Nanoparticles Meet Cell Membranes - American Chemical Society

Dec 16, 2013 - leaflet or within the hydrophobic core of the bilayer). If such ... will allow for the validation of experimental data and also the elu...
0 downloads 10 Views 2MB Size
Subscriber access provided by DUESSELDORF LIBRARIES

Feature

Nanoparticles Meet Cell Membranes: Probing Nonspecific Interactions using Model Membranes Kai Loon Chen, and Geoffrey D. Bothun Environ. Sci. Technol., Just Accepted Manuscript • Publication Date (Web): 16 Dec 2013 Downloaded from http://pubs.acs.org on December 22, 2013

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

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

Page 1 of 22

Environmental Science & Technology

1   2   3   4  

Nanoparticles Meet Cell Membranes: Probing Nonspecific

5  

Interactions using Model Membranes

6   7   8  

Feature Article

9   10   11  

Environmental Science & Technology

12  

Revised on December 10, 2013

13   14  

Kai Loon Chen*, † and Geoffrey D. Bothun*, ‡

15   16   17  



University, Baltimore, Maryland 21218-2686

18   19   20  

Department of Geography and Environmental Engineering, Johns Hopkins



Department of Chemical Engineering, University of Rhode Island, Kingston, Rhode Island 02881-2018

21   22   23  

*Co-corresponding authors: Kai Loon Chen, E-mail: [email protected], Phone: (410) 516-7095;

24  

Geoffrey D. Bothun, E-mail: [email protected], Phone: (401) 874-9518. 1   ACS Paragon Plus Environment

Environmental Science & Technology

25  

Abstract

26  

Nanotoxicity studies have shown that both carbon-based and inorganic engineered

27  

nanoparticles can be toxic to microorganisms. Although the pathways for cytotoxicity are

28  

diverse and dependent upon the nature of the engineered nanoparticle and the chemical

29  

environment, these studies have provided evidence that direct contact between nanoparticles and

30  

bacterial cell membranes is necessary for cell inactivation or damage, and may in fact be a

31  

primary mechanism for cytotoxicity. The propensities for nanoparticles to attach to and disrupt

32  

cell membranes are still not well understood due to the heterogeneous and dynamic nature of

33  

biological membranes.

34  

investigations of nanoparticle–membrane interactions. In this article, current and emerging

35  

experimental approaches to identify the key parameters that control the attachment of ENPs on

36  

model membranes and the disruption of membranes by ENPs will be discussed. This critical

37  

information will help enable the “safe-by-design” production of engineered nanoparticles that are

38  

non-toxic or biocompatible, and also allow for the design of antimicrobial nanoparticles for

39  

environmental and biomedical applications.

Model biological membranes can be employed for systematic

40   41  

TOC/Abstract Art

42   43   44   2   ACS Paragon Plus Environment

Page 2 of 22

Page 3 of 22

45  

Environmental Science & Technology

Introduction

46  

The fast-growing utilization of nanomaterials in diverse applications, including electronic

47  

devices, drug and gene delivery, consumer products, and environmental remediation, will

48  

undoubtedly lead to the release of these materials into the environment.1

49  

nanoparticles (ENPs) that are released into natural and engineered aquatic systems can undergo

50  

physical and chemical transformation, and the nature and degree of transformation are directly

51  

dependent on the solution chemistry, environmental conditions (e.g., sunlight and temperature),

52  

and constituents present in the environment.2 The transformation of ENPs will influence their

53  

mobility and transport,3 as well as their biological effects on microorganisms and higher

54  

organisms.4

Engineered

55  

While evidence from recent studies has shown that both carbon-based and inorganic

56  

ENPs can be toxic to microorganisms, the mechanisms for their cytotoxicity are varied and not

57  

always completely elucidated. Carbon nanotubes (CNTs) can cause cellular membrane damage

58  

upon direct CNT–membrane contact with microorganisms and hence result in their inactivation.5-

59  

8

60  

multiwalled CNTs (MWNTs) and the toxicity of the CNTs was attributed to a combination of

61  

physicochemical interactions with the cell membranes and oxidative stress.5, 6 Metallic SWNTs

62  

were recently discovered to be more toxic to Escherichia coli than semiconducting SWNTs and

63  

the adhesion of SWNTs to E. coli cells was necessary before the toxicity pathways occurred.7

64  

Individually dispersed SWNTs were more toxic to both gram-positive and gram-negative

65  

bacteria than SWNT aggregates and it was postulated that dispersed SWNTs were more effective

66  

in piercing cell membranes than SWNT aggregates.8 Graphene oxide nanosheets have also been

67  

reported to cause membrane and oxidative stress to E. coli upon attachment to the bacterial

68  

cells.9

Single-walled CNTs (SWNTs) were found to exhibit a higher level of toxicity compared to

69  

While direct contact between carbon-based ENPs and bacterial cells has been shown to

70  

result in membrane damage and cell inactivation, similar observations have been made for

71  

inorganic ENPs. The dissolution of silver nanoparticles (AgNPs) resulting in the release of Ag+

72  

ions is expected to be a key mechanism for their cytotoxicity.10 The attachment or close

73  

proximity of AgNPs to cell membranes will thus enhance the exposure of microorganisms to the

74  

ionic Ag+ species. AgNPs can also accumulate in the membranes of E. coli cells and cause pit

75  

formation in the cell walls.11

Similarly, ZnO nanoparticles were found to damage the 3   ACS Paragon Plus Environment

Environmental Science & Technology

76  

membranes of E. coli cells and internalization of the nanoparticles was observed through

77  

transmission electron microscopy (TEM).12

78  

Based on existing nanotoxicity studies, it is apparent that the attachment of ENPs to the

79  

membranes of microorganisms is the critical initial process that precedes the toxicity pathways.13

80  

For example, greater nanoparticle adhesion has been observed to correlate with greater cellular

81  

internalization.14-17 Because cell membranes are complex and dynamic and contain multiple

82  

components both within the membranes and on the membrane surface (such as phospholipid

83  

bilayers, proteins, extracellular polymeric substances, and lipopolysaccharides),18 it is

84  

advantageous to employ model biological membranes of known compositions to systematically

85  

investigate the key parameters that control the attachment of ENP attachment. Likewise, the role

86  

of the biophysical and chemical properties of cell membranes, as well as the physicochemical

87  

properties of ENPs, on the propensity for physical disruption or penetration of cell membranes is

88  

still not well understood. Model biological membranes have the potential to be used to elucidate

89  

the mechanisms for the disruption of membranes by ENPs.

