samples demonstrates their natural

242

occurrence in root tissue. Our findings of a significant total Ce accumulation in Triticum aestivum, however,

243

together with the low solubility of CeO2 indicate that at least some of the supplied CeO2-NPs are taken up in

244

particulate form. However, in case of Brassica napus, no significant discrimination between control and samples

245

exposed to CeO2 is possible. Total and particulate concentrations are closer in Brassica napus compared to the

246

cereals tested. In contrast to Brassica napus, Ce containing particles in Triticum aestivum account for the largest

247

portion of their total Ce-content with 13.7 and 80 mg kg-1 respectively in control and exposed plants. Thus, the

248

spICP-MS results for control and exposed samples give hints of an uptake of particulate Ce in Triticum aestivum,

249

but not in Brassica napus. However, ToF-SIMS analysis suggests an uptake of CeO2 by both, Brassica napus

250

and Triticum aestivum. The CeO2-specific ion CeO+ m/z 156 was detected in exposed, but not in control plants

251

(Figures 4, S4 and S5). The results demonstrate that the measurement of elements of a high abundance in the

252

environment is challenging. Uptake studies should therefore apply at least one different technique. Previously,

253

Lopez-Moreno et al.58 proved an uptake of CeO2-NPs by soybean using synchrotron X-ray absorption

254

spectroscopy. While in the case of Ag there was no significant difference in NP uptake by the different plant

255

species, the particulate fraction of CeO2 was significantly higher in Triticum aestivum compared to Brassica

256

napus. These differences are possibly due to the different solubility of Ag and CeO2 which seemingly holds also

257

true within the investigated plant root tissue sections. The results indicate material specific transformation,

258

bioavailability and toxicity of NPs. Furthermore, plant species and growth substrate were described to influence

259

these characters.59-60 The influence of physicochemical variations among plant species, e.g. hydraulic

260

conductivity or pore size of the cell wall on transport and accumulation of NPs has been described previously by

261

Judy and Bertsch.61 In addition to the plant species, environmental conditions may alter NPs stability, oxidation

262

state and precipitation and thus affect their reactivity and translocation inside the plant.2 Thus, the same NPs may

263

provoke different plant responses, depending on the species and conditions. The different particle accumulation

264

pattern of plant species observed in this study may therefore be explained by different physiological

ACS Paragon Plus Environment

Page 10 of 22

Page 11 of 22

Environmental Science & Technology

265

environments that impact on substance transportation. This might explain why there is a larger difference

266

between the uptake of the less soluble CeO2 in Triticum aestivum and Brassica napus , than for the more soluble

267

Ag.

268 269

Size distribution of CeO2- and Ag-NPs in planta

270

SpICP-MS and ToF-SIMS provide complementary information on different particle agglomerate sizes. Whereas

271

spICP-MS detects particles in the nm-range, ToF-SIMS gives evidence of the occurrence of larger agglomerates.

272

The size distributions for Ag- and CeO2-particles in Brassica napus roots measured with spICP-MS are given in

273

Figure 1. In case of CeO2-NPs, the size distribution is comparable between the control and exposed samples. In

274

Triticum aestivum, however, a clear shift towards larger particles could be observed in the exposed sample as

275

well as an overall increase in concentration. Considering a primary particle size of 28 nm for NM 212, our

276

results provide evidence for an agglomeration of the applied CeO2-NPs, either before or after uptake into the

277

root. Most of the Ag amount in the root samples appears to be ionic. However, some particles could also be

278

detected in the exposed sample (Table 2, Figure 1). While the tested material Ag NM300K owns a primary

279

particle size of 15 nm, a mean size of 60 nm was determined in the root samples (Figure 1), pointing to an

280

agglomeration as observed for CeO2. No difference was observed between the investigated plant species.

281

282 283

Figure 1: Detected spICP-MS signals and size distribution of Ag- and CeO2-NPs in Brassica napus root sections. a) 7.0 mg Ag-NP

284

(NM300K) /kg sludge per dm soil, Ag exposure group C2 and unexposed control, Ag exposure group C0, b) 50.0 mg CeO2-NP (NM212) /kg

285

sludge per dm soil, CeO2 exposure group C2 and unexposed control samples, CeO2 exposure group C0.

ACS Paragon Plus Environment

Environmental Science & Technology

286

A three dimensional distribution of Ag within a root section of Triticum aestivum was revealed by ToF-SIMS

287

analysis of a 100 µm x 100 µm tissue section of 3.5 µm thickness and is presented in Figure 3-f. The density of

288

the spots indicates agglomerates in the tissue, whereas a more refined Ag layer of lower intensity points to the

289

presence of ionic Ag. Analysis of agglomerates with regard to their size distribution revealed a number of 280

290

agglomerates in the total size range between 2 to > 281 µm². The highest number of agglomerates (126) was

291

found in a size range between 23 and 32 µm² (Figure 3-g). Based on these results, a presence of Ag in

292

particulate form is confirmed for the investigated tissue.

