Rotation and Retention Dynamics of Rod-Shaped ... - ACS Publications

Mar 29, 2019 - Colloid surface charge heterogeneity was incorporated into a three-dimensional trajectory model, which simulated particle translation a...
0 downloads 0 Views 920KB Size
Subscriber access provided by ALBRIGHT COLLEGE

Interfaces: Adsorption, Reactions, Films, Forces, Measurement Techniques, Charge Transfer, Electrochemistry, Electrocatalysis, Energy Production and Storage

Rotation and Retention Dynamics of Rod-Shaped Colloids with Surface Charge Heterogeneity in Sphere-in-Cell Porous Media Model ke li, and Huilian Ma Langmuir, Just Accepted Manuscript • DOI: 10.1021/acs.langmuir.9b00748 • Publication Date (Web): 29 Mar 2019 Downloaded from http://pubs.acs.org on March 29, 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 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

1

Rotation and Retention Dynamics of Rod-Shaped Colloids with Surface

2

Charge Heterogeneity in Sphere-in-Cell Porous Media Model

3 4

Ke Li, Huilian Ma

5

Department of Geology and Geophysics

6

University of Utah, Salt Lake City, UT 84112

7

Abstract

8 9

Colloid surface charge heterogeneity was incorporated into a three-dimensional trajectory model, which

10

simulated particle translation and rotation via a force/torque analysis, to study the transport and retention

11

dynamics of rod-shaped colloids over a wide size range in porous media under unfavorable conditions

12

(energy barriers to deposition exist). Our previous study1 for rod transport under favorable conditions

13

(lacking energy barriers) demonstrated that particle rotation due to the coupled effect of flow

14

hydrodynamics and Brownian rotation governed rod transport and retention. In this work, we showed that

15

the shape of a colloid affected both transport process and colloid-collector interactions, but shape alone

16

could not make rods to overcome energy barriers of over tens of kT for attachment under unfavorable

17

conditions. The location of colloid surface heterogeneity did not affect transport, but predominantly

18

affected colloid-surface interactions by influencing the likelihood of heterogeneity patches facing the

19

collector due to particle rotation. For surface heterogeneity located on the end(s) of a colloid, rods

20

displayed enhanced retention compared with spheres; for surface heterogeneity located on the middle

21

band, rods showed less retention compared with spheres. It was more effective to arrest a traveling rod

22

when surface heterogeneity was located on the end relative to the side, because the tumbling motion

23

greatly increased the likelihood of the end to intercept collector surfaces, and also because a rod would 

Corresponding author. E-mail: huilian. [email protected]; Tel: (801)585-5976; Fax: (801)581-7065.

1 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

24

experience less repulsion with an end-on orientation relative to the collector surface compared to a side-

25

on orientation due to curvature effect. The influences of particle aspect ratio on retention strongly

26

depended upon the location of colloid surface heterogeneity. Our findings demonstrated that rods had

27

distinct rotation and retention behaviors from spheres under conditions typically encountered in the

28

environment; thus, particle rotation should be considered when studying the transport process of non-

29

spherical colloids or spherical particles with inhomogeneous surface properties.

30 31

Introduction

32

Groundwater is one of the most important resources on earth. An enormous amount of groundwater is

33

needed for irrigation to grow food. It also supplies drinking water for more than half of the U.S.

34

population.2 Studying transport and deposition of colloids in saturated porous media is essential in

35

protecting and utilizing groundwater, for example, in the process to track contaminant fate and to produce

36

drinking water.3-4 Natural colloids in groundwater are often non-spherical in shape (e.g., rod-shaped

37

bacteria,5-7 plate-like clay particles8-9) and possess varied degrees of surface heterogeneities. General

38

filtration theory and many researches concerning colloid transport usually simplify natural colloids as

39

spheres with effective radii and evenly charged surfaces.10-13 Mounting evidences14-22 have demonstrated

40

that these approaches are inadequate in describing colloid transport and retention behaviors.

41 42

For the past several decades, many studies have contributed to understanding the influences of particle

43

shape on particle transport and adhesion, mainly in biomedical and aerodynamic fields.23-33 For instance,

44

rod-shaped particles within micron size range were investigated for targeted drug delivery due to their

45

better propensity to migrate towards blood vessel walls compared to spheres of equivalent volume.25-30, 34

46

Zhang et al. reported that particle aspect ratio played an important role in leading to deposition of rod-

2 ACS Paragon Plus Environment

Page 2 of 45

Page 3 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

47

shaped aerosols onto channel walls.35 Among such studies, some considered the influences of fluid

48

hydrodynamic interactions induced by non-spherical shape,25-28, 31 whereas others investigated the

49

translational and rotational dynamics of non-spherical particles in the Brownian regime.29, 36-38 However,

50

very few studies took into account of the combined influences of flow hydrodynamics and Brownian

51

rotation for colloids of non-spherical shape. Consequently, a comprehensive understanding of the effects

52

of particle shape on transport and deposition over a wide range of particle sizes (e.g., from nanometers to

53

tens of microns) is still lacking. In addition, many studies were carried out in parallel plate flow

54

chambers.29-32 Often times, those studies did not consider colloidal interactions,23, 25-34 which might be

55

justified by high ionic strength conditions associated with blood or lack of surface charges in air flow

56

environments. As a result, it is questionable that findings from such researches can directly translate to

57

colloidal filtration field involving porous media, where both colloids and collecting medium surfaces are

58

usually negatively charged. As of now, in colloid filtration field, only a few experimental studies15, 39-40

59

investigated the influences of particle shape on transport and retention. These experimental studies

60

demonstrated that rod-shaped particles generally exhibited higher retention compared to spheres of equal

61

volume.15, 39-40 However, the underlying mechanisms for the enhanced retention of rod-shaped particles

62

were not well understood. Further, inconsistent findings regarding the role of rod aspect ratio on retention

63

had been reported.15, 39-40 For example, Wang et al.39 observed that the slender, longer rod-shaped

64

bacterial cells attached to the micro-pore filter walls less than the corresponding shorter ones; whereas

65

Weiss et al.40 and Salerno et al.15 reported the opposite retention results. Also, the range of rod-shaped

66

particle size and aspect ratio examined from those experiments was limited.15, 39-40 In a very recent study,

67

we presented a three-dimensional particle transport model to investigate the effect of particle shape on

68

transport and compared the retention behaviors of rods versus spheres of equivalent volume in porous

69

media under favorable conditions (i.e., no energy barrier to retention).1 While particle rotation may be

70

irrelevant to spheres of uniform properties, it was critical for the transport and retention of rod-shaped

71

colloids.1 In shear flow, inertial rod-shaped particles underwent periodical tumbling motions when

72

external forces and Brownian motions were not considered (see Video S1 rendered from our simulated 3 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

73

trajectory in the parallel plate geometry in the Supporting Information), as predicted by Jeffery in 1922.41

74

The positions of rods also oscillated periodically in response to flow hydrodynamics,1, 42-45 which caused

75

rods to drift away from flow streamlines (Video S1 in the Supporting Information). Additionally, rotation

76

due to diffusion changed rod orientation constantly and greatly altered the trajectories of rod-shaped

77

particles under hydrodynamic effects. We also found that the rotation dynamics of rods varied with

78

particle size as well as fluid flow velocity. Our previous study on rod rotation and transport under

79

favorable conditions1 laid a foundation to continue the exploration of rod retention under unfavorable

80

conditions (i.e., repulsive energy barriers exist).

81 82

Predicting the attachment of colloids onto collector surfaces in the presence of an energy barrier remains

83

as a challenge. The conventional mean-field zeta-potential approach to represent repulsive colloid-surface

84

interactions yielded zero direct colloid attachment in porous media under unfavorable conditions;46-47

85

whereas significant colloid attachments have been observed from column experiments packed with clean

86

glass beads or quartz sand.21, 48-49 Many researchers incorporated the influences of surface heterogeneity,

87

such as roughness,50-55 chemical heterogeneities,21-22, 56-59 onto collector surfaces and greatly improved the

88

retention prediction of colloids (especially for spheres with uniform surface properties) relative to

89

experimental observations under unfavorable conditions. However, very few studies have investigated the

90

role of colloid surface heterogeneity on transport and retention.60-64 Until recently, Shave et al.60

91

experimented on nanoscale functionalized microspheres flowing in a shear flow and found that the

92

capture rate of particles onto chamber walls depended strongly on particle rotation, which controlled the

93

likelihood of a functionalized group on particle surface facing the collector surface.

94 95

For non-spherical colloids (e.g., rod-shaped ones), rotation becomes important not only in transporting

96

particles to the vicinity of a collector surface,1 but also in properly orienting them to regulate colloid-

4 ACS Paragon Plus Environment

Page 4 of 45

Page 5 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

97

surface interaction.6, 65 Since colloidal interactions scale with particle size, a rod may experience less

98

repulsion with an end-on orientation toward the collector surface due to the small radius of curvature of

99

its end. For instance, many rod-shaped bacteria have adhesins (ligands or pili) attached at one end of their

100

bodies;66-70 take Thiothrix nivea as an example, the fimbriated end of Thiothrix nivea usually initiated

101

contact with surrounding surfaces.67 Similarly, E. coli bacteria (rod-shaped) were observed to adhere to

102

polystyrene particles by one end over 90% of the times.66 The tumbling motion of a rod-shaped particle

103

may render its end to intercept with surrounding surfaces more frequently than the middle body sections.1

104

However, under completely unfavorable conditions, our simulations indicated that shape alone (even with

105

an end-on orientation for a rod) could not make the particle to overcome energy barriers above certain

106

height (e.g, over tens of kT) for attachment based on mean-field surface potential approach. This called

107

for the need to consider colloidal surface heterogeneity (here, we will incorporate surface charge

108

heterogeneity as a first attempt) to better understand bacterial adhesion. Because of particle rotation, the

109

location of charge heterogeneity over colloid surface becomes important.

