1 Large-scale maps of variable infection efficiencies in aquatic phage-

2 host model systems

3 Karin Holmfeldt1,2, Natalie Solonenko1,a, Cristina Howard-Varona1,a, Mario Moreno1,

4 Rex R. Malmstrom3, Matthew J. Blow3, Matthew B. Sullivan1,a,b

5 1Department of Molecular and Cellular Biology, University of Arizona, Tucson, AZ,

6 USA

7 2Centre for Ecology and Evolution in Microbial Model Systems, Department of Biology

8 and Environmental Sciences, Linnaeus University, Kalmar, Sweden

9 3DOE Joint Genome Institute, Walnut Creek, CA, USA

10 Current affiliation: Departments of aMicrobiology, and bCivil, Environmental and

11 Geodetic Engineering, The Ohio State University, Columbus, OH

12 * corresponding author: Karin Holmfeldt, Center for Ecology and Evolution in Microbial

13 Model Systems, Department of Biology and Environmental Sciences, Linnaeus

14 University, SE-391 82 Kalmar, Sweden. Phone nr +46-480-447310, Fax nr +46-480-

15 447305, email [email protected].

16

17 Running title: Variable phage-host infection interactions

1 18 Summary

19 Microbes drive ecosystem functioning, and their viruses modulate these impacts through

20 mortality, gene transfer, and metabolic reprogramming. Despite the importance of virus-

21 host interactions and likely variable infection efficiencies of individual phages across

22 hosts, such variability is seldom quantified. Here we quantify infection efficiencies of 38

23 phages against 19 host strains in aquatic Cellulophaga (Bacteroidetes) phage-host model

24 systems. Binary data revealed that some phages infected only one strain while others

25 infected 17, whereas quantitative data revealed that efficiency of infection could vary 10

26 orders of magnitude, even among phages within one population. This provides a baseline

27 for understanding and modeling intra-population host range variation. Genera specific

28 host ranges were also informative. For example, the Cellulophaga Microviridae, showed

29 a markedly broader intra- host range than previously observed in Escherichia coli

30 systems. Further, one phage genus, Cba41, was examined to investigate non-heritable

31 changes in plating efficiency and burst size that depended on which host strain it most

32 recently infected. While consistent with host modification of phage DNA, no differences

33 in nucleotide sequence or DNA modifications were detected, leaving the observation

34 repeatable, but the mechanism unresolved. Overall this study highlights the importance of

35 quantitatively considering replication variations in studies of phage–host interactions.

36

37 Introduction

38 Microbes drive the planet’s biogeochemical cycles (Falkowski et al., 2008) and

39 their viruses (called phages when infecting ) play important roles ranging from

40 mortality and nutrient cycling to gene transfer and metabolic reprograming (Fuhrman,

2 41 1999; Wommack and Colwell, 2000; Weinbauer, 2004; Suttle, 2005, 2007; Rohwer et al.,

42 2009; Breitbart, 2012; Hurwitz et al., 2013; Brum and Sullivan, 2015; Hurwitz et al.,

43 2015). Our knowledge of genome-wide viral diversity, in marine systems in particular, is

44 rapidly expanding through mining microbial datasets (Roux et al., 2014, Labonte et al.,

45 2015) and assembly from large-scale viral metagenomes (Brum et al., 2015). Such

46 advances mean that the field can now look forward to advancing from measuring gene

47 marker based diversity to assessing viral impacts. These impacts ultimately depend upon

48 a virus’ ability to infect a bacterium, which is a function of myriad levels of phage–

49 bacteria interactions from recognition to overcoming bacterial defenses (reviewed in

50 Hyman and Abedon, 2010; Labrie et al., 2010; Samson et al., 2013). The first step for

51 successful infection is recognition and attachment of the phage to the host receptors. This

52 involves a specific interaction between phage receptor-binding proteins, such as tail fiber

53 proteins among tailed phages, and bacterial cell surface receptors, which can include

54 surface proteins, polysaccharides, or lipopolysaccharides (Rakhuba et al., 2010). Through

55 sequence mutations phages can expand the range of hosts that they infect (host range),

56 likely by adopting new surface interaction capabilities and counteracting bacterial

57 defense mechanisms (Pepin et al., 2008; Paterson et al., 2010; Avrani et al., 2011). Once

58 inside the cell, phage DNA must avoid cleavage by restriction endonucleases (Arber,

59 1965a) or the CRISPR-Cas system (Barrangou et al., 2007), which can work together to

60 increase bacterial resistance to phages (Dupuis et al., 2013), as well as avoid abortive

61 infection systems (Snyder, 1995). Phages resist such defenses by altering their DNA to

62 escape restriction endonucleases (e.g. by methylation; Krüger and Bickle, 1983) or to

63 escape the CRISPR-Cas (Deveau et al., 2008) and the abortive infection system (Haaber

3 64 et al., 2009) (e.g., by mutations). Thus, a phage’s realized host range is dictated by the

65 ability to recognize hosts and avoid host defenses.

66 Such host ranges ultimately also dictate how phages impact the diversity of

67 bacterial communities (Weitz et al., 2013). Among cultured phages, most infect only

68 strains from a single or closely-related bacterial species (Wichels et al., 1998; Comeau et

69 al., 2005; Holmfeldt et al., 2007; Stenholm et al., 2008), though as assays are typically

70 constrained to strains of the species used for isolation, in situ host ranges may be broader.

71 Consistent with this, some phages are known to infect across bacterial genera (Sullivan et

72 al., 2003) or classes (Jensen et al., 1998). Our perception of the phage’s host range when

73 measured in the laboratory can also be affected by the choice of assay method, which

74 commonly includes plaque assays, spot tests (small amount of viruses applied on a

75 bacterial lawn), and liquid culture–based methods (Hyman and Abedon, 2010). Host

76 range assays of larger isolated phage–host systems (>10 phages compared to >10

77 bacteria) from natural communities have mainly been investigated using spot-tests

78 (Moebus and Nattkemper, 1981; Comeau et al., 2005; Holmfeldt et al., 2007; Stenholm

79 et al., 2008). However, spot-tests often overestimate phage host ranges (Kutter, 2009;

80 Holmfeldt et al., 2014; Khan Mirzaei and Nilsson, 2015), which leave plaque assays,

81 with their quantitative capability to determine the phage’s efficiency of plating (EOP;

82 Kutter, 2009) as the most compelling estimate of host range. The challenge for EOP

83 measurements is the time consuming nature of many plaque assays, which have left EOP

84 host range estimates limited to small data sets, using few (<10) hosts or phages (Jensen et

85 al., 1998; Holmfeldt et al., 2014; Silva et al., 2014).

4 86 To advance our understanding of phage–host interactions, we sought to

87 empirically evaluate the quantitative host range in the Cellulophaga baltica phage–host

88 system (Holmfeldt et al., 2007) by screening 38 phages against 19 host strains. From

89 these findings, a subset of the isolates was further explored to investigate host-driven

90 variation in phage replication within a single phage genus.

91

92 Results and Discussion

93 Quantitative host range variations among Cellulophaga phages

94 To assess the Cellulophaga phages’ efficiency of infection (EOI) on the different

95 host strains, their infectivity were investigated using plaque assays. All phages replicated

96 lytically with high EOI on at least one host. While some phages (Cba461, as well as

97 Cba401 and Cba183 investigated in Holmfeldt & Howard-Varona et al., 2014) contained

98 genes pointing towards possible lysogenic replication (Holmfeldt et al., 2013), no phages

99 capable of lysogeny have been observed, even though it was investigated in depth for the

100 Cba401 phages (Holmfeldt & Howard-Varona et al., 2014; Dang & Howard-Varona et

101 al., 2015). The number of host strains a phage could infect and the EOI varied greatly

102 across the phages (Fig. 1 and detailed in supp. Table 1). Specific phages (e.g.

103 siphoviruses φ12:1 and φ10:1) infected only one bacterial strain, while non-tailed ssDNA

104 phages within the Cba184 genus infected up to 17 of the 19 bacterial strains tested. When

105 multiple hosts were infected by a single phage, the EOI varied up to 10 orders of

106 magnitude across different hosts (e.g. φ17:1 on hosts #18 versus #17; φ47:1 on hosts

107 NN014873 versus NN014847). Large variations, both within strain specificity and EOI

108 between strains, could be seen among phages both within the same genus (e.g. Cba41)

5 109 and the same populations (e.g. φ48:1, φ18:4, and φ12a:1; Fig. 1 and supp. Table 2) as

110 defined by >95% average nucleotide identity (ANI) on >80% of their genes (Deng et al.,

111 2014; Brum et al., 2015).