293

Based on our results it can be concluded that the particles cover a broad size range, possibly in a bi-modal

294

distribution with the 1st mode at 40 or 60 nm, for CeO2 and Ag respectively (Figure 1), and the 2nd mode in the

295

µm range (Table 1 and Figure 3). The particles in the roots were all larger than the pristine NPs. It also has to be

296

considered that the sizes given here are assuming pure Ag or CeO2 particles. If the particles accumulate other

297

substances, the particle sizes would be under- or overestimated. Reasons for the increase of the particles may be

298

different and can most probably be explained by their solubility, but also by their natural occurrence in the soil

299

or root, respectively. For both, Ag and CeO2, an increase in particle size has already been observed in the

300

recovery samples. Thus it is conceivable that this increase is at least partly a consequence of the enzymatic

301

digestion. However, the detection of large Ag and CeO+ agglomerates within the root tissue with ToF-SIMS

302

provides proof that larger particles already occur in the root, as ToF-SIMS requires no digestion procedure. In

303

the case of Ag, a de novo formation of particles from soluble Ag is also very likely to occur after uptake. De

304

novo formation of Ag agglomerates from ionic suspensions has been shown in other contexts, e.g. the formation

305

of particulate Ag species during migration experiments with Ag containing textiles in artificial sweat.62 For the

306

less soluble CeO2, agglomeration inside the root seems to represent the most probable mechanism for particle

307

size increase. When particles interact with plants, they have to cross the cell wall as the first barrier. The wall,

308

which is composed of cellulose, permits the entry of small particles and restricts passage of the larger ones.63

309

According to Dietz and Herth,64 the size exclusion limit for the plant cell wall is between 5 and 20 nm.

310

Therefore, it can be assumed that the formation of most of the larger agglomerates occurred after uptake into the

311

root. However, some of the nano-sized particles have been reported to induce the formation of larger pores in the

312

cell wall which would enable the entry of larger particles.65

313 314

Localization, characterization and distribution of CeO2- and Ag-NPs in roots

315

Analyses of rhizodermis and exodermis were conducted with Ag as a case study using ToF-SIMS. The Ag

316

signal distribution pattern was recorded with ToF-SIMS within a root tissue section of Brassica napus (Figure

ACS Paragon Plus Environment

Page 12 of 22

Page 13 of 22

Environmental Science & Technology

317

2). While in the left hand side of the picture the rhizodermis was still present, it was removed in the right hand

318

side, showing Ag distribution patterns within the exodermis. Ag was not detected in or on the rhizodermis, but

319

was present in the exodermis. As ToF-SIMS is able to detect isotopes up to several µm in depth, this proves that

320

the Ag content detected with ICP-MS and spICP-MS arises from an uptake of Ag into the tissue and not from a

321

deposition outside the root. NPs have been found to penetrate the outer root cell layers, and accumulate later in

322

the membrane of the root epidermis from where they enter the xylem, and further translocated to the leaves.67

323

Even though in our study the NPs could be detected throughout the whole vascular system in roots, they were

324

not equally distributed along the central cylinder of the root. As shown for Ag and CeO2 in Triticum aestivum

325

(Figures 3 and 4) and Brassica napus (Figures S3 and S4 and S5), NP-agglomerates mainly accumulate in the

326

phloem layers around the central cylinder. Together with the presence of only few particles in the xylem, this

327

indicates an upward transport of absorbed NPs via the xylem, followed by a downward transport from the

328

vegetative parts of the plant via the phloem with subsequent accumulation. The transport via the xylem, which,

329

according to Aslani et al.66, serves as the most important vehicle in the distribution and translocation of NPs, has

330

already been proven in cucumber plants.67 Xylem and phloem mediated uptake, translocation and distribution

331

from root to shoot through the xylem and its reverse transport to root through the phloem was additionally

332

demonstrated by Wang et al.68 for nano copper oxide in Zea mays.

333

334 335

Figure 2: Reconstructed ion pictures showing the Ag signal from the Ag isotopes 106.9 u and 108.9 u: a) on the rhizodermis and b) on the

336

exodermis of a Brassica napus root exposed to 7.0 mg/kg sludge per dm soil. The white arrows indicate that Ag was not detected on the

337

surface of the rhizodermis. The upper panels show the Ag distribution of a root, where the rhizodermis (a, depicted with an arrow) was partly

338

removed (exodermis depicted in the picture as b). The lower panel shows a root, where the rhizodermis was removed completely and shows

339

the exodermis (b).

340 ACS Paragon Plus Environment

Environmental Science & Technology

341

342 343

Figure 3: Ag distribution in a longitudinal tissue section of Triticum aestivum exposed to Ag NP, ToF-SIMS total ion reconstruction from a

344

root tip, exposure group C2, 7.0 mg/kg sludge per dm soil: a) total ion picture of a root section (1500 µm x 500 µm), b) spectra from tissue

345

sections of an exposed root (upper panel, shown are both silver isotopes) and an untreated control root (lower panel, no silver isotope peaks

346

are present). c) ion reconstruction (500 µm x 500 µm) of the red square, marked in a. d) distribution of the reconstructed silver signal (in

347

green) from c. e) overlay of c and d, silver particles are visualized in green. f) 3D distribution of the silver nanoparticles, depicted in blue

348

(100 µm x 100 µm x 3.6 µm, length x width x depth) of the white square in e. g) size distribution of silver particles from f, size groups

349

depicted: Group I: 2 µm Ø , Group II: 3 - 12 µm Ø, Group III: 13 - 22 µm Ø, Group IV: 23 - 32 µm Ø, Group V: 33 - 42 µm Ø, Group VI:

350

43 - 62 µm Ø, Group VII: 63 - 82 µm Ø, Group VIII: 83 - 102 µm Ø, Group IX: 103 - 122 µm Ø, Group X: 123 - 142 µm Ø, Group XI >

351

142 µm Ø.

352 353 354 355 356 357 358 359 360 361 362 363 364 365 366

ACS Paragon Plus Environment

Page 14 of 22

Page 15 of 22

Environmental Science & Technology

367

368

e 369

Figure 4: CeO distribution in a longitudinal tissue section of Triticum aestivum, exposed to CeO2 NP, exposure group C2, 50 mg/kg sludge

370

per dm soil: a) ToF-SIMS total ion reconstruction of a root section (3.5 mm x 1.5 mm), b) spectra from tissue sections of an exposed root

371

(upper panel, the CeO+ peak, m/z 156, is present) and an untreated control root (lower panel, no CeO+ peaks are present. c) ion reconstruction

372

(500 µm x 500 µm) of the red square marked in a. d) 3D distribution of CeO2 NPs, depicted in red from the 3D ToF SIMS ion reconstruction

373

(100 µm x 100 µm x 3.6 µm, length x width x depth) of the red square in c). e) 2D enlarged section (20 µm x 20 µm) from the 3D

374

reconstruction, CeO2 particles are depicted in red.