110 111

In this work, we will extend our particle trajectory model1 previously developed for rod-shaped (prolate

112

spheroid) particle transport to incorporate colloid surface charge heterogeneity. This model predicts both

113

the translational and rotational motions of an ellipsoidal particle based on an analysis of all the forces and

114

torques acting on it. From this trajectory model analysis, the retention probability, attachment distribution

115

and orientation of heterogeneous rod particles will be derived and compared to those of heterogeneous

116

spheres of equivalent volume. Simulations will be performed primarily in the well-known Happel sphere-

117

in-cell porous media model (Figure 1). No heterogeneity will be considered on the Happel model

118

collector surfaces. To clearly demonstrate the orientation of retained colloids onto the collector surfaces,

119

we will also present corresponding simulation results performed in the parallel plate geometry, due to its

120

flat collector surfaces and simple flow field, under similar conditions as those in the Happel model.

121

However, we stress that in this work, it is not our focus to quantitatively compare colloidal retention from 5 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 45

122

these two geometrical models, despite that this study examines colloidal transport and retention at the

123

near-surface laminar flow regimes in both models. The specific questions we will address include: i) what

124

is the minimal colloid surface charge heterogeneity coverage required for particle capture? ii) how does

125

the location of charge heterogeneity over colloid surface affect the retention of rod-shaped particles? iii)

126

How does particle shape (i.e., aspect ratio) affect transport and retention of colloids with surface charge

127

heterogeneity?

128 129

Numerical Experimental Section

130

Particle Transport Model. A Lagrangian trajectory approach was adapted in this model to simulate the

131

time evolution of the three-dimensional (3D) translation and rotation of rod-shaped particles during

132

transport in shear flow. The translation and rotation of a rigid particle can be described by the following

133

equations of motion:

134

{

135

{

𝑑𝐮

𝑚 𝑑𝑡 = ∑𝐅

(𝟏)

𝑑𝐱 𝑑𝑡

(𝟐)

=𝐮

𝑑(𝐈 ∙ 𝛚) 𝑑𝑡 = 𝑑𝛀 𝑑𝑡 = 𝛚

∑𝐓

(𝟑) (𝟒)

136

where t is time, m is the mass of the particle, 𝐈 is the moment of inertia of the particle; x and u are the linear

137

position and velocity vectors of the particle, respectively;  and 𝛚 are the respective angular position and

138

velocity vectors of the particle; and ∑𝐅 and ∑𝐓 represent the total forces and torques acting on the particle,

139

respectively. Integrating all the forces acting on a particle yields particle linear velocity; likewise, integrating

140

all the torques acting on a particle yields particle rotational velocity.1, 14, 49, 71-73 Among the forces a particle

141

experienced, the following deterministic forces were accounted for in this research: virtual mass force, gravity,

142

hydrodynamic fluid drag, shear lift, colloidal forces (van der Waals, electrostatic double layer) and non-

143

colloidal (steric) forces. In this work, Brownian motions (including translation and rotation) were modeled

6 ACS Paragon Plus Environment

Page 7 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

144

separately as a stochastic process for each degree of freedom using Einstein’s equations.74 The torques acting

145

on the particle are derived from the forces that do not pass through the particle center of mass, e.g., colloidal

146

forces and hydrodynamic fluid drag. Detailed descriptions on how to solve the equations of motion, account

147

for all the forces and torques, and incorporate Brownian translation and rotation can be found in our previous

148

work.1 In this work, simulations were primarily performed in the Happel model (Figure 1), which represents

149

porous media as a collection of spherical collectors each enclosed in a concentric spherical fluid shell. The

150

flow field within the Happel model can be described analytically,75-76 and a representation of the laminar flow

151

field in the Happel model was also provided in Figure S1 in the Supporting Information.

152 153

Because of dilute particle concentration and small particle sizes (in the range of nanometers to microns)

154

investigated, the fluid field disturbance due to the presence of particles was neglected, meaning that the

155

velocity at the particle center was represented by the fluid velocity at that location in the absence of the

156

particle. Likewise, particle-particle interactions or aggregations were ignored as a result of the assumption of

157

dilute particle concentration. In this study, we used the universal hydrodynamic functions (𝑓1,𝑓2,𝑓3,𝑓4) that

158

were based on the equivalent hydrodynamic Stokes sphere assumption to approximate the wall effects of

159

ellipsoid particles (please refer to our previous work and other references for details).1, 13-14, 71

160 161

Coordinate Systems. To represent the 3D rotation and translation of ellipsoidal particles, the following

162

coordinate systems are needed (Figure 2): (1) an inertial frame with its origin often located on the

163

collector, 𝐱 = [x,y,z]; (2) a particle frame with its origin at the center-of-mass of the particle and its axes

164

being the particle principal axes, 𝐱 = [x,y,z]; (3) a co-moving frame, 𝐱 = [x,𝑦,z], with its origin at the

165

center-of-mass of the particle and its axes parallel to the corresponding axes of the inertial frame.

166

Transformation of the coordinates between the particle and co-moving frames can be achieved by the

167

orthonormal transformation matrix 𝐀 given by Goldstein:77

168

1 ― 2(𝑞22 + 𝑞23) 2(𝑞1𝑞2 + 𝑞3𝑞0) 2(𝑞1𝑞3 ― 𝑞2𝑞0) 𝐀 = 2(𝑞2𝑞1 ― 𝑞3𝑞0) 1 ― 2(𝑞23 + 𝑞21) 2(𝑞2𝑞3 + 𝑞1𝑞0) , 2(𝑞3𝑞1 + 𝑞2𝑞0) 2(𝑞3𝑞2 ― 𝑞1𝑞0) 1 ― 2(𝑞21 + 𝑞22)

[

]

7 ACS Paragon Plus Environment

(5)

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 45

169

where q0, q1, q2 and q3 are the four Euler parameters. The particle position can be described by tracking its

170

center-of-mass in the inertial frame. The particle frame is mainly used to describe the rotational motions

171

of the particles by following the three Euler angles (i.e., ,  and ) (Figure 2) or four Euler parameters

172

(i.e., quaternions q0, q1, q2, q3). Euler angles and Euler parameters are also interconvertible.1

173 174

Rotation Dynamics of an Ellipsoid in Shear Flow. Euler angles are convenient for visualization. Thus,

175

they were used to assign particle initial orientation and to record the final orientation of a particle. Euler

176

parameters are frequently employed in the simulations, since the use of Euler angles would involve a

177

large number of trigonometric calculations, which are computationally expensive. As a consequence, the

178

quaternions would evolve with time according to the angular velocity and describe the time evolution of

179

particle orientation:78

180

[ ][ 𝑑𝑞0/𝑑𝑡 𝑑𝑞1/𝑑𝑡 1 𝑑𝑞2/𝑑𝑡 = 2 𝑑𝑞3/𝑑𝑡

]

― 𝑞1𝜔𝑥 ― 𝑞2𝜔𝑦 ― 𝑞3𝜔𝑧 𝑞0𝜔𝑥 ― 𝑞3𝜔𝑦 + 𝑞2𝜔𝑧 𝑞3𝜔𝑥 + 𝑞0𝜔𝑦 ― 𝑞1𝜔𝑧 , ― 𝑞2𝜔𝑥 + 𝑞1𝜔𝑦 + 𝑞0𝜔𝑧

(6)

181

where 𝜔𝑥, 𝜔𝑦, and 𝜔𝑧 were the components of particle angular velocity vector in the particle frame. This

182

equation illustrates how eq 4 was implemented in our simulations.

183 184

The angular velocity of a moving ellipsoidal particle is governed by the torques acting on the particle. For

185

a homogeneously dense particle, only forces that do not pass through the center-of mass of the particle

186

would generate torques, for example, hydrodynamic fluid drags and colloidal forces. As a result, for an

187

ellipsoid moving in a shear flow, eq 3 can be decomposed as1, 35, 79 𝑑𝜔𝑥

188

𝐼𝑥 𝑑𝑡 ― 𝜔𝑦𝜔𝑧(𝐼𝑦 ― 𝐼𝑧) = 𝑇ℎ𝑥 + 𝑇𝑐𝑜𝑙𝑙 𝑥 ,

189

𝐼𝑦 𝑑𝑡 ― 𝜔𝑧𝜔𝑥(𝐼𝑧 ― 𝐼𝑥) = 𝑇ℎ𝑦 + 𝑇𝑐𝑜𝑙𝑙 𝑦 ,

𝑑𝜔𝑦

8 ACS Paragon Plus Environment

(7)

Page 9 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

𝑑𝜔𝑧

𝐼𝑧 𝑑𝑡 ― 𝜔𝑥𝜔𝑦(𝐼𝑥 ― 𝐼𝑦) = 𝑇ℎ𝑧 + 𝑇𝑐𝑜𝑙𝑙 𝑧 ,

190 191 192

where the components of the moment of inertia are (using z-axis as the major axis here): 𝐼𝑥 = 𝐼𝑦 = (𝛽2 + 1)𝑎2 5

𝑚, 𝐼𝑧 =

2𝑎2 5 𝑚,

a is the semi-minor axis of the ellipsoid and  is the aspect ratio (equal to major

193

𝑐𝑜𝑙𝑙 𝑐𝑜𝑙𝑙 axis divided by minor axis); 𝑇ℎ𝑥,𝑇ℎ𝑦, 𝑇ℎ𝑧 and 𝑇𝑐𝑜𝑙𝑙 𝑥 ,𝑇𝑦 , 𝑇𝑧 are the components of torques due to

194

hydrodynamic drag and colloidal forces in the particle frame, respectively.