112 Combining this host range data with that of the two previously investigated

113 Cellulophaga podophage genera (Cba401 and Cba183, supp. Table 1) (Holmfeldt &

114 Howard-Varona et al., 2014), the possible connection between host range and phage

115 family was examined. A statistical difference in host range was detected when comparing

116 phages with different taxonomic affiliation (Kruskal-Wallis, p-value=0.0003, χ2=21.30,

117 df=4). The difference was seen as a broader host range for phages within Myoviridae,

118 Microviride, and ssDNA phages (genus Cba482) when compared to Siphoviridae (Dunn-

119 test, p-value 0.0307, 0.0002, and 0.0359 respectively) (Fig. 1 and supp. Table 1). The

120 Microviridae phages also had a broader host range than Podoviridae phages (Dunn-test,

121 p-value 0.0220). That phages belonging to Myoviridae in general have a broader host

122 range than other Caudovirales phages has previously been noted among cyanophages

123 (Sullivan et al., 2003). However, the broad host range among the Cellulophaga

124 Microviridae phages contrasts previous findings (Muniesa et al., 2003; Sillankorva et al.,

125 2011) and is discussed below. While this and other studies (Sullivan et al., 2003;

126 Karumidze et al., 2013) have been able to show possible connection between phage

127 and host range, the patterns are not universal (Karumidze et al., 2012;

128 Holmfeldt & Howard-Varona et al., 2014) and generalizations regarding host-range

129 breadth within and between phage families are difficult to predict. What patterns are

130 observed are also dependent upon the scale of the host range assay, as seen for the

131 Cellulophaga phages comparing host ranges of the smaller set of podophages previously

6 132 tested (Holmfeldt & Howard-Varona et al., 2014) with this more broad-scale study.

133 Further, there was a significantly broader host-range among phages isolated in 2005

134 compared to 2000 (Welch Two Sample t-test, p=0.0001, t=-4.255), though this might be

135 due to that broad host-range phage families (Myoviridae, Microviridae, and ssDNA

136 phages) mainly were isolated in 2005, while narrow host range phages (Siphoviridae)

137 were isolated in 2000 (Fig. 1). When comparing the possibility for the phages to infect

138 hosts that were isolated from the same water sample, the phages were more prone to

139 infect the hosts originating from the same water sample (i.e. phages from 2000 infecting

140 hosts isolated in 2000 compared to hosts isolated in 1994, Welch Two Sample t-test,

141 p=0.0405, t=-2.159). This differs from previous findings among cyanophages (Waterbury

142 & Valois, 1993), where host community was less susceptible to phages occurring in the

143 same water. The 2005 phages were as likely to infect the 1994 hosts as the 2000 hosts

144 (Welch Two Sample t-test, p=0.5523, t=0.598) and no hosts were isolated from the same

145 year and location as the 2005 phages.

146 From the host perspective, some strains were permissive to phage infection, while

147 others were not (Fig. 1). For example, 26 of the 38 phages tested were capable of

148 infecting host #18. These 26 phages represented 9 out of the 11 phage genera tested here

149 as well as 3 phages within the 2 podophage genera previously investigated (Holmfeldt &

150 Howard-Varona et al., 2014). In contrast, bacterium NN014850 was only infected by six

151 phages from two different genera (Fig. 1), as well as 3 phages among the 2 podophage

152 genera previously investigated (Holmfeldt & Howard-Varona et al., 2014). While

153 findings from other host range assays of closely related aquatic bacteria commonly show

154 strains within a species that are resistant or only infected by a single phage (Waterbury

7 155 and Valois, 1993; Sullivan et al., 2003; Comeau et al., 2005; Comeau et al., 2006;

156 Stenholm et al., 2008), our findings point toward a high susceptibility of these bacteria to

157 the isolated phages as even the least susceptible strain NN014850 can be infected by

158 phages from four different genera.

159 Comparing the 722 phage–host combinations analyzed in this quantitative host

160 range assay to the spot test-based assay performed previously (Holmfeldt et al., 2007),

161 24% of the combinations showed a different result. In 117 cases the spot-test had

162 recorded infectivity but no plaques were seen in this study. Conversely, in 54 cases

163 positive infections were recorded with this study while no infection had been recorded

164 with the spot-test. These contrasting findings of spot-tests versus plaque assays are

165 similar to those observed in other studies (Holmfeldt & Howard-Varona et al., 2014;

166 Khan Mirzaei and Nilsson, 2015). Assuming the differences are not attributable to

167 storage artifacts (the spot tests were performed 9 years ago), then it is probable that the

168 plaque assay method provides the most accurate measurement of phage host range as is

169 the consensus in the literature (Kutter, 2009; Khan Mirzaei and Nilsson, 2015). It is

170 thought that spot assays overestimate infection due to scoring inhibition and lysis from

171 without as viral replication, but may underestimate infection of low titers on low

172 infectivity hosts (Moebus, 1983; Hyman and Abedon, 2010).

173

174 Host range patterns among Cellulophaga Microviridae phages

175 Among the Cellulophaga phages, the small Microviridae phages within genus

176 Cba184 infected a larger number of bacterial strains than phages within Siphoviridae and

177 Podoviridae families as well as some phages within Myoviridae (Fig. 1). These

8 178 Cellulophaga phage findings contrast previous results in Escherichia coli where

179 Microviridae have narrower host ranges than Myoviridae and Siphoviridae (Muniesa et

180 al., 2003; Sillankorva et al., 2011). In φX174, the type Microviridae, the H protein

181 recognizes the host cell and mediates DNA entry (Jazwinski et al., 1975; Sun et al.,

182 2013). Unfortunately, the gene responsible for host recognition has not been identifiable

183 among the Cba184 phages (categorized as Microviridae based upon morphology and

184 distant functional homologs; Holmfeldt et al., 2013), even though 10 encapsidated

185 structural proteins have been experimentally determined.

186 There were also large variations in Cba184 strain EOIs in spite of the fact that the

187 Cba184 phage genomes were highly similar. For example, φ18:4 and φ48:1 vary in host

188 range and EOI within shared hosts (Fig. 1), and yet differ at only 11 synonymous and 1

189 non-synonymous substitutions across their 6.5 kb genomes (supp. Table S3). Single non-

190 synonymous substitutions are known to cause differences in host range, as previously

191 observed in ssDNA phage phiX174, dsRNA phage phi6 and dsDNA phage φΕΦ24Χ

192 (Cox and Putonti, 2010; Uchiyama et al., 2011; Ford et al., 2014). Here the non-

193 synonymous substitution belongs to a protein located in the phage particle, and thus,

194 presumed structural (Holmfeldt et al., 2013), though its specific function is still unknown.

195 Genetic manipulation of this protein, not yet available in this phage or host, could

196 someday test the hypothesis that such a substitution can drive the host range for the

197 Cba184 phages.

198

199 Replication variations among Cba41 phages

9 200 Two Cba41 phages, φ4:1 and φ17:2, had only minor differences in their genomes

201 as well as host range, making them interesting candidates for investigating the factors

202 regulating differences in infection among these Cellulophaga phages. In terms of host

203 ranges, both phages infected the same six host strains but φ17:2 infected two additional

204 strains at a significantly lower EOI (Fig. 1). The genomes of these phages had 98%

205 nucleotide (nt) identity (id) over the 96% of the genome that was shared (Fig. 2A), with

206 most (88%) genes >95% nt id, placing them within the same population (see Fig. 2B for

207 details). Further, 3 genes were shared with 90–95% nt id, 11 genes shared only a fraction

208 (14–92%) of the gene and with 87–100% nt id, and the remaining genes (10 for φ4:1 and

209 11 for φ17:2) lacking detectable similarity (e-value cut off e-3) at either the nt or amino

210 acid level. Despite the high sequence similarity between the phages, their genomes were

211 divergent enough that predictions about genetic determinants for the detected host range

212 and EOI variations were difficult to make.

