Marine Phages As Tracers - ACS Publications - American Chemical

Oct 7, 2016 - Bärbel Kiesel,. †. René Kallies,. †. Hauke Harms,. †,‡. Antonis Chatzinotas,. †,‡ and Lukas Y. Wick*,†. †. Helmholtz C...
0 downloads 0 Views 658KB Size
Subscriber access provided by CORNELL UNIVERSITY LIBRARY

Article

MARINE PHAGES AS TRACERS: EFFECTS OF SIZE, MORPHOLOGY AND PHYSICO-CHEMICAL SURFACE PROPERTIES ON TRANSPORT IN A POROUS MEDIUM Nawras Ghanem, Bärbel Kiesel, Rene Kallies, Hauke Harms, Antonis Chatzinotas, and Lukas Y. Wick Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b04236 • Publication Date (Web): 07 Oct 2016 Downloaded from http://pubs.acs.org on October 18, 2016

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

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

Page 1 of 27

Environmental Science & Technology

1

MARINE PHAGES AS TRACERS: EFFECTS OF SIZE, MORPHOLOGY AND PHYSICO-CHEMICAL

2

SURFACE PROPERTIES ON TRANSPORT IN A POROUS MEDIUM

3 4 5 6

Nawras Ghanem1, Bärbel Kiesel1, René Kallies1, Hauke Harms1,2, Antonis Chatzinotas1,2 and Lukas Y. Wick1*

7 8 9

1

10 11

2

Helmholtz Centre for Environmental Research - UFZ, Department of Environmental Microbiology, Permoserstraße 15, 04318 Leipzig, Germany.

German Centre for Integrative Biodiversity Research (iDiv) Halle-Jena-Leipzig, Deutscher Platz 5e, 04103 Leipzig, Germany

12 13

Running title: Marine phages as microbial tracers

14 15 16

Intended for: Environmental Science and Technology

17 18 19 20 21 22 23 24

*

Corresponding author: Helmholtz Centre for Environmental Research - UFZ. Department of Environmental Microbiology; Permoserstrasse 15; 04318 Leipzig, Germany. phone: +49 341 235 1316, fax: +49 341 235 45 1316, e-mail: [email protected].

25

ACS Paragon Plus Environment

1

Environmental Science & Technology

Page 2 of 27

26

Abstract

27

Although several studies examined the transport of viruses in the terrestrial systems only few studies

28

exist on marine phages (i.e. non-terrestrial viruses infecting marine host bacteria) as sensitively

29

detectable microbial tracers for subsurface colloid transport and water flow. Here, we systematically

30

quantified and compared for the first time the effects of size, morphology and physico-chemical

31

surface properties of six marine phages and two coliphages (MS2, T4) during transport in sand-filled

32

percolated columns. Phage-sand interactions were described by colloidal filtration theory and the

33

extended Derjaguin-Landau-Verwey-Overbeek approach (XDLVO), respectively. The phages

34

belonged to different families and comprised four phages never used in transport studies (i.e. PSA-

35

HM1, PSA-HP1, PSA-HS2 and H3/49). Phage transport was influenced by size, morphology and

36

hydrophobicity in an approximate order of size > hydrophobicity ≥ morphology. Two phages PSA-

37

HP1, PSA-HS2 (Podoviridae and Siphoviridae) exhibited similar mass recovery as commonly used

38

coliphage MS2 and were sevenfold better transported than known marine phage vB_PSPS-H40/1.

39

Differing properties of the marine phages may be used to trace transport of indigenous viruses,

40

natural colloids or anthropogenic nanomaterials and, hence, contribute to better risk analysis. Our

41

results underpin the potential role of marine phages as microbial tracer for transport of colloidal

42

particles and water flow.

43 44

One sentence brief. Differences in the properties of marine phages influence their transport in

45

porous media and underpin their potential as microbial tracers for reactive transport of colloidal

46

particles and water flow.