90  

In this article, we present commonly used models for biological membranes and highlight

91  

several techniques that can be employed to investigate the nonspecific interactions (or

92  

interactions that do not involve specific cell receptors) between ENPs and model cell membranes.

93  

We focus on experimental approaches that detect the attachment of ENPs on model membranes,

94  

as well as the physical disruption of model membranes by ENPs. Since the area of nanoparticle–

95  

membrane interactions is still relatively new, challenges related to these types of measurements

96  

and opportunities to further this field of study are also discussed.

97   98  

Models for Biological Membranes

99  

In addition to providing mechanical structure and separating intracellular and

100  

extracellular environments, the main role of the membrane is to provide an anisotropic fluid

101  

phase for supporting proteins and to regulate molecular transport into and out of the cell (e.g.,

102  

resisting water and controlling ion permeation). Figure 1 shows the main outer-most structural

103  

components, excluding integral and peripheral proteins, of bacterial, plant, and mammalian cell

104  

membranes. In general, all cell membranes exhibit negative charge and contain structural

105  

carbohydrate-rich layers anchored to a lipid bilayer membrane (intracellular cytoskeletons are

106  

not shown in Figure 1). They are not equilibrium structures, but rather dynamic structures that 4   ACS Paragon Plus Environment

Page 4 of 22

Page 5 of 22

Environmental Science & Technology

107  

undergo transient chemical and physical changes depending on the environment the cells are

108  

exposed to and the stage the cells are in within their lifecycle. Given the inherent complexity of

109  

cell membranes that spans multiple length scales, model cell membranes composed of natural or

110  

synthetic lipid bilayers are typically used to gain fundamental insight into membrane

111  

organization and structure. In addition to reducing complexity, using model membranes

112  

eliminates the effects of cell metabolism and growth. FIGURE 1

113   114  

Lipid bilayers have been used as model membranes to examine the effects of pollutants,

115  

such as surfactants, organics, and metal ions on membrane-related cytotoxicity.19 Model

116  

membranes can be formed with biologically-relevant lipids and used to analyze mechanical,

117  

thermodynamic, and kinetic information relating to how compounds partition into membranes

118  

and cause disruption. In bilayers, partitioning and disruption are often assessed by measuring

119  

changes in lipid phase behavior (i.e., thermal transitions between ordered and disordered lipid

120  

phases), curvature, elasticity, and permeability, which are tied to membrane structure, domain

121  

formation, and pore formation. Systematically varying membrane composition provides a

122  

hierarchical approach where interaction mechanisms can be determined in ‘simple’ models and

123  

then extended to increasingly ‘lifelike’ membranes. Model membrane techniques can be

124  

extended to ENP–membrane systems to determine how nanoparticle size, shape, and surface

125  

chemistry influence their interactions.

126  

ENPs have been shown to damage bacterial and mammalian membranes through

127  

membrane disruption and to internalize cells through passive or active membrane transport.20

128  

Cytotoxic events associated with nanoparticle–membrane interactions involve, for example,

129  

destabilization of membrane proteins and membrane leakage,15, 16 which, among other things,

130  

affects transmembrane pH and ion gradients. These properties depend critically on membrane

131  

order and structure, as well as the membrane’s resistance to disordering or restructuring. There is

132  

precedence to using model membranes to examine membrane disordering or restructuring, and

133  

this precedence has been extended to nanoparticle–membrane interactions.

134  

Geometrical and experimental considerations. As in cellular membranes, model

135  

membranes are composed of lipid bilayers that self-assemble via hydrophobic forces. Model

136  

membranes are generally prepared in spherical or planar geometries (Figure 2). Spherical

137  

geometries consist of freestanding (unsupported) lipid bilayer vesicles,21 supported vesicles that 5   ACS Paragon Plus Environment

Environmental Science & Technology

138  

are deposited intact on a substrate,22 or bilayers supported on spherical nanoparticles or

139  

microparticles that are dispersed in an aqueous phase.23, 24 Planar geometries include bilayers

140  

supported on planar substrates,25,

141  

deposited at air/water or solvent/water interfaces.28 The geometry employed depends on the

142  

types of measurements to be conducted (Table 1). Spherical vesicles provide high lipid

143  

concentrations to examine nanoparticle binding, lipid phase behavior, and membrane

144  

permeabilization using, for example, calorimetric or spectroscopic techniques. Varying vesicle

145  

size allows one to determine the effects of membrane curvature, which are related to membrane

146  

compressibility and elasticity,29 on modulating nanoparticle interactions. Spherical vesicles are

147  

also amenable to cryogenic transmission electron microscopy (cryo-TEM) techniques. In planar

148  

geometries, the lipid concentration is dictated by the size of the support, interface, or orifice. A

149  

distinct advantage of planar geometries is that they are amenable to a range of microscopy and

150  

spectroscopy techniques commonly employed in surface science.