+

375 376

Occurrence of NPs in cell nucleus and cytosol

377

ToF-SIMS analysis and 3D ion reconstruction were used to investigate the presence of CeO2-NPs in the

378

cytosolic fraction and the nucleus of Brassica napus root cells. The masses of adenine and guanine, m/z 136 and

379

152, were employed for identification of the nucleus shown in yellow in Figure 5a. The mass of m/z 58

380

(C3H8N+) was used as an indicator for the cytosolic fraction (Figure 5). CeO2-NPs were identified in the cytosol

ACS Paragon Plus Environment

Environmental Science & Technology

381

(Figure 5-a-3, small red circles left of the nucleus) but not in the nucleus (Figure 5-a-2). However, cerium

382

phosphate (CePO4+, ion m/z 235) agglomerates were identified in the nucleus (Figure 5-a-3, green encircled).

383 384

385 386 387

Figure 5: Reconstructed 3D ToF-SIMS image (15 µm x 15 µm x 4 µm) of a single cell from a Brassica napus root, exposed to CeO2,

388

exposure group C2, 50 mg/kg sludge per dm soil. 5a: Cell nucleus is shown in yellow as the sum of adenine (m/z 136) and guanine (m/z 152)

389

ions. Cytosol is shown in blue (m/z 58). CeO+ is depicted by black dots in 2 and red encircled in 3. CePO4+ is green encircled in 3.

390

5b: Mixed NP agglomerates identified in the cytosol, red: CeO2, light blue: Ca3PO4+, yellow: K3SO4+ NP-agglomerate (Ion m/z 213).

391 392

As CePO4 was only found in exposed roots but not in the controls, it is assumed that the formation of CePO4

393

results from ion release of CeO2-NPs with subsequent formation of CePO4 within the cell nucleus. This would

394

support the theory of partial solubility of CeO2 within plant tissue similar to processes observed within animal

395

tissue.13, 72 This may indicate that NPs, even if they are not absorbed directly into the cell nucleus, can dissolve

396

and reform particles with different chemical entities in such a way that the heavy metal reaches the cell nucleus.

397

Similar findings were reported by Zhang et al.73,35 in vacuoles of romaine lettuce and cucumber roots. They

398

provided proof of flocculent CePO4 particles in the intercellular spaces and vacuoles using scanning transmission

399

X-ray and electron microscopy and explained the function of the vacuole as a “phosphorus pool”. The storage of

400

heavy metals in vacuoles represents a detoxification mechanism of plants.74 Small amounts of Ce3+ were shown

401

to be released from CeO2 at the presence of organic acids and reducing substances in root exudates.75 From a

402

study investigating Lactuca plants, the authors suggested transformation processes to occur at the root surface.

403

The interaction between the NPs and root exudates at the nano-bio interface was assumed to be necessary for the

404

transformation of nano-CeO2 in plants. It is hypothesized, that a part of the Ce3+ released at the nano-root

405

interface is immobilized by the formation of CePO4 precipitates on the root surface, in intercellular spaces and

406

vacuoles.3−35 ACS Paragon Plus Environment

Page 16 of 22

Page 17 of 22

Environmental Science & Technology

407 408

Identification of mixed agglomerates in Brassica napus root samples

409

Mixed agglomerates of CeO2 with either calcium phosphate (Ca3PO4) or potassium sulphate (K3SO4) were

410

identified by ToF-SIMS in Brassica napus root samples (Figure 5-b). These results point to the presence of

411

hetero agglomerates in plant roots and indicate a common crystallization point for both agglomerates, with a

412

mixed identity of CeO2 and the corresponding salt material. It is assumed that each agglomerate grows from its

413

crystallization point in one specific direction only. This illustrates that NPs are altered during their life cycle in

414

plant tissue. The presence of mixed agglomerates may indicate different solubility rates of different

415

agglomerates. This is probably due to the fact that compounds generally dissolved in the plant vascular system

416

may associate with NPs. Therefore, on the one hand they may move to plant parts where they normally do not

417

occur or, on the other hand, they may dissolve differently than agglomerates of pristine NPs.

418

This alteration may further influence the composition of the plant tissue, which in turn may have adverse, but

419

possibly even positive effects on the plant´s metabolism. For example, Rossi et al.28 were able to confirm their

420

hypothesis that CeO2-NPs modify the formation of plant apoplastic root barriers, with the consequence of an

421

enhanced salt stress tolerance of Brassica napus following exposure to CeO2-NPs in combination with NaCl in a

422

growth chamber. An enhanced transport of Na+ to the shoots and a decreased accumulation of the element in the

423

roots were identified to advance physiological performance.

424 425

Whereas particle uptake has already been shown in laboratory experiments, mostly during their germination

426

process, our study proved that particle uptake in plant roots is an important issue also under realistic exposure

427

conditions. Even though no significant concentrations could be detected within the edible parts of the

428

investigated plants, the presented data are relevant for human exposure by consumption of root crops.