195 196

Heterogeneity Pattern Design. To determine how the location of charge heterogeneity over particle

197

surface affects retention, the following three pattern designs of heterogeneity patches were employed in

198

the simulations: heterogeneity on one-end, heterogeneity on two-ends, and heterogeneity on the middle

199

band (Figure 3). The heterogeneous patches were assigned with positive charge, whereas the rest of

200

particle surface was negatively charged (see Table 1 for zeta-potential values).

201 202

Colloidal Forces and Torques. The heterogeneity patches on the ellipsoidal surface call for a discretized

203

method to accurately account for the colloidal forces between an ellipsoid and the collector surface. The

204

surface element integration (SEI) method80-81 was employed here to calculate colloidal interactions, e.g.,

205

the van der Waals interactions and electrical double layers forces (Figure 4). At each separation distance,

206

the particle surface was discretized into small elements (e.g., size of 0.2 nm to 60 nm). Each particle

207

surface element was assigned a zeta potential value according to specified heterogeneity design pattern.

208

The interactions between each discretized element from the particle and the entire collector surface were

209

computed and then integrated over entire particle surface to obtain the overall interaction between the

210

particle and the collector.1 The SEI method is capable of studying colloidal interactions for particles of

211

arbitrary shape, arbitrary orientation and arbitrary heterogeneity pattern.

212 9 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

213

Similarly, to calculate torques, the interaction forces for each discretized particle surface element were

214

multiplied with the lever distance from the center-of-mass of the particle to the line connecting the

215

particle surface element to its projected element over the collector surface. This operation was then

216

repeated over entire particle surface to obtain the total torques produced from colloidal forces.1 In this

217

particular case, to use SEI method, one needs to find the closest approach distance between the ellipsoid

218

of an arbitrary orientation and the spherical collector surface, and the method to find such a distance can

219

be found in our previous study.1

220 221

Computational Algorithm and Simulation Details. A computer program for solving the time evolution

222

of the three-dimensional translation and rotation of an ellipsoidal particle with particle surface charge

223

heterogeneity in a shear field was developed. Briefly, at each time step, according to the total forces and

224

torques acting on each particle, particle linear velocity and angular velocity were calculated based on eq 1

225

and eq 3 (or eq 7 more specifically), respectively. Based on the obtained linear velocity (eq 2) and angular

226

velocity (eq 4, or more specifically eq 6), particle position vector (x,y,z) and orientation matrix 𝐀 (given

227

by q0, q1, q2, q3) were correspondingly determined. Detailed computational steps can be found in our

228

previous work.1

229 230

The two-step Adams scheme was used to advance the particle position based on eq 2. Gauss elimination

231

method was used to solve simultaneously the coupled equations of quaternions (eq 6) to obtain the new

232

orientation quaternions and also to solve the three components of eq 1 to obtain the new translational

233

velocity on each direction. For the coupled non-linear equations in eq 7, a mixed approach was used,

234

where the time derivatives on the left-hand sides of the equations were treated by forward differencing. In

235

each equation, the particle angular velocities that corresponded to the dependent variable in the time

10 ACS Paragon Plus Environment

Page 10 of 45

Page 11 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

236

derivative term was evaluated at the next time step, and the other angular velocity and velocity gradient

237

terms were evaluated at the current time step.1

238 239

We performed a series of simulations under unfavorable conditions (i.e., several different heterogeneity

240

patterns as shown in Figure 3) with ellipsoidal particles of different sizes (diameters between 200 nm –

241

6.6 µm) and shapes (aspect ratios from 1- 6) flowing within the Happel model under pore water velocity

242

of 5 m/day. For each simulation, approximately 2000 – 9000 particles were injected randomly (both in

243

position and orientation) into the model system as shown in Figure 1. Simulation parameter values were

244

provided in Table 1. The computation algorithms described above were carried out to obtain the time

245

evolution of particle translation and rotation. We considered particles were in “physical contact” with the

246

collector surface once the separation distances were within a roughness layer 𝛿,1, 80 which, depending

247

upon particle sizes, were typically between 2-10 nm for our simulations. Within this distance, particle-

248

collector contact forces and torques were evaluated to investigate the dynamic adhesion of rolling or

249

rotating particle.37 JFK theory82 was employed to obtain particle deformation area resulting from particle-

250

surface contact. Based on this deformation, a non-hydrodynamic adhesive torque (e.g., resulted from

251

gravity, colloidal forces) was obtained. A particle was considered to be attached when the adhesive torque

252

was greater than the driving hydrodynamic torque.1 Based on simulated trajectory outcome, e.g., attached,

253

exited, or remained in the system without direct attachment, collector efficiency (the extent of retention),

254

attachment rate and attachment orientation were obtained.

255 256

Validation of Simulation Model. We have validated our trajectory model in two steps: i) simulated

257

retention trends for spherical particles agreed with existing filtration theory under favorable conditions;1 ii)

258

simulated rotation patterns of rod-shaped particles in a linear shear flow ignoring any external forces and

259

diffusion matched well with Jeffery’s prediction.41 Note that, in the trajectory simulations, we can easily

11 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

260

turn off each mechanism including Brownian motion, and external forces and diffusion were turned off

261

for this particular simulation to match Jeffery’s theoretical conditions.

262 263

Results with Discussion

264

Colloidal retention onto collector surfaces is controlled by its ability to transport to the vicinity of

265

collector surfaces and the subsequent interactions between the colloid and the collector. The shape of a

266

colloid (e.g., rod-shaped) can greatly affect both the transport process and the colloid-surface interactions.

267

Under favorable conditions, since every collision of a colloid with the collector leads to attachment, the

268

collector efficiency is essentially the probability of particles colliding with the collector. Under

269

unfavorable conditions, not all collisions result in attachment; thus, the attachment efficiency (α) is

270

conventionally used to describe the probability of attachment upon collision. The attachment efficiency is

271

calculated by the ratio of simulated single collector efficiency under unfavorable condition to that under

272

favorable condition. Below, we will present and discuss how the location of heterogeneous patches,

273

particle aspect ratio, colloid size affect retention, distribution and orientation of attached colloids over the

274

collector surface under unfavorable conditions.

275 276

Effect of Heterogeneity Location over Colloid Surface on Retention

277

Heterogeneous Patches Located on One versus Two Ends. The attachment efficiencies () for the

278

heterogeneous patches located on one end versus two ends of colloids of 1 m size as a function of

279

colloid charge heterogeneity coverage (0 - 100) from the Happel model were provided in Figure 5

280

(porosity 0.36, grain diameter 390 µm, pore water velocity of 5 m/day). For all the rods (AR = 2 – 6) and

281

spheres (AR = 1) examined here, the minimum surface coverage required to capture a colloid for

282

attachment was the same for heterogeneity on one end versus two ends. Also, the retention trends for

283

heterogeneity on one end versus two ends as a function of colloid surface heterogeneity coverage were 12 ACS Paragon Plus Environment

Page 12 of 45

Page 13 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

284

very similar (Figure 5 a versus b). The similar retention behaviors for heterogeneity located on one end

285

versus two ends were also observed from other particle sizes (e.g., 6.6 m in diameter as in Figure S2 in

286

the Supporting Information). Nonetheless, since heterogeneity located at the one versus two ends resulted

287

in similar retention behaviors, in what follows we will only examine the cases where heterogeneous

288

patched were located at the two ends of a colloid.

289 290

Also, from Figure 5, rods needed smaller patch coverage to initiate retention compared to spheres. This

291

reflected that the tumbling motions of rods in shear flow, as illustrated in our previous work,1 increased

292

the likelihood of colloid end(s) to intercept the collector surface. In addition, due to the relatively smaller

293

curvature at the end, a rod would experience less repulsion compared to a sphere,6,

294

discuss more shortly.

65

which we will

295 296

Heterogeneous Patches Located on the Ends versus the Middle Band. The attachment efficiencies for

297

the heterogeneous patches located on the ends versus the middle band of colloids for three representative

298

sizes (200 nm, 1 m and 6 m in diameter) as a function of colloid charge heterogeneity coverage (0 -

299

100) from the Happel model were presented in Figure 6. For spheres (aspect ratio = 1) undergoing

300

three-dimensional rotation, there was no distinction between the ends and the middle. Consequently, the

301

location of heterogeneity patches on a sphere surface (ends versus middle) did not affect retention. But for

302

rods, the location of heterogeneous patches over colloid surface affected retention dramatically. For

303

example, for 200 nm colloids, when charge heterogeneity was located on the ends, as shown in Figure 6a,

304

rods (with aspect ratio of 2 – 6) required considerably smaller patches compared to spheres to initiate

305

attachment. However, when charge heterogeneity was located on the middle band of the colloids (Figure

306

6b), rods required significantly larger patches compared to spheres to start attachment. The opposite

307

effects of heterogeneity location (ends versus middle) on retention of rods relative to that of spheres were

308

observed for other colloid sizes as well (see Figure 6c - 6d for 1 m colloids, and 5e - 5f for 6.6 m

13 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

309

colloids).

310 311

In general, once colloid surface charge heterogeneity was greater than the minimum patch size (which

312

was calculated based on the minimum surface coverage times the total surface area of the particle)

313

required for retention, attachment efficiencies increased with increasing heterogeneity coverage, till

314

reaching retention for favorable conditions (where   1). This was true for both spheres and rods of all

315

the sizes examined. The smaller the colloid size, the faster the attachment efficiency transitioned from

316

completely unfavorable to favorable conditions (see Figure 6).

