Europe PMC

This website requires cookies, and the limited processing of your personal data in order to function. By using the site you are agreeing to this as outlined in our privacy notice and cookie policy.

Abstract 


Bacteriophages (phages) are critical players in the dynamics and function of microbial communities and drive processes as diverse as global biogeochemical cycles and human health. Phages tend to be predators finely tuned to attack specific hosts, even down to the strain level, which in turn defend themselves using an array of mechanisms. However, to date, efforts to rapidly and comprehensively identify bacterial host factors important in phage infection and resistance have yet to be fully realized. Here, we globally map the host genetic determinants involved in resistance to 14 phylogenetically diverse double-stranded DNA phages using two model Escherichia coli strains (K-12 and BL21) with known sequence divergence to demonstrate strain-specific differences. Using genome-wide loss-of-function and gain-of-function genetic technologies, we are able to confirm previously described phage receptors as well as uncover a number of previously unknown host factors that confer resistance to one or more of these phages. We uncover differences in resistance factors that strongly align with the susceptibility of K-12 and BL21 to specific phage. We also identify both phage-specific mechanisms, such as the unexpected role of cyclic-di-GMP in host sensitivity to phage N4, and more generic defenses, such as the overproduction of colanic acid capsular polysaccharide that defends against a wide array of phages. Our results indicate that host responses to phages can occur via diverse cellular mechanisms. Our systematic and high-throughput genetic workflow to characterize phage-host interaction determinants can be extended to diverse bacteria to generate datasets that allow predictive models of how phage-mediated selection will shape bacterial phenotype and evolution. The results of this study and future efforts to map the phage resistance landscape will lead to new insights into the coevolution of hosts and their phage, which can ultimately be used to design better phage therapeutic treatments and tools for precision microbiome engineering.

Free full text 


Logo of plosbiolLink to Publisher's site
PLoS Biol. 2020 Oct; 18(10): e3000877.
Published online 2020 Oct 13. https://doi.org/10.1371/journal.pbio.3000877
PMCID: PMC7553319
PMID: 33048924

High-throughput mapping of the phage resistance landscape in E. coli

Vivek K. Mutalik, Conceptualization, Data curation, Formal analysis, Funding acquisition, Investigation, Methodology, Project administration, Resources, Supervision, Validation, Visualization, Writing – original draft, Writing – review & editing,1,2,* Benjamin A. Adler, Formal analysis, Investigation, Methodology, Validation, Visualization, Writing – review & editing,2,3 Harneet S. Rishi, Formal analysis, Investigation, Methodology, Resources, Writing – review & editing,4,5 Denish Piya, Investigation, Methodology, Validation, Writing – review & editing,2,3 Crystal Zhong, Investigation,1 Britt Koskella, Investigation, Methodology, Writing – review & editing,6 Elizabeth M. Kutter, Methodology, Resources,7 Richard Calendar, Resources, Writing – review & editing,8 Pavel S. Novichkov, Data curation, Formal analysis, Investigation, Methodology, Software,1 Morgan N. Price, Data curation, Formal analysis, Investigation, Methodology, Software, Writing – review & editing,1 Adam M. Deutschbauer, Formal analysis, Investigation, Methodology, Resources, Validation, Writing – review & editing,1,2,9 and Adam P. Arkin, Conceptualization, Data curation, Formal analysis, Funding acquisition, Investigation, Methodology, Project administration, Supervision, Visualization, Writing – review & editing1,2,3,4,5,*
J. Arjan G. M. de Visser, Academic Editor

Associated Data

Supplementary Materials
Data Availability Statement

Abstract

Bacteriophages (phages) are critical players in the dynamics and function of microbial communities and drive processes as diverse as global biogeochemical cycles and human health. Phages tend to be predators finely tuned to attack specific hosts, even down to the strain level, which in turn defend themselves using an array of mechanisms. However, to date, efforts to rapidly and comprehensively identify bacterial host factors important in phage infection and resistance have yet to be fully realized. Here, we globally map the host genetic determinants involved in resistance to 14 phylogenetically diverse double-stranded DNA phages using two model Escherichia coli strains (K-12 and BL21) with known sequence divergence to demonstrate strain-specific differences. Using genome-wide loss-of-function and gain-of-function genetic technologies, we are able to confirm previously described phage receptors as well as uncover a number of previously unknown host factors that confer resistance to one or more of these phages. We uncover differences in resistance factors that strongly align with the susceptibility of K-12 and BL21 to specific phage. We also identify both phage-specific mechanisms, such as the unexpected role of cyclic-di-GMP in host sensitivity to phage N4, and more generic defenses, such as the overproduction of colanic acid capsular polysaccharide that defends against a wide array of phages. Our results indicate that host responses to phages can occur via diverse cellular mechanisms. Our systematic and high-throughput genetic workflow to characterize phage-host interaction determinants can be extended to diverse bacteria to generate datasets that allow predictive models of how phage-mediated selection will shape bacterial phenotype and evolution. The results of this study and future efforts to map the phage resistance landscape will lead to new insights into the coevolution of hosts and their phage, which can ultimately be used to design better phage therapeutic treatments and tools for precision microbiome engineering.

Introduction

Bacterial viruses, or bacteriophages (phages), are obligate parasites that infect specific bacterial strains. Phages represent the most abundant biological entities on earth and are key ecological drivers of microbial community dynamics, activity, and adaptation, thereby impacting environmental nutrient cycles, agricultural output, and human and animal health [18]. Despite nearly a century of pioneering molecular work, the mechanistic insights into phage specificity for a given host, infection pathways, and the breadth of bacterial responses to different phages have largely focused on a handful of individual bacterium-phage systems [913]. Bacterial sensitivity/resistance to phages is typically characterized using phenotypic methods such as cross-infection patterns against a panel of phages [1427] or by whole-genome sequencing of phage-resistant mutants [2833]. As such, our understanding of bacterial resistance mechanisms against phages remains limited, and the field is therefore in need of improved methods to characterize phage-host interactions, determine the generality and diversity of phage resistance mechanisms in nature, and identify the degree of specificity for each bacterial resistance mechanism across diverse phage types [13,25,26,3448].

Unbiased and comprehensive genetic screens that are easily transferable and scalable to new phage-host combinations would be highly valuable for obtaining a deeper understanding of phage infection pathways and phage resistance phenotypes [4954]. Such genome-scale studies applied to different phage-host combinations have the unique potential to identify commonalities or differences in phage resistance patterns and mechanisms [18,25,28,5557]. There have been few attempts to use genetic approaches for studying genome-wide host factors essential in phage infection. These loss-of-function (LOF) genetic screens broadly use bacterial saturation mutagenesis [49,54,5861] or an arrayed library of single-gene deletion strains for studying phage-host interactions [50,51,53,62,63]. Consequently, these studies have generally involved laborious experiments on relatively few phages and their hosts, and scaling the approach to characterize hundreds of phages is challenging.

To overcome these technological limitations, we have developed three high-throughput genetic technologies that enable fast, economical, and quantitative genome-wide screens for gene function, which are suitable for discovering host genes critical for phage infection and bacterial resistance. Random barcode transposon site sequencing (RB-TnSeq) allows genome-wide insertion mutagenesis leading to LOF mutations [64]; a pooled CRISPR interference (CRISPRi) approach, which allows partial inhibition of gene function via transcriptional inhibition [65]; and dual-barcoded shotgun expression library sequencing (Dub-seq) [66], which queries the effects of gene overexpression. All three technologies can be assayed across many conditions at low cost, as RB-TnSeq and Dub-seq use randomized DNA barcodes to assay strain abundance (BarSeq [67]), whereas quantification of the pooled CRISPRi strains only requires deep sequencing of the guide sequences.

In RB-TnSeq, genome-wide transposon insertion mutant libraries labeled with unique DNA barcodes are generated, and next-generation sequencing methods are used to map the transposon insertions and DNA barcode at loci in genomes. Although RB-TnSeq can be applied on a large scale across multiple bacteria through barcode sequencing [68], it is limited to nonessential genes. Partial LOF assays such as CRISPRi use a catalytic null mutant of the Cas9 protein (dCas9) guided by chimeric single-guide RNA (sgRNA) to programmably knock down gene expression, thereby allowing the probing of all genes (including essential genes) and more precise targeting of intergenic regions [65]. We have recently applied CRISPRi technology to systematically query the importance of approximately 13,000 genomic features of E. coli in different conditions [69]. CRISPRi technology has been extended to different organisms to study essential genes [7078] and has been recently applied to E. coli to uncover host factors involved in T4, 186, and λ phage infection [52]. Lastly, Dub-seq uses shotgun cloning of randomly sheared DNA fragments of a host genome on a dual-barcoded replicable plasmid and next-generation sequencing to map the barcodes to the cloned genomic regions. The barcode association with genomic fragments and genes contained on those fragments enables a parallelized gain-of-function (GOF) screen, as demonstrated in our recent work [66]. In contrast to LOF genetic screens, GOF screens to study gene dosage effects on phage resistance have not been broadly reported, except for a recent work on T7 phage using an E. coli single-gene overexpression library [51]. There are indications that enhanced gene dosage can be an effective way to search for dominant-negative mutants, antisense RNAs, or other regulatory genes that may block phage growth cycle [9,10,13,51,56,79]. Such GOF screens, when applied in a high-throughput format across diverse phages, may yield novel mechanisms of phage resistance that LOF screens may not uncover.

In this study, we employ these three technologies (RB-TnSeq, Dub-seq, and CRISPRi) as a demonstration of “portable” and “scalable” platforms for rapidly probing phage-host interaction mechanisms. Using E. coli K-12 strain and 14 diverse double-stranded DNA (dsDNA) phages, we show that our screens successfully identify known receptors and other host factors important in infection pathways, and they also yield additional novel loci that contribute to phage resistance. We validate some of these new findings by deleting or overexpressing individual genes and quantifying fitness in the presence of phage. Additionally, we used RB-TnSeq and Dub-seq to compare similarity and distinctiveness in phage resistance displayed by E. coli strains K-12 and BL21. The comparison of two historical lineages of E. coli allowed us to examine how strain-level divergence of genotype can lead to differential susceptibility in phage resistance. Finally, we discuss the implications and extensibility of our approaches and findings to other bacteria-phage combinations and how these datasets can provide a foundation for understanding phage ecology and engineering phage for therapeutic applications.

Results and discussion

Mapping genetic determinants of phage resistance using high-throughput LOF and GOF methods

Despite E. coli being a well-studied model organism [80,81], there are significant knowledge gaps regarding gene function [82] and phage interaction mechanisms [26,27,47,83,84]. Different serotypes of E. coli are also important pathogens with significant global threat and are crucial players in specific human-relevant ecologies [8587], leading to the question of whether strain variation is also important in predicting the response to phage-mediated selection or whether the mechanisms are likely to be conserved. Such mechanistic information is unavailable for not only pathogenic E. coli strains but also for widely used laboratory nonpathogenic strains such as E. coli K-12 and BL21 [8890]. Historically, both E. coli K-12 and B (ancestor of BL21) strains have been used in disparate phage studies and have provided foundational knowledge on phage physiology and growth [10,9193], though phage studies on E. coli BL21 are limited relative to those in K-12. Both K-12 BW25113 and BL21 lack functional CRISPR machinery and type I restriction-modification (R-M) system that function as common antiphage systems, and both strains are unique in their genomic content, physiology, and growth characteristics [8890], thereby serving as a valuable reference for comparing closely related host responses to the same phage selection pressures. To demonstrate the efficacy of our approaches to characterize phage resistance mechanisms and to compare their similarity and differences between two closely related laboratory model strains, we first applied high-throughput genetic screens to E. coli K-12 strains (BW25113 and MG1655) and then extended it to E. coli BL21 strain (Fig 1).

An external file that holds a picture, illustration, etc.
Object name is pbio.3000877.g001.jpg
Overview of high-throughput genome-wide screens.

We used barcoded LOF technologies (RB-TnSeq and CRISPRi) and a GOF technology (Dub-seq) in E. coli K-12 (BW25113 and MG1655) to screen for host factors important in phage infection and resistance. In E. coli BL21, we performed RB-TnSeq and Dub-seq (but not CRISPRi). We sourced 14 diverse E. coli phages with dsDNA genomes, belonging to Myoviridae, Podoviridae, and Siphoviridae families, and performed pooled fitness screens in both planktonic and solid agar formats. Disruption or overexpression of certain genes provide fitness to host in the presence of phages, and we monitor these changes by quantifying the abundance of the DNA barcode or sgRNA associated with each strain. The individual strain abundances are then converted to gene fitness scores (normalized log2 change in the abundance of mutants in that gene). CRISPRi, CRISPR interference; dsDNA, double-stranded DNA; Dub-seq, dual-barcoded shotgun expression library sequencing; GOF, gain-of-function; LOF, loss-of-function; MOI, multiplicity of infection; RB-TnSeq, random barcode transposon site sequencing; sgRNA, single-guide RNA.

We used a previously constructed E. coli K-12 BW25113 RB-TnSeq library [64] and defined an E. coli K-12 MG1655 CRISPRi library [69]. To study host gene dosage and overexpression effects on phage resistance, we used a prior reported GOF Dub-seq library of E. coli K-12 BW25113 [66]. We sourced 14 diverse E. coli phages with dsDNA genomes, belonging to Myoviridae, Podoviridae, and Siphoviridae families (within the order Caudovirales) (Fig 1). These phages include 11 canonical and well-studied coliphages, each having overlapping but distinct mechanisms of host recognition, entry, replication, and host lysis [94], and two recently reported phages (CEV1 and CEV2), which are known to kill pathogenic shiga toxin–producing E. coli (STEC, O157:H7), and one novel coliphage (LZ4). These 14 phages include T-series phages (T2, T3, T4, T5, T6, T7), N4, 186, λcI857, P1vir, P2, and novel isolates of T-like phages (T6-like LZ4, STEC infecting T4-like CEV1 and T5-like CEV2).

The barcoded LOF or GOF libraries were grown competitively in a single-pot assay and placed under phage-mediated selection. Any variation in fitness resulting from disruption or overexpression of specific genes in the presence of phage is therefore exposed to selection, allowing for evolutionary change within the population. We monitor these changes by quantifying the abundance of the DNA barcode associated with each strain. The individual strain abundances are then converted to gene fitness scores, which we define as the normalized log2 change in the abundance of mutants in that gene (Methods). A positive gene fitness score (“positive hit”) indicates the strain with deletion or overexpression of a gene realized an increase in relative fitness in the presence of a particular phage (i.e., strains with these genetic changes are more resistant to the phage and are enriched in our phage selection assay). A positive fitness score for a gene in our LOF assay indicates that its encoded product (for example, a phage receptor) is needed for successful phage infection, whereas a positive fitness score for a gene in our GOF assay indicates that its encoded product (for example, a repressor of phage receptor) prevents the phage infection cycle. Negative fitness values, which suggest reduced relative fitness, indicate that the gene(s) disruption or overexpression results in these strains being more sensitive to the phage than the typical strain in the library. Lastly, fitness scores near zero indicate no fitness change for the mutated or overexpressed gene under the assayed condition. Because of the strong selection of phage infection, we anticipated (and indeed observed) that genetically modified strains in our libraries with resistance to a phage would lead to very strong positive fitness values. Although these very strong positive phenotypes are readily interpretable, one consequence is that strains with relatively poor or neutral fitness scores will be swept from the population. Thus, intermediate resistance factors to phage infection will have similar low or negative fitness values as a neutral mutant. As such, most of our focus in this study is on the strong positive fitness scores (Methods). Challenges in identifying intermediate phage-resistant mutants in the presence of highly resistant phage receptor mutations are well appreciated in the field [95].