26

bilayers suspended across orifices,27 and monolayers

151  

FIGURE 2

152  

TABLE 1

153  

Once formed, intermolecular forces between lipid molecules govern the organization and

154  

phase behavior of a membrane, as well as the membrane’s integrity or stability against physical

155  

deformation. Membrane-active molecules can bind to the surface or partition into membranes

156  

and disrupt these forces. ENPs can also bind to membrane surfaces, but in this case the particle

157  

acts more like a solid surface and interacts locally with groups of lipid molecules or lipid

158  

domains. ENP–membrane interaction schemes are depicted in Figure 3. Once a nanoparticle

159  

adheres to a membrane, driven by intermolecular and surface forces, it can lead to lipid

160  

restructuring, domain formation, and local deformation (Figure 3A, B).30, 31 These processes can

161  

lead to membrane leakage due to transient voids or pore formation (Figure 3B),21, 32 passive

162  

membrane translocation of the ENP due to poration or invagination (Figure 3C),12, 20 or lipid

163  

extraction from the membrane (Figure 3D).28, 31 Experimentally, ENP binding can be examined

164  

using any model membrane configuration, while membrane leakage studies are limited to

165  

configurations that yield freestanding membranes (i.e., where aqueous reservoirs are in contact

166  

with both sides of the membrane, analogous to cellular membranes separating intracellular and

167  

extracellular fluids). However, unsupported freestanding membrane vesicles are the best

168  

candidates to examine simultaneous binding and leakage processes as solid supports or confined 6   ACS Paragon Plus Environment

Page 6 of 22

Page 7 of 22

Environmental Science & Technology

169  

geometries influence lipid organization and phase behavior, and restrict elastic membrane

170  

deformation. To further add to the complexity, these same properties that can be influenced by

171  

ENP binding (lipid organization, phase behavior, and elasticity) can also affect the nature and

172  

extent of ENP binding. While there are preferred model membrane geometries for determining

173  

ENP binding and membrane disruption, there is no single experimental approach that provides a

174  

complete picture of the processes. FIGURE 3

175   176   177  

Probing Interactions between Nanoparticles and Model Membranes:

178  

Attachment and Disruption

179  

Several recent nanotoxicity studies have shown that ENPs can adsorb and penetrate (or

180  

disrupt) bacterial and mammalian cell membranes, both processes likely to play important roles

181  

in NP toxicity.11, 33 The approach of ENPs towards cell membranes resulting in either their

182  

attachment or close proximity to the membrane surface is expected to be the critical step

183  

preceding the toxicity pathways that result in cell inactivation or damage (including nanoparticle

184  

uptake or penetration). However, beyond the obvious electrostatic interactions, very little is

185  

currently known about the factors that control the interactions between ENPs and cell

186  

membranes. This includes how these interactions act to resist membrane disruption by ENPs;

187  

how ENP binding influences intermolecular lipid interactions and membrane organization; and

188  

how the shape and orientation of isotropic ENPs with respect to cell membranes, as well as the

189  

nanoparticle surface chemistry, affect ENP–membrane interactions.

190  

Since biological membranes, as well as most ENPs, carry surface charges, Derjaguin–

191  

Landau–Verwey–Overbeek (DLVO) interactions,34 namely, electric-double layer and van der

192  

Waals interactions, likely play important roles in controlling the propensity of the ENPs to attach

193  

on cell membranes.35 A recent study has reported a correlation between electrostatic attraction

194  

and bacterial minimal inhibitory concentrations for ZnO.36 Furthermore, because phospholipid

195  

bilayers, a key component of the cell membranes, are extremely hydrophilic and undergo

196  

dynamic fluctuations,34, 37 repulsive hydration and undulation forces are expected to contribute to

197  

ENP–membrane interactions. A detailed discussion on these ENP–membrane interfacial forces

198  

is presented in Nel et al.’s article,18 while Negoda et al.38 have recently reviewed current

199  

experimental and theoretical methods to characterize ENP–membrane interactions. 7   ACS Paragon Plus Environment

In this

Environmental Science & Technology

Page 8 of 22

200  

feature article, we focus on complimentary techniques focused more at the biophysical level that

201  

can experimentally assess the mechanisms by which ENPs attach to and disrupt (or rupture)

202  

membranes.

203  

Attachment of ENPs on Model Membranes. Atomic force microscopy (AFM) has

204  

been used to examine the adsorption of ENPs on synthetic membranes.

Currently, most

205  

published work on ENP–model membrane interactions has reported the use of AFM to obtain a

206  

top image of supported lipid bilayers (SLBs; Figure 2B) that have been exposed to ENPs.33, 39, 40

207  

AFM can also be used to obtain a cross-sectional analysis of the SLBs to determine the degree of

208  

penetration of ENPs into the model membranes.39, 40

209  

Other than the AFM, optical tweezers have been employed to study ENP–membrane

210  

interactions. Rusciano et al.41 employed optical tweezers to trap phospholipid vesicles that were

211  

exposed to carbon ENPs. By performing Raman spectroscopy analysis, the authors were able to

212  

conclude from their measurements that the carbon ENPs can either partition within or penetrate

213  

the bilayers (they were unable to distinguish the two processes).41 However, the optical tweezer

214  

setup is relatively sophisticated and is still not commercially available. Moreover, it does not

215  

allow for the rapid and convenient analysis of the propensity of an ENP to adsorb or penetrate a

216  

model membrane.

217  

Since a tremendous amount of research on lipid bilayers has already been conducted with

218  

the quartz crystal microbalance with dissipation monitoring (QCM-D),25, 42 this technique holds a

219  

great promise in the measurements of the interactions of ENPs and model cell membranes.13, 22, 43

220  

The QCM-D is commercially available and it allows for the monitoring of the frequency and

221  

dissipation responses of a quartz crystal as adsorption of polymers, polyelectrolytes, or

222  

nanoparticles takes place on the crystal surface. By monitoring both signals, the mass and

223  

viscoelastic properties of the adsorbed layer can be determined real time. Hence, the QCM-D

224  

technique allows for the in situ detection of ENP adsorption on model membranes. Other than

225  

the ability for real-time measurements, the QCM-D technique is highly sensitive and can detect a

226  

mass change on the crystal of as low as tens of nanograms. Thus, this technique will be sensitive

227  

enough for ENP suspensions of low concentrations, as well as ENPs with low propensities to

228  

adsorb on cell membranes. Furthermore, only a small volume of ENP suspension (about a few

229  

milliliters) is required for this assay because the volume of the flow chamber of the QCM-D is

230  

only about 0.05 cm3. Also, because of the flow-through design of the QCM-D, this technique 8   ACS Paragon Plus Environment

Page 9 of 22

Environmental Science & Technology

231  

potentially can be automated and included as one of the components in a production/process train

232  

in a nanomaterial production plant to assess the nanomaterial’s propensity to adsorb on cell

233  

membranes.