213 To explore the variation in EOI for φ4:1 and φ17:2, we both ‘purified’ (i.e.

214 plaques had been picked and replicated on the alternative host 3 consecutive times) and

215 conducted ‘shuffling’ experiments by passaging the phages between the alternative host

216 strains #13 (φ4:1) and #18 (φ4:1 and φ17:2) and the original host strains #4 and #17 after

217 each round of plaquing (see Fig. 3A for details). These alternative host strains were

218 chosen given the 1–3 orders of magnitude reduced EOIs of φ4:1 and φ17:2 on these host

219 strains compared to the EOI of the phage on their original hosts (#4 and #17,

220 respectively; Fig. 1 and 3). In all cases, the passaged phages (φ4:1_13, φ4:1_18, and

221 φ17:2_18) had increased EOI on the host they were passaged on (#13 and #18,

222 respectively) compared to the ancestral phages (φ4:1 and φ17:2) infecting the same host

10 223 (t-test, φ4:1_13 versus φ4:1 on host #13: p=0.034; φ4:1_18 versus φ4:1 on host #18:

224 p=0.025; and φ17:2_18 versus φ17:2 on host #18: p=0.001, respectively; Fig. 3A–B). No

225 change in infectivity was seen for φ4:1_13 when infecting host #18 (t-test, p=0.103) or

226 φ4:1_18 when infecting host #13 (t-test, p=0.298) (Fig. 3B). The difference in EOI

227 occurred both immediately when the phage in a newly picked plaque was exposed to a

228 new host (shuffling; Fig. 3) and if the phage had been ‘purified’ (Fig. 3B–C). Thus it

229 appeared that passaging a phage on the low EOI host affected it, leading to increased EOI

230 on that host. However, when the passaged phage (e.g. 4:1_13) was passaged again

231 through the original host (e.g. 4) its EOI returned to the same as the ancestral phage (e.g.

232 4:1) (for example see Fig. 3A), suggesting that the change in EOI was a non-heritable

233 trait.

234 To investigate additional phenotypic differences among the Cba41 ancestral and

235 passaged phages, one-step growth curves were performed for φ4:1 on the three host

236 strains (#4, #13, and #18) (Fig. 4A–C) and for φ4:1_13 and φ4:1_18 on host #4 and the

237 host the phages had been propagated on (#13 and #18, respectively; φ4:1_13: Fig. 4D-E,

238 φ4:1_18: Fig. 4F-G). Concurring with the EOI results, the ancestral phage φ4:1 had a

239 reduced burst size on the alternative hosts #13 (p=0.023) and #18 (p=0.035) compared to

240 when infecting host #4. Further, the curves on the alternative hosts (13 and 18) did not

241 produce a clean step (Fig. 4B–C), pointing towards replication difficulties for φ4:1 on

242 those hosts. For the passaged phages, φ4:1_13 and φ4:1_18, the replication cycle had

243 improved for the infection on the host they had been propagated on (e.g. 4:1_13 on 13).

244 This meant a clearer “step” in the one-step growth curve (Fig. 4E and G) and an

245 increased burst size for φ4:1_13 versus φ4:1 when infecting #13 (27 vs 4, respectively,

11 246 p=0.038). As there was no difference in latent period, the passaged phages had a faster

247 replication rate when infecting the alternative hosts (burst size divided by latent period;

248 Keen, 2014) than when infecting the original host. Given the increased EOI, this contrasts

249 previous findings where extended host range is connected with reduced replication rate

250 (Duffy et al., 2006; Kesik-Szeloch et al., 2013; Keen, 2014). However, it should be noted

251 that these studies looked at host range differences between different host strains, not the

252 difference in EOI on the same host strain and in two cases different phages (Kesik-

253 Szeloch et al., 2013; Keen, 2014) instead of phages evolved from the same ancestor.

254

255 Genetic similarities among Cba41 phages

256 To detect possible genetic differences between the ancestral and passaged phages,

257 the genome of the passaged phages (φ4:1_13, φ4:1_18 and φ17:2_18) were sequenced

258 with Illumina. A first comparison of the assembled passaged phage genomes to the

259 ancestral phages using blastn revealed 100% nt id. Further detailed by recruiting the

260 passaged phage reads to the ancestral phage genomes (average coverage 300–550 fold)

261 revealed two possible indels: a missing T at a 10-T region (φ4:1_13, φ4:1_18, and

262 φ17:2_18) and a missing AT at a 8-AT region (φ4:1_13 and φ4:1_18). However, <4% of

263 the reads at either position had these indels and while it might represent true variation

264 within the phage population it is likely not connected to the noted replication variations

265 which ought to have been noted among the majority of reads. Thus, the difference in EOI

266 could not be connected with sequence mutations in the phages’ genomes. While

267 difference in infection driven by altered phage entry (e.g. Baranowski et al., 2001; Cox

268 and Putonti, 2010; Uchiyama et al., 2011; Ford et al., 2014) or CRISPR system

12 269 (Barrangou et al., 2007) function is explained by nucleotide substitutions, this is not

270 always the case for phages circumventing host restriction-modification systems (Krüger

271 and Bickle, 1983). We therefore hypothesized that the difference in EOI seen between the

272 passaged and ancestral phages could be due to differences in DNA modification. This

273 hypothesis was also strengthened by the non-heritable nature of the change in EOI seen in

274 the shuffling experiment (Fig. 3A), as the phage would only be given the correct

275 modification if it were provided by the last bacteria the phage infected (Luria, 1953;

276 Labrie et al., 2010).

277 Among phages, such non-sequence-derived, functionality changing (epigenetic)

278 switches are commonly caused by host-derived methylation of the phage DNA, which

279 protect progeny phages from the host restriction system during future infections (Arber,

280 1965b; Hattman, 2009). For example, just 1–2 DNA modifications can cause a 100-fold

281 EOI difference (Arber and Kuhnlein, 1967), which is similar to the difference in EOI that

282 is noted between the ancestral and passaged Cba41 phages (Fig. 3). To investigate the

283 possible methylation signals among the phages, they and the bacterial hosts genomes

284 were sequenced with PacBio (Table 1). Using the single-molecule, real time (SMRT)

285 sequencing technique, PacBio sequencing can be used as a high throughput method, with

286 high accuracy to define modified nucleotides (Flusberg et al., 2010; Fang et al., 2012;

287 Burgess, 2013; Schadt et al., 2013) and has previously been used to detect methylation

288 differences among coli- and lactococcal-phages (Fang et al., 2012; Murphy et al., 2014).

289 The bacterial hosts all contained different modifications ranging from no

290 modifications (strain #4) to 2–3 modifications in the remaining strains (Table 1).

291 However, the phage methylomes did not match any of those modifications. Instead,

13 292 among the five Cba41 phages, only one, identical methylation event was detected, a 6-

293 methyladenosine (m6A) in the motif CTYCAG, which is presumably the product of a

294 phage encoded methylase (Table 1). Thus, methylation based evasion of the host

295 restriction system is not the obvious explanation of the observed phenotypic changes.

296 While the SMRT technique is able to detect the methylations that are commonly

297 occurring among bacteria and phages (e.g. m6A and 4- and 5-methylcytosine), as well as

298 glucosylated 5-hydroxymethylcytosine, a common epigenetic driver among T-even

299 phages (Lehman and Pratt, 1960; Hattman and Fukasawa, 1963), sequencing coverage

300 (25–250x per strand depending on modification) does impact detection accuracy of

301 methylation signals (Biosciences, 2012). Here, the mean sequence coverage for bacterial

302 genomes was 30–54 fold, which is at the lower end of acceptable and suggests that some

303 modifications might not have been detected. However, for the phages, sequencing

304 coverage was >500 fold, so we expect that any known modifications present in the

305 majority of the population of sequenced phages would have been detected.

306 While considering both sequential and DNA-modification differences as possible

307 mechanisms for the detected host-interaction changes, no clear evidence could be

308 detected that this would drive the differences in EOI or burst size. For Cellulophaga

309 phage φ38:1, transcriptional differences within different hosts have been seen to regulate

310 replication differences for the phage (Howard-Varona et al., 2016). Thus, potential

311 mechanisms, including DNA modification not detected with PacBio, transcriptional, or

312 translational differences have to be taken into future consideration. Still, these infection

313 and replication measurements across phages and hosts provide important empirical data

314 for drawing the boundaries of host-to-host EOI differences.