RB-TnSeq identifies known receptors and host factors for all 14 phages

To perform genome-wide transposon-based LOF assays, we recovered a frozen aliquot of the E. coli K-12 RB-TnSeq library in lysogeny broth (LB [96]) to mid-log phase, collected a cell pellet for the “start,” and used the remaining cells to inoculate an LB culture supplemented with different dilutions of a phage in SM buffer. After 8 hr of phage infection in planktonic cultures, we collected the surviving phage-resistant strains or “end” samples (Methods). We also repeated these fitness assays on solid media by plating the library post phage adsorption, incubating the plates overnight, and collecting all surviving phage-resistant colonies. We hypothesized that, given the spatial structure and possibility of phage refuges, fitness experiments on solid media might provide a less stringent selection environment than in liquid pooled assays, such that less fit survivors could potentially be detected. For all assays, recovered genomic DNA from surviving strains was used as templates to PCR amplify the barcodes for sequencing (Methods). The strain fitness and gene fitness scores were then calculated as previously described [64].

In total, we performed 68 RB-TnSeq assays across 14 phages at varying multiplicity of infection (MOI) and 9 no-phage control assays (Methods). For planktonic assays, the gene fitness scores were reproducible across different phage MOIs (Fig 2A), with a median pairwise correlation of 0.90. Because of stronger positive selection in the presence of phages (relative to our typical RB-TnSeq fitness assays with stress compounds and in defined growth media [64,68]), to determine reliable effects across experiments we used stringent filters (gene fitness score ≥ 5, t-like statistic ≥ 5, and estimated standard error ≤ 2) (Methods). Across all replicate experiments, we identified 354 high-scoring hits, which represent 52 unique genes and 133 unique gene-phage combinations (some genes were linked with more than one phage) (S1 Table). In all phage experiments, there was at least one gene with a high-confidence effect. In the presence of some phages (for example, T5 and T6), we observed enrichment of strains with disruption in only one gene (encoding the phage receptor), whereas other phages (such as 186 and λ cI857 phages) showed enrichment with disruptions in multiple genes (Fig 2A).

An external file that holds a picture, illustration, etc.
Object name is pbio.3000877.g002.jpg
Heatmap of E. coli K-12 RB-TnSeq data for 14 dsDNA phages at different MOI.

(A) Top 36 genes with high-confidence effects and a gene fitness score of ≥6.5 in at least one phage assay are shown. The pooled fitness assays performed on solid agar plates are shown as stars. Yellow boxes highlight genes that encode known receptors for the marked phages. The underlying data for this figure can be found in S1 Data. (B) Schematic of E. coli K-12 LPS structure with associated enzymes involved in LPS core biosynthesis. Top-scoring candidates in the presence of a particular phage (at any MOI) are highlighted in purple by associating each enzymatic step with phages. dsDNA, double-stranded DNA; LPS, lipopolysaccharide; MOI, multiplicity of infection; RB-TnSeq, random barcode transposon site sequencing.

Phage infection initiates through an interaction of the phage with any bacterial cell surface–exposed molecules (“receptors”), which could be lipopolysaccharide (LPS) moieties, peptidoglycan, teichoic acids, capsules, and proteinaceous components. Any changes in levels or structure of these receptors can compromise efficient phage infection, thereby leading to an improvement in host fitness in the presence of phages. Consequently, receptor mutants are the most commonly found candidates that display phage resistance [49,50,53,58,95,97]. To confirm the validity of our approach, we looked for receptors recognized by many of the canonical phages used in this study for which there are published data available [50,51,9193,98101]. Indeed, we found high fitness scores (fitness score > 10, corresponding to >1,000-fold enrichment of transposon mutants) for multiple known phage receptors (Fig 2, S1S3 Tables). These included genes coding for protein receptors such as fadL (long-chain fatty acid transporter) for T2, ompC (outer membrane porin C) for T4, fhuA (ferrichrome outer membrane transporter) for T5, tsx (nucleoside-specific transporter) for T6, nfrA and nfrB (unknown function) for N4, and lamB (maltose outer membrane transporter) for λ. fhuA showed high fitness scores in the presence of CEV2 phage, indicating CEV2 has a similar receptor requirement as T5 [102]. We find tsx as the top scorer in the presence of novel LZ4 phage, thus appearing to have similar receptor requirement as T6 phage [103]. In addition to protein receptors, we also identified a few high-scoring genes that are known to interfere or regulate the expression of receptors, thereby impacting phage infection. For example, deletion of the EnvZ/OmpR two-component system involved in the regulation of ompC and gene products involved in regulation of lamB expression (cyaA, malI, malT) all show high fitness scores in the presence of T4 and λ phages, respectively (Fig 2A).

For phages utilizing LPS as their receptors, we found top scores for gene mutations within the waa gene cluster, which codes for enzymes involved in LPS core biosynthesis (Fig 2). For example, waaC, waaD, waaE, and lpcA/gmhA were the top scorers for T3 and T7 phages, whereas waaC, waaD, waaE, waaF, waaJ, waaO, waaQ, and lpcA/gmhA showed high fitness in the presence of P1, P2, and 186 phages (Fig 2A and 2B). Even though P1 and P2 phages have been studied for decades, the host factors important in their infection cycle are not fully characterized [93,104]. Our results show that all LPS core components are essential for an efficient P1, P2, and 186 phage infection. CEV1 phage seems to require both outer membrane porin OmpF and LPS core for efficient infection. In addition to ompF and its regulator the envZ/ompR two-component signaling system, a number of genes involved in the LPS core biosynthesis pathway (waaC, waaD, waaE, waaF, waaG, waaP, galU, and lpcA/gmhA) and a regulator of genes involved in biosynthesis, assembly, and export of LPS core (rfaH) all showed high fitness scores (>10) in the presence of CEV1 phage. Among other top-scoring hits, genes encoding a putative L-valine exporter subunit (ygaH) and a diguanylate cyclase (DGC), dgcJ, showed stronger fitness in the presence of N4 phage. Both YgaH and DgcJ were not previously known to be involved in N4 phage resistance. Finally, igaA/yrfF, which encodes a negative regulator of the Rcs phosphorelay pathway, shows strong fitness scores against eight phages, indicating its importance to general phage resistance. Though igaA is an essential gene in E. coli [105], our RB-TnSeq library contained 9 disruptions in igaA’s cytoplasmic domain, and these strains seem to tolerate the disruption (S1 Fig). It is known that the Rcs signal transduction pathway functions as an envelope stress response system that monitors cell surface composition and regulates a large number of genes involved in diverse functions including colanic acid synthesis and biofilm formation [106].

Compared to assays performed in planktonic cultures, a few additional gene mutants showed stronger fitness effects on plate assays. For example, trxA, encoding thioredoxin 1, is known to be essential for T7 phage propagation and scored high in the solid plate assays but not in our planktonic growth assays (Fig 2). Thioredoxin 1 is a processivity factor for T7 RNA polymerase, and it is reported that T7 phage can bind and lyse a trxA deletion strain, though T7 phage propagation is severely compromised [107]. Similarly, we observed higher fitness scores for galU in the presence of T4, P1, and λ phages, and dnaJ in the presence of 186, P1, and λ phages, on plates but not in broth. GalU catalyzes the synthesis of UDP-D-glucose, a central precursor for synthesis of cell envelope components, including LPS core, and it is known that the growth of P1 phage is compromised on a galU mutant [104]. dnaJ codes for a chaperone protein and dnaJ insertion mutants are known to inhibit the growth of λ phage [108110]. We also observed higher fitness scores for a number of genes involved in LPS biosynthesis (waaC, waaD, waaE, waaF, lpcA) in the presence of T4, LZ4, and λ phages when grown on solid plates. These results suggest that LPS components either play an important role in an efficient phage infection cycle or these LPS truncations lead to a destabilized membrane and probably decrease outer membrane protein levels via envelope stress response [106,111114]. A detailed description about many of the genes we identified in this study is provided in S1 Text. In summary, our results indicate how the abiotic environment can have an important influence on the host fitness and susceptibility and the type of resistance mechanism selected in the presence of different phages [115119]. While this manuscript was under review, a new work using transposon sequencing to screen receptors for T2, T4, T6, and T7 phages came out [120]. Our results are largely in agreement with this new study, with a few differences that point to difference in the mutant library size (17,100 insertions in 3,253 bacterial genes compared to our RB-TnSeq library with 152,018 barcoded insertions in 3,728 genes).

Overall, our RB-TnSeq LOF screen provided a number of top hits that agree with earlier reports and also yielded a set of novel genes whose role in phage infection was not previously known (S1S3 Tables). Later in this manuscript, we describe follow-up experiments with 18 of these top-scoring hits from the RB-TnSeq screen to validate their role in phage resistance.

A CRISPRi screen provides a deeper view of phage resistance determinants

We next employed a rationally designed E. coli K-12 MG1655 genome-wide CRISPRi library approach to look for bacterial essential and nonessential genes and genomic regulatory regions important in phage infection and to provide a complementary genetic screen to RB-TnSeq. This CRISPRi library comprises multiple sgRNAs targeting annotated genes, promoter regions, and transcription factor binding sites, a total of 13,000 target regions [69] (Methods). By directing dCas9 to different genomic regions via unique sgRNAs, CRISPRi enables the interrogation of genic and nongenic regions. For the CRISPRi assays, we recovered a frozen aliquot of the library, which was back diluted in fresh media, supplemented with dCas9 and gRNA inducers, and mixed with phage. We assayed T2, T3, T4, T5, T6, N4, 186, CEV1, CEV2, LZ4, and λ phages in planktonic cultures at an MOI of 1, recovered survivors after phage treatment for 3 hr, isolated plasmid DNA, and the variable gRNA region was PCR amplified and sequenced (Methods). During these pooled CRISPRi assays, strains that carry sgRNAs targeting a feature on E. coli genome important for phage infection (for example, a phage receptor) will increase in abundance and will have a positive fitness score. In total, we identified 542 genes (including 75 genes of unknown function), 94 promoter regions, and 44 transcription factor binding sites that show high fitness scores across all phages (sgRNAs with log2 fold-change ≥2 and false discovery rate [FDR]-adjusted p-value <0.05) (S4 Table, Methods).

To confirm our assay system correctly identifies host factors important in phage infection, we looked for genes that are known to impact phage infection and also are in agreement with high fitness scores in our RB-TnSeq results. Indeed, the top-scoring hits included sgRNAs that target both the genes coding for phage receptors and their promoter and transcription factor binding sites (Fig 3A, S4 Table). Our CRISPRi data also confirmed many of the top-scoring genes uncovered in RB-TnSeq screen (Fig 3A, S3 Table), thus validating the importance of these genes in specific phage infection pathways. For example, ompF was the top-scoring gene for CEV1 phage, and tsx showed high fitness scores in the presence of LZ4, whereas fhuA was the top scorer for phage CEV2. We also found that sgRNAs targeting dgcJ or its promoter showed high fitness scores in the presence of N4 phage, thereby confirming the RB-TnSeq data (Fig 2A). In agreement with RB-TnSeq data, we observed that knockdown of igaA yields resistance to 10/11 phages we assayed (Fig 3B–3D).

An external file that holds a picture, illustration, etc.
Object name is pbio.3000877.g003.jpg
CRISPRi fitness profiles of E. coli nonessential and essential genes in the presence of diverse phages.

(A) Heatmap of top-scoring nonessential genes across 11 phages. Yellow boxes highlight genes that encode phage receptors and are known to interfere with phage growth when down-regulated. Yellow stars indicate these data points are in agreement with RB-TnSeq results. (B) Heatmap of top-scoring essential genes across 11 phages. Yellow boxes highlight genes that are known to interfere with phage growth when down-regulated. (C) Box plot of fitness data for igaA-targeting sgRNAs across 11 different phages. Each data point in this plot is a specific sgRNA targeting igaA. (D) Genome browser plot for nudE-igaA locus with targeting sites for gRNAs and their fitness scores. The downward-facing triangles mean that the sgRNA targeted the nontemplate strand of the gene. Under each promoter, a vertical bar denotes the +1 for the promoter with the stem for the promoter starting at −60 relative to the transcription start site. The underlying data for this figure can be found in S1 Data. CRISPRi CRISPR interference; RB-TnSeq, random barcode transposon site sequencing; sgRNA, single-guide RNA.

Among disagreements between the RB-TnSeq and CRISPRi datasets, we found that our CRISPRi screen failed to return some of the highest-scoring genes uncovered in RB-TnSeq dataset. These include genes encoding the EnvZ/OmpR two-component system for T4 and CEV1 phages and YgaH and WecB for N4 phage. Here, we note that we have only one sgRNA targeting ygaH and no sgRNAs targeting wecB/nfrC region in our CRISPRi library. In addition to these genes, the contribution of LPS core biosynthesis genes in phage infection was less clear in our CRISPRi dataset. For example, both OmpF and LPS seem to be crucial for CEV1 infection from our RB-TnSeq dataset (Fig 2A), whereas the CRISPRi screen data showed high fitness score for ompF and not for all core LPS biosynthetic genes (Fig 3A). Nevertheless, we find high fitness scores for genes encoding LPS transport system (lptABC) and lipid A biosynthesis enzymes (Fig 3B). In summary, these results indicate that we might be missing few candidates in our CRISPRi screen (as compared to RB-TnSeq) because either some genes lack sufficient sgRNA coverage or even minimal expression of these genes is probably enough for phage infection.

One of the key advantages of CRISPRi is the ability to study the contribution of essential genes on phage infection. We found 11 essential genes (csrA, kdsC, lptA, lptB, lptC, lpxA, lpxC, nusG, secE, yejM, and tsf) that showed broad fitness advantage (fitness score ≥2 in more than one phage assay) when down-regulated (Fig 3B). None of these essential genes are present in our RB-TnSeq library, except for yejM, which has transposon insertions in the C-terminal portion (after 5 putative transmembrane helices) of the protein. The putative cardiolipin transporter encoded by yejM and its upstream neighbor yejL both show enhanced fitness in the presence of T3, T4, T6, CEV1, CEV2, and LZ4 phages in the CRISPRi dataset (Fig 3A and 3B). These genes have not been previously associated with phage resistance. Although the physiological role of cardiolipin is still emerging, it is known that cardiolipins play an important role in outer membrane protein translocation system and membrane biogenesis [121123]. A recent study showed that decreased cardiolipin levels activate Rcs envelope stress response [123]. Down-regulation of cardiolipin transport probably results in phage resistance via increased colanic acid biosynthesis, but further mechanistic studies are needed.

Although the fitness benefit phenotype conferred by most top-scoring essential genes in the presence of phages is challenging to interpret, some of these hits agree with previous work. For example, it is known that the down-regulation of genes involved in transcription antitermination (nusB, nusG) and Sec translocon subunit E (secE) shows high fitness in the presence of λ phage and is crucial for the phage growth cycle [124126]. lptABC, kdsC, and lpxAC are known to impact outer membrane biogenesis, LPS synthesis, and transport [113,114,127,128], whereas the RNA binding global regulator CsrA is known to be involved in carbohydrate metabolism and regulation of biofilm [129,130]. Down-regulation of these genes likely leads to pleiotropic effects leading to enhanced fitness in the presence of phages. Our CRISPRi screen also identified a number of E. coli tRNA-related genes showing enhanced fitness in the presence of diverse phages (S4 Table). How the down-regulation of host tRNA and tRNA modification genes impacts host fitness, phage growth, and infection cycle is not clear, although recent studies have shown increased aminoacyl-tRNA synthetase activities in the early phage infection cycle [131,132].

Finally, neither the RB-TnSeq nor CRISPRi screens found high scores for ompF in the presence of T2 phage and cmk in the presence of T7 phage (Figs (Figs2A2A and and3A).3A). Previous reports had indicated that OmpF serves as a secondary receptor to T2 (in addition to the primary receptor FadL) [133] and Cytidine monophosphate kinase (encoded by cmk) is an essential host factor in T7 phage infection [51]. We also did not observe significant fitness scores for host factors that may bias the rate of lysogeny of the two temperate phages (186 and P2), leading to resistance of those lysogens. Our genome-wide screens did not recapitulate these findings probably because these mutants provide an intermediate fitness in the presence of phages [134] and are swept from the population in the presence of highly resistant phage receptor mutations.