317 318

We have rendered the simulated 3D translational and rotational motions of representative rods with aspect

319

ratio 6 for the heterogeneous patches located on the ends (Video S2) versus the middle band (Video S3)

320

within the Happel model into videos and provided in the Supporting Information. One can observe from

321

these two videos that the near-surface rotation patterns for rod particles to become attached were different

322

for the heterogeneous patches located on the ends versus the middle band. For the attached rods with

323

heterogeneous patches on the middle band, one can see that they approached to the collector surface with

324

a side-on orientation and the plane of rotation was approximately parallel to the collector surface.

325

Whereas the attached rods with heterogeneous patches on the ends generally travelled near the surface

326

region with an end-on orientation and the plane of rotation was about normal to the collector surface.

327 328

Effect of Colloid Aspect Ratio on Retention under Unfavorable Conditions

329

As shown in Figure 6, the influences of aspect ratio on retention under unfavorable conditions strongly

330

depended upon the location of colloid surface charge heterogeneity. When charge heterogeneity was on

331

the ends of colloids, the minimum surface coverage required for attachment, as well as heterogeneous

332

coverage needed to transition from unfavorable to favorable retention, decreased with increasing aspect

333

ratio (Figures 6a, 6c, 6e). For instance, 200 nm rods with aspect ratio 6 started to attach at very small

14 ACS Paragon Plus Environment

Page 14 of 45

Page 15 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

334

heterogeneity coverage (< 0.01%); at around 0.1% heterogeneity coverage, these rods could reach

335

retention for favorable conditions (α ≈ 1). As the aspect ratio decreased, rods required more heterogeneity

336

coverage to initiate attachment (e.g., ~ 0.1% and 0.5% for AR = 3 and 2, respectively) and to reach

337

attachment for favorable conditions (e.g., ~ 0.2% and 1% for AR = 3 and 2, respectively). Spheres needed

338

the most heterogeneity coverage – started to attach at 4-6% heterogeneity coverage and behaved like

339

favorable condition at 6-11% heterogeneity coverage.

340 341

However, when charge heterogeneity was on the middle section of colloids, the minimum surface

342

coverage required for attachment, along with the heterogeneous coverage needed to transition from

343

unfavorable to favorable retention, increased with increasing aspect ratio (Figures 6b, 6d, 6f). For instance,

344

for 200 nm particles, spheres started to attach and reached favorable retention with the smallest

345

heterogeneity coverage (14% and 17% respectively); and then followed by rods with aspect ratio of 2 and

346

3. Rods with aspect ratio 6 started to attach and reached favorable retention with the biggest heterogeneity

347

coverage (73% and 86% respectively). Besides plotting the attachment efficiency as a function of surface

348

heterogeneity coverage as shown in Figure 6, a similar figure showing the collector efficiency as a

349

function of surface heterogeneity coverage was also provided in Figure S3 in the Supporting Information.

350

Additionally, the minimum raw patch sizes required for particle capture changed with particle size, aspect

351

ratio and heterogeneity patterns (see Table S1 in the Supporting Information), but the retention trends

352

versus minimum surface coverage discussed above also held when raw patch sizes were used (see Table

353

S1 and discussion thereafter in the Supporting Information).

354 355

When colloids were relatively far away from the collector surfaces, the location of charge heterogeneity

356

over colloid surface would not affect the transport behavior of rods or spheres. In other words, regardless

357

of the patches on the ends or middle, a colloid would have similar propensity to transport to the vicinity

15 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

358

of the collector surface. However, once a colloid was very close to the collector, the shape of colloid (i.e.,

359

aspect ratio here) and the location of colloid charge heterogeneity could greatly influence colloid-surface

360

interactions. Figure 7 illustrated the integrated DLVO (abbreviated after Derjaguin-Landau-Verwey-

361

Overbeek) interaction profiles computed from the SEI methods for 6.6 m colloids with the

362

heterogeneous patches on the ends versus the middle for various aspect ratios. When charge heterogeneity

363

was located on the ends, repulsive energy barriers decreased with increasing aspect ratio for a given

364

colloid charge heterogeneity coverage (Figure 7a). This was due to the decreasing radius of curvature

365

with increasing aspect ratio when a colloid approached the collector surface with an end-on orientation. In

366

addition, as the heterogeneity coverage increased, the energy barriers dropped faster for rods with larger

367

aspect ratio to reach no-barrier attachment conditions relative to spheres (Figure S4 (a) in the Supporting

368

Information). For this reason, rod particles with larger aspect ratio required less heterogeneity coverage to

369

initiate attachment and to reach favorable attachment. Likewise, when charge heterogeneity was located

370

on the middle band, repulsive energy barriers increased with increasing aspect ratio (Figure 7b), because

371

the radius of curvature increased with increasing aspect ratio when a colloid approached the collector

372

surface with a side-on orientation. Moreover, as the heterogeneity coverage increased, the energy barriers

373

decreased slower with increasing aspect ratio (Figure S4 (b) in the Supporting Information). Consequently,

374

relative to spheres, rods with higher aspect ratio required larger heterogeneous patches to start attachment

375

and reach favorable retention.

376 377

In Figure 6f, for large size rods with heterogeneous patches located on the middle band, we observed that

378

the attachment efficiencies reached a plateau, but well below favorable retention, over a wide percentage

379

of colloid surface coverage (SCOV). For example, for 6.6 m rods with aspect ratio 6,  was

380

approximately 0.25 when SCOV varied within 32 – 99%. Similar plateau in  was observed for 6.6 m

381

rods with aspect ratio of 3 and 2, but not for small rod sizes of any aspect ratio examined (Figure 6b and

382

6d). To be sure of this behavior, we also ran simulations for 6.6 m with aspect ratio 6 under the same 16 ACS Paragon Plus Environment

Page 16 of 45

Page 17 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

383

heterogeneous pattern and coverage as in Figure 6f within the parallel plate geometry and observed

384

similar plateau behavior in  (data not shown). This can be partly explained by the concept of zone of

385

influence from the electrostatic double layer interaction for particles of differing size and curvature. Since

386

electrostatic interaction decays exponentially with distance away from the collector, for a given colloid-

387

collector separation distance, the relative portion of heterogeneous patches in the middle band out of the

388

entire colloid surface area contributing to the electrostatic interaction was smaller for large size rods (e.g.,

389

6.6 m) compared to small size rods (e.g., 1 m) (Figure S5 in the Supporting Information). Another

390

reason for the observed plateau in  arose from particle rotation, which controlled the likelihood of the

391

rod middle section facing the collector surface. As demonstrated above (see Video S3 in the Supporting

392

Information), for heterogeneity located on the middle band, only those trajectories that happened to be a

393

side-on orientation near the collector surface were able to attach. This plateau suggested that the rod

394

retention within this SCOV range was rate-limited by particle transport and rotation, not by colloid-

395

collector interaction.

396 397

These behaviors indicated that it was easier for rods to become attached to the collector when the

398

attractive heterogeneity was on the end(s) than on the side. This could explain why many rod-shaped

399

bacteria have attachment ligands particularly on the end(s).83-87 A traveling and rotating rod-shaped

400

bacterium would intercept with surrounding surfaces more likely with its end first and subsequently

401

become attached with an end-on orientation.66 On the other hand, a study from Afrooz et al.88 on

402

aggregation behaviors of poly-acrylic acid-coated gold nano-spheres versus nano-rods (12 nm, aspect

403

ratio = 5) showed that nano-rods were less likely to form aggregates in low NaCl concentrations

404

compared to nano-spheres. The relative stability of these nano-rods in suspension primarily resulted from

405

the stronger electro-steric repulsion when rods approached each other with side-on orientation to

406

aggregate compared to spheres due to curvature effect.88 These observations qualitatively agreed with our

407

simulation results when the heterogeneous patches were on the middle section of rods. Moreover, as 17 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

408

mentioned earlier, contradictory results regarding the role of rod aspect ratio on retention had been

409

reported from experiments.15, 39-40 Our findings here elucidated that the influences of rod aspect ratio on

410

retention depended strongly on the location of colloid surface heterogeneity, which may provide a

411

possible explanation for those observed contradictions.