14 315

316 Ecological and theoretical implications of variations in phage–host interactions

317 With increasing availability of next-generation sequencing methods, viral ecology

318 is moving further towards a more sequencing-based research field, providing important

319 data regarding functionality, taxonomy, and even host interactions (e.g. Hurwitz et al.,

320 2013; Deng et al., 2014; Hurwitz et al., 2015; Roux et al., 2015). To define ecological

321 perspective to the immense amount of data, metrics for describing e.g. populations, on a

322 nucleotide level have been suggested (Deng et al., 2014) and later been used for powerful

323 ecological inferences (Brum et al., 2015; Brum and Sullivan, 2015). Within this study we

324 use the same described metrics (>95% ANI on >80% of genes) to define populations

325 within the Cellulophaga phage system and evaluate infection variations within the

326 populations (Fig 1, supp. table 2), which provides experimental data for future phage-host

327 studies of sequence derived data. For example, methods have recently been evaluated and

328 shown to be able to predict possible hosts for phage contigs derived from metagenomes

329 (Roux et al., 2015), but it would be important to include computational analyses for

330 quantitative species-specific variations.

331 Further, through reinvestigation of numerous non-quantitative host range datasets,

332 mathematical models recently provided insight into nested and modular phage–host

333 infection networks (Flores et al., 2011; Flores et al., 2013; Weitz et al., 2013). While

334 phage-host systems of closely related taxa showed nested networks, where the least

335 susceptible is infected by the generalist phages and most susceptible host is infected by

336 both specialist and generalist phages (Flores et al., 2011) modularity seem to be the norm

337 when investigating phage-host interactions across larger taxonomic boundaries (Flores et

15 338 al., 2013; Roux et al., 2015). These studies have highlighted the importance of including

339 strain-specificity variations among bacterial species when investigating the impact of

340 phage driven Lotka-Voltera-like oscillations (Jover et al., 2013; Thingstad et al., 2014;

341 Thingstad et al., 2015). Now, with the results from the Cellulophaga phage–host system

342 we posit that the next stage in advancing our understanding will come from developing

343 theory that considers empirical measurements of EOI with particular focus on variation

344 across different bacterial strains. Such theory is likely more critically needed for phage–

345 host systems where “restrictive” hosts are still infected by many phages (e.g.,

346 Cellulophaga, this study) than those that are infected by few or no phages (e.g.,

347 Prochlorococcus and Synechococcus, Sullivan et al 2003). Further, infection efficiency

348 and replication success may also change among the phages depending on which host the

349 phage most recently infected. For example, with the higher infection efficiency and

350 replication rate the passaged Cba41 phages would presumably provide a heightened

351 mortality pressure on the host community than the ancestral phages (Abedon, 2009). This

352 change in viral replication success could be an important fitness aspect for the phages co-

353 occurring with a large number of different host strains, which is the case for the

354 Cellulophaga phages. Regarding this study, the 4 hosts used for the Cba41 experiments

355 were isolated at the same sampling together with 5 additional C. baltica strains

356 (Holmfeldt et al., 2007). Though, it should be noted that the Cba41 phages were isolated

357 5 years later and there is no data regarding the dynamics of the host community at that

358 point of time. Preferably, to understand the implications regarding the impact of these

359 ancestral/passaged phages on their host community, experiments with mixes of both

360 phage types and different host strains would be performed. However, due to their

16 361 identical DNA sequences this would be difficult, as we at this time point do not have any

362 means to separate or enumerate the different phages after they have been mixed.

363 Besides having variations in EOI among genetically identical phages, we posit

364 that the difference in susceptibility among the bacterial hosts will also be difficult to

365 elucidate using common bacterial evolution and clustering genes, like the 16S rRNA

366 gene (Olsen et al., 1986). The C. baltica bacteria varied largely in phage susceptibility

367 (Fig. 1) between strains that were 100% identical on both their 16S rRNA gene and ITS

368 sequences (e.g. MM#3 and NN016046 in Fig. 1; Holmfeldt et al., 2007). This points

369 towards the rapid evolution of the phage susceptibility trait compared to the slow, long-

370 term evolution of 16S and ITS as previously noted among Prochlorococcus and their

371 phages (Coleman and Chisholm, 2007). As host susceptibility further may be connected

372 to substrate utilization, which has previously been seen in the Cellulophaga phage–host

373 system (Middelboe et al., 2009), this difference in host susceptibility may lead to

374 variations in the bacterial strains’ impact on nutrient cycling, neither of which would be

375 noted on the level of 16S or ITS sequence variations.

376

377 Conclusions

378 This study illustrates the intricate interactions occurring among different and

379 genetically identical phages when infecting strains of the same bacterial species (99–

380 100% similar on 16S) and how these interactions change depending on the last host that

381 was infected. It points out the importance of quantitative host range investigations of

382 aquatic phages to fully understanding the impact they will have on the host community.

383 However, investigations are heavily bottlenecked by the time consuming nature of plaque

17 384 assays and the lack of isolated phage–host systems, which is further complicated by

385 cultivation difficulties of aquatic bacteria (Rappé et al., 2002; Simu and Hagström, 2004).

386 While novel methods are underway for host range determination in wild populations with

387 reduced dependence on cultivation (Tadmor et al., 2011; Deng et al., 2012; Ohno et al.,

388 2012; Allers et al., 2013; Deng et al., 2014; Martinez-Garcia et al., 2014; Roux et al.,

389 2014; Lima-Mendez et al. 2015; Roux et al., 2015), this study points out the importance

390 of also bringing the quantitative efficiency of infection into the host range determination

391 of these culture independent methods. It will be important to get large-scale data on these

392 variations to increase our understanding of phage-host interactions and population

393 dynamics within natural environments.

394

395 Experimental procedures

396 Bacterial and phage growth procedures

397 Isolation procedures for the 19 bacterial strains and 38 phage isolates investigated

398 in this study were described previously (Holmfeldt et al., 2007). The Cellulophaga

399 baltica host strains were grown at room temperature (RT) on Zobell agar plates (1 g yeast

400 extract [Becton, Dickinson and Company (BD)], 5 g bacto-peptone [BD], 15 g agarose

401 [BD], and 15 g Sea Salts [Sigma] per liter). Single colonies were inoculated into MLB

402 liquid media (0.5 g yeast extract [BD], 0.5 g bacto-peptone [BD], 0.5 g casaminoacids

403 [Fisher], 3 mL glycerol [Millipore] and 15 g Sea Salts [Sigma] per liter). Cultures were

404 grown without agitation and experiments were performed on cultures that had reached

405 late exponential phase (108 cells mL-1; typically overnight).

18 406 Phages were grown using the top-agar plating technique (plaque assay)

407 (Sambrook et al., 1989). Briefly, phages were diluted in Marine Sodium Magnesium

408 (MSM) buffer (450 mM NaCl, 50 mM MgSO4x7H20, 50mM Tris base, pH8), mixed with

409 0.3 mL of bacterial over night culture and 3.5 mL molten soft agar (MSM containing

410 0.6% low melting point agarose), and dispersed on agar plates. Plates were incubated at

411 RT in the dark and plaques were visible after 1–2 days. For phage lysate, 5 mL of MSM

412 was added to fully lysed plates, the plates were shaken for at least 30 min at RT, and then

413 the lysate was collected, 0.2 µm filtered and stored in the dark at 4ºC.

414

415 Quantitative host range assay

416 Phages for host range assays were harvested from fully lysed plates where the

417 individual phages were grown on their original hosts. 10x serial dilutions of these lysates

418 were made to 10-10 in MSM buffer. 100 µL of undiluted lysate and each dilution were

419 then used for plaque assays (see above) on each of the 19 hosts within 2 weeks of lysate

420 preparation.

421

422 Isolation of mutant phages and shuffling assay

423 Passaged phages for φ4:1 were isolated from host #13 and #18 (new phage name:

424 φ4:1_13 and φ4:1_18) and for φ17:2 from host #18 (new phage name: φ17:2_18). A

425 single plaque was picked with a 100 µL pipet tip, dispersed in 1 mL MSM buffer,

426 incubated for 30 min at RT and 0.2 µm filtered. This was used as a phage stock for new

427 plaque assays as described above. The passaged phages were replicated in the same

428 manner 3 times before their efficiencies of infection were determined as described above.

19 429 Further, for shuffling assays, the newly picked phage was exposed (plaque assay) to a

430 different host each time a new plaque was picked.