In summary, CRISPRi served as a complementary screening technology to RB-TnSeq, validating many RB-TnSeq hits, and provided an avenue to study the role of essential genes and nongenic regions on phage infection.

Dub-seq enables parallel mapping of multicopy suppressors of phage infection

To study the effect of host factor gene dosage or overexpression on phage resistance, we performed competitive fitness assays using the E. coli K-12 Dub-seq library (expressed in the E. coli K-12 strain) in the presence of phages. This library is made up of randomly sheared 3-kb genomic DNA of E. coli K-12 BW25113 cloned into a dual-barcoded vector with the copy number of approximately 15–20 [66]. Similar to pooled fitness assays with the RB-TnSeq library, we recovered a frozen aliquot of the library to mid-log phase, collected a cell pellet for the initial sample, and used the remaining cells to inoculate an LB culture supplemented with different dilutions of a phage in SM buffer (Methods). Similar to LOF RB-TnSeq assays, after 8 hr of phage infection in planktonic cultures (and overnight incubation in case of plate assays), we collected the surviving phage-resistant strains, isolated plasmid DNA, and sequenced the DNA barcodes (Methods). In total, we performed 67 genome-wide GOF fitness assays in the presence of 13 different phages at varying MOIs, both in planktonic and solid plate assay format and 4 no-phage control experiments (Methods). Here, one genome-wide GOF fitness assay encompasses an ensemble of barcoded vectors with sheared 3-kb genomic region of the whole E. coli K-12 BW25113. Overall we identified 233 high-scoring positive hits for the E. coli K-12 Dub-seq screen made up of 129 unique phages-gene combinations and found more than five Dub-seq hits per phage that confer positive growth benefit (fitness score ≥ 4, FDR of 0.7, Methods, S5 Table). A positive fitness score for a gene in our Dub-seq assay indicates that the overexpression of that gene interferes with successful phage infection.

The growth benefit phenotypes we observe in Dub-seq assays may not be simply due to overexpression of genes encoded on genomic fragments but rather be due to other potential effects such as increased gene copy number (gene dosage), or other indirect dominant-negative effects may be playing a role [135140]. For example, overexpression or higher copy number of a regulatory region might lower the effective concentration of a transcription factor important in the regulation of phage receptor expression and may lead to a phage-resistant strain. To confirm whether our method captures such host factors that control the expression of a phage receptor, we looked for known regulators (Fig 4). Acetylesterase Aes, transcription repressor Mlc, and mal regulon repressor MalY are known to reduce the expression of mal regulon activator malT or prevent MalT’s activation of the λ phage receptor lamB [141146]. Indeed, our Dub-seq screen identified mlc, malY, and aes as the top-scoring genes in the presence of λ phage, confirming that Dub-seq can identify host factors involved in regulating the expression of a phage receptor (Fig 4A and 4B). We also found that overexpression of a gene encoding glucokinase (glk) and cyclic-AMP (cAMP) phosphodiesterase (cpdA) showed enhanced fitness in the presence of λ phage (S2 Fig). Glk has been proposed to inhibit mal regulon activator malT [147], whereas CpdA hydrolyzes cAMP and negatively impacts cAMP-CRP–regulated gene expression of lamB [9,148,149].

An external file that holds a picture, illustration, etc.
Object name is pbio.3000877.g004.jpg
Dub-seq screening data for 13 dsDNA phages.

(A) Heatmap of Dub-seq data for 13 dsDNA phages at different MOI and 30 genes with large fitness benefits. Overexpression or higher dosage of these genes interferes with the phage infectivity cycle and impart fitness benefits to the host. Only genes with high-confidence effects and gene fitness score of ≥4 in at least one phage assay are shown. Yellow boxes highlight genes that are known to show resistance when overexpressed. The pooled fitness assays performed on solid plate agar are marked with stars. The underlying data for this figure can be found in S1 Data. (B to D) Dub-seq viewer plots for high-scoring mlc- (B), pdeL- (C), and ygbE-containing (D) fragments in the presence of λ, N4, and T4 phages, respectively. Red lines represent fragments covering highlighted genes completely (start to stop codon), whereas gray-colored fragments either cover the highlighted gene partially or do not cover the highlighted gene completely. Additional Dub-seq viewer plots are provided in S2S4 Figs. dsDNA, double-stranded DNA; Dub-seq, dual-barcoded shotgun expression library sequencing; MOI, multiplicity of infection.

One of the top candidates that broadly enhanced host fitness in the presence of diverse phages is transcriptional activator RcsA (that increases colanic acid production by inducing capsule synthesis gene cluster [106]). Genomic fragments with rcsA showed the highest gene score of +12 to +16 in most experiments (47/51 assays shown in Fig 4A). In addition, we identified growth advantages with dozens of genes when overexpressed in the presence of specific phages. For example, we found genomic fragments encoding pdeO (dosP), pdeR (gmr), pdeN (rtn), pdeL (yahA), pdeC (yjcC), pdeB (ylaB), pdeI (yliE), ddpX, flhD, and yhbJ/rapZ all confer resistance to N4 phage (Fig 4A and 4C, S4 Fig). Except for ddpX, flhD, and yhbJ/rapZ, which encode D-Ala-D-Ala dipeptidase involved in peptidoglycan biosynthesis, flagellar transcriptional regulator, and an RNase adaptor protein, respectively, all others encode cyclic di-GMP (c-di-GMP) phosphodiesterases (PDEs). The PDEs are a highly conserved group of proteins in bacteria that catalyze the degradation of c-di-GMP, a key secondary signaling molecule involved in biofilm formation, motility, virulence, and other cellular processes [150152]. The fitness benefit of increased dosage of PDEs is presumably mediated by their degradation of c-di-GMP. We infer that high levels of c-di-GMP are required for phage N4 infection, although the mechanism is unclear. Incidentally, increased dosage of pdeN (rtn) is known to confer resistance to N4 phage, albeit with unknown mechanism [153,154]. In addition to the N4 phage hits, we found that overexpression of ygbE and ompF showed high fitness scores in the presence of T4 phages, overexpression of small RNA micF showed high fitness scores for CEV1 phage, and overexpression of ompT gives high fitness scores in the presence of T3 and T7 phages (Fig 4).

Following earlier work on phage T7 [51], this is the first global survey of how host gene overexpression/dosage impacts resistance to diverse phages belonging to three families. Although we do not understand all of the mechanisms leading to phage resistance, a few of these hits are consistent with the known biology of phage receptors. For example, expression of outer membrane porins ompC and ompF are regulated reciprocally by ompR, and increased ompF level reduces expression of the T4 phage receptor ompC [155157]. Similarly, increased expression of the sRNA micF causing resistance to phage CEV1 (S2 Fig) is consistent with a report that elevated levels of micF specifically down-regulates ompF (CEV1 receptor) [158,159]. As micF is encoded within the intergenic region of rcsD (activator of Rcs pathway and colanic acid biosynthesis) and ompC and also contains OmpR operator sites, the resistance-causing Dub-seq fragments containing micF could be acting via a combination of effects that cannot be resolved in our screen. Finally, overexpression of the lit gene within the defective prophage element e14, shows high fitness in the presence of T6, CEV1, CEV2, 186, λ cI857, and LZ4 phages, but only on plate-based assays (Fig 4, S3 and S4 Figs). Overexpression of lit is known to interfere with the T4 phage growth [160], though we did not observe high fitness scores in the presence of T4 phage.

In summary, we identified 129 multicopy suppressors of phage infection that encode diverse functions, and our results indicate that enhanced host fitness (phage resistance) can occur via diverse cellular mechanisms. In the following section, we describe follow-up experiments with 13 of these top-scoring hits from the Dub-seq screen to validate their role in phage resistance.

Experimental validation of LOF and GOF screen hits

To validate the phage resistance phenotypes observed in our LOF screens, we sourced 6 individual deletion strains lpcA, galU, ompF, dgcJ, ygaH, and dsrB from E. coli K-12 BW25113 Keio library [105] and disrupted igaA in the E. coli K-12 BW25113 strain (S1 Fig). We then determined the efficiency of plating (EOP) for eight phages (Fig 5). In addition, we performed gene complementation experiments using E. coli K-12 ASKA plasmids [161] to check if the plating defect can be restored (Fig 5, S5 Fig). To validate the hits identified in our Dub-seq GOF screens, we moved 12 individual plasmids expressing rcsA, ygbE, yedJ, flhD, mtlA, pdeB, pdeC, pdeI, pdeL, pdeN, pdeO, and pdeR into E. coli K-12 and tested the EOP of 11 phages (total 36 phage-gene combinations) (Fig 5, Methods). We present these results below.

An external file that holds a picture, illustration, etc.
Object name is pbio.3000877.g005.jpg
Experimental validations of top-scoring gene hits in LOF and GOF screens.

(A) EOP experiments for LOF screen hits using Keio [105,161] library strains. Gene complementation data is presented in S2 Fig. (B) EOP experiments for GOF screen hits with ASKA plasmid library [161] expressing genes (shown as +gene names) in the presence of different phages. We used no IPTG or 0.1 mM IPTG for inducing expression of genes from ASKA plasmid. We used the BW25113 strain with an empty vector for estimating EOP. The plaque morphology or EOP of T3, T7, P2, and 186 phages on lpcA deletion strain indicated inefficient infection. The plating defect was restored to normal when mutants were complemented with a plasmid expressing the respective deleted genes indicating LPS core as the receptor for these phages (S5 Fig). EOP, efficiency of plating; GOF, gain-of-function; LOF, loss-of-function; LPS, lipopolysaccharide.

CEV1 phage requires both OmpF and LPS for K-12 infection

ompF, lpcA/gmhA, and additional genes involved in LPS biosynthesis and transport showed the highest fitness scores in the presence of CEV1 phage in the RB-TnSeq and CRISPRi data (Figs (Figs2A2A and and3A).3A). In agreement with our LOF screen data, we observed severely reduced EOP of CEV1 on ompF and lpcA deletion strains compared to the control BW25113 strain (Fig 5A, S5 Fig). These plating efficiency defects could be reverted when the respective deleted genes were expressed in trans (S5 Fig). These results indicate that CEV1 infection proceeds by recognizing both OmpF and LPS core, and loss of either ompF or LPS disruption leads to a resistance phenotype.

Overexpression of colanic acid biosynthesis pathway reduces sensitivity to diverse phages

The Rcs signaling pathway is one of the well-studied signaling pathways in bacteria and is known to regulate a large number of genes involved in diverse functions, including synthesis of colanic acid and biofilm formation in response to perturbations in the cell envelope [106]. Down-regulation of igaA, a negative regulator of the Rcs signaling pathway, causes resistance to infection by T4, λ, and 186 phages [52] as well as resistance to T7 phage [51]. We found that activation of the Rcs pathway (by either disruption in igaA [yrfF] or overexpression of rcsA) confers resistance to all of the phage we studied. We also observed a high fitness score for sgRNA targeting dsrB (a gene of unknown function located downstream of rcsA) in CRISPRi screens in the presence of N4 phage (Fig 3A). In agreement with our Dub-seq screen data, we observed the EOP of T2, T3, T4, T5, T6, 186, λ, CEV1, CEV2, LZ4, and N4 phages is compromised when rcsA is overexpressed (Fig 5B). Similarly, a dsrB deletion strain showed severe N4 phage plating defects, confirming the high fitness scores in our CRISPRi screen (Fig 5A).

Despite igaA’s reported essentiality [162,163], our RB-TnSeq library contained 9 disruptions in igaA’s cytoplasmic domain, which also overlapped with sgRNA target sites in our CRISPRi screen (Figs (Figs3A3A and and5A5A and S1 Fig). To validate that this domain is indeed dispensable for strain viability and also important in phage resistance, we successfully reconstructed the igaA insertion mutant (Methods). This mutant strain displayed a mucoidy phenotype indicative of increased activation of colanic acid biosynthesis via the Rcs signaling pathway [106,164]. We observed defective plaque morphologies with T3, T5, T6, T7, P2, 186, CEV1, and N4 phages on the igaA insertion mutant, indicating inefficient infection and the plating defect could be reversed by supplying the wild-type igaA gene in trans (Fig 5A, S5 Fig). To gain further insight into the phage resistance mechanism, we performed RNA sequencing (RNA-seq) analysis on the igaA disruption mutant (Methods). We found that multiple components of the Rcs pathway (including rcsA itself) were up-regulated, with 24 genes from the capsular biosynthesis-related operons wca and yjb significantly up-regulated (log2FC > 2, adjusted p-value < 0.001) (Fig 6A and 6B, S6 and S7 Tables). These results indicate that the igaA disruption mutant uncovered in this work activates colanic acid biosynthesis, leads to a mucoidy phenotype, and may be interfering with phage infection by blocking phage receptor accessibility. The IgaA residues 18–164 we mutate in this work overlap with the N-terminal cytoplasmic domain of IgaA that inhibits Rcs signaling in the absence of stress [165]. Recent studies have highlighted that the activation of the Rcs pathway and capsular biosynthesis are essential to survive in diverse environmental contexts, membrane damage, and stress-inducing conditions, including those by antibiotic treatment [106,166]. A number of earlier studies have also highlighted the generation of the mucoid phenotype that probably provides a fitness advantage by blocking phage adsorption [16,167170]. Our results suggest this might be a generalized mechanism that provides cross-resistance to diverse phages.

An external file that holds a picture, illustration, etc.
Object name is pbio.3000877.g006.jpg
RNA-seq analysis to gain insights into phage resistance mechanisms.

(A) Up-regulation of EPS biosynthesis genes observed in igaA disruption mutant relative to wt (N = 3). (B) Schematic of the Rcs phosphorelay pathway. Mutations in igaA (shown as stars) identified in this work activate Rcs signaling pathway and induce rcsA expression, colonic acid, and EPS biosynthesis pathway. Disruption in igaA or overexpression of rcsA show a mucoidy phenotype and broad resistance to different phages. (C) Down-regulation of ompC transcript and up-regulation of arnBCA operon observed during ygbE overexpression relative to wt (N = 3). (D) Schematic of ompF and ompC expression regulation via EnvZ-OmpR and YgbE. Overexpression of ygbE down-regulates ompC expression (via unknown mechanism) and up-regulates genes involved in lipid A modification, probably the reason for resistance to phage T4. (E) RNA-seq data of pdeL overexpression showed no down-regulation of N4 phage receptor genes (nfrA and nfrB) and no up-regulation of genes involved in EPS or biofilm. (F) Schematic of c-di-GMP pathway with dgcJ deletion or overexpression of one of 7 PDEs encoding genes (representing decreased c-di-GMP levels) show a high fitness score in the presence of N4 phage via an unknown mechanism that is independent of expression of N4 phage receptor and genes involved EPS biosynthesis. In (A), (C), and (E) plots, purple filled data points are adjusted p-value < 0.001 and abs(log2FC) > 2. Blue filled is nonsignificant data points. The dashed lines are effect size thresholds of greater than 4-fold. The underlying data for this figure can be found in S1 Data. c-di-GMP, cyclic di-GMP; EPS, exopolysaccharide; IM, inner membrane; LPS, lipopolysaccharide; OM, outer membrane; PDE, phosphodiesterase; RNA-seq, RNA sequencing; wt, wild type.