412 413

Effect of Colloid Size on Retention under Unfavorable Conditions

414

The size of a colloid (e.g., rod or sphere) greatly affects particle transport process as well as colloid-

415

surface interactions. The simulated single collector efficiencies () for rods (aspect ratio = 6) and their

416

counterpart spheres of equivalent volume were plotted for several representative sizes (0.1, 0.2, 1, 3, 6.6

417

µm in diameter) under different charge heterogeneity coverages at pore water velocity of 5 m/day with

418

the Happel model in Figure 8. For spheres, when surface charge heterogeneity coverage reached 30% or

419

above, the retention trends were similar to those under favorable conditions (Figure 8a), where retention

420

was at minimum for size ~1 – 3 µm, as expected from colloidal filtration theories.10, 71, 76, 89-90 As colloid

421

heterogeneity coverage decreased (e.g., 3.5%, 1.5%), the simulated collector efficiencies decreased and

422

the decreased amount in  strongly depended upon sphere size. For small size spheres (e.g., 0.1 and 0.2

423

m), our model predicted no retention because 3.5% coverage or lower was below the minimal patch

424

coverage needed to initialize attachment (see Figure 6a). For spheres from 1 to 6.6 µm, a local minimum

425

in retention was still observed around 1 – 3 µm, but the drop in  compared to favorable conditions was

426

more pronounced around this minimum (e.g., at 1.5%). It is worth to note that the gaps in retention,

427

especially around sphere sizes corresponding to local minimum attachment, between favorable and

428

unfavorable conditions have been observed by experiments.21, 52 For example, in an impinging jet cell to

429

study colloid retention with carboxylate-modified polystyrene latex fluorescent microspheres, Rasmuson

430

et al.52 reported that the minimal retention occurring at 1.1 µm microspheres dropped by two orders of

431

magnitude from favorable to unfavorable conditions, exhibiting the biggest gap compared with other

18 ACS Paragon Plus Environment

Page 18 of 45

Page 19 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

432

sphere sizes. Our simulation results in Figure 8a can be explained by the colloidal forces between patched

433

colloids with the collector surface. To illustrate this, we compared the DLVO energy profiles for the

434

sphere sizes examined above (see Figure S6 in the Supporting Information) at a heterogeneity coverage

435

0.25%, and found that spheres around 3 µm in diameter, which corresponded to the size with the largest

436

gap in retention in Figure 8a, were subjected to the biggest energy barrier compared to other sizes. The

437

0.25% coverage was chosen because at this coverage, an energy barrier existed for all the sphere sizes

438

examined, even when the heterogeneous patch was facing the collector surface. In other words, at 1.5% or

439

3.5% colloid surface coverages, retention may occur when the heterogeneous patch faced toward the

440

collector at close contact due to lack of repulsion; but retention definitely would not occur when the patch

441

faced away from the collector surface. Essentially, retention for heterogeneously patched spheres

442

reflected the probability of the patch(s) facing the collector surface when spheres were transported near

443

collector surfaces, indicating the importance of particle rotation on retention.

444 445

For rods, when colloid surface charge heterogeneity coverage (located at the ends of rods) reached 6% or

446

above, the retention trends were similar to those under favorable conditions, where the dependency of 

447

showed a “W-shape” dependency on rod size (Figure 8b) – for particle size greater than 2 µm, the

448

attachments of rods increased with increasing size; for particle size smaller than 200 nm, the attachment

449

of rods decreased with increasing size; whereas for rod particle size from 200 nm to 2 µm, the attachment

450

first increased then decreased with increasing size (“hump” region).1 As colloid heterogeneity coverage

451

decreased, e.g., < 0.01% coverage with patchy ends, retention of rod-shaped particles still followed the

452

trend as observed under favorable condition, but shifted to smaller  values. As described above, when

453

charge heterogeneity was located at the ends of rods, only a very small coverage was required to initiate

454

retention due to the small radius of curvature for an end-on orientation. Thus, for rod-shaped particles

455

with heterogeneous patches on the end(s) under unfavorable conditions, retention was largely limited by

19 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

456

the rotation dynamics as observed under favorable conditions,1 which controlled the rate of particles

457

being transported to the vicinity of collector surfaces.

458 459

Attachment Location and Orientation

460

Rod-shaped particles were observed to attach all over the Happel model surface (Figure 9a). In general,

461

more colloids were retained at the upstream collector surface relative to downstream. For heterogeneity

462

located on the ends versus on the middle band, the basic patterns of attachment locations were similar.

463

However, rod-shaped particles had higher tendency to attach at the downstream relative to spheres,

464

similar to what we have observed under favorable conditions.1 This was due to the drifting effect of rod

465

particles, which enabled rod particles to travel further downstream to attach. Also, small size particles

466

(e.g., 0.2 µm rods or spheres) were observed to more likely travel downstream for attachment compared

467

to large size particles (e.g., 6.6 µm rods or spheres), which was due to diffusion and was well discussed

468

by the general filtration theory.10-13

469 470

With regard to attachment orientation, for all the unfavorable conditions investigated (e.g., different

471

heterogeneous patterns over colloidal surfaces), both spheres and rods attached via the attractive

472

heterogeneity patches (the orientation distribution were shown for rods in Figure 9). Since it is easier to

473

visualize in the parallel plate geometry, we also performed simulations in that geometry under the same

474

conditions as those in the Happel model and presented the simulated orientation distribution patterns for

475

attached rods (aspect ratio 6) in Figure 9(b). The parameters for the parallel plate geometry were set with

476

same channel width (31.3 µm to fluid envelope, very narrow) and average pore water velocity as those in

477

the Happel model. Hence, aside from the curved collector surface from the Happel model, simulation

478

results obtained from the parallel plate geometry under similar conditions should allow us to clearly

479

visualize the orientations of attached rod colloids, as shown in Figure 9(b). One can see clearly that, in the 20 ACS Paragon Plus Environment

Page 20 of 45

Page 21 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

480

parallel plate model, rod-shaped particles with heterogeneity on the ends adhered to the collector surface

481

with one end, whereas rods with heterogeneity on the middle band attached via a side-on orientation. Note

482

that, this observation did not contradict with previous statements wherein rod-shaped particles preferred

483

an end-on orientation for initial attachment, because for the same heterogeneity coverage (e.g. before

484

reaching favorable conditions), the attachment efficiencies for rods with heterogeneity on the ends were

485

often several orders of magnitude greater than those for rods with heterogeneity on the middle band. We

486

stress here that in this work, we do not intend to quantitatively compare the attachment efficiency or

487

collector efficiency obtained from the Happel model with those from the parallel plate flow chamber. The

488

main reason we included the parallel plate geometry was to validate our trajectory model development for

489

rods against Jeffery’s theory which was done in simple shear flow,41and to clearly illustrate rod

490

attachment orientation, due to its simple flow field and flat collector surfaces. That being said, since this

491

work focused on the near-surface laminar flow regions (less than 50 m away from the walls) from both

492

Happel and parallel plate models, the trends on attachment efficiency obtained from the Happel model

493

should qualitatively apply to the parallel plate geometries, based on our preliminary simulation data from

494

the latter (data not shown).

495 496

Effects of Initial Position and Orientation on Rod Retention

497

Particle initial position and orientation greatly affected the retention and orientation of rods, and this

498

effect depended upon model geometry dimensions relative to particle size. As described above, it is easier

499

to visualize rod orientation in the parallel plate geometry, so the effects of particle initial conditions on

500

retention will be discussed primarily in that geometry (unless stated otherwise). For the narrow and short

501

channel explored here (e.g., half channel height of 31.3 m, channel length of ~ 1 mm, similar to the

502

Happel model dimension), large size retained rods (e.g., 6.6 m) with heterogeneity located on the ends

503

had a wide range of initial orientations (e.g., Euler angle , which characterizes the angle between the rod

21 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

504

major axis and the normal to the flat collector surface in the parallel plate geometry, varied from 0 to ),

505

whereas large size retained rods with heterogeneity located on the middle band tended to have an initial

506

orientation close to side-on relative to the channel walls (e.g., Euler angle  was close to /2). However,

507

small size retained rods (e.g., 1.0 m) with either side-on or end-on attachment orientation had a wide

508

range of initial orientations, regardless of the locations of heterogeneity (data not shown). Based on this

509

observation, the channel height and/or channel length were increased by 10 times, we then observed that

510

large size retained rods (e.g., 6.6 m) with heterogeneity located on the middle band no longer showed

511

any preferable initial orientations. Moreover, the effect of initial position on retention depended upon the

512

channel length. For instance, when the channel length was set to 1 mm (similar to the travel length in the

513

Happel model), rods with initial positions within ~10 um distance from the wall (limiting trajectory

514

distance) could become attached. This limiting trajectory distance increased to ~20 um when channel

515

length was increased to 3 mm. However, further increase in channel length would not increase the

516

limiting trajectory distance any more for the conditions examined here. As described earlier and in our

517

previous paper,1 rods tended to rotate in shear flow, and their positions would oscillate during rotation,

518

which led to the drift of rods across flow streamlines toward the channel walls. If confined within a

519

narrow and/or short channel, the oscillation and drifting effects of rods may not be fully observable.

520 521

Likewise, in the Happel model, rods injected close to the center of the injection plane had a higher

522

possibility to become attached, whereas particles injected outside of the limiting injection area were not

523

going to attach (Figure 1). In the Happel model, the limiting trajectories beyond which particles wouldn’t

524

attach were similar for rods for the two heterogeneity patterns (ends versus middle band). This reflected

525

the fact that the location of heterogeneity over the colloidal surface did not affect the rotational and

526

oscillational motions (shown in Video S1 in the Supporting Information) of the rods when far away from

527

the collector surface. But we did not observe any preferential initial orientation for the attached rods with

528

middle-band heterogeneity from the Happel model, which was probably due to the coupled effects from 22 ACS Paragon Plus Environment

Page 22 of 45

Page 23 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

529

the complicated flow patterns as well as collector curvature in the Happel model relative to those in the

530

parallel plate geometry. For the retained rods with heterogeneity located on the ends, no preferential

531

initial orientation was observed from both geometric models.

532 533

Summary and Conclusions

534

In this work, we extended our three-dimensional particle trajectory model to investigate the influences of

535

particle shape and location of colloid surface charge heterogeneity on retention in porous media under

536

unfavorable conditions for a wide range of particle sizes. Our model accounted for the combined

537

influences of flow hydrodynamics, Brownian rotation and translation, colloid-surface interaction on the

538

rotation and retention dynamics of prolate spheroid particles during transport process. Retention of a

539

colloid onto collector surfaces was governed by its ability to transport to the vicinity of collector surfaces

540

and the subsequent interactions between the colloid and the collector. The shape of a colloid greatly

541

affected both transport process and colloid-surface interactions. The location of colloid heterogeneous

542

patches did not affect transport, but predominantly affect colloid-surface interactions by influencing the

543

likelihood of the heterogeneity patches facing the collector due to particle rotation.