234  

Using the approach of Richter et al.,25,

26

Yi and Chen22 recently assembled a SLB

235  

composed of zwitterionic 1,2-dioleoyl-sn-glyero-3-phosphocholine (DOPC) in a QCM-D to

236  

investigate the effects of solution chemistry on the adsorption of oxidized MWNTs on the model

237  

biological membranes. They demonstrated that the presence of Ca2+ cations, as well as low pH

238  

conditions, will allow for favorable deposition of MWNTs on DOPC membranes through

239  

electrostatic attraction. These findings are consistent with observations made by other groups

240  

that electric-double layer interactions play an important role in controlling the attachment of

241  

ENPs on membranes.23, 24 While the QCM-D can measure the adsorption of ENPs to a SLB, it

242  

does not provide information about the location of the adsorbed ENPs with respect to the bilayer

243  

(e.g., on the outer leaflet or within the hydrophobic core of the bilayer). If such information is

244  

needed, AFM imaging39, 40 or cryo-TEM (discussed in later section) can be used to locate the

245  

ENPs on or in the bilayer.

246  

Disruption of Model Membranes by ENPs. Since the conditions which favor ENP

247  

adsorption on model membranes may not necessarily favor ENP penetration, experiments should,

248  

ideally, be designed to study both processes independently. One of the most common methods

249  

currently employed to test the propensity for ENPs to disrupt model membranes is the dye-

250  

leakage assay.21, 44 This method involves the preparation of an aqueous mixture comprising lipid

251  

vesicles (or liposomes) that encapsulate a fluorescent dye and ENPs that are being evaluated

252  

(Figure 2A). The fluorescence intensity is monitored and an increase in the measurements would

253  

indicate a leakage of dyes from within the vesicles due to the disruption of the bilayers by the

254  

ENPs. Moghadam et al.21 employed this technique to demonstrate that positively charged ENPs

255  

have a higher propensity than negatively charged ENPs to disrupt DOPC vesicles. The same

256  

technique was also employed by Shi et al.44 to show that SWNTs stabilized with sodium

257  

dodecylbenzenesulfonate (SDBS) do not disrupt egg L-α-phosphatidylcholine (egg-PC) vesicles.

258  

Phospholipid vesicles can be used with the QCM-D to evaluate the propensity for an ENP

259  

to disrupt model membranes. Yi and Chen22 developed an assay that involved the assembly of a

260  

supported vesicular layer (SVL, Figure 2D) on a QCM-D crystal surface and the subsequent

261  

exposure of the SVL to the ENPs of interest under flow-through conditions. If the vesicles were 9   ACS Paragon Plus Environment

Environmental Science & Technology

262  

to be disrupted by the ENPs, the solution that was initially encapsulated in the vesicles will be

263  

released into the bulk solution, hence resulting in a decrease in the deposited mass on the crystal

264  

surface (which will be reflected by an increase in the crystal frequency response). Similar to the

265  

findings of Shi et al.,44 Yi and Chen22 showed that MWNTs do not cause any noticeable damage

266  

to DOPC vesicles. One advantage of the QCM-D assay is that it does not require the use of a

267  

fluorescent dye. Furthermore, it enables the monitoring of changes in the viscoelastic properties

268  

of the SVL upon contact with the ENPs. The viscoelastic parameters, namely, viscosity and

269  

shear modulus, can be derived by using the Voigt-based model45 to fit the frequency and

270  

dissipation responses collected by the QCM-D.

271  

Electrophysiological measurement is an emerging approach to detect the disruption of

272  

model membranes by ENPs. This technique involves the measurements of the electrical/ionic

273  

conductance across an unsupported, planar lipid bilayer that is formed across an orifice (Figure

274  

2E).38 The disruption or penetration of a lipid bilayer by ENPs can result in the formation of

275  

pores and hence, lead to an increase in the electrical conductance across the bilayer. The

276  

advantage of electrophysiological measurements is that the measurements are highly sensitive to

277  

changes in electrical conductance and thus, can detect minute perturbations of the membrane.

278  

The measurements, however, require an elaborate experimental setup, which is not yet

279  

commercially available. Corredor et al.27 conducted electrophysiological measurements which

280  

showed that MWNTs have the ability to disrupt a planar DOPC bilayer resulting in an increase in

281  

the current through the bilayer. The differing results regarding the propensity of CNTs to disrupt

282  

model cell membranes22, 27, 44 are not surprising considering the preparation methods for the

283  

CNTs were different in the experiments performed by the various groups. Instead, they highlight

284  

the pressing need for more in-depth studies of the role of surface chemistry of CNTs (and ENPs

285  

in general) on the propensity of the nanomaterials to disrupt cell membranes. Furthermore, the

286  

use of a combination of the techniques discussed above to investigate a single ENP–membrane

287  

system will allow for the validation of experimental data and also the elucidation of the

288  

mechanisms for membrane disruption by ENPs.

289  

Visualization of ENP–Membrane Interactions.

Direct visualization of ENP–

290  

membrane binding and membrane disruption can be achieved through cryo-TEM. An advantage

291  

of this technique is that it can be performed on unsupported spherical vesicles (Figure 2A),

292  

which can exhibit elastic deformations and rupture. Chen and Bothun31 have recently 10   ACS Paragon Plus Environment

Page 10 of 22

Page 11 of 22

Environmental Science & Technology

293  

demonstrated the effect of ENP size (anionic iron oxide nanoparticles) on the binding and

294  

disruption of oppositely charged membranes. Small ENPs (16 nm hydrodynamic diameter) were

295  

shown to bind to membrane vesicles without causing the rupture of vesicles. In contrast, larger

296  

ENPs (30 nm hydrodynamic diameter) led to local changes in membrane curvature and

297  

significant vesicle rupture. Differences in the nature of the membrane interaction for the two

298  

ENPs were based upon the balance between membrane bending elasticity (2kb), which reflects

299  

how the membrane resists deformation, and the adhesion energy (Eadh, per area). This balance

300  

can be used to determine the critical nanoparticle radius, R2 = 2kb/Eadh, that causes deformation.46

301  

The total adhesion energy provided by larger ENPs (i.e. > 20 nm diameter), EadhR2, was enough

302  

to overcome the energy penalty for bending the membrane around the nanoparticle, which

303  

effectively led to the extraction of lipids from the membranes. Le Bihan et al.20 utilized this same

304  

concept of critical nanoparticle radius to analyze ENP invagination into membrane vesicles

305  

driven solely by membrane adhesion. Finally, cryo-TEM has also been used to examine the

306  

interactions between CNTs and vesicles, where it was shown that zwitterionic vesicles bound

307  

readily to MWNTs.22, 44 In this case the vesicles remained intact, indicating that there was

308  

limited contact area and/or weak ENP-membrane adhesion. An important point to note is that

309  

since the vitrified film has to be sufficiently thin (ca. 200–500 nm) in order for electrons to pass

310  

through, the size of the vesicles that can be observed is limited by the sample thickness.