431

432 One-step growth curves

433 One-step growth curves measuring properties of the infection were done with

434 MOI of 0.1 to minimize multiple phage infections per cell. For this, 108 cells (late

435 exponential phase) and 107 phages were mixed for a total volume of 1 mL. After 15 min

436 of bacterial–phage adsorption time, new infections were stopped by diluting the bacteria–

437 phage mix 1000-fold in MLB medium. The first sample was immediately obtained and

438 termed “time 0” and from there on, samples were periodically taken from the bacterial–

439 phage mix. Phages were enumerated by plaque assays on the host used for the one-step.

440 Sampling was carried out every 30 or 60 min for a total of 8–10 h depending on the

441 phage–host pair. In order to calculate the burst size, phage samples collected included

442 both free (extracellular) as well as total (intracellular and extracellular) phages. For the

443 former, a 500 µL sample of the diluted infection was centrifuged at 10 000 x g for 5 min

444 to pellet the cells and 100 µL of the supernatant, containing only free phages, was

445 analyzed through plaque assay. For the total phages, the plaque assay was done with 100

446 µL taken directly from the 1000-fold dilution flask containing both intra- and

447 extracellular phages. Plates were left overnight in the dark at RT and plaques from two

448 consecutive dilutions were counted to obtain number of plaque forming units (pfu) per

449 mL. The burst size was calculated using the free phage values at the plateau after the

450 burst and subtracting the free phages before the burst from them, and then dividing that

20 451 number by the number of phages that were able to infect (the difference between the total

452 and free fraction of phages during the latent period).

453

454 DNA extraction, sequencing and downstream genome analyses

455 Large-scale phage concentrates (3–10 fully lysed plates) were precipitated with

456 polyethylene glycol (PEG 8000, Fisher) to use for DNA extraction. Briefly, 6.5 % (w/v)

457 of NaCl was added to the phage lysate, incubated on ice for 1 h–ON at 4ºC, and

458 centrifuged (11 000 x g, 10 min). To the supernatant 10% (w/v) PEG was added,

459 incubated 1 h–ON at 4ºC, and centrifuged (10 000 x g, 10 min). After the pellet had air-

460 dried it was resuspended in 1 mL MSM. DNA from the dsDNA phages was extracted

461 using the Wizard DNA purification kit (Promega) according to the manufacturer’s

462 recommendations. DNA from liquid ON bacterial cultures was extracted using the

463 DNeasy Blood and Tissue kit (Qiagen) according to the manufacturer’s

464 recommendations.

465

466 DNA sequencing and genome analyses

467 The genomes for the Cellulophaga phages were retrieved from GenBank and

468 genes of closely related phages were compared using BLASTn, to identify the number of

469 shared genes with >95% nucleotide identity.

470 Passaged phages for φ4:1 and φ17:2 were sequenced commercially using Illumina

471 HiSeq 2000 (http://uagc. arl.arizona.edu). Phage genomes derived from the Illumina

472 reads were assembled using the Newbler assembly software package (454 Life Sciences,

473 Branford, CT) with all settings set to default. Illumina reads for the passaged phages were

21 474 recruited to the ancestral phage genomes using Bowtie2 (version 2.2.5; Langmead and

475 Salzberg, 2012) with default settings, converted with Samtools (version 1.2; Li et al.,

476 2009) and visualized with Tablet (version 1.14.11.07; Milne et al., 2013). Genomes and

477 proteins were compared using blastn and blastp and genomes were aligned using

478 progressive Mauve (Darling et al., 2010).

479 All φ4:1 and φ17:2 phages (both ancestral and passaged) and the bacteria they had

480 been replicated on were sequenced using PacBio SMRT sequencing. SMRT sequencing

481 was performed at the DOE Joint Genome Institute using the Pacific Biosciences RS

482 instrument with C2 chemistry (Travers et al., 2010). The average SMRT sequence

483 coverage per genome ranged from 32–53X for the host strains and 532–554X for the

484 phages. DNA modification detection and motif analysis was performed using the PacBio

485 SMRT analysis platform (protocol version=2.2.0 method=RS Modification and Motif

486 Analysis.1, http://www.pacb.com/devnet/code.html). Briefly, short reads and sequencing

487 adapters were removed using SFilter. Filtered reads were aligned to the reference genome

488 using BLASR(v1) (Chaisson and Tesler, 2012). Modified sites were then identified

489 through kinetic analysis of the aligned DNA sequence data (Flusberg et al., 2010) and

490 motifs identified using MotifFinder (v1)2.

491

492 Statistics

493 The connection between host range and phage family was tested using a Kruskal-

494 Wallis test with a Dunn-test Post hoc with Bonferroni correction and the connection

495 between phage and host sampling year and location vs. host range was tested using

496 Welch Two Sample t-test using R (version 3.2.1). In these analyses the phages

22 497 investigated in Holmfeldt & Howard-Varona et al., 2014 was included. For host year and

498 location, the 1994 hosts were grouped together and the 2005 host was removed as only

499 one bacterium had been obtained from that location. Students-t tests were performed on

500 triplicates when comparing efficiency of infection and burst size of the passaged phages

501 compared to the ancestral phages using SPSS Statistics (version 22).

502

503 Nucleotide sequence accession numbers

504 The assembled genomes for the passaged phages has been deposited in GenBank

505 and assigned the accession numbers KT962245-962247. Bacterial genomes have

506 GenBank accession numbers NZ_ATLH01000000, NZ_ATLG01000000, and

507 NZ_CP009976.

508

509 Acknowledgments

510 We thank Emelie Nilsson for statistical assistance. We thank reviewers for

511 insightful comments improving the manuscript. The study was sponsored by postdoctoral

512 fellowships from the Swedish Research Council (623-2012-1395 and 623-2010-6548) to

513 KH, and Gordon and Betty Moore Foundation awards (GBMF2631 and GBMF3790) to

514 MBS. Work conducted by the U.S. Department of Energy Joint Genome Institute was

515 supported by the DOE Office of Science (DE-AC02-05CH11231).

516 Conflict of interest

517 The authors declare no conflict of interest.

518

519 References

23 520 Abedon, S.T. (2009) Impact of phage properties on bacterial survival. In Contemporary

521 trends in bacteriophage research. Adams, H.T. (ed). New York: Nova Science

522 Publishers.

523 Allers, E., Moraru, C., Duhaime, M.B., Beneze, E., Solonenko, N., Barrero-Canosa, J. et

524 al. (2013) Single-cell and population level viral infection dynamics revealed by

525 phageFISH, a method to visualize intracellular and free viruses. Environ Microbiol 15:

526 2306-2318.

527 Arber, W. (1965a) Host-controlled modification of bacteriophage. Annu Rev Microbiol

528 19: 365-378.

529 Arber, W. (1965b) Host specificity of DNA produced by Escherichia coli V . The role of

530 methionine in the production of host specificity. J Mol Biol 11: 247-256.

531 Arber, W., and Kuhnlein, U. (1967) Mutationeller verlust B-spezifischer restriktion des

532 bakteriophagens fd. Pathol Microbiol (Basel) 30: 946-952.

533 Avrani, S., Wurtzel, O., Sharon, I., Sorek, R., and Lindell, D. (2011) Genomic island

534 variability facilitates Prochlorococcus-virus coexistence. Nature 474: 604-608.

535 Baranowski, E., Ruiz-Jarabo, C.M., and Domingo, E. (2001) Evolution of cell

536 recognition by viruses. Science 292: 1102-1105.

537 Barrangou, R., Fremaux, C., Deveau, H., Richards, M., Boyaval, P., Moineau, S. et al.

538 (2007) CRISPR provides acquired resistance against viruses in prokaryotes. Science 315:

539 1709-1712.

540 Biosciences, P. (2012). Detecting DNA base modifications using single molecule, real-

541 time sequencing [WWW document]. URL http://www.pacb.com/wp-

24 542 content/uploads/2015/09/WP_Detecting_DNA_Base_Modifications_Using_SMRT_Sequ

543 encing.pdf

544 Breitbart, M. (2012) Marine viruses: Truth or dare. Annu Rev Mar Sci 4: 5.1-5.24.

545 Brum, J.R., and Sullivan, M.B. (2015) Rising to the challenge: accelerated pace of

546 discovery transforms marine virology. Nat Rev Microbiol 13: 147-159.

547 Brum, J.R., Ignacio-Espinoza, J.C., Roux, S., Doulcier, G., Acinas, S.G., Alberti, A. et al.

548 (2015) Ocean plankton. Patterns and ecological drivers of ocean viral communities.