429

Characterization of NP fate remains an urgent need for the depiction of processes in biological tissues. The

430

understanding of these mechanisms represents the basis for further hazard and risk assessment considering

431

realistic environmental conditions. SpICP-MS and ToF-SIMS were found to be suitable instruments for

432

providing the necessary information. However, it became also clear, that in case of Ce, an element that occurs

433

frequently in the environment, ICP-MS and spICP-MS measurements are challenging due to the high

434

background. Imaging techniques such as ToF-SIMS that are able to record the specific chemical composition of

435

particulate materials provide a suitable complementation. Beyond the characterization of particulate matter as

436

such, they are able to analyze the alteration of other substances in the vicinity of NPs. Such an analysis may give

437

indications of possible metabolic activities.

ACS Paragon Plus Environment

Environmental Science & Technology

438 439

Acknowledgement

440

This study was funded by the Federal Ministry of Education and Research within the project DENANA

441

(03X0152). We thank Aaron Katz and Ajay Vikram Singh for proofreading the manuscript.

442 443

References

444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491

1. 2. 3. 4. 5. 6. 7. 8.

9. 10. 11. 12.

13.

14. 15. 16. 17.

Pastrana, H.; Avila, A.; Tsai, C. S. J., Nanomaterials in Cosmetic Products: the Challenges with regard to Current Legal Frameworks and Consumer Exposure. Nanoethics 2018, 12 (2), 123-137. Klebowski, B.; Depciuch, J.; Parlinska-Wojtan, M.; Baran, J., Applications of Noble Metal-Based Nanoparticles in Medicine. Int J Mol Sci 2018, 19 (12). Adawia, H. I. N., M.A.; Reed, J.M.; Vance; M.E.; Feitshansd, I.L.; Bickford, L.R.; Lewinski, N.A., Nano-enabled personal care products: Current developments in consumer safety. Nanoimpact 2018, 11, 170-179. Nowack, B.; Ranville, J. F.; Diamond, S.; Gallego-Urrea, J. A.; Metcalfe, C.; Rose, J.; Horne, N.; Koelmans, A. A.; Klaine, S. J., Potential scenarios for nanomaterial release and subsequent alteration in the environment. Environmental Toxicology and Chemistry 2012, 31 (1), 50-59. Wiesner, M. R.; Lowry, G. V.; Jones, K. L.; Hochella, M. F.; Di Giulio, R. T.; Casman, E.; Bernhardt, E. S., Decreasing Uncertainties in Assessing Environmental Exposure, Risk, and Ecological Implications of Nanomaterials. Environmental Science & Technology 2009, 43 (17), 6458-6462. Sharmaa, V. K. S., C.M.; Guo, B.; Pillai, S.; Parsons, J.G.; Wang, C.; Yan;, B.; Mag, X., Interactions between silver nanoparticles and other metal nanoparticles under environmentally relevant conditions: A review. Sci. Tot. Environ. 2019, 653, 1042-1051. Ma, C. X.; White, J. C.; Dhankher, O. P.; Xing, B. S., Metal-Based Nanotoxicity and Detoxification Pathways in Higher Plants. Environmental Science & Technology 2015, 49 (12), 7109-7122. Laux, P.; Riebeling, C.; Booth, A. M.; Brain, J. D.; Brunner, J.; Cerrillo, C.; Creutzenberg, O.; EstrelaLopis, I.; Gebel, T.; Johanson, G.; Jungnickel, H.; Kock, H.; Tentschert, J.; Tlili, A.; Schaffer, A.; Sips, A. J. A. M.; Yokel, R. A.; Luch, A., Challenges in characterizing the environmental fate and effects of carbon nanotubes and inorganic nanomaterials in aquatic systems. Environ Sci-Nano 2018, 5 (1), 48-63. Tripathi, D. K.; Shweta; Singh, S.; Singh, S.; Pandey, R.; Singh, V. P.; Sharma, N. C.; Prasad, S. M.; Dubey, N. K.; Chauhan, D. K., An overview on manufactured nanoparticles in plants: Uptake, translocation, accumulation and phytotoxicity. Plant Physiology and Biochemistry 2017, 110, 2-12. Cornelis, G.; Pang, L. P.; Doolette, C.; Kirby, J. K.; McLaughlin, M. J., Transport of silver nanoparticles in saturated columns of natural soils. Sci Total Environ 2013, 463, 120-130. Schlich, K.; Hund-Rinke, K., Influence of soil properties on the effect of silver nanomaterials on microbial activity in five soils. Environmental Pollution 2015, 196, 321-330. Wijnhoven, S. W. P.; Peijnenburg, W. J. G. M.; Herberts, C. A.; Hagens, W. I.; Oomen, A. G.; Heugens, E. H. W.; Roszek, B.; Bisschops, J.; Gosens, I.; Van de Meent, D.; Dekkers, S.; De Jong, W. H.; Van Zijverden, M.; Sips, A. J. A. M.; Geertsma, R. E., Nano-silver - a review of available data and knowledge gaps in human and environmental risk assessment. Nanotoxicology 2009, 3 (2), 109-U78. Laux, P.; Riebeling, C.; Booth, A. M.; Brain, J. D.; Brunner, J.; Cerrillo, C.; Creutzenberg, O.; EstrelaLopis, I.; Gebel, T.; Johanson, G.; Jungnickel, H.; Kock, H.; Tentschert, J.; Tlili, A.; Schaffer, A.; Sips, A.; Yokel, R. A.; Luch, A., Biokinetics of Nanomaterials: the Role of Biopersistence. NanoImpact 2017, 6, 69-80. Shen, C. X.; Zhang, Q. F.; Li, J. A.; Bi, F. C.; Yao, N., Induction of Programmed Cell Death in Arabidopsis and Rice by Single-Wall Carbon Nanotubes. American Journal of Botany 2010, 97 (10), 1602-1609. Khot, L. R.; Sankaran, S.; Maja, J. M.; Ehsani, R.; Schuster, E. W., Applications of nanomaterials in agricultural production and crop protection: A review. Crop Prot 2012, 35, 64-70. Rico, C. M., Peralta-Videa, J.R., Gardea-Torresdey, J.L., Chemistry, biochemistry of nanoparticles, and their role in antioxidant defense system in plants. Springer International Publishing (2015): 2015. Tripathi, D. K., Singh, S.., Singh, V.P., Prasad, S.M., Chauhan, D.K., Dubey N.K., Silicon Nanoparticles More Efficiently Alleviate Arsenate Toxicity than Silicon in Maize Cultiver and Hybrid Differing in Arsenate Tolerance. Front. Environ. Sci. 2016, https://doi.org/10.3389/fenvs.2016.00046. ACS Paragon Plus Environment