311   312  

Challenges, opportunities, and recommendations

313  

Establishing links between model and ‘real’ membranes. The goal of any model

314  

membrane study is to mimic cellular membrane behavior and to provide fundamental biophysical

315  

insight into membrane-related phenomena that can be used to explain physiological responses. In

316  

nanotechnology environmental health and safety (EHS) research, such information would help

317  

identify the specific role of nanoparticle–membrane interactions in cytotoxicity, and would

318  

provide criteria for designing or selecting nanomaterials with that exhibit minimal (or no)

319  

membrane disruption. A first step towards improving our understanding of nanoparticle–

320  

membrane interactions requires transitioning from simple homogenous model membranes to

321  

more life-like (and more complex) heterogeneous membranes. This includes (i) using

322  

multicomponent membranes composed of biologically-relevant lipids, (ii) using membranes

323  

reconstituted from cell membrane extracts, and, specific to bacterial membranes, (iii) using 11   ACS Paragon Plus Environment

Environmental Science & Technology

324  

membranes that incorporate polysaccharide or peptidoglycan surface coatings and bacterial lipids

325  

(Figure 1A, B). Multicomponent membranes could include mixtures of zwitterionic and anionic

326  

lipids, which would provide variable surface charge density, or lipids with different tail lengths

327  

or degree of saturation, which would provide membranes that contain co-existing phases or

328  

domains (e.g., sterol-rich domains). These approaches would allow one to determine if

329  

nanoparticles preferentially interact with specific lipids or lipid structures. Reconstituted

330  

membranes could be an ideal platform to study such interactions; however, the studies would

331  

need to include analyses of lipid composition and lipid organization to be complete.

332  

Accounting for environmental transformations. The second transition needed is from

333  

nanoparticles dispersed in “clean” media to nanoparticles dispersed in environmentally relevant

334  

media containing proteins, ions, and/or organic molecules (e.g., natural organic matter).2 The

335  

presence of these components is known to lead to the formation of a nanoparticle “corona” that

336  

alters nanoparticle surface chemistry and aggregation state (Figure 4). Recent research indicates

337  

that protein adsorption onto nanoparticles (carboxylated polystyrene NPs in serum) reduces cell

338  

adhesion and subsequent nanoparticle uptake.13 Adding to the complexity is the observation that

339  

ENPs with mixed hydrophilic/hydrophobic surface coatings can rearrange these coatings to

340  

maximize hydrophobic matching with the membrane and achieve penetration.47 Therefore, it is

341  

not just the composition of the coating in biological or natural environments, but the dynamic

342  

nature of this coating to restructure and adapt. FIGURE 4

343   344  

Complexities associated with predicting ENP–membrane interactions. Over the last

345  

decade, there have been many insightful studies conducted to determine the effects of ENP–

346  

membrane binding on membrane disruption. Fundamental information on ENP–membrane

347  

interactions has been gained primarily using model membranes. However, common platforms or

348  

techniques to determine ENP–membrane interactions have not been developed, and predictive

349  

frameworks for determining the membrane activity of an ENP are lacking. Addressing these

350  

issues is no small task given the inherent complexity and diversity of both biological membranes

351  

and ENPs.

352  

Attachment (adhesion) represents the first step for nanoparticle–membrane interactions.

353  

Electrostatic interactions have proven to be critical for ENP adhesion to and disruption of

354  

model35 and intact12, 48 cellular membranes. Classic approaches using DLVO or extended DLVO 12   ACS Paragon Plus Environment

Page 12 of 22

Page 13 of 22

Environmental Science & Technology

355  

theories seem to capture the attachment process and can be adjusted based on nanoparticle size

356  

and surface functionality. However, these theories do not take into account membrane surface

357  

tension or membrane bending rigidity, which influence the adhesion process and the extent of

358  

membrane deformation (after adhesion).49 These factors could be accounted for by varying

359  

membrane lipid composition, which would change the mechanical properties of the membrane.

360  

Adhesion and membrane deformation will drive lipid reorganization and govern the capacity for

361  

transient void or pore formation. Lipid reorganization is also dependent upon lipid composition,

362  

namely, the presence of charged lipids, sterols, and lipids with varying degrees of tail saturation.

363  

Collectively, these processes occur due to changes in intermolecular interactions between

364  

membrane components that arise due to nanoparticle adhesion. High adhesion energies due to

365  

strong ENP–membrane surface interactions alter these intermolecular interactions and can lead

366  

to membrane disruption. Connecting ENP adhesion to changes in intermolecular membrane

367  

interactions, which may be assessed through calorimetry, spectroscopy, or AFM, is another

368  

critical step in predicting ENP membrane activity.

369  

At this point, studies are still underway to determine the possible (and dominant)

370  

mechanisms of ENP–membrane interactions and the extent to which they contribute to

371  

nanotoxicology. Model membrane studies have provided critical insight into these mechanisms

372  

and have demonstrated the importance of adhesion. Relatively few studies, however, have

373  

connected adhesion with changes in membrane function. This connection and the transition to

374  

heterogeneous membranes and environmentally-transformed ENPs should represent the next

375  

stage in ENP–membrane interactions studies.

376   377   378  

Biography Kai Loon Chen is an Assistant Professor in the Department of Geography and

379  

Environmental Engineering at the Johns Hopkins University.