47

KEYWORDS: Marine phage, transport, colloid, bacteria, microbial tracer, XDLVO.

ACS Paragon Plus Environment

2

Page 3 of 27

48

Environmental Science & Technology

TOC / Abstract Art

49

50 51

ACS Paragon Plus Environment

3

Environmental Science & Technology

Page 4 of 27

52

INTRODUCTION

53

Bacteriophages (or short: phages) are viruses that infect bacteria. In contrast to human- and animal

54

health-related microorganisms, viruses infecting environmental microbes in particular in terrestrial

55

system are less studied.1,2,3 Next to their impact on their hosts and on manifold ecosystem

56

functions,4,5,6 their suitability as markers for water transport in the subsurface has been considered.7,8

57

In 1980 Keswick and Gerba suggested to use bacteriophages as tracers because of similar movement

58

and survival as animal viruses.9 Many phages (especially coli-phages such as MS2) have been tested

59

since then either as potential surrogates for viral pathogens10,11 or as a tool to trace the flow of water

60

in hydrology.12,13 An awkwardness of this usage was their differing adsorption behavior from the

61

human adenoviruses (hAdVs)14 or the Norwalk virus15 and the isolation of both, coli-phages and

62

their hosts from groundwater.9,16 Moreover, strains of E. coli are listed as pathogenic bacteria of

63

concern.17 Despite a large number of studies examining the transport of phages in the terrestrial

64

subsurface, only very few studies evaluated marine phages, most often phage H40/1,18,19,20,21 as

65

biological tracers.22 Marine phages are highly suitable for transport studies because they are virtually

66

absent in the terrestrial ecosystem23 and non-pathogenic. Additionally they allow for highly sensitive

67

detection22 due to the possibility to apply as many as 1015 phages (~1 g) in tracer experiments and to

68

detect as little as one or two phages per mL of recovered water,22 via its specific interaction with the

69

host bacterium.

70

First studies by Gerba and Goyal proposed that viruses can be grouped by their adhesion behavior as

71

mimic for their ability to interact with soil.10 Follow-up studies however rather focused on the effect

72

of environmental rather than viral factors on virus transport in soil (e.g. heterogeneity of the soil24,25;

73

soil textures and conditions26,27,28,29 or chemical factors).30,31,32,33,34,35 Data on the effect of virus

74

properties on their transport, however, are still scarce.36 A few studies aimed at predicting transport

75

in relation to physical viral properties such as the isoelectric point,37 a combination of isoelectric

ACS Paragon Plus Environment

4

Page 5 of 27

Environmental Science & Technology

76

point and size,38 or the different surface charge, polarity and topography of viral protein structures.39

77

Yet no such study on marine phages is known. In the frame of the Collaborative Research Centre

78

AquaDiva (http://www.aquadiva.uni-jena.de/) our study aimed at selecting and characterizing

79

different marine phages and at studying their transport in porous media.40 Although inherently not

80

being as inert and conservative as ideal chemical tracers may be, we here consider marine phages as

81

sensitively detectable microbial tracers17 for colloidal transport and advective water flow. We

82

hypothesized that size, morphology and physico-chemical surface properties control their transport

83

and, hence, their appropriateness as surrogates for transport of colloidal particles. For this purpose,

84

we systematically investigated the transport of six marine phages in addition with two commonly

85

applied coli-phages in transport studies (T4 and MS2) using sand-filled laboratory columns. The

86

selected phages belonged to four different virus families (Myoviridae, Siphoviridae, Podoviridae and

87

Leviviridae) and thus differed in their size, morphology and physicochemical properties. The set of

88

phages included four marine phages that never had been examined for their suitability as tracers

89

(PSA-HM1, PSA-HP1, PSA-HS2 and H3/49). The phages’ physico-chemical properties, were

90

assessed by dynamic light scattering and water contact angle analysis, size and morphology

91

information were obtained from literature data, while phage transport and phage-sand interactions

92

were described by colloidal filtration theory and the extended Derjaguin-Landau-Verwey-Overbeek

93

approach (XDLVO), respectively.