549 Science 348: 1261498.

550 Burgess, D.J. (2013) Epigenetics: Bacterial DNA methylation gets SMRT. Nat Rev Genet

551 14: 4.

552 Chaisson, M.J., and Tesler, G. (2012) Mapping single molecule sequencing reads using

553 basic local alignment with successive refinement (BLASR): application and theory. Bmc

554 Bioinformatics 13: 238.

555 Coleman, M.L., and Chisholm, S.W. (2007) Code and context: Prochlorococcus as a

556 model for cross-scale biology. Trends Microbiol 15: 398-407.

557 Comeau, A.M., Buenaventura, E., and Suttle, C.A. (2005) A persistent, productive, and

558 seasonally dynamic vibriophage population within Pacific oysters (Crassostrea gigas).

559 Appl Environ Microbiol 71: 5324-5331.

560 Comeau, A.M., Chan, A.M., and Suttle, C.A. (2006) Genetic richness of vibriophages

561 isolated in a coastal environment. Environ Microbiol 8: 1164-1176.

562 Cox, J., and Putonti, C. (2010) Mechanisms responsible for a PhiX174 mutant's ability to

563 infect Escherichia coli by phosphorylation. J Virol 84: 4860-4863.

25 564 Dang, V.T., Howard-Varona, C., Schwenck, S., and Sullivan, M.B. (2015) Variably lytic

565 infection dynamics of large Bacteroidetes podovirus phi38:1 against two Cellulophaga

566 baltica host strains. Environ Microbiol 17: 4659-4671.

567 Darling, A.E., Mau, B., and Perna, N.T. (2010) progressiveMauve: multiple genome

568 alignment with gene gain, loss and rearrangement. PLoS One 5: e11147.

569 Deng, L., Gregory, A., Yilmaz, S., Poulos, B.T., Hugenholtz, P., and Sullivan, M.B.

570 (2012) Contrasting strategies of viruses that infect photo- and hetero- trophic bacteria, as

571 revealed by viral tagging. mBio 3: e00373-00312.

572 Deng, L., Ignacio-Espinoza, J.C., Gregory, A.C., Poulos, B.T., Weitz, J.S., Hugenholtz,

573 P., and Sullivan, M.B. (2014) Viral tagging reveals discrete populations in

574 Synechococcus viral genome sequence space. Nature 513: 242-245.

575 Deveau, H., Barrangou, R., Garneau, J.E., Labonte, J., Fremaux, C., Boyaval, P. et al.

576 (2008) Phage response to CRISPR-encoded resistance in Streptococcus thermophilus. J

577 Bacteriol 190: 1390-1400.

578 Duffy, S., Turner, P.E., and Burch, C.L. (2006) Pleiotropic costs of niche expansion in

579 the RNA bacteriophage Phi 6. Genetics 172: 751-757.

580 Dupuis, M.E., Villion, M., Magadan, A.H., and Moineau, S. (2013) CRISPR-Cas and

581 restriction-modification systems are compatible and increase phage resistance. Nat

582 Commun 4: 2087.

583 Falkowski, P.G., Fenchel, T., and Delong, E.F. (2008) The microbial engines that drive

584 Earth's biogeochemical cycles. Science 320: 1034-1039.

26 585 Fang, G., Munera, D., Friedman, D.I., Mandlik, A., Chao, M.C., Banerjee, O. et al.

586 (2012) Genome-wide mapping of methylated adenine residues in pathogenic Escherichia

587 coli using single-molecule real-time sequencing. Nat Biotechnol 30: 1232-1239.

588 Flores, C.O., Valverde, S., and Weitz, J.S. (2013) Multi-scale structure and geographic

589 drivers of cross-infection within marine bacteria and phages. ISME J 7: 520-532.

590 Flores, C.O., Meyer, J.R., Valverde, S., Farr, L., and Weitz, J.S. (2011) Statistical

591 structure of host-phage interactions. Proc Natl Acad Sci U S A 108: E288-297.

592 Flusberg, B.A., Webster, D.R., Lee, J.H., Travers, K.J., Olivares, E.C., Clark, T.A. et al.

593 (2010) Direct detection of DNA methylation during single-molecule, real-time

594 sequencing. Nat Methods 7: 461-465.

595 Ford, B.E., Sun, B., Carpino, J., Chapler, E.S., Ching, J., Choi, Y. et al. (2014) Frequency

596 and fitness consequences of bacteriophage phi6 host range mutations. PLoS One 9:

597 e113078.

598 Fuhrman, J.A. (1999) Marine viruses and their biogeochemical and ecological effects.

599 Nature 399: 541-548.

600 Haaber, J., Rousseau, G.M., Hammer, K., and Moineau, S. (2009) Identification and

601 characterization of the phage gene sav, involved in sensitivity to the lactococcal abortive

602 infection mechanism AbiV. Appl Environ Microbiol 75: 2484-2494.

603 Hattman, S. (2009) The first recognized epigenetic signal: DNA glucosylation of T-even

604 bacteriopages. Epigenetics 4: 150-151.

605 Hattman, S., and Fukasawa, T. (1963) Host-induced modification of T-even phages due

606 to defective glucosylation of their DNA. Proc Natl Acad Sci U S A 50: 297-300.

27 607 Holmfeldt, K., Middelboe, M., Nybroe, O., and Riemann, L. (2007) Large variabilities in

608 host strain susceptibility and phage host range govern interactions between lytic marine

609 phages and their Flavobacterium hosts. Appl Environ Microbiol 73: 6730-6739.

610 Holmfeldt, K., Howard-Varona, C., Solonenko, N., and Sullivan, M.B. (2014)

611 Contrasting genomic patterns and infection strategies of two co-existing Bacteroidetes

612 podovirus genera. Environ Microbiol 16: 2501-2513.

613 Holmfeldt, K., Solonenko, N., Shah, M., Corrier, K., Riemann, L., Verberkmoes, N.C.,

614 and Sullivan, M.B. (2013) Twelve previously unknown phage genera are ubiquitous in

615 global oceans. Proc Natl Acad Sci U S A 110: 12798-12803.

616 Howard-Varona, C., Roux, S., Dore, H., Solonenko, N.E., Holmfeldt, K., Markillie,

617 L.M., Orr, G., and Sullivan, M.B. (2016) Regulation of infection efficiency in a globally

618 abundant marine Bacteriodetes virus. ISME J accepted.

619 Hurwitz, B.L., Hallam, S.J., and Sullivan, M.B. (2013) Metabolic reprogramming by

620 viruses in the sunlit and dark ocean. Genome Biol 14: R123.

621 Hurwitz, B.L., Brum, J.R., and Sullivan, M.B. (2015) Depth-stratified functional and

622 taxonomic niche specialization in the 'core' and 'flexible' Pacific Ocean Virome. ISME J

623 9: 472-484.

624 Hyman, P., and Abedon, S.T. (2010) Bacteriophage host range and bacterial resistance.

625 Adv Appl Microbiol 70: 217-248.

626 Jazwinski, M.S., Lindberg, A.A., and Kornberg, A. (1975) The gene H spike protein of

627 bacteriophages phiX174 and S13. I. Functions in the phage-receptor recognition and in

628 transfection. Virology 66: 283-293.

28 629 Jensen, E.C., Schrader, H.S., Rieland, B., Thompson, T.L., Lee, K.W., Nickerson, K.W.,

630 and Kokjohn, T.A. (1998) Prevalence of broad-host-range lytic bacteriophages of

631 Sphaerotilus natans, Escherichia coli, and Pseudomonas aeruginosa. Appl Environ

632 Microbiol 64: 575-580.

633 Jover, L.F., Cortez, M.H., and Weitz, J.S. (2013) Mechanisms of multi-strain coexistence

634 in host-phage systems with nested infection networks. J Theor Biol 332: 65-77.

635 Karumidze, N., Kusradze, I., Rigvava, S., Goderdzishvili, M., Rajakumar, K., and

636 Alavidze, Z. (2013) Isolation and characterisation of lytic bacteriophages of Klebsiella

637 pneumoniae and Klebsiella oxytoca. Curr Microbiol 66: 251-258.