His research focuses on

380  

environmental applications and implications of nanotechnology, interactions between engineered

381  

nanomaterials and biological membranes, and membrane filtration processes for water treatment

382  

and purification.

383  

Geoffrey D. Bothun is an Associate Professor in the Department of Chemical

384  

Engineering at the University of Rhode Island. His research focuses on the role of nanoparticle-

385  

membrane interactions in nanotoxicology and nanomedicine, biological membrane adaptation to 13   ACS Paragon Plus Environment

Environmental Science & Technology

386  

membrane-active solutes and surfaces, and nanoparticle-based environmental remediation

387  

technologies.

388   389  

Acknowledgments

390  

This material is based upon work supported by Semiconductor Research Corporation

391  

(award 425-MC-2001, project 425.041) and the National Science Foundation under Grant No.

392  

CBET-1055652.

393  

14   ACS Paragon Plus Environment

Page 14 of 22

Page 15 of 22

Environmental Science & Technology

394  

References

395   396   397   398   399   400   401   402   403   404   405   406   407   408   409   410   411   412   413   414   415   416   417   418   419   420   421   422   423   424   425   426   427   428   429   430   431   432   433   434   435   436   437   438   439  

1. Nowack, B.; Bucheli, T. D., Occurrence, behavior and effects of nanoparticles in the environment. Environ Pollut 2007, 150, (1), 5-22. 2. Lowry, G. V.; Gregory, K. B.; Apte, S. C.; Lead, J. R., Transformations of Nanomaterials in the Environment. Environ Sci Technol 2012, 46, (13), 6893-6899. 3. Petosa, A. R.; Jaisi, D. P.; Quevedo, I. R.; Elimelech, M.; Tufenkji, N., Aggregation and Deposition of Engineered Nanomaterials in Aquatic Environments: Role of Physicochemical Interactions. Environ Sci Technol 2010, 44, (17), 6532-6549. 4. Klaine, S. J.; Alvarez, P. J. J.; Batley, G. E.; Fernandes, T. F.; Handy, R. D.; Lyon, D. Y.; Mahendra, S.; McLaughlin, M. J.; Lead, J. R., Nanomaterials in the environment: Behavior, fate, bioavailability, and effects. Environmental Toxicology and Chemistry 2008, 27, (9), 1825-1851. 5. Kang, S.; Pinault, M.; Pfefferle, L. D.; Elimelech, M., Single-walled carbon nanotubes exhibit strong antimicrobial activity. Langmuir 2007, 23, (17), 8670-8673. 6. Kang, S.; Herzberg, M.; Rodrigues, D. F.; Elimelech, M., Antibacterial effects of carbon nanotubes: Size does matter. Langmuir 2008, 24, (13), 6409-6413. 7. Vecitis, C. D.; Zodrow, K. R.; Kang, S.; Elimelech, M., Electronic-Structure-Dependent Bacterial Cytotoxicity of Single-Walled Carbon Nanotubes. Acs Nano 2010, 4, (9), 5471-5479. 8. Liu, S. B.; Wei, L.; Hao, L.; Fang, N.; Chang, M. W.; Xu, R.; Yang, Y. H.; Chen, Y., Sharper and Faster "Nano Darts" Kill More Bacteria: A Study of Antibacterial Activity of Individually Dispersed Pristine Single-Walled Carbon Nanotube. Acs Nano 2009, 3, (12), 38913902. 9. Liu, S. B.; Zeng, T. H.; Hofmann, M.; Burcombe, E.; Wei, J.; Jiang, R. R.; Kong, J.; Chen, Y., Antibacterial Activity of Graphite, Graphite Oxide, Graphene Oxide, and Reduced Graphene Oxide: Membrane and Oxidative Stress. Acs Nano 2011, 5, (9), 6971-6980. 10. Xiu, Z. M.; Zhang, Q. B.; Puppala, H. L.; Colvin, V. L.; Alvarez, P. J. J., Negligible Particle-Specific Antibacterial Activity of Silver Nanoparticles. Nano Lett 2012, 12, (8), 42714275. 11. Sondi, I.; Salopek-Sondi, B., Silver nanoparticles as antimicrobial agent: a case study on E-coli as a model for Gram-negative bacteria. J Colloid Interf Sci 2004, 275, (1), 177-182. 12. Brayner, R.; Ferrari-Iliou, R.; Brivois, N.; Djediat, S.; Benedetti, M. F.; Fievet, F., Toxicological impact studies based on Escherichia coli bacteria in ultrafine ZnO nanoparticles colloidal medium. Nano Lett 2006, 6, (4), 866-870. 13. Lesniak, A.; Salvati, A.; Santos-Martinez, M. J.; Radomski, M. W.; Dawson, K. A.; Aberg, C., Nanoparticle Adhesion to the Cell Membrane and Its Effect on Nanoparticle Uptake Efficiency. J Am Chem Soc 2013, 135, (4), 1438-1444. 14. Peetla, C.; Labhasetwar, V., Effect of Molecular Structure of Cationic Surfactants on Biophysical Interactions of Surfactant-Modified Nanoparticles with a Model Membrane and Cellular Uptake. Langmuir 2009, 25, (4), 2369-2377. 15. Leroueil, P. R.; Hong, S.; Mecke, A.; Baker, J. R., Jr.; Orr, B. G.; Banaszak Holl, M. M., Nanoparticle interaction with biological membranes: does nanotechnology present a Janus face? Acc Chem Res 2007, 40, (5), 335-42. 16. Leroueil, P. R.; Berry, S. A.; Duthie, K.; Han, G.; Rotello, V. M.; McNerny, D. Q.; Baker, J. R.; Orr, B. G.; Holl, M. M. B., Wide varieties of cationic nanoparticles induce defects in supported lipid bilayers. Nano Letters 2008, 8, (2), 420-424. 17. Yuan, H.; Li, J.; Bao, G.; Zhang, S., Variable nanoparticle-cell adhesion strength regulates cellular uptake. Phys Rev Lett 2010, 105, (13), 138101. 15   ACS Paragon Plus Environment