Overexpression of YgbE confers resistance to phage T4 and down-regulates OmpC

YgbE is a DUF3561-containing inner membrane protein with no known function [171]. In Dub-seq screens, ygbE shows highest fitness scores in the presence of T4 phage and our EOP data confirm that T4 phage shows strong plating defects when ygbE is overexpressed (Fig 5B, EOP of 7E-6). To gain insight into the mechanism of phage resistance, we performed RNA-seq analysis on a ygbE overexpression strain. Differential expression analysis of this strain revealed strong down-regulation of ompC (log2FC = −5.7, adjusted p-value [double less-than sign] 0.001), the primary T4 phage receptor (Fig 6C and 6D). In addition, we also noticed a strong down-regulation of 26 genes (log2FC = −4.5 adjusted p-value [double less-than sign] 0.001) related to flagella structure including RNA polymerase sigma 28 (sigma F) factor, and strong up-regulation (log2FC = 4.5 adjusted p-value [double less-than sign] 0.001) of arnBCADT operon involved lipid A modification (Fig 6C and 6D, S6 and S7 Tables). The down-regulation of ompC and up-regulation of LPS modification genes is in agreement with the observed phage resistance phenotype of ygbE (Figs 4A, 4D and and5B),5B), but the mechanism of ompC down-regulation in the ygbE overexpression strain remains to be determined.

c-di-GMP is required for infection by phage N4

c-di-GMP is a key bacterial secondary signaling molecule involved in regulation of diverse cellular functions. The top eight hits in our LOF and GOF screens for N4 phage resistance included enzymes that catalyze synthesis and degradation of c-di-GMP. Diguanylate cyclase J (encoded by dgcJ) is involved in the biosynthesis of c-di-GMP and is one of the top scorers in both RB-TnSeq and CRISPRi LOF screens, whereas the seven c-di-GMP-specific PDEs that are involved in degradation of c-di-GMP showed the highest fitness scores in the Dub-seq screen. Though the signaling network of c-di-GMP is complex, deletion of DGCs or overexpression of PDEs is known to reduce c-di-GMP levels, inhibit biofilm formation, reduce biosynthesis of curli, and increase motility [150152,172]. The high fitness scores for dgcJ in RB-TnSeq and CRISPRi screens is intriguing considering it is one of the 12 DGCs encoded on E. coli K-12 genome [173,174], and none of the other DGCs show phenotypes in our screens. Similarly, E. coli K-12 genome codes for 13 PDEs in total [173,174], and we find six of these PDEs show a phage resistance phenotype when overexpressed. Our EOP estimations with N4 phage showed severe plating defect (EOP < 8E-8) on pdeO (dosP), pdeR (gmr), pdeN (rtn), pdeL (yahA), pdeB (ylaB), and pdeI (yliE) overexpressing strains and minor plating defect on dgcJ mutant (EOP of 0.8) (Fig 5A and 5B). The plating defect of N4 phage on dgcJ::kan could be reverted back when the dgcJ was provided in trans (S5 Fig).

To gain insight into how the overexpression of specific genes might affect phage infection pathways, we performed differential RNA-seq experiments (standard growth conditions, in the absence of phage) on five c-di-GMP PDE (pdeL, pdeB, pdeC, pdeN, and pdeO) overexpressing strains, each of which shows resistance to N4 phage (Methods). RNA-seq experiments on the pdeL overexpression strain revealed large changes in the E. coli transcriptome relative to the wild-type BW25113 strain, with 103 genes significantly up-regulated (log2FC > 2, q < 0.001) and 109 genes significantly down-regulated (log2FC < −2, q < 0.001) (Fig 6E and 6F, S6 and S7 Tables). N4 phage receptor genes nfrA and nfrB were not differentially expressed. Similar to pdeL overexpression, overexpression of the other four PDEs also did not show substantially different expression of nfrA and nfrB, suggesting that N4 resistance phenotype in these instances does not appear to be via transcriptional regulation of the N4 phage receptor genes (S6 and S7 Tables). Further studies are underway to gain more insights on how c-di-GMP levels influence N4 phage infection cycle.

Validation of other top-scoring genes displaying N4 phage resistance phenotype

ygaH is one of the four top-scoring RB-TnSeq candidates in the presence of N4 phage (Fig 2A). YgaH is predicted to be a L-valine exporter subunit [171,175] and had not been previously associated with N4 phage infection. In agreement with its strong fitness scores in RB-TnSeq data, our EOP estimations on ygaH deletion strain confirmed severe plating defects on N4 phage (Fig 5A). Although the role of YgaH in N4 phage infection pathway is unclear, Dub-seq fragments encoding mprA (a negative regulator of multidrug efflux pump genes), a gene downstream of ygaH, also shows an N4 phage resistance phenotype (Fig 4, S3 Fig), further demonstrating the importance of this region in N4 phage infection.

Among other top hits in the GOF Dub-seq screen, we observed defective plating of N4 phage on lrhA and rapZ (yhbJ) overexpression strains (Fig 5B, S3 Fig). lrhA codes for a transcription regulator of genes involved in the synthesis of type 1 fimbriae, motility, and chemotaxis [176] and has not been previously linked to N4 phage infection. rapZ codes for an RNase adaptor protein that negatively regulates the expression level of glucosamine-6-phosphate synthase (GlmS) [176,177]. GlmS catalyzes the first rate-limiting step in the amino sugar pathway supplying precursors for assembly of the cell wall and the outer membrane. Though the role of the essential gene glmS in N4 phage infection is unclear, the N4 phage resistance phenotype shown by rapZ multicopy expression (which down-regulates glmS expression) agrees with our CRISPRi screen data wherein knockdown of both glmS and glmU within glmUS operon show strong fitness in the presence of N4 phage (Fig 3C). These results illustrate the power of using these combination technologies for studying phage infection.

Finally, candidates that showed high fitness scores in our Dub-seq screen but failed to demonstrate strong phage plating defects include strains overexpressing flhD or mtlA in the presence of N4 phages and yedJ in the presence of multiple phages (Fig 4A, S6 Fig). We speculate other entities such as sRNA coding regions or transcription factor binding sites on Dub-seq fragments encoding these gene loci may be essential for phage resistance phenotype (S6 and S7 Figs).

Extending genome-wide screens to E. coli BL21 strain

To compare the phage resistance phenotypes we observed in E. coli K-12 to a closely related host, we chose E. coli BL21 strain as our alternate model system. The genomes of BL21 and K-12 strains have 99% bp identity over 92% of their genomes, interrupted by deletions or disruptions (by mobile elements) [8890]. For example, in comparison to K-12, BL21 has a disruption or deletions in the colanic acid pathway, biofilm formation, flagella formation (a 21-gene cluster including flagella-specific fliA sigma factor), genes involved in Rcs signaling pathway (rcsA, rcsB, rcsC), the LPS core gene cluster (BL21 forms truncated LPS core with only two hexose units compared to normal five hexose units in K-12, Fig 2B), and is deficient in Lon and OmpT proteases, which are important components of the protein homeostasis network [8890]. In addition, BL21 also lacks ompC and nfrAB genes that code for T4 phage and N4 phage receptors in E. coli K-12, respectively [90,178,179].

To investigate BL21 genes essential for phage growth, we constructed an RB-TnSeq library made up of approximately 97,000 mutants (Methods). For fitness assays, we used the same set of phages that were assayed with K-12 library except phages N4 and 186. Both N4 and 186 phages do not infect BL21 because of a lack of N4 phage receptors (NfrA and NfrB) and truncated LPS that probably limits 186 binding. We performed 53 pooled fitness experiments using the BL21 RB-TnSeq library in both liquid and solid plate assay format at varying MOI and nine no-phage control experiments. In total, we identified 115 high-scoring hits, made up of 50 unique phage-gene combinations and representing 32 unique genes (S8 Table). All 12 phages have at least one high-confidence hit. Largely, the BL21 LOF data are in agreement with our K-12 results for T3, T5, T6, T7, CEV2, LZ4, λ, P1, and P2 phages, especially the high-scoring genes that either code the phage receptor or its regulators (Fig 7).

An external file that holds a picture, illustration, etc.
Object name is pbio.3000877.g007.jpg
Genome-wide screens in E. coli BL21 strain.

(A) Heatmap of BL21 LOF RB-TnSeq data for 12 dsDNA phages at a single MOI, and selected genes with high-confidence fitness benefits are shown. (B) Heatmap of GOF BL21 Dub-seq data for 12 dsDNA phages with high-confidence fitness benefit. Fitness scores of ≥4 in at least one phage assay are shown. These assays were performed in planktonic culture. Yellow stars indicate these data points are in agreement with RB-TnSeq, CRISPRi, and Dub-seq data for E. coli K-12. The underlying data for this figure can be found in S1 Data. CRISPRi, CRISPR interference; dsDNA, double-stranded DNA; Dub-seq, dual-barcoded shotgun expression library sequencing; GOF, gain-of-function; LOF, loss-of-function; MOI, multiplicity of infection; RB-TnSeq, random barcode transposon site sequencing.

The key differences between BL21 and K-12 LOF fitness data are the host factors important in T2, T4, and CEV1 phage infection. For example, ompC and its regulator EnvZ/OmpR two-component system and genes involved LPS core biosynthesis and showed high fitness scores in the presence of T4 phage in K-12 LOF screens, whereas only genes involved in the LPS core biosynthesis (waaG, waaF, and waaQ) were important in BL21 screens (Fig 7). Lack of ompC in the BL21 strain probably alleviates the need for the EnvZ/OmpR two-component system for T4 infection. The absolute requirement for LPS R-core structure for T4 phage growth on BL21 is in agreement with the earlier reports [91,92,179181]. Similarly, CEV1 phage, which showed a strict requirement for OmpF and full-length LPS in K-12 screens (Fig 2), seems to require only OmpF in the BL21 infection cycle (Fig 7). This suggests that CEV1 phages can tolerate truncated LPS of BL21 but not that of K-12 (Fig 2). Because of the OmpF requirement, CEV1 growth on BL21 also showed strict dependence on the EnvZ/OmpR two-component system, a key regulator of ompF expression. Finally, in distinction to K-12 data, our BL21 RB-TnSeq data indicate that T2 phage probably binds preferentially to truncated LPS R-core in BL21 (waaF and waaG showed high fitness in our screen). Furthermore, this effect seems to be independent of FadL. We confirmed this observation by measuring the EOP on a BL21 fadL deletion strain. The absence of a FadL requirement for T2 growth on BL21 is intriguing considering its 100% nucleotide identity with K-12 fadL. It is possible that either the conformational integrity of FadL is compromised in the absence of full-length LPS or that T2 may be recognizing more than one outer membrane protein [95,182] (S1 Text). These results are consistent with earlier observations on the difficulty in isolating T2-resistant mutants in E. coli B cells [15].

Finally, to investigate whether increased gene copy number of host factors interferes with phage growth in BL21, we constructed a BL21 Dub-seq library. This library is made up of randomly sheared BL21 genomic DNA cloned into a dual-barcoded vector and is made up of a total 65,788 unique 3-kb fragments (Methods, S8 Fig). We then screened the BL21 Dub-seq library in a variant of BL21 as the host (Methods). From 24 pooled fitness experiments in planktonic cultures and on solid media in the presence of 12 phages, we identified 39 high-scoring candidates (fitness score ≥ 4, FDR of 0.74, Methods). Other than a few top-scoring candidates in the presence of λ phage, the BL21 dataset was considerably different from K-12 dataset (S9 Table).

Some of the top-scoring hits in BL21 Dub-seq screen showed broad resistance to many different phages whereas some were phage specific. For example, BL21 Dub-seq fragments with mlc (dgsA) gave higher fitness in the presence of T2, T4, T6, T7, λ, and P1 phages. This resistance phenotype of Mlc in the presence of λ phage is consistent with its known regulatory role. Mlc, a global regulator of carbohydrate metabolism, is known to negatively regulate the maltose regulon and mannose permease system [183] (via regulating the lamB expression activator MalT), both of which are known to play a crucial role in phage λ DNA penetration and infection [184187]. Another top-scoring candidate glgC showed higher fitness in the presence of T7 and P2 phages. Overexpression of glgC (which encodes Glucose-1-phosphate adenylyltransferase) is known to increase glycogen accumulation [188,189] and titrate out the global carbon storage regulator CsrA [190]. We speculate that the interaction of GlgC and CsrA probably impacts biofilm formation [130,191193] and leads to alternations in the LPS profile [194,195], leading to phage resistance phenotype. In agreement to our results for E. coli K-12 BW25113 and MG1655 fitness assays, we did not observe enrichment of host factors that may enhance the rate of lysogeny of 186 and P2 phages.

Among the candidates that show phage specific resistance phenotype, we observed that overexpression of argG showed a fitness score of 3.7 in the presence of T4 phages (S9 Table). Argininosuccinate synthetase enzyme (encoded by argG) catalyzes the penultimate step of arginine biosynthesis and has not been associated with phage resistance phenotype before. However, early studies have indicated the inactivating effect of arginine on T4 phages [196]. Finally, Dub-seq fragments encoding ferrous iron uptake system (FeoB) and putative heme trafficking protein (YdiE) yield strong fitness in the presence of T5 phage and T5-like CEV2 phage (Fig 7). It is known that increased uptake of ferrous iron increases Fur-ferrous iron occupancy and Fur-mediated repression of T5 phage receptor fhuA [197201]. Most of these top-scoring candidates were missing in the K-12 dataset, probably because of strong selection for rcsA overexpressing strains in all of our K-12 Dub-seq experiments. This highlights the importance of studying how even closely related hosts can have nuanced interactions with the same bacteriophage.

Summary and conclusions

We applied unbiased high-throughput LOF and GOF screening methods to two different E. coli strains to map the landscape of genetic determinants important in host-phage interactions. We demonstrate how these methods can rapidly identify phage receptors and both novel and previously described non-phage-receptor-related host factors involved in resistance across a wide panel of dsDNA phage types. By using LOF RB-TnSeq and CRISPRi methods, we extensively map both nonessential and essential host genes along with nongenic regions such as promoter and transcription factor binding sites implicated in phage infection and resistance. The Dub-seq methodology uncovers dozens of multicopy suppressors that encode diverse functions and point to a myriad of ways how host gene dosage can influence phage resistance. This global survey of host factors that play an important role in phage growth across two widely studied E. coli strains provides a detailed view of cross-resistance patterns for diverse phages and will be a rich dataset for deeper biological insights and machine learning approaches.

Our data from both LOF and GOF screens show consistency across a range of MOI and also suggest that different assay formats (solid and liquid) allow increased discovery of diverse phage resistance mechanisms. These assay platforms could be expanded to phage-banks made up of hundreds of phages at just one MOI and may be sufficient to rapidly discover the phage receptors in the target host. In addition, extending the screening methods to closely related but different strains would be highly valuable. For example, we used two laboratory E. coli strains that have rough LPS architecture where core oligosaccharide is the terminal part of the LPS. However, it is known that the genetic and structural diversity of LPS and the repeat structure O-polysaccharide attached to LPS (to form smooth LPS) is very large in pathogenic and environmental isolates of Enterobacteriaceae and may impact phage infectivity [48,58,202205]. Although we used nonpathogenic E. coli in this work, we did study two phages (CEV1 and CEV2) that infect pathogenic O157:H7 strains and identified key host factors important for growth of these phages. Our study provides an opportunity to compare these host factor requirements for CEV1 and CEV2 with their structurally similar non-pathogen-associated phages T4 and T5, respectively. The genetic screens presented in this work and future extension to diverse E. coli serogroups may aid in filling the knowledge gap on phage interaction with different antigens and its impact on phage infectivity and resistance.