638 Karumidze, N., Thomas, J.A., Kvatadze, N., Goderdzishvili, M., Hakala, K.W.,

639 Weintraub, S.T. et al. (2012) Characterization of lytic Pseudomonas aeruginosa

640 bacteriophages via biological properties and genomic sequences. Appl Microbiol

641 Biotechnol 94: 1609-1617.

642 Keen, E.C. (2014) Tradeoffs in bacteriophage life histories. Bacteriophage 4: e28365.

643 Kesik-Szeloch, A., Drulis-Kawa, Z., Weber-Dabrowska, B., Kassner, J., Majkowska-

644 Skrobek, G., Augustyniak, D. et al. (2013) Characterising the biology of novel lytic

645 bacteriophages infecting multidrug resistant Klebsiella pneumoniae. Virol J 10: 100.

646 Khan Mirzaei, M., and Nilsson, A.S. (2015) Isolation of phages for phage therapy: a

647 comparison of spot tests and efficiency of plating analyses for determination of host

648 range and efficacy. PLoS One 10: e0118557.

649 Krüger, D.H., and Bickle, T.A. (1983) Bacteriophage survival: multiple mechanisms for

650 avoiding the deoxyribonucleic-acid restriction systems of their hosts. Microbiol Rev 47:

651 345-360.

29 652 Kutter, E. (2009) Phage host range and efficiency of plating. Methods Mol Biol 501: 141-

653 149.

654 Labonte, J.M., Swan, B.K., Poulos, B., Luo, H., Koren, S., Hallam, S.J. et al. (2015)

655 Single-cell genomics-based analysis of virus-host interactions in marine surface

656 bacterioplankton. ISME J 9: 2386-2399.

657 Labrie, S.J., Samson, J.E., and Moineau, S. (2010) Bacteriophage resistance mechanisms.

658 Nat Rev Microbiol 8: 317-327.

659 Langmead, B., and Salzberg, S.L. (2012) Fast gapped-read alignment with Bowtie 2. Nat

660 Methods 9: 357-U354.

661 Lehman, I.R., and Pratt, E.A. (1960) On the structure of the glucosylated hydroxymethyl

662 cytosine nucleotides of coliphages T2, T4 and T6. J Biol Chem 235: 3254-3259.

663 Li, H., Handsaker, B., Wysoker, A., Fennell, T., Ruan, J., Homer, N. et al. (2009) The

664 Sequence Alignment/Map format and SAMtools. Bioinformatics 25: 2078-2079.

665 Lima-Mendez, G., Faust, K., Henry, N., Decelle, J., Colin, S., Carcillo, F. et al. (2015)

666 Top-down determinants of community structure in the global plankton interactome.

667 Science in press.

668 Luria, S.E. (1953) Host-induced modifications of viruses. Cold Spring Harb Symp Quant

669 Biol 18: 237-244.

670 Martinez-Garcia, M., Santos, F., Moreno-Paz, M., Parro, V., and Anton, J. (2014)

671 Unveiling viral-host interactions within the 'microbial dark matter'. Nature

672 Communications 5: 4542.

30 673 Middelboe, M., Holmfeldt, K., Riemann, L., Nybroe, O., and Haaber, J. (2009)

674 Bacteriophages drive strain diversification in a marine Flavobacterium: implications for

675 phage resistance and physiological properties. Environ Microbiol 11: 1971-1982.

676 Milne, I., Stephen, G., Bayer, M., Cock, P.J.A., Pritchard, L., Cardle, L. et al. (2013)

677 Using Tablet for visual exploration of second-generation sequencing data. Brief

678 Bioinform 14: 193-202.

679 Moebus, K. (1983) Lytic and inhibition responses to bacteriophages among marine

680 bacteria, with special reference to the origin of phage-host systems. Helgol Meeresunters

681 36: 375-391.

682 Moebus, K., and Nattkemper, H. (1981) Bacteriophage sensitivity patterns among

683 bacteria isolated from marine waters. Helgol Meeresunters 34: 375-385.

684 Muniesa, M., Moce-Llivina, L., Katayama, H., and Jofre, J. (2003) Bacterial host strains

685 that support replication of somatic coliphages. Antonie Van Leeuwenhoek 83: 305-315.

686 Murphy, J., Klumpp, J., Mahony, J., O'Connell-Motherway, M., Nauta, A., and van

687 Sinderen, D. (2014) Methyltransferases acquired by lactococcal 936-type phage provide

688 protection against restriction endonuclease activity. BMC Genomics 15: 831.

689 Ohno, S., Okano, H., Tanji, Y., Ohashi, A., Watanabe, K., Takai, K., and Imachi, H.

690 (2012) A method for evaluating the host range of bacteriophages using phages

691 fluorescently labeled with 5-ethynyl-2 '-deoxyuridine (EdU). App Microbiol Biotechnol

692 95: 777-788.

693 Olsen, G.J., Lane, D.J., Giovannoni, S.J., Pace, N.R., and Stahl, D.A. (1986) Microbial

694 ecology and evolution: a ribosomal RNA approach. Annu Rev Microbiol 40: 337-365.

31 695 Paterson, S., Vogwill, T., Buckling, A., Benmayor, R., Spiers, A.J., Thomson, N.R. et al.

696 (2010) Antagonistic coevolution accelerates molecular evolution. Nature 464: 275-278.

697 Pepin, K.M., Domsic, J., and McKenna, R. (2008) Genomic evolution in a virus under

698 specific selection for host recognition. Infect Genet Evol 8: 825-834.

699 Rakhuba, D.V., Kolomiets, E.I., Dey, E.S., and Novik, G.I. (2010) Bacteriophage

700 receptors, mechanisms of phage adsorption and penetration into host cell. Pol J Microbiol

701 59: 145-155.

702 Rappé, M.S., Connon, S.A., Vergin, K.L., and Giovannoni, S.J. (2002) Cultivation of the

703 ubiquitous SAR11 marine bacterioplankton clade. Nature 418: 630-633.

704 Rohwer, F., Prangishvili, D., and Lindell, D. (2009) Roles of viruses in the environment.

705 Environ Microbiol 11: 2771-2774.

706 Roux, S., Hallam, S., Woyke, T., and Sullivan, M.B. (2015) Viral dark matter and virus–

707 host interactions resolved from publicly available microbial genomes. eLife 4: e08490.

708 Roux, S., Hawley, A.K., Beltran, M.T., Scofield, M., Schwientek, P., Stepanauskas, R. et

709 al. (2014) Ecology and evolution of viruses infecting uncultivated SUP05 bacteria as

710 revealed by single-cell- and meta- genomics. eLife 3: e03125.

711 Sambrook, J., Fritsch, E.F., and Maniatis, T. (1989) Molecular cloning: a laboratory

712 manual: Cold Spring Harbor Laboratory Press.

713 Samson, J.E., Magadan, A.H., Sabri, M., and Moineau, S. (2013) Revenge of the phages:

714 defeating bacterial defences. Nat Rev Microbiol 11: 675-687.

715 Schadt, E.E., Banerjee, O., Fang, G., Feng, Z., Wong, W.H., Zhang, X. et al. (2013)

716 Modeling kinetic rate variation in third generation DNA sequencing data to detect

717 putative modifications to DNA bases. Genome Res 23: 129-141.

32 718 Sillankorva, S., Oliveira, D., Moura, A., Henriques, M., Faustino, A., Nicolau, A., and

719 Azeredo, J. (2011) Efficacy of a broad host range lytic bacteriophage against E. coli

720 adhered to urothelium. Curr Microbiol 62: 1128-1132.

721 Silva, Y.J., Costa, L., Pereira, C., Mateus, C., Cunha, A., Calado, R. et al. (2014) Phage

722 therapy as an approach to prevent Vibrio anguillarum infections in fish larvae production.

723 Plos One 9: e114197.

724 Simu, K., and Hagström, Å. (2004) Oligotrophic bacterioplankton with a novel singel-

725 cell life strategy. Appl Environ Microbiol 70: 2445-2451.

726 Snyder, L. (1995) Phage-exclusion enzymes: a bonanza of biochemical and cell biology

727 reagents? Mol Microbiol 15: 415-420.

728 Stenholm, A.R., Dalsgaard, I., and Middelboe, M. (2008) Isolation and characterization

729 of bacteriophages infecting the fish pathogen Flavobacterium psychrophilum. Appl

730 Environ Microbiol 74: 4070-4078.

731 Sullivan, M.B., Waterbury, J.B., and Chisholm, S.W. (2003) Cyanophages infecting the

732 oceanic cyanobacterium Prochlorococcus. Nature 424: 1047-1051.