544 545

Under unfavorable conditions, particles (rods or spheres) attached via attractive heterogeneity. For rod-

546

shaped colloids, it was more effective for retention to occur when heterogeneities were located on the

547

end(s) than on the middle band. This indicated that the initial retention of rod particles was controlled by

548

particle rotation, since the end of a rod had higher probability to intercept the collector surface during

549

rotation for attachment than its sides. In addition, a rod would experience much less repulsion with an

550

end-on orientation toward the collector surface than a side-on orientation due to curvature effect. This

551

observation could explain why rod-shaped bacteria have tethered ligands located on the end(s),83-87 and

552

why rod-shaped bacteria would intercept with surrounding surfaces more likely with its end first to 23 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

553

establish initial contact.66 The preferential attachment by the end(s) of a rod-shaped particle could have

554

important implications for rod-shaped bacteria (or engineered rods) transport and application in biological

555

or colloidal science fields.

556 557

The influences of particle aspect ratio on retention under unfavorable conditions strongly depended upon

558

the location of colloid surface charge heterogeneity. For a given colloid surface coverage, retention

559

increased with increasing aspect ratio when charge heterogeneity was on the end(s) of colloids; whereas

560

retention decreased with increasing aspect ratio when charge heterogeneity was located on the middle

561

band of colloids. These interesting transport and retention behaviors for rod-shaped colloids could not be

562

captured if rods were simplified as “effective spheres” or assumed with “homogeneous surface”. Our

563

findings could provide useful guidance on engineering rod-shaped spheroids of differing aspect ratios

564

with appropriate surface properties to achieve desired transport and retention results. More importantly,

565

our study demonstrated that when studying the transport process of non-spherical colloids or spherical

566

particles with inhomogeneous surface properties – the typical scenarios encountered in the environment,

567

particle rotation becomes very important and needs to be considered.

568 569

Supporting Information. Additional information includes laminar flow field in the Happel model,

570

energy profiles for spheres and rods with patched surfaces, and short videos for the tumbling and

571

oscillational motion of rods in the parallel plate geometry, and transport of rods with different

572

heterogeneity patterns in the Happel geometry. This material is available free of charge via the Internet at

573

http://pubs.acs.org.

574

24 ACS Paragon Plus Environment

Page 24 of 45

Page 25 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

575

Acknowledgements. This article was based upon work supported by the National Science

576

Foundation Hydrologic Science Programs (EAR 1521421). Any opinions, findings, and conclusions or

577

recommendations expressed in this material are those of the authors and do not necessarily reflect the

578

views of the National Science Foundation. We are grateful for the technical and facility support provided

579

at the Center for High Performance Computing at the University of Utah. We thank Professor William P.

580

Johnson and his research group for kindly sharing their computational resources and for their technical

581

help in many subjects. We also thank Professors Yusong Li, Mathias Schubert and their research groups

582

for their constructive inputs throughout our meetings together.

25 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

583

References Cited

584

1.

585

Favorable Conditions. Langmuir 2018, 34 (9), 2967-2980.

586

2.

587

unconventional gas extraction: the critical need for field‐based hydrogeological research. Groundwater 2013, 51 (4),

588

488-510.

589

3.

590

conventional banana cultivation in Costa Rica: A disconnect between protecting ecosystems and consumer health.

591

Sci. Total Environ. 2018, 613, 1250-1262.

592

4.

593

polyacrylonitrile hollow fiber membranes: Spinning, characterization and application in fluoride removal from

594

groundwater. Chem. Eng. J. 2018,. 334, 41-53

595

5.

596

bacteria. FEMS Microbiol. Lett. 1997, 148 (2), 227-231.

597

6.

Young, K. D., The selective value of bacterial shape. Microbiol. Mol. Biol. Rev. 2006, 70 (3), 660-703.

598

7.

Young, K. D., Bacterial morphology: why have different shapes? Curr. Opin. Microbiol. 2007, 10 (6), 596-

599

600.

600

8.

601

1959, 6, 220-236.

602

9.

Carman, P. C., Permeability of saturated sands, soils and clays. J Agric Sci. 1939, 29 (2), 262-273.

603

10.

Yao, K.-M.; Habibian, M. T.; O'Melia, C. R., Water and waste water filtration: Concepts and applications.

604

Environ. Sci. Technol. 1971, 5 (11), 1105-1112.

605

11.

606

physicochemical filtration in saturated porous media. Environ. Sci. Technol. 2004, 38 (2), 529-536.

607

12.

608

engineered flow conditions. Water Resour. Res. 2011, 47 (5), W05543.

Li, K.; Ma, H., Deposition Dynamics of Rod-Shaped Colloids during Transport in Porous Media under

Jackson, R.; Gorody, A.; Mayer, B.; Roy, J.; Ryan, M.; Van Stempvoort, D., Groundwater protection and

Mendez, A.; Castillo, L. E.; Ruepert, C.; Hungerbuehler, K.; Ng, C. A., Tracking pesticide fate in

Karmakar, S.; Bhattacharjee, S.; De, S., Aluminium fumarate metal organic framework incorporated

Cooper, S.; Denny, M. W., A conjecture on the relationship of bacterial shape to motility in rod-shaped

Kahn, A., Studies on the Size and Shape of Clay Particles in Aqueous Suspension1. Clays Clay Miner.

Tufenkji, N.; Elimelech, M., Correlation equation for predicting single-collector efficiency in

Nelson, K. E.; Ginn, T. R., New collector efficiency equation for colloid filtration in both natural and

26 ACS Paragon Plus Environment

Page 26 of 45

Page 27 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

609

13.

610

and retention in zones of flow stagnation. Environ. Sci. Technol. 2007, 41 (4), 1279-1287.

611

14.

612

porous media. Environ. Sci. Technol. 2009, 43 (22), 8573-8579.

613

15.

614

Environ. Sci. Technol. 2006, 40 (20), 6336-6340.

615

16.

616

saturated porous media. J. Environ. Qual. 2010, 39 (2), 500-508.

617

17.

618

prediction of particle deposition. Langmuir 2009, 25 (12), 6856-6862.

619

18.

620

Colloid Interface Sci. 2011, 165 (2), 91-101.

621

19.

622

velocities. Int. J. Multiphase Flow 1997, 23 (1), 155-182.

623

20.

624

experimentally-observed non-straining colloid retention mechanisms in porous media in the presence of energy

625

barriers. Langmuir 2011, 27 (24), 14982-14994.

626

21.

627

colloid retention on soda lime glass in the presence of energy barriers. Langmuir 2014, 30 (19), 5412-21.

628

22.

629

Unfavorable Mineral Surfaces Using Representative Discrete Heterogeneity. Langmuir 2015, 31 (34), 9366-78.

630

23.

Mitragotri, S.; Lahann, J., Physical approaches to biomaterial design. Nat Mater 2009, 8 (1), 15-23.

631

24.

Lee, K. J.; Yoon, J.; Lahann, J., Recent advances with anisotropic particles. Curr. Opin. Colloid Interface

632

Sci. 2011, 16 (3), 195-202.

633

25.

634

filaments versus spherical particles in flow and drug delivery. Nature nanotechnol 2007, 2 (4), 249-255.

635

26.

636

interactions of nanoparticles in vascular-targeted drug delivery. Mol. Membr. Biol. 2010, 27 (4-6), 190-205.

Johnson, W. P.; Li, X.; Yal, G., Colloid retention in porous media: Mechanistic confirmation of wedging

Ma, H.; Pedel, J.; Fife, P.; Johnson, W. P., Hemispheres-in-cell geometry to predict colloid deposition in

Salerno, M. B.; Flamm, M.; Logan, B. E.; Velegol, D., Transport of rodlike colloids through packed beds.

Liu, Q.; Lazouskaya, V.; He, Q.; Jin, Y., Effect of particle shape on colloid retention and release in

Wang, P.; Keller, A. A., Natural and engineered nano and colloidal transport: Role of zeta potential in

Drelich, J.; Wang, Y. U., Charge heterogeneity of surfaces: Mapping and effects on surface forces. Adv.

Gavze, E.; Shapiro, M., Particles in a shear flow near a solid wall: Effect of nonsphericity on forces and

Ma, H.; Pazmino, E.; Johnson, W. P., Surface heterogeneity on hemispheres-in-cell model yields all

Pazmino, E.; Trauscht, J.; Dame, B.; Johnson, W. P., Power law size-distributed heterogeneity explains

Trauscht, J.; Pazmino, E.; Johnson, W. P., Prediction of Nanoparticle and Colloid Attachment on

Geng, Y.; Dalhaimer, P.; Cai, S. S.; Tsai, R.; Tewari, M.; Minko, T.; Discher, D. E., Shape effects of

Huang, R. B.; Mocherla, S.; Heslinga, M. J.; Charoenphol, P.; Eniola-Adefeso, O., Dynamic and cellular

27 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

637

27.

638

triggered release into nanoscale drug carriers. Expert Opin Drug Deliv. 2010, 7 (4), 479-95.

639

28.

640

Controlled Geometry. Iubmb Life 2011, 63 (8), 596-606.

641

29.

642

of design in nanotechnology for drug delivery. Ther. Deliv. 2012, 3 (2), 181-194.

643

30.

644

Sci. 2015, 125, 20-24.

645

31.

646

Decuzzi, P., The effect of shape on the margination dynamics of non-neutrally buoyant particles in two-dimensional

647

shear flows. J. Biomech. 2008, 41 (10), 2312-18.

648

32.