94 95

EXPERIMENTAL PROCEDURES

96

Material and Methods

97

Phages and phage assay: Six lytic marine phages were selected (Table 1) and obtained together

98

with their host strains from different sources: Phages PSA-HP1, PSA-HS2 and PSA-HM1 were

99

kindly provided by Dr. B. M. Duhaime (University of Michigan, USA); phage H3/49 by Dr. E.

ACS Paragon Plus Environment

5

Environmental Science & Technology

Page 6 of 27

100

Roine (University of Helsinki, Finland); phage VB_PSPS-H40/1 and VB_PSPS-H6/1 by Dr. J. Zopfi

101

(University of Basel, Switzerland). In addition to marine phages two well-characterized, non-marine

102

phages (phage T4 and MS2) were used as controls.41,42 Phages T4 and MS2 and their host E. coli

103

were purchased from Deutsche Sammlung von Mikroorganismen und Zellkulturen GmbH (DSMZ,

104

Germany). Except for Schewanella baltica that grew in 100% nutrient broth at 10°C, all marine

105

bacterial strains were grown at room temperature using dilute (50%) 2216E medium.43Phage titration

106

was performed using a modified spotting plaque assay technique. Deviating from Adams44 the top

107

layer of the counting agar plates contained solely a bacterial suspension, to which phage lysates (5

108

µL) were spotted in triplicate. No significant differences to results obtained with the technique by

109

Adams were found in the range of 7 – 22 (±3) plaque forming units (PFU) 5 µL-1. All plates

110

containing the phage host pairs were incubated overnight at room temperature (except 10°C for

111

H3/49 and 37°C for phages MS2 and T4).

112

Phage propagation and purification: Phages were propagated on their hosts using the double agar-

113

layer technique and high phage concentrations to reach completely lysis plates. Phage particles were

114

gathered after confluent lysis of bacteria by adding 5 mL of SM buffer (100 mM NaCl, 8 mM

115

MgSO4 x 7H2O, 50 mM Tris-HCl; pH = 7) and a few drops of chloroform with mild shaking for 2 h

116

at room temperature. The supernatant was recovered and cell debris was removed by centrifugation

117

at 10,000g for 15 min. The supernatant was mixed with an equal volume of chloroform and

118

centrifuged at 5,000g for 5 min to further purify the phage particles. The phage particle-containing

119

supernatant was filtrated through 0.22 µm polyvinylidene fluoride (PVDF) CHROMAFIL®

120

membrane filters45 and the phage suspension was stored at 4°C.

121

Genome sequencing of phages: Genomes of all marine phages were sequenced in order to either

122

confirm characteristics of phages described recently i.e. PSA-HM1, PSA-HP1, PSA-HS2 and H3/49

123

(in particular to exclude nucleotide changes in the genomes that may occur during phage

124

propagation) or to characterize phages whose genome was still unknown i.e. vB_PSPS-H40/1 and

ACS Paragon Plus Environment

6

Page 7 of 27

Environmental Science & Technology

125

vB_PSPS-H6/1. DNA was extracted using the protocol of Thurber et al.46 and shotgun genome

126

sequencing was performed on an Illumina MiSeq system using standard Illumina protocols for

127

generation of paired 150-bp reads. Resulting sequencing reads were de novo assembled using

128

Geneious assembler (Geneious R9, Biomatters, Auckland, New zealand) and alignments were

129

performed with ClustalW.47 Genome sequences of PSA-HM1, PSA-HP1, PSA-HS2 and H3/49 were

130

100% identical with the corresponding sequences available at GenBank (KF302034, KF302037,

131

KF302036 and KJ018214). Sequences of vB_PSPS-H40/1 and vB_PSPS-H6/1 (GenBank acc. Nos.

132

KU747973 and KX257490) confirmed these two phages belong to the Siphoviridae family.