Page 18 of 22

Page 19 of 22

492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552

Environmental Science & Technology

18.

19.

20. 21. 22. 23. 24. 25.

26.

27. 28. 29.

30.

31.

32. 33. 34. 35. 36.

Krause, B.; Meyer, T.; Sieg, H.; Kastner, C.; Reichardt, P.; Tentschert, J.; Jungnickel, H.; Estrela-Lopis, I.; Burel, A.; Chevance, S.; Gauffre, F.; Jalili, P.; Meijer, J.; Bohmert, L.; Braeuning, A.; Thunemann, F.; Emmerling, F.; Fessard, V.; Laux, P.; Lampen, A.; Luch, A., Characterization of aluminum, aluminum oxide and titanium dioxide nanomaterials using a combination of methods for particle surface and size analysis. Rsc Adv 2018, 8 (26), 14377-14388. Sieg, H.; Braeuning, C.; Kunz, B.; Kaestner, C.; Jalili, P.; Krause, B. C.; Boehmert, L.; Lichtenstein, D.; Burel, A.; H., J.; Tentschert, J.; Laux, P.; Braeuning, A.; Gauffre, F.; Fessard, V.; Thunemann, A. F.; Luch, A.; Lampen, A., Uptake and molecular impact of aluminum-containing nanomaterials on intestinal human Caco-2 cells. Nanotoxicology 2018, accepted. Hossain, Z.; Mustafa, G.; Sakata, K.; Komatsu, S., Insights into the proteomic response of soybean towards Al2O3, ZnO, and Ag nanoparticles stress. Journal of Hazardous Materials 2016, 304, 291-305. Creutzenberg, O.; Bellmann, B.; Korolewitz, R.; Koch, W.; Mangelsdorf, I.; Tillmann, T.; Schaudien, D., Change in agglomeration status and toxicokinetic fate of various nanoparticles in vivo following lung exposure in rats. Inhalation Toxicology 2012, 24 (12), 821-830. Donaldson, K.; Poland, C. A., Nanotoxicity: challenging the myth of nano-specific toxicity. Current Opinion in Biotechnology 2013, 24 (4), 724-734. Krug, H. F., Nanosafety Research-Are We on the Right Track? Angewandte Chemie-International Edition 2014, 53 (46), 12304-12319. Gebel, T.; Foth, H.; Damm, G.; Freyberger, A.; Kramer, P. J.; Lilienblum, W.; Rohl, C.; Schupp, T.; Weiss, C.; Wollin, K. M.; Hengstler, J. G., Manufactured nanomaterials: categorization and approaches to hazard assessment. Arch Toxicol 2014, 88 (12), 2191-211. Hund-Rinke, K.; Baun, A.; Cupi, D.; Fernandes, T. F.; Handy, R.; Kinross, J. H.; Navas, J. M.; Peijnenburg, W.; Schlich, K.; Shaw, B. J.; Scott-Fordsmand, J. J., Regulatory ecotoxicity testing of nanomaterials - proposed modifications of OECD test guidelines based on laboratory experience with silver and titanium dioxide nanoparticles. Nanotoxicology 2016, 10 (10), 1442-1447. Laux, P.; Tentschert, J.; Riebeling, C.; Braeuning, A.; Creutzenberg, O.; Epp, A.; Fessard, V.; Haas, K. H.; Haase, A.; Hund-Rinke, K.; Jakubowski, N.; Kearns, P.; Lampen, A.; Rauscher, H.; Schoonjans, R.; Stormer, A.; Thielmann, A.; Muhle, U.; Luch, A., Nanomaterials: certain aspects of application, risk assessment and risk communication. Arch Toxicol 2018, 92 (1), 121-141. Rossi, L.; Zhang, W.; Schwab, A. P.; Ma, X., Uptake, Accumulation, and in Planta Distribution of Coexisting Cerium Oxide Nanoparticles and Cadmium in Glycine max (L.) Merr. Environ Sci Technol 2017, 51 (21), 12815-12824. Rossi, L.; Zhang, W. L.; Ma, X. M., Cerium oxide nanoparticles alter the salt stress tolerance of Brassica napus L. by modifying the formation of root apoplastic barriers. Environmental Pollution 2017, 229, 132-138. Dan, Y. B.; Zhang, W. L.; Xue, R. M.; Ma, X. M.; Stephan, C.; Shi, H. L., Characterization of Gold Nanoparticle Uptake by Tomato Plants Using Enzymatic Extraction Followed by Single-Particle Inductively Coupled Plasma-Mass Spectrometry Analysis. Environmental Science & Technology 2015, 49 (5), 3007-3014. Lopez-Moreno, M. L.; de la Rosa, G.; Hernandez-Viezcas, J. A.; Peralta-Videa, J. R.; GardeaTorresdey, J. L., X-ray Absorption Spectroscopy (XAS) Corroboration of the Uptake and Storage of CeO2 Nanoparticles and Assessment of Their Differential Toxicity in Four Edible Plant Species. Journal of Agricultural and Food Chemistry 2010, 58 (6), 3689-3693. Zhao, L.; Peralta-Videa, J. R.; Varela-Ramirez, A.; Castillo-Michel, H.; Li, C.; Zhang, J.; Aguilera, R. J.; Keller, A. A.; Gardea-Torresdey, J. L., Effect of surface coating and organic matter on the uptake of CeO2 NPs by corn plants grown in soil: Insight into the uptake mechanism. J Hazard Mater 2012, 225226, 131-8. Rui, Y.; Zhang, P.; Zhang, Y.; Ma, Y.; He, X.; Gui, X.; Li, Y.; Zhang, J.; Zheng, L.; Chu, S.; Guo, Z.; Chai, Z.; Zhao, Y.; Zhang, Z., Transformation of ceria nanoparticles in cucumber plants is influenced by phosphate. Environ Pollut 2015, 198, 8-14. Mirzajani, F.; Askari, H.; Hamzelou, S.; Schober, Y.; Rompp, A.; Ghassempour, A.; Spengler, B., Proteomics study of silver nanoparticles toxicity on Oryza sativa L. Ecotoxicol Environ Saf 2014, 108, 335-9. Mejias, F. J. R.; Lopez-Haro, M.; Gontard, L. C.; Cala, A.; Fernandez-Aparicio, M.; Molinillo, J. M. G.; Calvino, J. J.; Macias, F. A., A Novel Electron Microscopic Characterization of Core/Shell Nanobiostimulator Against Parasitic Plants. Acs Appl Mater Inter 2018, 10 (3), 2354-2359. Zhang, P.; Ma, Y. H.; Liu, S. T.; Wang, G. H.; Zhang, J. Z.; He, X.; Zhang, J.; Rui, Y. K.; Zhang, Z. Y., Phytotoxicity, uptake and transformation of nano-CeO2 in sand cultured romaine lettuce. Environmental Pollution 2017, 220, 1400-1408. Komaty, S.; Letertre, M.; Dang, H. D.; Jungnickel, H.; Laux, P.; Luch, A.; Carrie, D.; MerdrignacConanec, O.; Bazureau, J. P.; Gauffre, F.; Tomasi, S.; Paquin, L., Sample preparation for an optimized