649

adhesion of drug carriers in blood vessels depend on their shape: A study using model synthetic microvascular

650

networks. J. Controlled Release 2010, 146 (2, Sp. Iss. SI), 196-200.

651

33.

652

Reynolds number flows. J. Aerosol Sci 2012, 45, 1-18.

653

34.

654

micro/nanoparticles to the endothelium in human blood flow. Biomaterials 2013, 34 (23), 5863-71.

655

35.

656

turbulent channel flows. Int. J. Multiphase Flow 2001, 27 (6), 971-1009.

657

36.

Clift, R.; Grace, J. R.; Weber, M. E., Bubbles, drops, and particles. Courier Corporation: 2005.

658

37.

Mortensen, P. H.; Andersson, H. I.; Gillissen, J. J. J.; Boersma, B. J., On the orientation of ellipsoidal

659

particles in a turbulent shear flow. Int. J. Multiphase Flow 2008, 34 (7), 678-83.

660

38.

661

bacterial cell-cell and cell-surface scattering. Proc. Natl. Acad. Sci. U. S. A. 2011, 108 (27), 10940-5.

662

39.

663

through micropore membrane filters. Environ. Sci. Technol. 2008, 42 (17), 6749-6754.

Caldorera-Moore, M.; Guimard, N.; Shi, L.; Roy, K., Designer nanoparticles: incorporating size, shape and

Pillai, J. D.; Dunn, S. S.; Napier, M. E.; DeSimone, J. M., Novel Platforms for Vascular Carriers with

Liu, Y.; Tan, J. T.; Thomas, A.; Ou-Yang, D.; Muzykantov, V. R., The shape of things to come: importance

Chen, J.; Clay, N. E.; Park, N.-h.; Kong, H., Non-spherical particles for targeted drug delivery. Chem. Eng.

Gentile, F.; Chiappini, C.; Fine, D.; Bhavane, R. C.; Peluccio, M. S.; Cheng, M. M. C.; Liu, X.; Ferrari, M.;

Doshi, N.; Prabhakarpandian, B.; Rea-Ramsey, A.; Pant, K.; Sundaram, S.; Mitragotri, S., Flow and

Tian, L.; Ahmadi, G.; Wang, Z.; Hopke, P. K., Transport and deposition of ellipsoidal fibers in low

Thompson, A. J.; Mastria, E. M.; Eniola-Adefeso, O., The margination propensity of ellipsoidal

Zhang, H.; Ahmadi, G.; Fan, F.-G.; McLaughlin, J. B., Ellipsoidal particles transport and deposition in

Drescher, K.; Dunkel, J.; Cisneros, L. H.; Ganguly, S.; Goldstein, R. E., Fluid dynamics and noise in

Wang, Y.; Hammes, F.; Duggelin, M.; Egli, T., Influence of size, shape, and flexibility on bacterial passage

28 ACS Paragon Plus Environment

Page 28 of 45

Page 29 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

664

40.

665

bacteria in porous media. Environ. Sci. Technol. 1995, 29 (7), 1737-1740.

666

41.

667

1922, 102 (715), 161-179.

668

42.

669

application to aerosol transport in microgravity. J. Fluid Mech. 1998, 371, 59-79.

670

43.

671

laminar flows. Nanotechnology 2009, 20 (49), 495101.

672

44.

673

Phys. Fluids 1998, 10 (1), 86-100.

674

45.

675

channel. J. Fluid Mech. 1994, 271, 1-16.

676

46.

677

colloid transport in geochemically heterogeneous porous media. Environ. Sci. Technol. 2000, 34 (11), 2143-2148.

678

47.

679

Modeling and measurements. Environ. Sci. Technol. 1996, 30 (11), 3284-3293.

680

48.

681

velocity, and grain angularity. Environ. Sci. Technol. 2006, 40 (24), 7725-7731.

682

49.

683

presence of energy barriers, and implications for inferred mechanisms from indirect observations. Water Res. 2010,

684

44 (4), 1158-1169.

685

50.

686

Technol. 1990, 24 (10), 1528-1536.

687

51.

688

bacterial deposition. Environ. Sci. Technol. 2002, 36 (2), 184-189.

689

52.

690

heterogeneity: Closing the gap between unfavorable and favorable colloid attachment conditions. Environ. Sci.

691

Technol. 2017, 51 (4), 2151-2160.

Weiss, T. H.; Mills, A. L.; Hornberger, G. M.; Herman, J. S., Effect of bacterial cell shape on transport of

Jeffrey, G. B., The motion of ellipsoidal particles immersed in a viscous fluid. Proc. R. Soc. London, Ser. A

Gavze, E.; Shapiro, M., Motion of inertial spheroidal particles in a shear flow near a solid wall with special

Lee, S.-Y.; Ferrari, M.; Decuzzi, P., Shaping nano-/micro-particles for enhanced vascular interaction in

Broday, D.; Fichman, M.; Shapiro, M.; Gutfinger, C., Motion of spheroidal particles in vertical shear flows.

Huang, P. Y.; Feng, J.; Joseph, D. D., The turning couples on an elliptic particle settling in a vertical

Elimelech, M.; Nagai, M.; Ko, C.-H.; Ryan, J. N., Relative insignificance of mineral grain zeta potential to

Johnson, P. R.; Sun, N.; Elimelech, M., Colloid transport in geochemically heterogeneous porous media:

Tong, M.; Johnson, W. P., Excess colloid retention in porous media as a function of colloid size, fluid

Johnson, W. P.; Pazmino, E.; Ma, H., Direct observations of colloid retention in granular media in the

Elimelech, M.; O'Melia, C. R., Kinetics of deposition of colloidal particles in porous media. Environ. Sci.

Shellenberger, K.; Logan, B. E., Effect of molecular scale roughness of glass beads on colloidal and

Rasmuson, A.; Pazmino, E.; Assemi, S.; Johnson, W. P., Contribution of nano-to microscale roughness to

29 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

692

53.

693

heterogeneous surfaces. J. Colloid Interface Sci. 2011, 353 (1), 87-97.

694

54.

695

between a sphere and a nano-patterned plate. Colloid Surf A Physicochem Eng Asp. 2008, 318 (1-3), 45-52.

696

55.

697

1998, 14 (12), 3365-75.

698

56.

699

micron-scale adhesion: Crossover from heterogeneity-dominated to mean-field behavior. J. Colloid Interface Sci.

700

2009, 337 (2), 396-407.

701

57.

702

with montmorillonite. Applied Clay Science 2006, 34 (1-4), 105-124.

703

58.

704

Surface ionization state and nanoscale chemical composition of UV-irradiated poly (dimethylsiloxane) probed by

705

chemical force microscopy, force titration, and electrokinetic measurements. Langmuir 2007, 23 (10), 5430-5438.

706

59.

707

heterogeneity in bacterial surface hydrophobicity. Langmuir 2008, 24 (9), 4944-4951.

708

60.

709

Controlled Capture in Shear Flow. ACS applied materials & interfaces 2018.

710

61.

711

(9), 2833-2838.

712

62.

713

Boundaries can steer active Janus spheres. Nat Commun. 2015, 6, 8999-Article No.: 8999.

714

63.

715

bacteria through sand columns at low flow rates. Environ. Sci. Technol. 1998, 32 (14), 2100-2105.

716

64.

717

reuse of wastewater. Environ. Sci. Technol. 2001, 35 (9), 1798-1805.

718

65.

719

Bioeng. 1982, 24 (11), 2527-2537.

Bendersky, M.; Davis, J. M., DLVO interaction of colloidal particles with topographically and chemically

Martines, E.; Csaderova, L.; Morgan, H.; Curtis, A. S. G.; Riehle, M. O., DLVO interaction energy

Bhattacharjee, S.; Chun-Han, K.; Elimelech, M., DLVO interaction between rough surfaces. Langmuir

Duffadar, R.; Kalasin, S.; Davis, J. M.; Santore, M. M., The impact of nanoscale chemical features on

Tombácz, E.; Szekeres, M., Surface charge heterogeneity of kaolinite in aqueous suspension in comparison

Song, J.; Duval, J. F.; Cohen Stuart, M. A.; Hillborg, H.; Gunst, U.; Arlinghaus, H. F.; Vancso, G. J.,

Dorobantu, L. S.; Bhattacharjee, S.; Foght, J. M.; Gray, M. R., Atomic force microscopy measurement of

Shave, M. K.; Kalasin, S.; Ying, E.; Santore, M. M., Nanoscale Functionalized Particles with Rotation-

Gao, Y.; Yu, Y., Macrophage Uptake of Janus Particles Depends upon Janus Balance. Langmuir 2015, 31

Das, S.; Garg, A.; Campbell, A. I.; Howse, J.; Sen, A.; Velegol, D.; Golestanian, R.; Ebbens, S. J.,

Simoni, S. F.; Harms, H.; Bosma, T. N. P.; Zehnder, A. J. B., Population heterogeneity affects transport of

Redman, J. A.; Grant, S. B.; Olson, T. M.; Estes, M. K., Pathogen filtration, heterogeneity, and the potable

Powell, M. S.; Slater, N. K., Removal rates of bacterial cells from glass surfaces by fluid shear. Biotechnol.

30 ACS Paragon Plus Environment

Page 30 of 45

Page 31 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

720

66.

721

Escherichia coli to polystyrene particles. Appl. Environ. Microbiol. 2003, 69 (11), 6515-6519.

722

67.

723

169 (12), 5877-5879.

724

68.

725

Science 2002, 296 (5576), 2229-2232.

726

69.

727

pneumoniae to glass surfaces. J. Bacteriol. 1983, 153 (1), 1-5.

728

70.