Our results highlight that phage infectivity depends on the host cellular physiology and, in particular, membrane characteristics of the host imparted by LPS and outer membrane protein biogenesis pathways [113,114,206]. For instance, our results indicate that the disruptions in LPS (for example, deletions or down-regulation of waa genes Figs Figs2A,2A, ,3A3A and and7A),7A), LPS transport pathways (for example, down-regulation of lptABC, Fig 3B), and signal resembling membrane stress (for example, disruption in igaA in Figs Figs2A2A and and3B;3B; overexpression of rcsA, ompF, micF, in Fig 4A) can influence phage infectivity (additional discussion in S1 Text). We also observed strain-specific similarities and differences in the phage infectivity patterns, as well as environmentally context dependent differences. These results indicate the importance of studying phage resistance across a diverse set of biotic and abiotic conditions and how they impact the cross-resistance and cross-sensitivity in closely and distantly related organisms. For example, gaining insights into conditions under which a bacterium would actively overexpress the colanic acid biosynthesis pathway may help us to develop therapies to overcome the generalized phage defense mechanism.

The strong positive fitness scores we observe due to the selection pressure during these pooled fitness experiments are both a strength and limitation of our methods. The strength of these screens is that we can rapidly identify host factors crucial in phage infectivity cycle because they display a very strong fitness score when disrupted (for example, a phage receptor and its regulators in LOF screen). This strong positive selection, however, limits the detection of intermediate phage resistance routes (for example, effect of trxA and cmk deletion on T7 phage growth), whose disruption may lead to strains with relatively poor or neutral fitness and will be swept from the population. We addressed this limitation by using solid agar plate assays, where direct competition among resistant types may be reduced. Future fitness assays can be extended to droplet microfluidics platforms [207] to uncover intermediate fitness phenotypes and retain the high-throughput scalability of these approaches. We note here that the relative fitness of these lab-generated strains may be distinct from the likelihood that their cognate mutant will naturally evolve under phage selection pressure [208]. In addition, we also did not identify strong effects for bacterial defense mechanisms (for example, R-M and CRISPR systems). In both E. coli K-12 BW25113 and BL21, CRISPR and type I R-M systems are disrupted but other R-M systems such as McrA, McrBC, and Mrr are intact [88,90,105]. Other than strains with high-scoring mcrB (encodes subunit of McrBC system, S1 Text, S3 Fig) in the presence of T2 phage, we did not see high fitness scores for other restriction systems in our Dub-seq fitness screens. We speculate that the expression of restriction enzyme subunit has to be at optimum level to counter the phage infectivity cycle and also to compete with high fitness strains (such as strains with higher RcsA expression) in the pooled fitness assays. One way to improve the detection of overexpression gene phenotypes is to use phage mutants that lack anti-R-M systems and nonessential regions. For example, T7 phage mutants lacking gene 0.3, gene 1.2, and region 0.3–1.4 were used earlier to uncover the overexpression phenotype of hsdR (constitutes the restriction subunit of EcoKI), dgt (dGTPase), and udk (encodes for uridine/cytidine kinase) on T7 phage growth, respectively [51,209].

The screening technologies presented in this work are scalable to study phage resistance mechanisms in diverse organisms. Two key considerations in this regard that need to be accounted for are (1) the availability and extendibility of genetic tools to the organisms of interest and (2) the cost and time in creating the RB-TnSeq, CRISPRi, and Dub-seq libraries. Based on our experience, the cost of creation of these libraries falls in the range of 4,00012,000 per library, and it takes about 3 mo on average for building fully characterized libraries. The newly developed sequencing technologies and extendibility of genetic tools to nonmodel organisms can further reduce the time and cost of generating the LOF and GOF libraries [70,210,211]. Once the libraries are built, the running cost of these genome-wide assays across 96 conditions cost about $10 per assay and serve as a standardized reagent tool for in-house experiments or for sharing and comparing data across different laboratories. The ease and economics of these genetic screens enable extendibility of phage resistance assays in diverse conditions that simulate the natural environment and may provide valuable insights on host fitness and host-phage coevolutionary dynamics under more ecologically relevant conditions. For example, recent studies highlighted the evidence of subdued phage resistance in the natural environment, probably because of the fitness cost associated with resistance mechanisms [97,208,212216]. In addition, these methods have the capability to identify fitness costs associated with broadly seen phage resistance phenotypes in a competitive natural environment and thus improve our understanding of microbial ecology in general [13,119,213,214,217219]. Such systems-level insights will be valuable both in uncovering new mechanisms in host-phage interaction and perhaps in developing different design strategies for targeted microbial community interventions, engineering highly virulent or extended host-range phages and rationally formulated phage cocktails for therapeutic applications [97,212,215,220233]. Alternatively, identifying phage resistance determinants may also enable engineering of bacterial strains with phage defense systems crucial in a number of bioprocesses such as in the dairy industry [234,235], biocontainment strategies for bioproduction industry [236,237], or facilitation of bacterial vaccine discovery and development [238240].

There is a clear applied interest in utilizing combinations/cocktails of phages to regulate or eliminate bacterial populations due to the reduced likelihood of evolved multiphage resistance [97,221,225]. However, designing such cocktails relies on a better understanding of cross-resistance among phages [28,61,241,242]. In particular, identification of phages that differ in their receptor use or against which cross-resistance is unlikely to evolve would allow for better design of such therapies [23,37,242]. Moreover, identifying phages that select for resistance that have interrelated phenotypic consequences with, for example, antibiotic sensitivity is a recent advancement in the field that could directly benefit from these screening approaches [33,53,97,243]. By combining fitness datasets for phages and antibiotics or phage-antibiotic combination therapies [244246], such screens could provide an avenue for performing high-throughput search for genetic trade-offs or “evolutionary traps” [33,53,97,243] that could provide a much-needed solution to overcome the antibiotic-resistance pandemic.

Methods

Bacterial strains and growth conditions

The primers and plasmids used in this study are listed in S10 and S11 Tables, respectively. The bacterial strains, phages, and their sources are listed in S12 Table. Phage plaque-forming units/ml and MOIs used in each experiment are listed in the S13 Table. All plasmid manipulations were performed using standard molecular biology techniques [247]. All enzymes were obtained from New England Biolabs (NEB) and oligonucleotides were received from Integrated DNA Technologies (IDT). Unless noted, all strains were grown in LB supplemented with appropriate antibiotics at 37 °C in the Multitron shaker. All bacterial strains were stored at −80 °C for long-term storage in 15% sterile glycerol (Sigma). The genotypes of E. coli strains used in the assays include BW25113 (K-12 lacI+rrnBT14 Δ(araB–D)567 Δ(rhaD–B)568 ΔlacZ4787(::rrnB-3) hsdR514 rph-1), E. coli MG1655 (K-12 F–λ–ilvG–rfb-50 rph-1) and E. coli BL21 (B F–ompT gal dcm lon hsdSB(rB–mB–) [malB+]K-12(λS)). The genetic and phenotypic differences between these strains are well documented [8890,105,248]. We note here that all three strains lack an active CRISPR system. Both E. coli K-12 BW25113 and BL21 have disruption in the type I R-M system (known as EcoKI in E. coli K-12 BW25113 and EcoBI in E. coli Bl21), encoded by the hsdRMS genes, but have functional McrA, McrBC, and Mrr restriction systems, encoded by mcrA, mcrBC, and mrr [90]. E coli MG1655 has functional EcoKI, McrA, McrBC, and Mrr restriction systems.

Bacteriophages and propagation

The bacteriophages used in this study and their sources are listed in S12 Table. All phages except P2 phage and N4 phage were propagated on E. coli BW25113 strain. To propagate P2 phage and N4 phage, we used E. coli C and E. coli W3350 strains, respectively. We used a host-range mutant of T3 (from our in-house phage stock that can grow on both E. coli K-12 and BL21 strains), mutant λ phage (temperature-sensitive mutant allele cI857) [9,249], and a strictly virulent strain of P1 phage (P1vir) [250] that favors a lytic phage growth cycle. We followed standard protocols for propagating phages [231]. Phage titer was estimated by spotting 2 μl of a 10-fold serial dilution of each phage in SM buffer (Teknova) on a lawn of E. coli BW25113 via top agar overlay method using 0.7% LB agar. SM buffer was supplemented with 10 mM calcium chloride and magnesium sulphate (Sigma). We routinely stored phages as filter-sterilized (0.22 μm) lysates at 4 °C.

Construction of BL21 RB-TnSeq library

We created the E. coli BL21-ML4 transposon mutant library by conjugating E. coli BL21 strain with E. coli WM3064 harboring the pKMW3 mariner transposon vector library (APA752) [64]. We grew E. coli BL21 at 30 °C to mid-log-phase and combined equal cell numbers of BL21 and donor strain APA752, conjugated them for 5 hr at 30 °C on 0.45-m nitrocellulose filters (Millipore) overlaid on LB agar plates containing diaminopimelic acid (DAP) (Sigma). The conjugation mixture was then resuspended in LB and plated on LB agar plates with 50 μg/ml kanamycin to select for mutants. After 1 d of growth at 30 °C, we scraped the kanamycin-resistant (kanR) colonies into 25 ml LB, determined the OD600 of the mixture, and diluted the mutant library back to a starting OD600 of 0.2 in 250 ml of LB with 50 μg/ml kanamycin. We grew the diluted mutant library at 30 °C for a few doublings to a final OD600 of 1.0, added glycerol to a final volume of 15%, made multiple 1-ml −80 °C freezer stocks, and collected cells for genomic DNA extraction. To link random DNA barcodes to transposon insertion sites, we isolated the genomic DNA from cell pellets of the mutant libraries with the DNeasy kit (Qiagen) and followed published protocol to generate Illumina compatible sequencing libraries [64,68]. We then performed single-end sequencing (150 bp) with the HiSeq 2500 system (Illumina). Mapping the transposon insertion locations and the identification of their associated DNA barcodes was performed as described previously [64]. Of 4,195 protein-coding genes in E. coli BL21, our BL21 RB-TnSeq library has fitness estimates for 3,083. Twelve independent strains were used to compute fitness for the typical protein-coding gene.

Construction of BL21 Dub-seq library

To construct E. coli BL21 Dub-seq library, we used the dual-barcoded pFAB5526 plasmid library with a kanamycin resistance marker (https://benchling.com/s/seq-1Gkg3lrrSno4EF0Ye11k). The Dub-seq backbone plasmid pFAB5526 is the same in design and genetic composition to the original pFAB5491 Dub-seq plasmid (https://benchling.com/s/seq-39Hoh4d1AResilOUPVJ9) [64,66] except for the kanamycin resistance marker and a mobility gene present on pFAB5526. We mapped the dual barcodes of the pFAB5526 library via the Barcode-Pair sequencing (BPseq) protocol [64,66]. We sequenced BPseq samples on HiSeq 2500 system with 150-bp single-end runs. We then cloned 3 kbp of E. coli BL21 genomic fragments between UP and DOWN barcodes by ligating the end-repaired genomic fragments with PmlI restriction digested pFAB5526, electroporating the ligation into E. coli DH10B cells (NEB) and selecting the transformants on LB agar plates supplemented with kanamycin (50 μg/ml). We scraped the kanR colonies into 25 ml LB and diluted the transformant mixture to a starting OD600 of 0.2 in 250 ml of LB with 50 μg/ml kanamycin and grew the library at 30 °C for few doublings to a final OD600 of 1.0. Finally, we added glycerol to a final volume of 15%, made multiple 1-ml −80 °C freezer stocks, and collected cells for plasmid DNA extraction (Qiagen). Next, we mapped the cloned genomic fragment and its pairings with neighboring dual barcodes via Barcode Association with Genome fragment sequencing (BAGseq) [64,66]. We sequenced the BAGseq samples on HiSeq 2500 system with 150-bp single-end runs using the reported sample preparation steps [64,66]. The data processing of BPseq and BAGseq steps was performed using Dub-seq python library with default setting (https://github.com/psnovichkov/DubSeq), as detailed earlier [64,66]. We mapped E. coli BL21 Dub-seq library to E. coli BL21-DE3 genome sequence [88]. BL21 Dub-seq library was then electroporated into BL21DE3C43 strain and the transformants were collected into 25 ml LB. The transformant mixture was then diluted to a starting OD600 of 0.2 in 250 ml of LB with 50 g/ml kanamycin and grew the library at 30 °C for a few doubling to a final OD600 of 1.0. Finally, we added glycerol to a final volume of 15%, made multiple 1-ml −80 °C freezer stocks. These stocks were further used for pooled fitness assays as described below. The BL21 Dub-seq library is made up of 65,788 unique barcoded fragments. The average fragment size of the library is 2.5 kb and the majority of fragments covered 2–3 genes in their entirety (S8 Fig). Similar to E. coli BW25113 Dub-seq library [64,66], BL21 Dub-seq library covers 85% of genes from start to stop codon by at least five independent genomic fragments, and 97% of all genes are covered by at least one fragment. In total, 132 genes are not covered in their entirety by any Dub-seq fragment.

Competitive growth experiments with RB-TnSeq library

A single aliquot (1 ml) of a mutant library was thawed, inoculated into 25 ml of medium supplemented with kanamycin (50 μg/ml), and grown to OD600 of 0.6–0.8 at 37 °C. After the mutant library recovered and reached mid-log phase, we collected cell pellets as a common reference for BarSeq (termed time-zero or start samples) and used the remaining cells to set up competitive mutant fitness assays in the presence of different phages at different MOI. For performing planktonic culture assays, we diluted the recovered mutant library stock to a starting OD600 of 0.04 in 2X LB media (350 μl) and mixed in equal volume (350 μl) of phages diluted in phage dilution buffer. We also set up control “no-phage” competitive mutant fitness assays wherein we replaced phages with simply the phage dilution buffer. The mutant library experiments were grown in the wells of a 48-well microplate (700 μl per well). We grew the microplates in Tecan Infinite F200 readers with orbital shaking and OD600 readings every 15 min for 8 hr. After the experiment, survivors were collected, pelleted, and stored at −80 °C prior to genomic DNA extraction. For solid plate-based assays, we incubated the mixture of phage and diluted mutant library at room temperature for 15 min and then plated the mixture on LB agar supplemented with kanamycin plates and incubated at 37 °C overnight. The next day, the resistant colonies were scraped, resuspended in 1 ml LB media, and pelleted. Assay pellets were typically stored at −80 °C prior to genomic DNA extraction. The 8-hr assay period was decided based on our preliminary time-course experiment results for a specific phage at different MOIs, wherein we took intermittent samples between 2 and 10 hr, processed as detailed below and found to yield consistent top-scoring hits. In addition, assay samples after 8 hr provided sufficient DNA for the sample processing step.

Competitive growth experiments with Dub-seq library

Similar to RB-TnSeq competitive fitness assays, a single aliquot of the E. coli Dub-seq library (E. coli BW25113 library expressed in E. coli BW25113 or E. coli BL21 Dub-seq library in E. coli BL21DE3C43 strain) was thawed, inoculated into 25 ml of LB medium supplemented with appropriate antibiotics (chloramphenicol 30 μg/ml for E. coli BW25113 Dub-seq library and kanamycin [50 μg/ml] for E. coli BL21 Dub-seq library), and grown to OD600 of 0.6–0.8. At mid-log phase, we collected cell pellets as a common reference for BarSeq (termed start or time-zero samples), and we used the remaining cells to set up competitive fitness assays in the presence of different phages at different MOI. For performing planktonic culture assays, we diluted the recovered Dub-seq library stock to a starting OD600 of 0.04 in 2X LB media and mixed in equal volume (350 μl) with phages diluted in phage dilution buffer. We also set up control “no-phage” competitive mutant fitness assays wherein we replaced phages with simply the phage dilution buffer. The Dub-seq library cultures were grown in the wells of a 48-well microplate (700 μl per well) and grew the microplates in Tecan Infinite F200 readers with orbital shaking and OD600 readings every 15 min for 3–8 hr. After the experiment, survivors were collected, pelleted, and stored at −80 °C prior to plasmid DNA extraction. For solid plate-based assays, we incubated the mixture of phage and diluted Dub-seq library at room temperature for 15 min, and then we plated the mixture on LB agar plates supplemented with appropriate antibiotics and incubated at 37 °C overnight. Next day, the resistant colonies were scraped, resuspended in 1ml LB media, and pelleted. Assay pellets were typically stored at −80 °C prior to plasmid DNA extraction. We also performed these assays using E. coli strains with empty plasmid (used in Dub-seq library creation).