133

Characterization of surface properties. The zeta-potentials (ζ) were approximated from the

134

electrophoretic mobility measured by Doppler electrophoretic light scattering analysis (Zetamaster,

135

Malvern Instruments, Malvern, UK) of phage suspension in 100 mM phosphate buffer (PB; 0.87 g L-

136

1

137

contact angles of water θw, formamide θf and methylene iodide θmi were measured using a DSA 100

138

drop-shape analysis system (Krüss GmbH, Hamburg, Germany) as described earlier48. Phage lawns

139

were prepared by deposition of phage lysate of high concentration (i.e. 109 - 1010 PFU mL-1) onto

140

Whatmann inorganic AnoDisc filter (0.02 µm, diameter: 25 mm). The filters were then dried for 2 h

141

and contact angles were measured by applying 3–5 droplets on a filter using at least two filters per

142

liquid tested (i.e. n ≥ 6 measurements).

143

Column deposition experiments: All experiments were performed in triplicates at 25°C in heat-

144

sterilized vertical percolation columns (i.d.: 1 cm; L.: 10 cm). These were made of borosilicate glass

145

and confined at the bottom by a glass frit (pore size: 100 – 160 µm). The columns were wet packed

146

in 100 mM PB with clean, sterile commercial quartz sand (Euroquarz-group) with a d50 median

147

particle size of 0.31 mm, a porosity of ≈ 0.4 (estimated gravimetrically), and a total pore volume

148

(PV) of 2.86 mL. Prior to the filling, the sand was washed with deionized water, heated at 400°C for

149

3 h and then allowed to cool down under sterile conditions. Unfilled columns showed no retention of

K2HPO4, 0.68 g L-1 KH2PO4; pH = 7). In order to calculate the surface free energies of the phages,

ACS Paragon Plus Environment

7

Environmental Science & Technology

Page 8 of 27

150

any of the phages. Sand-filled columns were allowed to equilibrate by flushing at least 8 PV of clean

151

PB. Then homogenized phage suspensions (106-107 PFU mL-1 in 100 mM PB) were circulated with a

152

peristaltic pump at a hydraulic flow rate of 0.7 × 10-4 m s-1 (19.8 mL h-1) from the top to the bottom

153

of the columns, i.e. at flow conditions lower than in coarse gravel aquifers (23 - 44 m day-1).32 To

154

characterize the flow conditions prevailing in the columns and as reference to evaluate the arrival

155

time of studied phages. Bromide tracer tests were conducted in triplicate using 0.05 g L-1 KBr.

156

Bromide concentrations in the outflow of the columns were quantified by ion chromatography

157

(Dionex ics-2000; Dionex Corporation, USA). Phage transport and deposition under saturated

158

conditions were determined by comparing the PFU concentrations of the influent and effluent as

159

described earlier:27,49 about 8 PV of phage suspensions in PB were applied to the columns, the

160

columns then rinsed with ≈ 4 PV with phage-free PB to assess the reversibility of phage adhesion

161

and, finally, with ≈ 4 PV of pure water (Milli-Q) to study the effect of electrostatic repulsion on

162

phage adhesion. The colloidal stability of the phage suspensions over 3 h (i.e. the duration of column

163

transport experiments) was quantified in independent batch experiments using identical buffer

164

composition and temperature. PFU in samples taken over time was followed assuming that any

165

aggregate forms only a single plaque. Colloidal stability was calculated as PFU relative to the initial

166

number. Phage lysate was prepared in concentrations of 108 - 1010 PFU mL-1 using PB buffer and

167

stored in glass vials at room temperature under static condition (i.e. no agitation). Samples were

168

taken at the beginning of the experiment and after 3 h to evaluate the colloidal stability and viability

169

of the phages.