733 Sun, L., Young, L.N., Zhang, X., Boudko, S.P., Fokine, A., Zbornik, E. et al. (2013)

734 Icosahedral bacteriophage PhiX174 forms a tail for DNA transport during infection.

735 Nature 505: 432-435.

736 Suttle, C.A. (2005) Viruses in the sea. Nature 437: 356-361.

737 Suttle, C.A. (2007) Marine viruses - major players in the global ecosystem. Nat Rev

738 Microbiol 5: 801-812.

739 Tadmor, A.D., Ottesen, E.A., Leadbetter, J.R., and Phillips, R. (2011) Probing individual

740 environmental bacteria for viruses by using microfluidic digital PCR. Science 333: 58-62.

33 741 Thingstad, T.F., Pree, B., Giske, J., and Vage, S. (2015) What difference does it make if

742 viruses are strain-, rather than species-specific? Front Microbiol 6: 320.

743 Thingstad, T.F., Våge, S., Storesund, J.E., Sandaa, R.A., and Giske, J. (2014) A

744 theoretical analysis of how strain-specific viruses can control microbial species diversity.

745 Proc Natl Acad Sci U S A 111: 7813-7818.

746 Travers, K.J., Chin, C.S., Rank, D.R., Eid, J.S., and Turner, S.W. (2010) A flexible and

747 efficient template format for circular consensus sequencing and SNP detection. Nucleic

748 Acids Res 38: e159.

749 Uchiyama, J., Takemura, I., Satoh, M., Kato, S., Ujihara, T., Akechi, K. et al. (2011)

750 Improved adsorption of an Enterococcus faecalis bacteriophage phiEF24C with a

751 spontaneous point mutation. PLoS ONE 6: e26648.

752 Waterbury, J.B., and Valois, F.W. (1993) Resistance to co-occurring phages enables

753 marine Synechococcus communities to coexist with cyanophages abundant in seawater.

754 Appl Environ Microbiol 59: 3393-3399.

755 Weinbauer, M.G. (2004) Ecology of prokaryotic viruses. FEMS Microbiol Rev 28: 127-

756 181.

757 Weitz, J.S., Poisot, T., Meyer, J.R., Flores, C.O., Valverde, S., Sullivan, M.B., and

758 Hochberg, M.E. (2013) Phage-bacteria infection networks. Trends Microbiol 21: 82-91.

759 Wichels, A., Biel, S.S., Gelderblom, H.R., Brinkhoff, T., Muyzer, G., and Schütt, C.

760 (1998) Bacteriophage diversity in the North Sea. Appl Environ Microbiol 64: 4128-4133.

761 Wommack, K.E., and Colwell, R.R. (2000) Virioplankton: viruses in aquatic ecosystems.

762 Microbiol Mol Biol Rev 64: 69-114.

763

34 764 Tables

765 Table 1. Methylations detected in the bacterial and phage genomes. Modified base marked bold.

Organis motif string modification # detected # genome Mean Mean Mean Objective m type Score IpdRatio Coverage Score #4 No modifications detected #13 GAGNNNNNRTGT1 m6A 290 315 59.2 7.1 31.9 15936.7 #13 ACAYNNNNNCTC1 m6A 288 315 57.4 6.1 32.1 15248.3 #13 AGGAAT m6A 3686 4071 57.5 5.5 32.4 193870.1 #17 ACAGNNNNNGGA2 m6A 471 479 66.9 8.6 34.2 31052.7 #17 TCCNNNNNCTGT2 m4C 302 479 43.3 3.1 37.9 8640.6 #17 GAAGNNNNNGRTY3 m6A 717 751 59.8 7.8 33.1 41106.7 #17 RAYCNNNNNCTTC3 m6A 707 751 59.3 5.7 33.1 39710.5 #17 CTAGNNNNRTC4 m6A 166 566 41.4 5.0 37.8 2277.2 #17 GAYNNNNCTAG4 m6A 121 566 40.9 3.8 38.7 1235.0 #18 CAGNNNNNRTATT5 m6A 508 511 81.9 7.3 49.9 41391.1 #18 AATAYNNNNNCTG5 m6A 507 511 81.3 6.4 49.8 40923.5 #18 CCAGTV modified_base 608 2935 40.0 2.9 47.4 5898.4 #18 CCTGGDY m4C 175 1014 41.2 2.6 53.6 1482.5 φ 4:1 CTYCAG m6A 169 169 763.9 8.1 554.3 129104.0 φ 4:1_13 CTYCAG m6A 169 169 720.7 7.8 531.6 121791.0 φ4:1_18 CTYCAG m6A 169 169 648.5 7.3 532.8 109595.0 φ17:2 CTYCAG m6A 170 171 654.2 7.5 522.9 110628.5 φ17:2_18 CTYCAG m6A 171 171 686.2 7.5 549.0 117347.0 766 1-5Marks motif pairs.

35 767 Figure legends

768 Figure 1. Efficiency of infection for 38 Cellulophaga phages (rows) against 19 C.

769 baltica strains (columns). Infection is shown as relative number of Plaque forming

770 units (PFU) ml-1 compared to the infection with highest efficiency per phage, where

771 darker color corresponds to higher efficiency of infection. The phages are divided into

772 their respective families and genera according to (Holmfeldt et al., 2013) and

773 (Holmfeldt et al., 2007) for φ39:2, φ40:2, φ50:1, φ73:1, and φ19:4. Phages are

774 arranged with the genus that infects the highest number of bacteria at the top and

775 smallest number of bacteria at the bottom. Bacteria are arranged with the bacterium

776 that can be infected with the highest number of phages to the left and the smallest

777 number of phages to the right. Phage families and genera displayed on the left; M)

778 Myoviridae, P) Podoviridae, S) Siphoviridae, Mi) Microviridae, and ss) unclassified

779 single stranded DNA phage. Sampling location (bacteria only) and year is presented

780 at the bottom and to the left. All phages were sampled at the same location, Öresund.

781 Host location correspond to O) Öresund, S) Svaneke, E) Ellekilde Hage, and B) Baltic

782 Proper. 1-6 represent phages of the same population based on >95% ANI on >80% of

783 genes.

784

785 Figure 2. Genome comparison of ancestral Cba4 phages φ4:1 and φ17:2. A)

786 Alignment using progressive Mauve where a similarity plot gives the degree of

787 similarity between the genomes and the height of the plot is proportional to the

788 average nucleotide identity. Below each similarity plot colored annotated genomes

789 are presented. B) Pie chart showing the proportion of shared genes at different

790 nucleotide identity levels.

791

36 792 Figure 3. Efficiency of infection of the ancestral and passaged phages presented as

793 percentage of the efficiency of infection on the original hosts. A) Example of

794 shuffling experiment of phage φ4:1 infecting hosts #4 and #13 with efficiency of

795 infection given as percentage of efficiency of infection on host #4. Efficiency of

796 infection for B) phages originating from phage φ4:1 and C) phages originating from

797 phage φ17:2. Bar graphs show mean of three replicates of plaque assays of purified

798 phages and two replicates of shuffled phages ± standard deviation.

799

800 Figure 4. One-step growth curves of A) φ4:1 infecting #4, B) φ4:1 infecting #13, C)

801 φ4:1 infecting #18, D) φ4:1_13 infecting #4, E) φ4:1_13 infecting #13, F) φ4:1_18

802 infecting #4, and G) φ4:1_18 infecting #18. Phage abundance is presented as

803 normalized to average total phage abundance before burst. Black squares represent

804 total phage abundance and grey squares represent free phage abundance. Burst size

805 and latent period (in h) ± standard deviation (n=3) is presented in each graph.

806

807 Supplemental information

808 Supplemental Table 1. Efficiency of plating for 44 Cellulophaga phages (rows)

809 against 19 C. baltica strains (columns). Infection is shown as relative number of PFU

810 ml-1 compared to the infection on the host with highest efficiency per phage. The

811 rows and columns are structured as in Figure 1 with the addition of the 6 phages

812 investigated in Holmfeldt and Howard-Varona et al., (2014).

813

814 Supplemental Table 2. Genes shared (%) at >95% average nucleotide identity

815 among closely related Cellulophaga phages divided by genus. A) Cba184, B)

816 Cba131, C) CbaSM, D) Cba41, E) Cba 181.

37 817

818 Supplemental Table 3. Synonymous and non-synonymous DNA changes in the

819 genomes for two phages within the genus Cba481.

38