BarSeq of RB-TnSeq and Dub-seq pooled fitness assay samples

We isolated genomic DNA from RB-TnSeq library samples using the DNeasy Blood and Tissue kit (Qiagen). Plasmid DNA from the Dub-seq library samples was extracted either individually using the Plasmid miniprep kit (Qiagen) or in 96-well format with a QIAprep 96 Turbo miniprep kit (Qiagen). We performed 98 °C BarSeq PCR protocol as described previously [64,66] with the following modifications. BarSeq PCR in a 50 μl total volume consisted of 20 μmol of each primer and 150–200 ng of template genomic DNA or plasmid DNA. For the HiSeq4000 runs, we used an equimolar mixture of BarSeq_P2 primers along with new Barseq3_P1 primers. The BarSeq_P2 primer contains the tag that is used for demultiplexing by Illumina software, and the new Barseq3_P1 primer contained an additional sequence to verify that it came from the expected sample. The new Barseq3_P1 primer contains the same sequence as BarSeq_P1 reported earlier with 1–4 N’s (which varies with the index) followed by the reverse (not the reverse complement) of the 6-nucleotide index sequence [251]. This modification to earlier BarSeq PCR protocol was done to eliminate the barcode bleed-through problem in sequencing and also to aid in cluster and sample discrimination on the HiSeq4000. All experiments done on the same day and sequenced on the same lane are considered as a “set.” Equal volumes (5 μl) of the individual BarSeq PCRs were pooled, and 50 μl of the pooled PCR product was purified with the DNA Clean and Concentrator kit (Zymo Research). The final BarSeq library was eluted in 40 μl water. The BarSeq libraries were sequenced on Illumina HiSeq4000 instrument with 50 SE runs. We usually multiplexed 96 Barseq samples per lane for both RB-TnSeq and Dub-seq library samples.

Data processing and analysis of BarSeq reads

RB-TnSeq fitness data were analyzed as previously described [64] with additional filters as presented below. Briefly, the fitness value of each strain (an individual transposon mutant) is the normalized log2(strain barcode abundance at end of experiment/strain barcode abundance at the start of the experiment). The fitness value of each gene is the weighted average of the fitness of its strains. We applied filters on experiments such that the mean reads per gene is ≥10. As we have reported earlier [64], in a typical experiment without stringent positive selection, gene fitness score (fit) >1 and associated t-like statistic >4 suffices to give a low rate of false positives. Because of the stringent positive selection in phage assays, our standard quality metrics reported earlier [64] were not suitable. Because most of the sequencing reads were from a handful of phage-resistant mutants in the population, the majority of the strains in the library did not have enough reads to accurately calculate fitness values. To determine suitable filters, we compared fitness data between two halves of each gene. As we have several insertions in most genes, we can compute fitness values for the two halves of a gene separately and then plot the first half and second half fitness values for each gene (that has sufficient coverage) against each other as described earlier [64]. These plots are available in the figshare file https://doi.org/10.6084/m9.figshare.11413128. Phage T2 assays on solid agar showed poor consistency and were dropped out of the analysis. We observed from the first half and second half fitness plots that fitness values <5 were often not consistent. To reduce false positives, we required that fit ≥ 5; t ≥ 5; standard error = fit/t ≤ 2; and fit ≥ maxFit − 8, where maxFit is the highest fitness value of any gene in an experiment. The limit of maxFit − 8 was chosen based on the fitness score of positive controls (host factors that are known to interfere with phage growth when deleted) and from experimental validations of new hits. The fitness data for all strains in the RB-TnSeq library are provided at https://doi.org/10.6084/m9.figshare.11413128.

For the Dub-seq library, fitness data were analyzed using Barseq script from the Dub-seq python library with default settings as previously described [66]. From a reference list of barcodes mapped to the genomic regions (BPSeq and BAGseq), and their counts in each sample (BarSeq), we estimate fitness values of each genomic fragment (strain) using fscore script from the Dub-seq python library. The fscore script identifies a subset of barcodes mapped to the genomic regions that are well represented in the time-zero samples for a given experiment set. We require that a barcode to have at least 10 reads in at least one time-zero (sample before the experiment) sample to be considered a valid barcode for a given experiment set. Then the fscore script calculates fitness score (normalized ratio of counts between the treatment sample and sum of counts across all time-zero samples) only for the strains with valid barcodes. The fitness scores calculated for all Dub-seq fragments, we estimate a fitness score for each individual gene gscore that is covered by at least one fragment as detailed earlier using nonnegative least squares regression [66]. The nonnegative regression determines if the high fitness of the fragments covering the gene is due to a particular gene or its nearby gene, and avoids overfitting. We applied the same data filters as reported earlier [66] to ensure that the fragments covering the gene had a genuine benefit. Briefly, we identify a subset of the effects to be reliable, if the fitness effect was large relative to the variation between start samples (|score| ≥ 2); the fragments containing the gene showed consistent fitness (using a t test); and the number of reads for those fragments was sufficient for the gene score to have little noise. As in RB-TnSeq BarSeq analysis, Dub-seq data also showed strong positive selection in the presence of phages, where few strains accounting for the most reads and displaying very high fitness scores. To reduce false positives, we applied an additional filter of gene fitness score ≥4 threshold. This threshold cutoff was chosen by analyzing the fragment fitness scores against the number of supported reads and whether the cutoff value agrees with experimental validations carried out in this work. Finally, we estimate the FDR for high-confidence effects as detailed earlier [66]. Briefly, to estimate the number of hits in the absence of genuine biological effects, we randomly shuffled the mapping of barcodes to fragments, recomputed the mean scores for each gene in each experiment, and identified high-confidence effects as for the genuine data. We repeated the shuffle procedure 10 times. The complete dataset of fragment and gene scores are provided at https://doi.org/10.6084/m9.figshare.11838879.v2. Dub-seq data for E. coli strain with empty plasmid (used in Dub-seq library creation) were consistent with E. coli with no plasmid, indicating we do not see any plasmid-mediated fitness effects. Phage P2 assays were excluded from the E. coli K-12 Dub-seq study, as the sequencing runs did not pass the data analysis filters because of low sequencing reads per assay. We used the Dub-seq viewer tool from the Dub-seq python library (https://github.com/psnovichkov/DubSeq) to generate regions of the E. coli chromosome covering fragments (gene-browser mode). The fitness data for all strains in the Dub-seq library are provided at https://doi.org/10.6084/m9.figshare.11838879.v2.

E. coli MG1655 CRISPRi library

The design and construction of E. coli MG1655 CRISPRi library and the overall quality of sgRNAs used are detailed elsewhere [69]. Briefly, the E. coli MG1655 CRISPRi library is made up of 32,992 unique sgRNAs targeting 4,457 genes (including small RNA genes, insertion elements, and prophages), 7,442 promoter regions, and 1,060 transcription factor binding sites. The sgRNA library is driven by pBAD promoter (induced by arabinose) and is expressed from a high copy plasmid (ColE1 replication origin) in E. coli K-12 MG1655 strain harboring a genomically encoded aTc-inducible dCas9 gene.

To perform pooled fitness experiments using E. coli MG1655 CRISPRi library, a single aliquot of the library was thawed, inoculated into 5 ml of LB medium supplemented with carbenicillin, kanamycin, and glucose and grown to OD600 of 0.5. At mid-log phase, we collected cell pellets as an initial time point (time-zero) of the library and diluted the remaining culture in induction media (LB broth with arabinose [0.1%], aTc [200 ng/mL], carbenicillin 100 μg/mL, and kanamycin 30 μg/mL) to initiate dCas9 and sgRNA expression for about six doublings (to OD600 about 0.5). Finally, we diluted the library to OD600 of 0.1 in 2X LB supplemented with induction media and incubated 350 μl of the library with 350 μl of diluted phage stocks (MOI of 1) in 5-ml test tubes at room temperature for 15 min. We also set up a control “no-phage” culture wherein we simply mixed the phage dilution buffer with the library. We then moved these competitive fitness assay cultures to 37 °C with shaking (200 rpm) for 2 hr, collected the entire cell pellet after 2 hr, and stored at −80 °C prior to plasmid DNA extraction to isolate the library. The plasmid library was extracted using QIAprep Spin Miniprep Kit, used for PCR to generate Illumina sequencing samples, and sequenced on Illumina HiSeq2500 instrument with 100 SE runs. Only samples that yielded more than 2 million reads (average library read depth of approximately 60) were used in the downstream analysis.

Illumina reads were quality filtered, trimmed, and mapped using the same procedure as earlier [69] to generate strain abundances (i.e., read counts) for each sgRNA library member (recall that each strain in the library is uniquely determined by the 20-bp variable region of the sgRNA it harbors). The fitness of each sgRNA library member was calculated as the log2 fold-change in abundance of the sgRNA after the experiment versus before the experiment using edgeR [252,253]. For a given enrichment comparison, sgRNAs with fewer than 10 read counts in each replicate of the time-zero and end of experiment samples were filtered out of the analysis. In the edgeR pipeline, each sample was normalized for sequencing depth using the edgeR function calcNormFactors, pseudocounts and dispersions were calculated using the edgeR functions estimateCommonDisp and estimateTagwiseDisp, and the log2 fold-change and corresponding p-values were calculated using the edgeR function exactTest, which is based on a quantile-adjusted conditional maximum likelihood (qCML) method. FDR-adjusted p-values were calculated using the Benjamini-Hochberg method. We performed a post hoc analysis on these fitness scores and picked a threshold of fitness ≥2 and FDR-adjusted p-value <0.05 for new hits of interest based on the satisfaction of these criteria by both positive controls (known phage receptors for example fadL for T2 phage and waa operon for 186 phage) and validated hits (for example, dsrB in the presence of N4 phage). Finally, gene fitness scores were calculated by taking the median fitness scores of (filtered) sgRNAs targeting a given gene. The fitness data for all strains in the CRISPRi library are provided at https://doi.org/10.6084/m9.figshare.11859216.

Construction of igaA mutant strains

We created our igaA genetic disruption mutant through a modified recombineering method using pSIM5 [66,254] and approximate insertion site based off of RB-TnSeq mapping (S1 Fig). Primers were designed to amplify a kanR selection marker with 50-bp homology to the insertion site of interest corresponding to a kanR insertion at 51 bp internal to igaA. PCRs were generated and gel-purified through standard molecular biology techniques [247] and stored at −20 °C until use. Deletions were performed by incorporating the above dsDNA template into the BW25113 genome through standard pSIM5-mediated recombineering methods. First, a temperature-sensitive recombineering vector, pSIM5, was introduced into BW25113 through standard electroporation protocols and grown with chloramphenicol at 30 °C. Recombination was performed through electroporation with an adapted pSIM5 recombineering protocol. Post recombination, clonal isolates were streaked onto kanamycin plates without chloramphenicol at 37 °C to cure the strain of pSIM5 vector, outgrown at 37 °C, and stored at −80 °C until use. A detailed protocol is available upon request. Disruptions were verified by colony PCR followed by Sanger sequencing at the targeted locus and 16S regions. The igaA mutant map is given at https://benchling.com/s/seq-pZEspfGx8K9tYzixzMoJ.

Experimental validation of individual phage resistance phenotypes

To validate the phage resistance phenotypes from both LOF and GOF screens, we performed EOP assays on select E. coli mutants from the Keio collection [105] and overexpression ASKA library [105,161], respectively. The complete list of primers, plasmids, and strains used in validation assays are provided in S10S12 Tables. All deletion strains and plasmids used in this work were confirmed via Sanger sequencing. An isogenic deletion mutant in E. coli BL21 for fadL was generated by phage P1–mediated transduction of kanamycin resistance from individual Keio mutants [105]. The gene deletion strains from Keio library and the BL21 transductants were cultured in LB media or LB agar supplemented with kanamycin (25 μg/ml), whereas ASKA strains were cultured in LB media or LB agar supplemented with chloramphenicol (30 μg/ml). The plasmids from ASKA library were extracted using QIAprep Miniprep kit (Qiagen), and electroporated into E. coli BW25113 strains for GOF validations or respective Keio mutants for complementation of LOF phenotypes. The bottom agar was supplemented with 0.1 mM IPTG to induce protein production from ASKA plasmids. Based on toxicity associated with gene overexpression, IPTG levels (0–0.1 mM) were adjusted and cultures were grown overnight at 30 °C.

Phages were quantified by spot titration method. Two microliters of serially 10-fold diluted phages were spotted on a solidified lawn of approximately 4 ml 0.5% top agar inoculated with 100 μl of a fresh overnight bacterial culture and incubated overnight at either 30 °C or 37 °C. The EOP was calculated as the ratio of the number of plaques on mutant or overexpression strain to the number of plaques on the parental strains (BW25113 or BL21). The EOPs were calculated at least by two biological replicates. S5 Fig lists all EOP estimation numbers and plaque images.

RNA-seq experiments

Transcriptomes were collected and analyzed for strains BW25113 (N = 3), knockout mutants for igaA (N = 3), and overexpression strains for pdeB (N = 1), pdeC (N = 1), pdeL (N = 3), pdeN (N = 1), pdeO (N = 1), and ygbE (N = 3). All cultures for RNA-seq experiments were performed on the same day from unique overnights and subsequent outgrowths.

Strains were grown overnight in LB with an appropriate selection marker at 30 °C. Strains were diluted to OD600 approximately 0.02 in 10 mL LB with the appropriate selection marker and, for overexpression strains, 0.1 mM IPTG. Cultures were grown at 30 °C at 180 rpm until they reached an OD600 0.3–0.4. Samples were collected as follows: 400 μL of culture was added to 800 μL RNAProtect (Qiagen), incubated for 5 min at room temperature, and centrifuged for 10 min at 5,000g. RNA was purified using RNeasy RNA isolation kit (Qiagen) and quantified and quality-assessed by Bioanalyzer. Library preparation was performed by the Functional Genomics Laboratory (FGL), a QB3-Berkeley Core Research Facility at UC Berkeley. Illumina Ribo-Zero rRNA Removal Kit was used to deplete ribosomal RNA. Subsequent library preparation steps of fragmentation, adapter ligation, and cDNA synthesis were performed on the depleted RNA using the KAPA RNA HyperPrep kit (KK8540). Truncated universal stub adapters were used for ligation, and indexed primers were used during PCR amplification to complete the adapters and to enrich the libraries for adapter-ligated fragments. Samples were checked for quality on an Agilent Fragment Analyzer. Sequencing was performed at the Vincent Coates Sequencing Center on a HiSeq4000 using 100PE runs.

RNA-seq data analysis

For all RNA-seq experiments, analyses were performed through a combination of KBase [255] and custom jupyter notebook-based methods. Briefly, Illumina reads were trimmed using Trimmomatic [256] v0.36 and assessed for quality using FASTQC. Trimmed reads were mapped to the E. coli BW25113 genome (NCBI accession: CP009273) with HISAT2 [257]. Alignments were quality-assessed with BAMQC. From this alignment, transcripts were assembled and abundance-estimated with StringTie [258]. For experiments with sufficient biological replicates, tests for differential expression were performed on normalized gene counts by DESeq2 (negative binomial generalized linear model) [259]. Additional analyses for all experiments were performed in Python3 and visualized employing matplotlib and seaborn packages. For analyses involving DESeq2, conservative thresholds were employed for assessing differentially expressed genes. Genes were considered differentially expressed if they possessed a Bonferroni-corrected p-value below a threshold of 0.001 and an absolute log2 fold-change greater than 2. For supplemental analyses without tests for differential expression, data were analyzed by comparing log2 fold-changes of transcripts per kilobase million (TPM) counts. The complete datasets are provided in S6 and S7 Tables.