170 171

Theory

172

Calculation of phage collision efficiency ( ), deposition rate coefficient (Kd) and mass recovery

173

(M). Filtration of phages was approximated using the colloidal filtration theory50 in columns

ACS Paragon Plus Environment

8

Page 9 of 27

Environmental Science & Technology

174

assuming spherical phages with properties as given in Table 1. For a detailed explanation of the

175

calculations, the reader is referred to Tufenkji & Elimelech 2004,51 the filtration equation of

176

Rajagopalan and Tien52 and the SI, respectively. Therein, the  was defined as the ratio of the

177

experimental single-collector removal efficiency ( ; f. eq. S1) to the predicted single-collector

178

contact efficiency ( ; eq. S2)51. The  is a function of four dimensionless groups: the

179

aspect ratio NR, the Peclet number NPe, the van der Waals number NvdW, and the gravitational number

180

Ngr. 53,54 The deposition rate coefficient (Kd, h-1 ) relates to the attachment efficiency ( )55,51 by eq. 1

181

 

 = ()

(1)

! ɳ#

182

where vP is the pore water velocity (cm min-1), L the column length in cm, ac the mean radius of the

183

sand (0.29 mm), $ the porosity of the column packing, and ɳ0 the single collector efficiency

184

determined as given by Tufenkji and Elimelech, 2004.51 The Kd is calculated by eq. 2

185

%& =

!

'

+

ln * +# -

(2)

,

186

where C0 is the initial concentration of the phages (PFU mL-1) and Ct is the effluent concentration

187

calculated as the average of > 10 steady state data points of the phages’ breakthrough curves (PFU,

188

mL-1) as represented in present study.55 The mass recovery (M) was calculated as the ratio of phages

189

in the effluent relative to that of the influent as inferred from the difference of inlet and outlet phage

190

concentration as described by eq. 3

191

∑ + ∆

. = ∑ + , ∆ ∗ 100

(3)

#

192

Prediction of the effect of XDLVO phage-sand interaction energy. Phages were assumed to

193

follow the principles of colloid chemistry.50 As Lewis acid-base interactions are known to play an

194

important role in the total interaction energy between phages and surfaces,56 the extended XDLVO

195

interaction energies between phages and sand particles. Therein the total interaction energy 345678

ACS Paragon Plus Environment

9

Environmental Science & Technology

Page 10 of 27

196

is the sum of the electrostatic repulsion (3956 ), the Lifshitz-van der Waals (36: ) and the acid-base

197

(3;< ) interaction energy. While 3956 and 36: are functions of the separation distance h (nm)

198

between two surfaces57,58, 3;< compares the energy status between attached and nonattached

199

situations (eq. 4):

200

345678 (ℎ) = 3;< +3956 (ℎ) + 36: (ℎ)

(4)

201

As the phages studied are far smaller than sand particles we applied a sphere-plate geometry.56 3?@'

202

3'A and 3BC were calculated as described by Chrysikopoulos and Syngouna, 2012.56 For 3BC and

203

3'A calculations and in contrast to Chrysikopoulos and Syngouna, 2012 we used experimentally

204

determined Gibbs free energies and Hamaker constants of all phages, respectively. The Gibbs free

205

energies (Table S1) and Hamaker constants were calculated using the surface free energies of studied

206

phages applying eqs. S8 and S4, whereas surface free energy calculations were based on measured

207

contact angles (θ) of phages using water, formamide and methylene iodide as liquids and the Young

208

equation.59

209 210

RESULTS

211

Physico-chemical surface properties, size and morphology of phages

212

Surface charge and hydrophobicity of the six marine phages and the two coli-phages were

213

determined by their zeta-potentials (ζ) and water contact angles, respectively (Fig. S1A). A broad

214

contact angle range from 38o to 95o was observed (Table 1). The phages could be clustered into three

215

groups: (i) poorly hydrophobic phages (≈ 40o: PSA-HM1, PSA-HP1 and PSA-HS2), (ii) moderately

216

hydrophobic phages (≈ 53-61o: vB_PSPS-H40/1, vB_PspS-H6/1 and H3/49), and (iii) hydrophobic