ACS Paragon Plus Environment

Environmental Science & Technology

553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614

37.

38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49.

50. 51. 52. 53.

54. 55. 56. 57. 58.

extraction of localized metabolites in lichens: Application to Pseudevernia furfuracea. Talanta 2016, 150, 525-530. Meyer, T. V., T.; Sieg, H.; Boehmert, L.; Kunz, B. M.; Krause, B.; Jalili, P.; Hoogeveen, S.; Chevance, S.; Gauffre, F.; Burel, A.; Jungnickel, H.; Tentschert, J.; P. Laux, P.; Luch, A.; Braeuning, A.; Lampen, A.; Meijer, J.; Fessard, V.; Estrela-Lopis, I, Simultaneous Quantification and Visualization of Titanium Dioxide Nanomaterial Uptake at Single Cell Level in an In Vitro Model of the Human Small Intestine. Small Methods 2019, accepted for publication. Fletcher, J. S.; Lockyer, N. P.; Vaidyanathan, S.; Vickerman, J. C., TOF-SIMS 3D biomolecular imaging of Xenopus laevis oocytes using buckminsterfullerene (C-60) primary ions. Anal Chem 2007, 79 (6), 2199-2206. Booth, A.; Storseth, T.; Altin, D.; Fornara, A.; Ahniyaz, A.; Jungnickel, H.; Laux, P.; Luch, A.; Sorensen, L., Freshwater dispersion stability of PAA-stabilised-cerium-oxide nanoparticles and toxicity towards Pseudokirchneriella subcapitata. Sci Total Environ 2015, 505, 596-605. Jungnickel, H.; Pund, R.; Tentschert, J.; Reichardt, P.; Laux, P.; Harbach, H.; Luch, A., Time-of-flight secondary ion mass spectrometry (ToF-SIMS)-based analysis and imaging of polyethylene microplastics formation during sea surf simulation. Sci Total Environ 2016, 563, 261-266. Schlich, K.; Hoppe, M.; Kraas, M.; Fries, E.; Hund-Rinke, K., Ecotoxicity and fate of a silver nanomaterial in an outdoor lysimeter study. Ecotoxicology 2017, 26 (6), 738-751. Du, W. C.; Gardea-Torresdey, J. L.; Ji, R.; Yin, Y.; Zhu, J. G.; Peralta-Videa, J. R.; Guo, H. Y., Physiological and Biochemical Changes Imposed by CeO2 Nanoparticles on Triticum aestivum: A Life Cycle Field Study. Environmental Science & Technology 2015, 49 (19), 11884-11893. Piccinno, F.; Gottschalk, F.; Seeger, S.; Nowack, B., Industrial production quantities and uses of ten engineered nanomaterials in Europe and the world. Journal of Nanoparticle Research 2012, 14 (9). Kim, J. S.; Kuk, E.; Yu, K. N.; Kim, J. H.; Park, S. J.; Lee, H. J.; Kim, S. H.; Park, Y. K.; Park, Y. H.; Hwang, C. Y.; Kim, Y. K.; Lee, Y. S.; Jeong, D. H.; Cho, M. H., Antimicrobial effects of silver nanoparticles. Nanomedicine-Nanotechnology Biology and Medicine 2007, 3 (1), 95-101. Panacek, A.; Kvitek, L.; Prucek, R.; Kolar, M.; Vecerova, R.; Pizurova, N.; Sharma, V. K.; Nevecna, T.; Zboril, R., Silver colloid nanoparticles: Synthesis, characterization, and their antibacterial activity. Journal of Physical Chemistry B 2006, 110 (33), 16248-16253. Deshmukh, S. P.; Patil, S. M.; Mullani, S. B.; Delekar, S. D., Silver nanoparticles as an effective disinfectant: A review. Mater Sci