729

aeruginosa type IV pilus ATPases involved in twitching motility. J. Bacteriol. 2005, 187 (3), 829-839.

730

71.

731

hemispheres-in-cell model. Langmuir 2010, 26 (3), 1680-1687.

732

72.

733

arguments used to infer straining as the mechanism of colloid retention in porous media. Environ. Sci. Technol. 2011,

734

45 (9), 3831-3832.

735

73.

736

retention in porous media in the absence of energy barriers. Environ. Sci. Technol. 2011, 45 (19), 8306-8312.

737

74.

738

ellipsoidales. Le Journal de Physique et Le Radium 1934, Serie VII, Tome V (10), 497-511.

739

75.

740

Particles. AlChE J. 1958, 4 (2), 197-201.

741

76.

742

model. AlChE J. 1976, 22 (3), 523-533.

743

77.

Goldstein, H., Classical Mechanics. 2nd Ed. ed.; Addison-Wesley: Reading, MA, 1980.

744

78.

Hughes, P. C., Spacecraft Attitude Dynamics. Wiley: New York, 1986.

745

79.

Zhang, H.; Ahmadi, G.; Asgharian, B., Transport and Deposition of Angular Fibers in Turbulent

746

Channel Flows. Aerosol Sci. Technol. 2007, 41, 529-548.

Jones, J. F.; Feick, J. D.; Imoudu, D.; Chukwumah, N.; Vigeant, M.; Velegol, D., Oriented adhesion of

Larkin, J. M.; Nelson, R., Mechanism of attachment of swarm cells of Thiothrix nivea. J. Bacteriol. 1987,

Hogan, D. A.; Kolter, R., Pseudomonas-Candida interactions: an ecological role for virulence factors.

Feldner, J.; Bredt, W.; Kahane, I., Influence of cell shape and surface charge on attachment of Mycoplasma

Chiang, P.; Habash, M.; Burrows, L. L., Disparate subcellular localization patterns of Pseudomonas

Ma, H.; Johnson, W. P., Colloid retention in porous media of various porosities: Predictions by the

Johnson, W. P.; Ma, H.; Pazmino, E., Straining credibility: a general comment regarding common

Ma, H.; Pazmino, E. F.; Johnson, W. P., Gravitational settling effects on unit cell predictions of colloidal

Perrin, P. F., Mouvement Brownien d'un ellipsoide (I). Dispersion dielectrique pour des molecules

Happel, J., Visous Flow in Multiparticle Systems: Slow Motion of Fluids Relative to Beds of Spherical

Rajagopalan, R.; Tien, C., Trajectory analysis of deep-bed filtration with the sphere-in-cell porous media

31 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

747

80.

748

patchy surfaces with nanoscale electrostatic heterogeneity. J. Colloid Interface Sci. 2008, 326 (1), 18-27.

749

81.

750

flat surface. Colloids Surf. Physicochem. Eng. Aspects 2000, 165 (1-3), 143-156.

751

82.

752

London, Ser. A (Mathematical and Physical Sciences) 1971, 324 (1558), 301-13.

753

83.

754

Fluid Mech. 2007, 588, 437-462.

755

84.

756

Rotational Brownian Motion. Phys. Rev. Lett. 2009, 103 (7).

757

85.

758

Accumulation of swimming bacteria near a solid surface. Physical Review E 2011, 84 (4).

759

86.

760

region. Exp. Fluids 2016, 57 (2).

761

87.

762

Nature 1972, 239, 500-504.

763

88.

764

rods: The shape of gold nanoparticles influences aggregation and deposition behavior. Chemosphere 2013, 91 (1),

765

93-98.

766

89.

767

direct numerical simulation of colloids. Langmuir 2005, 21 (6), 2173-2184.

768

90.

769

physicochemical filtration in saturated porous media. Environ. Sci. Technol. 2004, 38 (2), 529-536.

Duffadar, R. D.; Davis, J. M., Dynamic adhesion behavior of micrometer-scale particles flowing over

Bhattacharjee, S.; Chen, J. Y.; Elimelech, M., DLVO interaction energy between spheroidal particles and a

Johnson, K. L.; Kendall, K.; Roberts, A. D., Surface energy and the contact of elastic solids. Proc. R. Soc.

Ishikawa, T.; Pedley, T. J., Diffusion of swimming model micro-organisms in a semi-dilute suspension. J.

Li, G. L.; Tang, J. X., Accumulation of Microswimmers near a Surface Mediated by Collision and

Li, G. L.; Bensson, J.; Nisimova, L.; Munger, D.; Mahautmr, P.; Tang, J. X.; Maxey, M. R.; Brun, Y. V.,

Lee, S. J.; Go, T.; Byeon, H., Three-dimensional swimming motility of microorganism in the near-wall

Berg, H. C.; Brown, D. A., Chemotaxis in Escherichia Coli analysed by Three-Dimensional Tracking.

Afrooz, A. N.; Sivalapalan, S. T.; Murphy, C. J.; Hussain, S. M.; Schlager, J. J.; Saleh, N. B., Spheres vs.

Nelson, K. E.; Ginn, T. R., Colloid filtration theory and the happel sphere-in-cell model revisited with

Tufenkji, N.; Elimelech, M., Correlation equation for predicting single-collector efficiency in

770

32 ACS Paragon Plus Environment

Page 32 of 45

Page 33 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

771

Langmuir

Table 1. Parameters Used in Lagrangian Trajectory Simulation parameter

value

collector diameter, 𝑎𝑔

390 µm

porosity, 𝜀

0.36

pore water velocity, v

5 m/day

particle density, 𝜌𝑝

1055 kg/m3

fluid density, 𝜌𝑓

998 kg/m3

dynamic fluid viscosity, 𝜇

9.98 × 10 ―4 kg ∙ m/s

Hamaker constant, 𝐴𝐻

3.84 × 10 ―21 𝐉

Ionic strength, IS

1 mM

colloid zeta potential, 𝜉𝑝𝑜

-20 mV

collector zeta potential, 𝜉𝑐

-53.5 mV

colloidal heterogeneity potential, 𝜉𝑝

20 mV

absolute temperature, T

298.2 K

time step, ∆𝑡

1-100 MRT

aspect ratio, 𝛽

1-6

shear rate at the wall

1.96 s ―1

Peclet number, Pe

0.1~100

Reynolds number, Re

< 0.001

772

33 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

773

Figure Captions.

774

Figure 1. Schematic of the Happel model. Particles were injected through the particle injection plane, and

775

only particles injected through the limiting injection area that were able to attach.

776

Figure 2. Coordinate systems and Euler angles used to describe particle translation and rotation.

777

Figure 3. Heterogeneity pattern designs over particle surfaces: (a) on one end, (b) on two ends, and (c) on

778

the middle. Particle surface heterogeneous coverage was varied from 0-100% of the total colloid surface

779

area.

780

Figure 4. Schematic of the SEI method. The local unit normal (n) of an ellipsoid surface element and its

781

local distance (h) to the surface were required to compute the interactions between this element and the

782

collector. The closest separation distance between the ellipsoid and the collector was represented by H.

783

Blue patches represented the heterogeneity location over particle surface.

784

Figure 5. Simulated attachment efficiencies for heterogeneity located on (a) one end versus (b) two ends

785

as a function of surface heterogeneity coverage (SCOV) and aspect ratio for 1 µm in diameter particles.

786

AR1, AR2, AR3 and AR6 represented ellipsoids with an aspect ratio of 1 (red circle), 2 (blue triangle), 3

787

(orange cross), and 6 (purple square), respectively. Simulation conditions were provided in Table 1. The

788

statistical errors for our simulation data depended on the total numbers of simulated particle trajectories

789

and the numbers of particles attached and typically were less than ±0.05 in α.

790

Figure 6. Simulated attachment efficiencies for heterogeneity located on the ends (a, c, e) versus on the

791

middle band (b, d, f) as a function of colloid surface heterogeneity coverage (SCOV) and aspect ratio for

792

representative particle sizes: (a-b) 200 nm; (c-d) 1 µm; (e-f) 6.6 µm. AR1, AR2, AR3 and AR6

793

represented ellipsoids with an aspect ratio of 1 (red circle), 2 (blue triangle), 3 (orange cross), and 6

794

(purple square), respectively. Simulation conditions were provided in Table 1. Note that the x-axis scales

795

for (a, c, e) and (b, d, f) were different. The statistical errors were the same as described in Figure 5.

34 ACS Paragon Plus Environment

Page 34 of 45

Page 35 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

796

Figure 7. Integrated DLVO interaction energy profiles from the SEI method as a function of colloid-

797

collector separation distance when colloid surface heterogeneity was located (a) on the ends versus (b) on

798

the middle band for 6.6 µm (in diameter) ellipsoids of different aspect ratios: AR1 - sphere (purple), AR2

799

- aspect ratio 2 rod (red), AR3 - aspect ratio 3 rod (blue) and AR6 - aspect ratio 6 rod (green). Blue

800

patches represented colloid surface charge heterogeneity. These energy profiles were generated when the

801

heterogeneous patches were facing the collector with an end-on orientation (a, 0.004% coverage) or side-

802

on orientation (b, 2.6% coverage). Other parameter values were identical as Table 1 unless specified

803

otherwise.

804

Figure 8. Simulated single collector efficiencies () as a function of particle size under unfavorable

805

conditions with representative heterogeneity surface coverage (SCOV): (a) spheres colloids - 1.5% (red

806

circle) , 3.5% ( green square), 30% (orange triangle), 100% (blue diamond), and representative theory

807

predictions from MHJ prediction71 (blue line); (b) aspect ratio 6 rods -