217

phages (≈ 84-95o: T4 and MS2). In the transport buffer (pH = 7) all phages had slightly negative

218

surface charges as inferred from similar zeta potentials (ζ = -11 – -18 mV). Phage morphology and

ACS Paragon Plus Environment

10

Page 11 of 27

Environmental Science & Technology

219

size varied according to previous own work60,61 and literature data (cf. Table 1). Belonging to

220

different families, the phages represented four different morphologies: contractile tail (Myoviridae),

221

noncontractile long tail (Siphoviridae), noncontractile short tail (Podoviridae), and icosahedral

222

without tail (Leviviridae). Phage sizes ranged from 25 to 200 nm (Table 1).

223

Phage transport in saturated percolated columns

224

In order to determine transport and deposition characteristics of the phages, experiments were

225

conducted using sand-packed columns under continuous, saturated flow conditions. Each experiment

226

was divided in three phases: after flushing of 7 PV of dilute phage suspensions in PB, 4 PV of phage-

227

free PB followed by a switch in the column inflow to deionized water were applied. The last two

228

phases served to assess the reversibility of phage adhesion and to study if electrostatic repulsion

229

would lead to phage detachment, respectively. Plots of relative phage densities in the outflow (C/C0)

230

against time (normalized to the number of PV exchanged) showed about 0.5 of the C0 at 1 PV and

231

were similar to the behavior of KBr tracer (Fig. S2). One exception was for the phage H3/49 which

232

exhibited rapid attachment to sand and reached quasi-equilibrium (Fig. S2). After about 2 PV a

233

leveling off of the breakthrough curves at plateaus of different heights were observed (Fig. S2).

234

Calculated maximal coverages of the sand surface of < 0.2% were calculated and excluding

235

substantial changes of the collector surface properties. Subsequent flushing with phage-free PB

236

resulted in poor tailing pointing at irreversible attachment of the phages. Changing the inflow from

237

PB to deionized water led 0 - 6% of the attached phages to flush out (Table 1); i.e. only partially

238

reverted the deposition of the phages despite of a clearly increased electrostatic repulsion between

239

the sand and the phage surfaces.

240

efficiencies (eq. 1) and deposition rate coefficients (eq. 2) of all phages. It further shows the

241

following trend of the different phages’ mass recoveries: MS2 > PSA-HP1 > PSA-HS2 > PSA-HM1

242

> vB_PspS-H6/1 > vB_PspS-H40/1 > T4 > H3/49. Conspicuously, the smallest phages MS2 and

243

PSA-HP1 exhibited the highest mass recovery regardless of their hydrophilicity or hydrophobicity.

62,63

Table 1 summarizes the mass recoveries (eq. 3), collision

ACS Paragon Plus Environment

11

Environmental Science & Technology

Page 12 of 27

244

Deposition rate coefficients and calculated collision efficiencies ranged from 0.21 to 0.02 h-1 and

245

0.34 to 0.01, respectively. Phages with fast deposition rates and high collision affinities (i.e.

246

vB_PspS-H40/1, T4 and H3/49) thereby exhibited low mass recoveries (Table 1). Hydrophobic and

247

moderately hydrophobic phages exhibited lower collision efficiencies (Fig. S2B) and deposition rate

248

coefficients (Fig. S2Dand thus a lower mass recovery (Fig. 1) than poorly hydrophobic phages. One

249

exception such latter trend was phage MS2 where the small size of MS2 seems to determine its

250

transport behavior. The mass recoveries of the poorly hydrophobic phages varied by the order of:

251

PSA-HP1 (Podoviridae) > PSA-HS2 (Siphoviridae) > PSA-HM1 (Myoviridae) (Fig. 1). Data from

252

control experiments revealed high colloidal stability and no loss of phage viability in the transport

253

buffer used (Table 1).