Eng C Mater Biol Appl 2019, 97, 954-965. EPA, U. S., Toxicological Review of Cerium Oxide and Cerium Compounds. EPA/635/R-08/002F. http://www.epa.gov/iris/toxreviews/1018tr.pdf. Sept. 2009. Garcia, T.; Solsona, B.; Taylor, S. H., Nano-crystalline ceria catalysts for the abatement of polycyclic aromatic hydrocarbons. Catal Lett 2005, 105 (3-4), 183-189. Clar, J. G.; Platten, W. E.; Baumann, E. J.; Remsen, A.; Harmon, S. M.; Bennett-Stamper, C. L.; Thomas, T. A.; Luxton, T. P., Dermal transfer and environmental release of CeO2 nanoparticles used as UV inhibitors on outdoor surfaces: Implications for human and environmental health. Sci Total Environ 2018, 613, 714-723. Sendra, M.; Volland, M.; Balbi, T.; Fabbri, R.; Yeste, M. P.; Gatica, J. M.; Canesi, L.; Blasco, J., Cytotoxicity of CeO2 nanoparticles using in vitro assay with Mytilus galloprovincialis hemocytes: Relevance of zeta potential, shape and biocorona formation. Aquat Toxicol 2018, 200, 13-20. German seage sludge allowance regulation. 1992, §6 Absatz 1. Singh, C., Friedrichs, S., Ceccone, G., Gibson, N., Jensen, K.A., Levin, M., Goenaga Infante, H., Carlander, D., Rasmussen, K., NM-series of Representative Manufactured Nanomaterials - Cerium Dioxide, NM-211, NM-212, NM-213. Characterisation and test item preparation. 2014. Klein, C. L., Comero, S., Stahlmecke, B., Romazanov, J., Kuhlbusch, T.A.J., Van Doren, E., De Temmerman P.-J., Mast, J., Wick, P., Krug, H., Locoro, G., Hund-Rinke, K., Kördel, W., Friedrichs, S., Maier, G., Werner, J., ; Linsinger, T., Gawlik, B.M. NM-Series of Representative Manufactured Nanomaterials - NM-300 Silver - Characterisation, Stability, Homogeneity; 2011. Wagener, S.; Dommershausen, N.; Jungnickel, H.; Laux, P.; Mitrano, D.; Nowack, B.; Schneider, G.; Luch, A., Textile Functionalization and Its Effects on the Release of Silver Nanoparticles into Artificial Sweat. Environmental Science & Technology 2016, 50 (11), 5927-5934. Lee, S.; Bi, X. Y.; Reed, R. B.; Ranville, J. F.; Herckes, P.; Westerhoff, P., Nanoparticle Size Detection Limits by Single Particle ICP-MS for 40 Elements. Environmental Science & Technology 2014, 48 (17), 10291-10300. Senoner, M.; Unger, W. E. S., SIMS imaging of the nanoworld: applications in science and technology. J. Anal. Atom. Spectrom. 2012, 27, 1050-1068. Holzweber, M.; Shard, A. G.; Jungnickel, H.; Luch, A.; Unger, W. E. S., Dual Beam Organic Depth Profiling using Large Argon Cluster Ion Beams. Surf Interface Anal 2014, 46, 936-939. Lopez-Moreno, M. L.; de la Rosa, G.; Hernandez-Viezcas, J. A.; Castillo-Michel, H.; Botez, C. E.; Peralta-Videa, J. R.; Gardea-Torresdey, J. L., Evidence of the Differential Biotransformation and ACS Paragon Plus Environment

Page 20 of 22

Page 21 of 22

615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671

Environmental Science & Technology

59. 60. 61. 62. 63. 64. 65. 66. 67.

68. 69. 70. 71.

72.

73. 74. 75. 76.

77.