Supporting information

S1 Fig

Map of igaA mutants.

igaA mutants and their overall fitness scores from E. coli K-12 RB-TnSeq screen (S1 Table). Schematic of Rcs phosphorelay with the predicted topology of IgaA [165] is shown at the top. RB-TnSeq mutants in IgaA are mapped to the predicted topology of IgaA (top) and also mapped on to igaA nucleotide sequence, with mutant position and fitness scores. Mutants A, B, and D were constructed and their mucoidy and phage resistance phenotype were confirmed (mutant A data are shown in Fig 5A). Though full-length deletion of igaA has not been possible, our results indicate that disruption between amino acid 22 and 151 is dispensable. Strain with igaA mutant A was further subjected to EOP and RNA-seq analysis presented in the main text. The underlying data for this figure can be found in S1 Data. EOP, efficiency of plating; RB-TnSeq, random barcode transposon site sequencing; RNA-seq, RNA sequencing.

(TIF)

S2 Fig

Dub-seq viewer plots for high-scoring E. coli K-12 BW25113 genomic fragments in the presence of phages.

Following top candidates are shown: high-scoring fragments encoding micF and degP for CEV1 phage; aes, cpdA, malY, glk, sdiA, and gltP for λ phage cI857. Red lines represent fragments covering highlighted genes completely (start to stop codon), and gray-colored fragments either cover the highlighted gene partially or do not cover the highlighted gene completely. The underlying data for this figure can be found in S1 Data. Dub-seq, dual-barcoded shotgun expression library sequencing.

(TIF)

S3 Fig

Dub-seq viewer plots for high-scoring E. coli K-12 BW25113 genomic fragments in the presence of phages.

Following top candidates are shown: high-scoring fragments encoding mcrB for T2 phage; ompT and glgC for T3 phage; ompF for T4 phage; lit for T6 phage; yjcC (pdeC), mprA, ddpX, yhbJ (rapZ), ylaB (pdeB) for N4 phage. Red lines represent fragments covering highlighted genes completely (start to stop codon), whereas gray-colored fragments either cover the highlighted gene partially or do not cover the highlighted gene completely. The underlying data for this figure can be found in S1 Data. Dub-seq, dual-barcoded shotgun expression library sequencing.

(TIF)

S4 Fig

Dub-seq viewer plots for high-scoring E. coli K-12 BW25113 genomic fragments in the presence of phages.

Following top candidates are shown: high-scoring fragments encoding rtn(pdeN), yliE (pdeI), gmr (pdeR), dosP (pdeO), and lrhA for N4 phage; gltS, gltJ, and gltP for 186 phage; mcrB and lit for LZ4 phage. Red lines represent fragments covering highlighted genes completely (start to stop codon), whereas gray-colored fragments either cover the highlighted gene partially or do not cover the highlighted gene completely. The underlying data for this figure can be found in S1 Data. Dub-seq, dual-barcoded shotgun expression library sequencing.

(TIF)

S5 Fig

Estimation of EOP with Keio deletion strains and genetic complementation.

(A and B) EOP experiments with Keio strains and ASKA plasmid complementation of respective genes (indicated by +p) in the presence of different phages. We used no IPTG or 0.1 mM IPTG for inducing expression of genes from ASKA plasmid. All experiments were in E. coli K-12 BW25113 strain. EOP, efficiency of plating.

(TIF)

S6 Fig

Examples of candidates that showed high fitness scores in our Dub-seq screen but failed to show such stronger effect in EOP validation experiments.

(A) EOP experiments with ASKA plasmid expressing genes (shown as +gene names) in the presence of different phages. We used no IPTG and 0.1 mM IPTG for inducing expression of genes from ASKA plasmid. We used E. coli K-12 BW25113 strain with an empty vector for EOP calculations. (B) Dub-seq viewer plots for high-scoring fragments (red bars) encoding yedJ, along with neighboring genes (gray bars) with fitness score in the presence of T4 phage on the y-axis. The Ecocyc operon [171] view for yedJ regions is on the bottom. The genomic fragments encoding yedJ also encode a small RNA rseX (RNA suppressor of extracytoplasmic stress protease) that binds to RNA binding protein Hfq (a global regulator) and specifically targets ompA and ompC mRNAs. Our library does not have genomic fragments that can resolve rseX contribution to fitness independently. Overexpression of RseX has also been shown to increase biofilm formation [260], indicating the role of yedJ-rseX locus on resistance to T4 phage and other phages. The underlying data for this figure can be found in S1 Data. Dub-seq, dual-barcoded shotgun expression library sequencing; EOP, efficiency of plating.

(TIF)

S7 Fig

Dub-seq and RB-TnSeq data uncovers flhD upstream loci important in N4 phage infection.

(A) Dub-seq viewer plots for high-scoring fragments (red bars) encoding flhD and upstream region along with neighboring genes (gray bars) with fitness score on the y-axis. Overexpression of flhD failed to demonstrate strong phage plating defects in our EOP validation experiments (S6A Fig). (B) The Ecocyc operon view of the regulatory region upstream of flhD with multiple transcription factor binding sites [171]. (C) Zoom-in view of the upstream region of flhD operon with arrows identifying the location of 13 RB-TnSeq insertion mutants that have fitness score >10 in the presence of N4 phage. These results indicate the role of flhD regulatory region on N4 phage growth. We speculate that these TnSeq mutants of flhD regulatory region may not be overexpressing flhD, as our EOP experiments failed to validate flhD overexpression as the cause of N4 phage resistance. The underlying data for this figure can be found in S1 Data. Dub-seq, dual-barcoded shotgun expression library sequencing; RB-TnSeq, random barcode transposon site sequencing.

(TIF)

S8 Fig

Description of E. coli BL21 Dub-seq library.

(A) The fragment insert size distribution in the E. coli BL21 Dub-seq library. (B) Cumulative distribution plot showing the percentage of genes in the E. coli BL21 genome (y-axis) covered by a number of independent genomic fragments (x-axis). (C) The distribution of the number of genes that are completely covered (start to stop codon) per genomic fragment in the E. coli BL21 Dub-seq library. (D) Genome coverage of E. coli BL21 Dub-seq library in 10,000-kB windows mapped to E. coli BL21-DE3. The underlying data for this figure can be found in S1 Data. Dub-seq, dual-barcoded shotgun expression library sequencing.

(TIF)

S1 Table

RB-TnSeq E. coli K-12 dataset.

RB-TnSeq, random barcode transposon site sequencing.

(XLSX)

S2 Table

Summary of known and new high fitness score hits per phage per screen at any MOI.

MOI, multiplicity of infection.

(PDF)

S3 Table

Literature table: Mapping RB-TnSeq, CRISPRi, and Dub-seq hits to reported phage resistance data.

CRISPRi, CRISPR interference; Dub-seq, dual-barcoded shotgun expression library sequencing; RB-TnSeq, random barcode transposon site sequencing.

(PDF)

S4 Table

CRISPRi E. coli K-12 dataset.

CRISPRi, CRISPR interference.

(XLSX)

S5 Table

Dub-seq E. coli K-12 dataset.

Dub-seq, dual-barcoded shotgun expression library sequencing.

(XLSX)

S6 Table

RNA-seq dataset.

RNA-seq, RNA sequencing.

(XLSX)

S7 Table

RNA-seq DESeq2_Summary.

RNA-seq, RNA sequencing.

(XLSX)

S8 Table

RB-TnSeq E. coli BL21 dataset.

RB-TnSeq, random barcode transposon site sequencing.

(XLSX)

S9 Table

Dub-seq E. coli BL21 dataset.

Dub-seq, dual-barcoded shotgun expression library sequencing.

(XLSX)

S10 Table

List of primers.

(XLSX)

S11 Table

List of plasmids.

(XLSX)

S12 Table

List of strains: Phage and bacteria.

(XLSX)

S13 Table

MOI per phage per assay.

MOI, multiplicity of infection.

(XLSX)

S1 Text

Detailed information on host factors important in phage infection and top-scoring hits for each screen per phage.

(DOCX)

S1 Data

The underlying data for all figures.

(XLSX)

Acknowledgments

Authors would like to thank Studier F. William, Lucia B. Rothman-Denes, Ian J. Molineux, Jason J. Gill, and Paul Turner for sharing reagents and helpful discussions.

Abbreviations

cAMPcyclic-AMP
c-di-GMPcyclic di-GMP
CRISPRiCRISPR interference
DGCdiguanylate cyclase
dsDNAdouble-stranded DNA
Dub-seqdual-barcoded shotgun expression library sequencing
EOPefficiency of plating
EPSexopolysaccharide
FDRfalse discovery rate
GlmSglucosamine-6-phosphate synthase
GOFgain-of-function
LBlysogeny broth
LOFloss-of-function
LPSlipopolysaccharide
MOImultiplicity of infection
PDEphosphodiesterase
RB-TnSeqrandom barcode transposon site sequencing
R-Mrestriction-modification
RNA-seqRNA sequencing
sgRNAsingle-guide RNA

Funding Statement

This project was funded by the Microbiology Program of the Innovative Genomics Institute, Berkeley (to VKM, AMD, and APA). The initial concepts for this project were funded by ENIGMA, a Scientific Focus Area Program at the Lawrence Berkeley National Laboratory, supported by the US Department of Energy, Office of Science, Office of Biological and Environmental Research under contract DE-AC02-05CH11231 (to VKM, AMD, and APA). The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

Data Availability

Sequencing data have been uploaded to the Sequence Read Archive under BioProject accession number PRJNA645443 [https://www.ncbi.nlm.nih.gov/bioproject/PRJNA645443/]. Supplementary Tables with complete CRISPRi data are deposited here: https://doi.org/10.6084/m9.figshare.11859216.v1. In addition, the complete data from RB-TnSeq experiments are deposited here: https://doi.org/10.6084/m9.figshare.11413128. And Dub-seq experiments are deposited here: https://doi.org/10.6084/m9.figshare.11838879.v2. The underlying data for all figures are provided in supporting information file S1 Data.