Environmental Science & Technology

440   441   442   443   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  

18. Nel, A. E.; Madler, L.; Velegol, D.; Xia, T.; Hoek, E. M.; Somasundaran, P.; Klaessig, F.; Castranova, V.; Thompson, M., Understanding biophysicochemical interactions at the nano-bio interface. Nat Mater 2009, 8, (7), 543-57. 19. Beney, L.; Gervais, P., Influence of the fluidity of the membrane on the response of microorganisms to environmental stresses. Appl. Microbiol. Biotechnol. 2001, 57, 34-42. 20. Le Bihan, O.; Bonnafous, P.; Marak, L.; Bickel, T.; Trepout, S.; Mornet, S.; De Haas, F.; Talbot, H.; Taveau, J. C.; Lambert, O., Cryo-electron tomography of nanoparticle transmigration into liposome. J Struct Biol 2009, 168, (3), 419-425. 21. Moghadam, B. Y.; Hou, W. C.; Corredor, C.; Westerhoff, P.; Posner, J. D., Role of Nanoparticle Surface Functionality in the Disruption of Model Cell Membranes. Langmuir 2012, 28, (47), 16318-16326. 22. Yi, P.; Chen, K. L., Interaction of Multiwalled Carbon Nanotubes with Supported Lipid Bilayers and Vesicles as Model Biological Membranes. Environ Sci Technol 2013, 47, 57115719. 23. Hou, W. C.; Moghadam, B. Y.; Corredor, C.; Westerhoff, P.; Posner, J. D., Distribution of Functionalized Gold Nanoparticles between Water and Lipid Bilayers as Model Cell Membranes. Environ Sci Technol 2012, 46, (3), 1869-1876. 24. Hou, W. C.; Moghadam, B. Y.; Westerhoff, P.; Posner, J. D., Distribution of Fullerene Nanomaterials between Water and Model Biological Membranes. Langmuir 2011, 27, (19), 11899-11905. 25. Richter, R.; Mukhopadhyay, A.; Brisson, A., Pathways of lipid vesicle deposition on solid surfaces: A combined QCM-D and AFM study. Biophys J 2003, 85, (5), 3035-3047. 26. Richter, R. P.; Him, J. L. K.; Tessier, B.; Tessier, C.; Brisson, A. R., On the kinetics of adsorption and two-dimensional self-assembly of annexin A5 on supported lipid bilayers. Biophys J 2005, 89, (5), 3372-3385. 27. Corredor, C.; Hou, W. C.; Klein, S. A.; Moghadam, B. Y.; Goryll, M.; Doudrick, K.; Westerhoff, P.; Posner, J. D., Disruption of model cell membranes by carbon nanotubes. Carbon 2013, 60, 67-75. 28. Peetla, C.; Labhasetwar, V., Biophysical characterization of nanoparticle-endothelial model cell membrane interactions. Mol Pharm 2008, 5, (3), 418-429. 29. Cevc, G.; Marsh, D., Phospholipid bilayers: Physical principles and models. John Wiley and Sons, Inc.: New York, 1987. 30. Wang, B.; Zhang, L. F.; Bae, S. C.; Granick, S., Nanoparticle-induced surface reconstruction of phospholipid membranes. Proc. Nat. Acad. Sci. USA 2008, 105, (47), 1817118175. 31. Chen, Y.; Bothun, G. D., Cationic Gel-Phase Liposomes with "Decorated" Anionic SPIO Nanoparticles: Morphology, Colloidal, and Bilayer Properties. Langmuir 2011, 27, (14), 86458652. 32. Li, S.; Malmstadt, N., Deformation and poration of lipid bilayer membranes by cationic nanoparticles. Soft Matter 2013, 9, (20), 4969-4976. 33. Leroueil, P. R.; Hong, S. Y.; Mecke, A.; Baker, J. R.; Orr, B. G.; Holl, M. M. B., Nanoparticle interaction with biological membranes: Does nanotechnology present a janus face? Accounts of Chemical Research 2007, 40, (5), 335-342. 34. Israelachvili, J., Intermolecular and Surface Forces. Academic Press: London, England, 1991.

16   ACS Paragon Plus Environment

Page 16 of 22

Page 17 of 22

485   486   487   488   489   490   491   492   493   494   495   496   497   498   499   500   501   502   503   504   505   506   507   508   509   510   511   512   513   514   515   516   517   518   519   520   521   522   523   524   525  