Genotoxicity of ZnO and CeO2 Nanoparticles on Soybean (Glycine max) Plants. Environmental Science & Technology 2010, 44 (19), 7315-7320. Arruda, S. C. C.; Silva, A. L. D.; Galazzi, R. M.; Azevedo, R. A.; Arruda, M. A. Z., Nanoparticles applied to plant science: A review. Talanta 2015, 131, 693-705. Rico, C. M.; Majumdar, S.; Duarte-Gardea, M.; Peralta-Videa, J. R.; Gardea-Torresdey, J. L., Interaction of Nanoparticles with Edible Plants and Their Possible Implications in the Food Chain. Journal of Agricultural and Food Chemistry 2011, 59 (8), 3485-3498. Judy, J. D.; Bertsch, P. M., Bioavailability, Toxicity, and Fate of Manufactured Nanomaterials in Terrestrial Ecosystems. Advances in Agronomy, Vol 123 2014, 123, 1-64. Mitrano, D. M.; Rimmele, E.; Wichser, A.; Erni, R.; Height, M.; Nowack, B., Presence of Nanoparticles in Wash Water from Conventional Silver and Nano-silver Textiles. Acs Nano 2014, 8 (7), 7208-7219. Rastogi, A.; Zivcak, M.; Sytar, O.; Kalaji, H. M.; He, X. L.; Mbarki, S.; Brestic, M., Impact of Metal and Metal Oxide Nanoparticles on Plant: A Critical Review. Frontiers in Chemistry 2017, 5. Dietz, K. J.; Herth, S., Plant nanotoxicology. Trends in Plant Science 2011, 16 (11), 582-589. Kurepa, J.; Paunesku, T.; Vogt, S.; Arora, H.; Rabatic, B. M.; Lu, J. J.; Wanzer, M. B.; Woloschak, G. E.; Smalle, J. A., Uptake and Distribution of Ultrasmall Anatase TiO2 Alizarin Red S Nanoconjugates in Arabidopsis thaliana. Nano Letters 2010, 10 (7), 2296-2302. Aslani, F.; Bagheri, S.; Julkapli, N. M.; Juraimi, A. S.; Hashemi, F. S. G.; Baghdadi, A., Effects of Engineered Nanomaterials on Plants Growth: An Overview. Scientific World Journal 2014. Ma, Y. H.; He, X.; Zhang, P.; Zhang, Z. Y.; Ding, Y. Y.; Zhang, J. Z.; Wang, G. H.; Xie, C. J.; Luo, W. H.; Zhang, J.; Zheng, L. R.; Chai, Z. F.; Yang, K., Xylem and Phloem Based Transport of CeO2 Nanoparticles in Hydroponic Cucumber Plants. Environmental Science & Technology 2017, 51 (9), 5215-5221. Wang, Z. Y.; Xie, X. Y.; Zhao, J.; Liu, X. Y.; Feng, W. Q.; White, J. C.; Xing, B. S., Xylem- and Phloem-Based Transport of CuO Nanoparticles in Maize (Zea mays L.). Environmental Science & Technology 2012, 46 (8), 4434-4441. Brison, J.; Benoit, D. S. W.; Muramoto, S.; Robinson, M.; Stayton, P. S.; Castner, D. G., ToF-SIMS imaging and depth profiling of HeLa cells treated with bromodeoxyuridine. Surface and Interface Analysis 2011, 43 (1-2), 354-357. Draude, F.; Galla, S.; Pelster, A.; Tentschert, J.; Jungnickel, H.; Haase, A.; Mantion, A.; Thunemann, A. F.; Taubert, A.; Luch, A.; Arlinghaus, H. F., ToF-SIMS and Laser-SNMS analysis of macrophages after exposure to silver nanoparticles. Surface and Interface Analysis 2013, 45 (1), 286-289. Haase, A.; Arlinghaus, H. F.; Tentschert, J.; Jungnickel, H.; Graf, P.; Mantion, A.; Draude, F.; Galla, S.; Plendl, J.; Goetz, M. E.; Masic, A.; Meier, W.; Thunemann, A. F.; Taubert, A.; Luch, A., Application of Laser Postionization Secondary Neutral Mass Spectrometry/Time-of-Flight Secondary Ion Mass Spectrometry in Nanotoxicology: Visualization of Nanosilver in Human Macrophages and Cellular Responses. Acs Nano 2011, 5 (4), 3059-3068. Graham, U. M.; Tseng, M. T.; Jasinski, J. B.; Yokel, R. A.; Unrine, J. M.; Davis, B. H.; Dozier, A. K.; Hardas, S. S.; Sultana, R.; Grulke, E. A.; Butterfield, D. A., In Vivo Processing of Ceria Nanoparticles inside Liver: Impact on Free-Radical Scavenging Activity and Oxidative Stress. Chempluschem 2014, 79 (8), 1083-1088. Zhang, P.; Ma, Y. H.; Zhang, Z. Y.; He, X.; Zhang, J.; Guo, Z.; Tai, R. Z.; Zhao, Y. L.; Chai, Z. F., Biotransformation of Ceria Nanoparticles in Cucumber Plants. Acs Nano 2012, 6 (11), 9943-9950. Zenk, M. H., Heavy metal detoxification in higher plants - A review. Gene 1996, 179 (1), 21-30. Zhang, P.; Ma, Y. H.; Zhang, Z. Y.; He, X.; Li, Y. Y.; Zhang, J.; Zheng, L. R.; Zhao, Y. L., Speciesspecific toxicity of ceria nanoparticles to Lactuca plants. Nanotoxicology 2015, 9 (1), 1-8. Klein, C.L.; Comero, S.; Stahlmecke, B.; Romazanov, J.; Kuhlbusch, T.A.J.; VanDoren, E.; DeTemmerman, P.J.; Mast, J.; Wick, P.; Krug, H.; Locoro, G.; Hund-Rinke, K.; Koerdel, W.; Friedrichs, S.; Maier, G.; Werner, J.; Linsinger, Th.; Gawlik, B.M. NM-Series of Representative Manufactured Nanomaterials NM-300 Silver Characterisation, Stability, Homogeneity. JRC Scientific and Technical Report, European Commission Joint Research Centre, Institute for Health and Consumer Protection: Luxembourg, 2011; p 86. Singh, C.; Friedrichs, S.; Ceccone, G.; Gibson, P.; Jensen, K.A.; Levin, M.; Goenaga-Infante, H.; Carlander, D.; Rasmussen, K. Cerium Dioxide, NM-211, NM-212, NM-213. Characterisation and test item preparation. JRC Scientific and Technical Report, European Commission, http://dx.doi.org/10.2788/80203. EUR 26649.

672 673

ACS Paragon Plus Environment

Environmental Science & Technology

674

Supporting information

675 676

-

Scheme for the analysis of plant tissues with ICP-MS, spICP-MS and ToF-SIMS

677

-

Detection and characterization of CeO2-NPs in Triticum aestivum root

678

-

Ag distribution in a longitudinal tissue section of Brassica napus exposed to Ag-NP

679

-

CeO+ distribution in a longitudinal Brassica napus tissue, exposed to CeO2-NP

680

-

Particle size and crystallinity data of the tested nanomaterials

ACS Paragon Plus Environment

Page 22 of 22