254

Calculation of the phage-sand interaction energies

255

Phage-sand interaction energy profiles were calculated using the extended DLVO (XDLVO) theory

256

(Fig. S3, Table 2) based on the sphere-plate model.56 For the experimental conditions applied, the

257

interaction energy profiles indicated repulsive interactions for MS2, PSA-HP1, PSA-HS2 and PSA-

258

HM1 (Φmax1 = 0.003–0.37 kbT) and attractive interactions for vB_PspS-H6/1, vB_PspS-H40/1, H3/49

259

and T4 phages (Φmax1 = -0.046 – -0.011 kbT) (Fig. S3, Table 2). The two phages with the highest

260

mass recovery (PSA-HP1, MS) exhibited the highest Φmax1 (53). Although the XDLVO calculations

261

indicated the existence of secondary energy minima (Φmin2) at h ≈ 12 nm, their depths for all phages

262

were shallow ( hydrophobicity ≥ morphology. Likely due to size exclusion effects69,70 smaller

282

phages (≤ 77 nm) were best transported with their transport efficiency being independent of the

283

surface hydrophobicity, as seen from similar mass recovery of the poorly hydrophobic PSA-HP1 (M

284

= 75%) and the hydrophobic MS2 phage (M = 82%). Good transport arising from low collision

285

efficiency of small phages is also in line with comparisons of differently sized phi X 174, MS2, and

286

T4 phages showing better retention of large phages in rapid sand filtration42 as well as with a report

287

of efficient transport of the MS2 phage.71,42,72 It also agrees with a previous report demonstrating that

288

phages of < ≈ 60 nm size are less retained than larger phages37 and that rod shaped particles (such as

289

long tailed phages) have higher retention and lower reversibility than spherical particles due to

290

orientation effects during transport in saturated porous media.73 Such findings confirm our results

291

that the morphology of phages of similar hydrophobicity (PSA-HP1, PSA-HS2 and PSA-HM1 or

ACS Paragon Plus Environment

13

Environmental Science & Technology

Page 14 of 27

292

vB_PspS-H6/1 and H3/49 (Fig. 1)) is highly influential. Comparison of PSA-HP1, PSA-HS2 and

293

PSA-HM1 shows that the short-tail phage PSA-HP1 is less retained than the noncontractile long tail

294

phage PSA-HS2, or the contractile tail phage PSA-HM1. Due to probable interferences of size and

295

morphology we observed the following trend of phage mass recovery: Leviviridae (icosahedral, no

296

tail) > Podoviridae (noncontractile short tail) > Siphoviridae (noncontractile long tail) ≥ Myoviridae

297

(contractile tail). This suggests that morphology controls the transport of phages as also has been

298

described for the human adenovirus (HAdV).74 Small-sized MS2 phages excluded, we found a

299

negative correlation of water contact angle and mass recovery (Fig. 1) underpinning the effect of

300

hydrophobic interactions on the deposition and removal of phages36,75,76,39; in particular at water

301

contact angles of > 65o.56

302 303

Phage-sand interaction energies and phage deposition

304

Using the XDLVO approach, we calculated the distance-dependent interaction energies between the

305

phages and the sand surface (Table 2 and Fig. S3). This model estimates the interaction energy as a

306

function of surface to surface distance h for a phage approaching a sand grain (eq. 4). It is

307

characterized by three distinct interaction energies: the primary minimum (Φmin1) as the deep energy

308

at small h from the sorbent surface, the secondary minimum (Φmin2) as the shallow energy at

309

relatively large distance, and the maximum energy barrier to attachment and detachment (Φmax1).50

310

Although phages do not have smooth surfaces and are composed of various protein macromolecules

311

and hence are far from being ideal colloidal particles, we found good agreement between calculated

312

interaction energy profiles and our experimental results which were also shown in other studies

313

under similar conditions.56 The XDLVO calculations proposed the existence of secondary energy

314

minima (Φmin2) at h ≈ 12 nm; their depths, however, were shallow (