Environmental Science & Technology

35. Xiao, X. Y.; Montano, G. A.; Edwards, T. L.; Allen, A.; Achyuthan, K. E.; Polsky, R.; Wheeler, D. R.; Brozik, S. M., Surface Charge Dependent Nanoparticle Disruption and Deposition of Lipid Bilayer Assemblies. Langmuir 2012, 28, (50), 17396-17403. 36. Feris, K.; Otto, C.; Tinker, J.; Wingett, D.; Punnoose, A.; Thurber, A.; Kongara, M.; Sabetian, M.; Quinn, B.; Hanna, C.; Pink, D., Electrostatic Interactions Affect NanoparticleMediated Toxicity to Gram-Negative Bacterium Pseudomonas aeruginosa PAO1. Langmuir 2010, 26, (6), 4429-4436. 37. Marra, J.; Israelachvili, J., Direct Measurements of Forces between Phosphatidylcholine and Phosphatidylethanolamine Bilayers in Aqueous-Electrolyte Solutions. Biochemistry-Us 1985, 24, (17), 4608-4618. 38. Negoda, A.; Liu, Y.; Hou, W.-C.; Corredor, C.; Moghadam, B. Y.; Musolff, C.; Li, L.; Walker, W.; Westerhoff, P.; Mason, A. J.; Duxbury, P.; Posner, J. D.; Worden, R. M., Engineered nanomaterial interactions with bilayer lipid membranes: screening platforms to assess nanoparticle toxicity. Int. J. Biomedical Nanoscience and Nanotechnology 2013, 3, 52-83. 39. Roiter, Y.; Ornatska, M.; Rammohan, A. R.; Balakrishnan, J.; Heine, D. R.; Minko, S., Interaction of nanoparticles with lipid membrane. Nano Letters 2008, 8, (3), 941-944. 40. Spurlin, T. A.; Gewirth, A. A., Effect of C-60 on solid supported lipid bilayers. Nano Lett 2007, 7, (2), 531-535. 41. Rusciano, G.; De Luca, A. C.; Pesce, G.; Sasso, A., On the interaction of nano-sized organic carbon particles with model lipid membranes. Carbon 2009, 47, (13), 2950-2957. 42. Keller, C. A.; Kasemo, B., Surface specific kinetics of lipid vesicle adsorption measured with a quartz crystal microbalance. Biophys J 1998, 75, (3), 1397-1402. 43. Zhang, X. F.; Yang, S. H., Nonspecific Adsorption of Charged Quantum Dots on Supported Zwitterionic Lipid Bilayers: Real-Time Monitoring by Quartz Crystal Microbalance with Dissipation. Langmuir 2011, 27, (6), 2528-2535. 44. Shi, L.; Shi, D. C.; Nollert, M. U.; Resasco, D. E.; Striolo, A., Single-Walled Carbon Nanotubes Do Not Pierce Aqueous Phospholipid Bilayers at Low Salt Concentration. J Phys Chem B 2013, 117, (22), 6749-6758. 45. Voinova, M. V.; Rodahl, M.; Jonson, M.; Kasemo, B., Viscoelastic acoustic response of layered polymer films at fluid-solid interfaces: Continuum mechanics approach. Phys Scripta 1999, 59, (5), 391-396. 46. Lipowsky, R.; Dobereiner, H. G., Vesicles in contact with nanoparticles and colloids. Europhys Lett 1998, 43, (2), 219-225. 47. Van Lehn, R. C.; Alexander-Katz, A., Penetration of lipid bilayers by nanoparticles with environmentally-responsive surfaces: simulations and theory. Soft Matter 2011, 7, (24), 1139211404. 48. Zhang, L. L.; Jiang, Y. H.; Ding, Y. L.; Povey, M.; York, D., Investigation into the antibacterial behaviour of suspensions of ZnO nanoparticles (ZnO nanofluids). J Nanopart Res 2007, 9, (3), 479-489. 49. Yi, X.; Shi, X. H.; Gao, H. J., Cellular Uptake of Elastic Nanoparticles. Phys Rev Lett 2011, 107, (9).

526   527  

 

528   17   ACS Paragon Plus Environment

Environmental Science & Technology

529  

Table 1. General Properties of Model Membrane Configurations. Membrane

Sample

Lipid

configuration

Common experimental techniques

concentration

Spherical vesicles

Dispersed in aqueous

High and

Calorimetry; spectroscopy; electronb

(Figure 2A, D)

solutions or

variable or low

and optical microscopy;

deposited on supports

and defined by

microbalanced

support Supported planar

Deposited at defined

Low and

Spectroscopy; atomic force, electron,c

bilayers (Figure 2B)

interfaces

defined by

and optical microscopy; microbalance

Interfacial

support or

Spectroscopy; atomic force,d Brewster

monolayers (Figure

interfacial areaa

angle, electron,d and optical

2C)

microscopy; microbalance

Unsupported planar

Deposited across

Low and

bilayers (Figure 2E)

orifices

defined by orifice area

530  

a

531  

manipulated by compression.

532  

b

533  

c

534  

d

Spectroscopy; atomic force microscopy; electrophysiology a

Low relative to dispersed vesicles. Monolayer concentrations determined by interfacial area and Requires cryogenic sample preparation.

Limited to scanning electron microscopy. Requires deposition onto a planar solid substrate.

535  

18   ACS Paragon Plus Environment

Page 18 of 22

Page 19 of 22

Environmental Science & Technology

536   (A) Gram-positive bacteria

(B) Gram-negative bacteria outer lipid membrane containing anionic lipopolysaccharides

peptidoglycan layer anchored by anionic lipoteichoic acids

peptidoglycan layer anchored by lipoproteins

lipid membrane inner lipid membrane

(C) Plant cell (D) Mammalian cell

pectin cellulose microfibrils cross-linked with hemicellulose

carbohydrates

lipid membrane

lipid membrane

537   538   539  

FIGURE 1. Membrane structure, excluding integral and peripheral proteins, of (A) gram-

540  

positive bacteria, (B) gram-negative bacteria, (C) plant cells, and (D) mammalian cells.

541  

19   ACS Paragon Plus Environment

Environmental Science & Technology

Page 20 of 22

542   (A) Spherical vesicles

(B) Supported planar bilayers water

(D) Vesicles on planar support water water water

solid support water

543  

water

(C) Interfacial monolayers air or solvent

water water

solid support (E) Unsupported planar bilayers water

water

water

544   545  

FIGURE 2. Model membrane configurations: (A) spherical vesicles (dispersed), (B) supported

546  

planar bilayers, (C) interfacial monolayers, (D) vesicles on planar supports, and (E) unsupported

547  

planar bilayers (i.e., spanning an orifice). Spherical vesicles (A) can consist of freestanding

548  

bilayers, bilayers supported on microparticles or nanoparticles, or intact vesicles deposited on

549  

supports (D).

550  

20   ACS Paragon Plus Environment

Page 21 of 22

Environmental Science & Technology

551   (A) Nanoparticle adhesion (binding) at membrane/water interface

-

+

-

+

-

+

+

(C) Passive (adhesive) membrane translocation

+ +

- - -

domain formation

(D) Adhesive lipid extraction

(B) Membrane restructuring and leakage

domains

pores

552   553  

FIGURE 3. Effect of ENP–membrane interactions on membrane organization and translocation.

554  

The red regions denote the membrane segments that are locally affected by nanoparticle binding.

555  

21   ACS Paragon Plus Environment

Environmental Science & Technology

Page 22 of 22

556  

+

+ Corona formation

-

-

+

Aggregation

+ +

+

proteins

- + 557  

ions organic molecules

+

+

558  

FIGURE 4. Environmental transformation leads to the formation of an adsorbed nanoparticle

559  

corona, which alters the colloidal stability of the nanoparticles. The composition of the corona

560  

depends upon the ‘native’ nanoparticle surface coating and the water chemistry.

22   ACS Paragon Plus Environment