References

1. Breitbart M, Rohwer F. Here a virus, there a virus, everywhere the same virus? Trends in Microbiology. 2005:278–284. 10.1016/j.tim.2005.04.003 [Abstract] [CrossRef] [Google Scholar]
2. Wilhelm SW, Suttle CA. Viruses and Nutrient Cycles in the Sea. BioScience. 1999:781–788. 10.2307/1313569 [CrossRef] [Google Scholar]
3. Koskella B, Taylor TB. Multifaceted Impacts of Bacteriophages in the Plant Microbiome. Annu Rev Phytopathol. 2018;56: 361–380. 10.1146/annurev-phyto-080417-045858 [Abstract] [CrossRef] [Google Scholar]
4. Shkoporov AN, Clooney AG, Sutton TDS, Ryan FJ, Daly KM, Nolan JA, et al. The Human Gut Virome Is Highly Diverse, Stable, and Individual Specific. Cell Host Microbe. 2019;26: 527–541.e5. 10.1016/j.chom.2019.09.009 [Abstract] [CrossRef] [Google Scholar]
5. Mirzaei MK, Maurice CF. Ménage à trois in the human gut: interactions between host, bacteria and phages. Nature Reviews Microbiology. 2017:397–408. 10.1038/nrmicro.2017.30 [Abstract] [CrossRef] [Google Scholar]
6. Suttle CA. Marine viruses—major players in the global ecosystem. Nature Reviews Microbiology. 2007:801–812. 10.1038/nrmicro1750 [Abstract] [CrossRef] [Google Scholar]
7. De Sordi L, Lourenço M, Debarbieux L. The Battle Within: Interactions of Bacteriophages and Bacteria in the Gastrointestinal Tract. Cell Host Microbe. 2019;25: 210–218. 10.1016/j.chom.2019.01.018 [Abstract] [CrossRef] [Google Scholar]
8. Shkoporov AN, Hill C. Bacteriophages of the Human Gut: The “Known Unknown” of the Microbiome. Cell Host Microbe. 2019;25: 195–209. 10.1016/j.chom.2019.01.017 [Abstract] [CrossRef] [Google Scholar]
9. Casjens SR, Hendrix RW. Bacteriophage lambda: Early pioneer and still relevant. Virology. 2015;479–480: 310–330. 10.1016/j.virol.2015.02.010 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
10. Karam JD, Drake JW. Molecular biology of bacteriophage T4. American Society for Microbiology; 1994.
11. Molineux I. T7 Bacteriophages In: Encyclopedia of Molecular Biology. John Wiley and Sons;2002. [Google Scholar]
12. Heller KJ. Molecular interaction between bacteriophage and the gram-negative cell envelope. Archives of Microbiology. 1992:235–248. 10.1007/bf00245239 [Abstract] [CrossRef] [Google Scholar]
13. De Smet J, Hendrix H, Blasdel BG, Danis-Wlodarczyk K, Lavigne R. Pseudomonas predators: understanding and exploiting phage-host interactions. Nat Rev Microbiol. 2017;15: 517–530. 10.1038/nrmicro.2017.61 [Abstract] [CrossRef] [Google Scholar]
14. Sandulache R, Prehm P, Kamp D. Cell wall receptor for bacteriophage Mu G(+). J Bacteriol. 1984;160: 299–303. 10.1128/JB.160.1.299-303.1984 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
15. Demerec M, Fano U. Bacteriophage-Resistant Mutants in Escherichia Coli. Genetics. 1945;30: 119–136. [Europe PMC free article] [Abstract] [Google Scholar]
16. Hancock RE, Reeves P. Bacteriophage resistance in Escherichia coli K-12: general pattern of resistance. J Bacteriol. 1975;121: 983–993. 10.1128/JB.121.3.983-993.1975 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
17. Chirakadze I, Perets A, Ahmed R. Phage Typing. Methods in Molecular Biology. 2009:293–305. 10.1007/978-1-60327-565-1_17 [Abstract] [CrossRef] [Google Scholar]
18. Bondy-Denomy J, Qian J, Westra ER, Buckling A, Guttman DS, Davidson AR, et al. Prophages mediate defense against phage infection through diverse mechanisms. ISME J. 2016;10: 2854–2866. 10.1038/ismej.2016.79 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
19. Moller AG, Lindsay JA, Read TD. Determinants of Phage Host Range in Staphylococcus Species. Appl Environ Microbiol. 2019;85 10.1128/AEM.00209-19 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
20. Trudelle DM, Bryan DW, Hudson LK, Denes TG. Cross-resistance to phage infection in Listeria monocytogenes serotype 1/2a mutants. Food Microbiol. 2019;84: 103239 10.1016/j.fm.2019.06.003 [Abstract] [CrossRef] [Google Scholar]
21. Kauffman KM, Hussain FA, Yang J, Arevalo P, Brown JM, Chang WK, et al. A major lineage of non-tailed dsDNA viruses as unrecognized killers of marine bacteria. Nature. 2018:118–122. 10.1038/nature25474 [Abstract] [CrossRef] [Google Scholar]
22. Porter NT, Hryckowian AJ, Merrill BD, Gardner JO, Singh S, Sonnenburg JL, et al. Multiple phase-variable mechanisms, including capsular polysaccharides, modify bacteriophage susceptibility in Bacteroides thetaiotaomicron. bioRxiv [Preprint]. 2019. [cited 2020 Jan 1]. 10.1101/521070 [CrossRef] [Google Scholar]
23. Schooley RT, Biswas B, Gill JJ, Hernandez-Morales A, Lancaster J, Lessor L, et al. Development and Use of Personalized Bacteriophage-Based Therapeutic Cocktails To Treat a Patient with a Disseminated Resistant Acinetobacter baumannii Infection. Antimicrob Agents Chemother. 2017;61 10.1128/AAC.00954-17 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
24. Dedrick RM, Guerrero-Bustamante CA, Garlena RA, Russell DA, Ford K, Harris K, et al. Engineered bacteriophages for treatment of a patient with a disseminated drug-resistant Mycobacterium abscessus. Nat Med. 2019;25: 730–733. 10.1038/s41591-019-0437-z [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
25. Dedrick RM, Jacobs-Sera D, Bustamante CAG, Garlena RA, Mavrich TN, Pope WH, et al. Prophage-mediated defence against viral attack and viral counter-defence. Nat Microbiol. 2017;2: 16251 10.1038/nmicrobiol.2016.251 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
26. Mathieu A, Dion M, Deng L, Tremblay D, Moncaut E, Shah SA, et al. Virulent coliphages in 1-year-old children fecal samples are fewer, but more infectious than temperate coliphages. Nat Commun. 2020;11: 378 10.1038/s41467-019-14042-z [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
27. Bourdin G, Navarro A, Sarker SA, Pittet A-C, Qadri F, Sultana S, et al. Coverage of diarrhoea-associated Escherichia coli isolates from different origins with two types of phage cocktails. Microb Biotechnol. 2014;7: 165–176. 10.1111/1751-7915.12113 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
28. Wright RCT, Friman V-P, Smith MCM, Brockhurst MA. Cross-resistance is modular in bacteria–phage interactions. PLoS Biol. 2018:e2006057 10.1371/journal.pbio.2006057 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
29. Denes T, den Bakker HC, Tokman JI, Guldimann C, Wiedmann M. Selection and Characterization of Phage-Resistant Mutant Strains of Listeria monocytogenes Reveal Host Genes Linked to Phage Adsorption. Applied and Environmental Microbiology. 2015:4295–4305. 10.1128/AEM.00087-15 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
30. Leprohon P, Gingras H, Ouennane S, Moineau S, Ouellette M. A genomic approach to understand interactions between Streptococcus pneumoniae and its bacteriophages. BMC Genomics. 2015;16: 972 10.1186/s12864-015-2134-8 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
31. Duerkop BA, Huo W, Bhardwaj P, Palmer KL, Hooper LV. Molecular Basis for Lytic Bacteriophage Resistance in Enterococci. MBio. 2016;7 10.1128/mBio.01304-16 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
32. Sumrall ET, Shen Y, Keller AP, Rismondo J, Pavlou M, Eugster MR, et al. Phage resistance at the cost of virulence: Listeria monocytogenes serovar 4b requires galactosylated teichoic acids for InlB-mediated invasion. PLoS Pathog. 2019;15: e1008032 10.1371/journal.ppat.1008032 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
33. Altamirano FG, Forsyth JH, Patwa R, Kostoulias X, Trim M, Subedi D, et al. Bacteriophages targeting Acinetobacter baumannii capsule induce antimicrobial resensitization. bioRxiv [Preprint]. 2020. [cited 2020 Mar 1]. 10.1101/2020.02.25.965590 [CrossRef] [Google Scholar]
34. Goldfarb T, Sberro H, Weinstock E, Cohen O, Doron S, Charpak‐Amikam Y, et al. BREX is a novel phage resistance system widespread in microbial genomes. The EMBO Journal. 2015:169–183. 10.15252/embj.201489455 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
35. Bernheim A, Sorek R. The pan-immune system of bacteria: antiviral defence as a community resource. Nat Rev Microbiol. 2019. 10.1038/s41579-019-0278-2 [Abstract] [CrossRef] [Google Scholar]
36. Ofir G, Sorek R. Contemporary Phage Biology: From Classic Models to New Insights. Cell. 2018;172: 1260–1270. 10.1016/j.cell.2017.10.045 [Abstract] [CrossRef] [Google Scholar]
37. Young R, Gill JJ. MICROBIOLOGY. Phage therapy redux—What is to be done? Science. 2015;350: 1163–1164. 10.1126/science.aad6791 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
38. Rostøl JT, Marraffini L. (Ph)ighting Phages: How Bacteria Resist Their Parasites. Cell Host Microbe. 2019;25: 184–194. 10.1016/j.chom.2019.01.009 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
39. Samson JE, Magadán AH, Sabri M, Moineau S. Revenge of the phages: defeating bacterial defences. Nature Reviews Microbiology. 2013:675–687. 10.1038/nrmicro3096 [Abstract] [CrossRef] [Google Scholar]
40. Campbell A. The future of bacteriophage biology. Nat Rev Genet. 2003;4: 471–477. 10.1038/nrg1089 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
41. Weitz JS, Poisot T, Meyer JR, Flores CO, Valverde S, Sullivan MB, et al. Phage-bacteria infection networks. Trends Microbiol. 2013;21: 82–91. 10.1016/j.tim.2012.11.003 [Abstract] [CrossRef] [Google Scholar]
42. Moller AG, Lindsay JA, Read TD. Determinants of Phage Host Range in Staphylococcus Species. Applied and Environmental Microbiology. 2019. 10.1128/aem.00209-19 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
43. de Jonge PA, de Jonge PA, Nobrega FL, Brouns SJJ, Dutilh BE. Molecular and Evolutionary Determinants of Bacteriophage Host Range. Trends in Microbiology. 2019:51–63. 10.1016/j.tim.2018.08.006 [Abstract] [CrossRef] [Google Scholar]
44. Dy RL, Richter C, Salmond GPC, Fineran PC. Remarkable Mechanisms in Microbes to Resist Phage Infections. Annu Rev Virol. 2014;1: 307–331. 10.1146/annurev-virology-031413-085500 [Abstract] [CrossRef] [Google Scholar]
45. Nobrega FL, Vlot M, de Jonge PA, Dreesens LL, Beaumont HJE, Lavigne R, et al. Targeting mechanisms of tailed bacteriophages. Nature Reviews Microbiology. 2018:760–773. 10.1038/s41579-018-0070-8 [Abstract] [CrossRef] [Google Scholar]
46. van Houte S, Buckling A, Westra ER. Evolutionary Ecology of Prokaryotic Immune Mechanisms. Microbiol Mol Biol Rev. 2016;80: 745–763. 10.1128/MMBR.00011-16 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
47. Brüssow H. Bacteriophage-host interaction: from splendid isolation into a messy reality. Curr Opin Microbiol. 2013;16: 500–506. 10.1016/j.mib.2013.04.007 [Abstract] [CrossRef] [Google Scholar]
48. Broeker NK, Barbirz S. Not a barrier but a key: How bacteriophages exploit host’s O-antigen as an essential receptor to initiate infection. Molecular Microbiology. 2017:353–357. 10.1111/mmi.13729 [Abstract] [CrossRef] [Google Scholar]
49. Christen M, Beusch C, Bösch Y, Cerletti D, Flores-Tinoco CE, Del Medico L, et al. Quantitative Selection Analysis of Bacteriophage [var phi]CbK Susceptibility in Caulobacter crescentus. J Mol Biol. 2016;428: 419–430. 10.1016/j.jmb.2015.11.018 [Abstract] [CrossRef] [Google Scholar]
50. Maynard ND, Birch EW, Sanghvi JC, Chen L, Gutschow MV, Covert MW. A forward-genetic screen and dynamic analysis of lambda phage host-dependencies reveals an extensive interaction network and a new anti-viral strategy. PLoS Genet. 2010;6: e1001017 10.1371/journal.pgen.1001017 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
51. Qimron U, Marintcheva B, Tabor S, Richardson CC. Genomewide screens for Escherichia coli genes affecting growth of T7 bacteriophage. Proc Natl Acad Sci U S A. 2006;103: 19039–19044. 10.1073/pnas.0609428103 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
52. Rousset F, Cui L, Siouve E, Becavin C, Depardieu F, Bikard D. Genome-wide CRISPR-dCas9 screens in E. coli identify essential genes and phage host factors. PLoS Genet. 2018;14: e1007749 10.1371/journal.pgen.1007749 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
53. Chan BK, Sistrom M, Wertz JE, Kortright KE, Narayan D, Turner PE. Phage selection restores antibiotic sensitivity in MDR Pseudomonas aeruginosa. Scientific Reports. 2016. 10.1038/srep26717 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
54. Bohm K, Porwollik S, Chu W, Dover JA, Gilcrease EB, Casjens SR, et al. Genes affecting progression of bacteriophage P22 infection in Salmonella identified by transposon and single gene deletion screens. Mol Microbiol. 2018;108: 288–305. 10.1111/mmi.13936 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
55. Howard-Varona C, Hargreaves KR, Solonenko NE, Markillie LM, White RA 3rd, Brewer HM, et al. Multiple mechanisms drive phage infection efficiency in nearly identical hosts. ISME J. 2018;12: 1605–1618. 10.1038/s41396-018-0099-8 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
56. Roucourt B, Lavigne R. The role of interactions between phage and bacterial proteins within the infected cell: a diverse and puzzling interactome. Environmental Microbiology. 2009:2789–2805. 10.1111/j.1462-2920.2009.02029.x [Abstract] [CrossRef] [Google Scholar]
57. Doron S, Melamed S, Ofir G, Leavitt A, Lopatina A, Keren M, et al. Systematic discovery of antiphage defense systems in the microbial pangenome. Science. 2018:eaar4120 10.1126/science.aar4120 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
58. Cowley LA, Low AS, Pickard D, Boinett CJ, Dallman TJ, Day M, et al. Transposon Insertion Sequencing Elucidates Novel Gene Involvement in Susceptibility and Resistance to Phages T4 and T7 in O157. MBio. 2018;9 10.1128/mBio.00705-18 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
59. Chatterjee A, Willett JLE, Nguyen UT, Monogue B, Palmer KL, Dunny GM, et al. Parallel genomics uncover novel enterococcal-bacteriophage interactions. mBio.2020; 11 10.1128/mBio.03120-19 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
60. Pickard D, Kingsley RA, Hale C, Turner K, Sivaraman K, Wetter M, et al. A genomewide mutagenesis screen identifies multiple genes contributing to Vi capsular expression in Salmonella enterica serovar Typhi. J Bacteriol. 2013;195: 1320–1326. 10.1128/JB.01632-12 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
61. Shin H, Lee J-H, Kim H, Choi Y, Heu S, Ryu S. Receptor diversity and host interaction of bacteriophages infecting Salmonella enterica serovar Typhimurium. PLoS ONE. 2012;7: e43392 10.1371/journal.pone.0043392 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
62. Filippov AA, Sergueev KV, He Y, Huang X-Z, Gnade BT, Mueller AJ, et al. Bacteriophage-resistant mutants in Yersinia pestis: identification of phage receptors and attenuation for mice. PLoS ONE. 2011;6: e25486 10.1371/journal.pone.0025486 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
63. Piya D, Lessor L, Koehler B, Stonecipher A, Cahill J, Gill JJ. Genome-wide screens reveal Escherichia coli genes required for growth of T1-like phage LL5 and V5-like phage LL12. Scientific Reports. 2020. 10.1038/s41598-020-64981-7 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
64. Wetmore KM, Price MN, Waters RJ, Lamson JS, He J, Hoover CA, et al. Rapid quantification of mutant fitness in diverse bacteria by sequencing randomly bar-coded transposons. MBio. 2015;6: e00306–15. 10.1128/mBio.00306-15 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
65. Qi LS, Larson MH, Gilbert LA, Doudna JA, Weissman JS, Arkin AP, et al. Repurposing CRISPR as an RNA-guided platform for sequence-specific control of gene expression. Cell. 2013;152: 1173–1183. 10.1016/j.cell.2013.02.022 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
66. Mutalik VK, Novichkov PS, Price MN, Owens TK, Callaghan M, Carim S, et al. Dual-barcoded shotgun expression library sequencing for high-throughput characterization of functional traits in bacteria. Nat Commun. 2019;10: 308 10.1038/s41467-018-08177-8 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
67. Smith AM, Heisler LE, Mellor J, Kaper F, Thompson MJ, Chee M, et al. Quantitative phenotyping via deep barcode sequencing. Genome Research. 2009:1836–1842. 10.1101/gr.093955.109 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
68. Price MN, Wetmore KM, Jordan Waters R, Callaghan M, Ray J, Liu H, et al. Mutant phenotypes for thousands of bacterial genes of unknown function. Nature. 2018:503–509. 10.1038/s41586-018-0124-0 [Abstract] [CrossRef] [Google Scholar]
69. Rishi HS, Toro E, Liu H, Wang X, Qi LS, Arkin AP. Systematic genome-wide querying of coding and non-coding functional elements in E. coli using CRISPRi. bioRxiv [Preprint]. 2020. [cited 2020 Mar 10]. 10.1101/2020.03.04.975888 [CrossRef] [Google Scholar]
70. Peters JM, Koo B-M, Patino R, Heussler GE, Hearne CC, Qu J, et al. Enabling genetic analysis of diverse bacteria with Mobile-CRISPRi. Nat Microbiol. 2019;4: 244–250. 10.1038/s41564-018-0327-z [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
71. Rock JM, Hopkins FF, Chavez A, Diallo M, Chase MR, Gerrick ER, et al. Programmable transcriptional repression in mycobacteria using an orthogonal CRISPR interference platform. Nature Microbiology. 2017. 10.1038/nmicrobiol.2016.274 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
72. Liu X, Gallay C, Kjos M, Domenech A, Slager J, Kessel SP, et al. High‐throughput CRISPRi phenotyping identifies new essential genes in Streptococcus pneumoniae. Molecular Systems Biology. 2017:931 10.15252/msb.20167449 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
73. Cui L, Vigouroux A, Rousset F, Varet H, Khanna V, Bikard D. A CRISPRi screen in E. coli reveals sequence-specific toxicity of dCas9. Nature Communications. 2018. 10.1038/s41467-018-04209-5 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
74. Wang T, Guan C, Guo J, Liu B, Wu Y, Xie Z, et al. Pooled CRISPR interference screening enables genome-scale functional genomics study in bacteria with superior performance. Nat Commun. 2018;9: 2475 10.1038/s41467-018-04899-x [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
75. Hogan AM, Zisanur AS, Lightly TJ, Cardona ST. A Broad-Host-Range CRISPRi Toolkit for Silencing Gene Expression in Burkholderia. ACS Synthetic Biology. 2019:2372–2384. 10.1021/acssynbio.9b00232 [Abstract] [CrossRef] [Google Scholar]
76. Müh U, Pannullo AG, Weiss DS, Ellermeier CD. A Xylose-Inducible Expression System and a CRISPR Interference Plasmid for Targeted Knockdown of Gene Expression in Clostridioides difficile. Journal of Bacteriology. 2019. 10.1128/jb.00711-18 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
77. Qu J, Prasad NK, Yu MA, Chen S, Lyden A, Herrera N, et al. Modulating Pathogenesis with Mobile-CRISPRi. J Bacteriol. 2019;201 10.1128/JB.00304-19 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
78. Guzzo M, Castro LK, Reisch CR, Guo MS, Laub MT. A CRISPR Interference System for Efficient and Rapid Gene Knockdown in Caulobacter crescentus. MBio. 2020;11 10.1128/mBio.02415-19 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
79. Häuser R, Blasche S, Dokland T, Haggård-Ljungquist E, von Brunn A, Salas M, et al. Bacteriophage protein-protein interactions. Adv Virus Res. 2012;83: 219–298. 10.1016/B978-0-12-394438-2.00006-2 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
80. Blount ZD. The unexhausted potential of E. coli. Elife. 2015;4: e05826. [Europe PMC free article] [Abstract] [Google Scholar]
81. Zimmer C. Microcosm: E. coli and the New Science of Life. Vintage; 2008.
82. Ghatak S, King ZA, Sastry A, Palsson BO. The y-ome defines the 35% of Escherichia coli genes that lack experimental evidence of function. Nucleic Acids Res. 2019;47: 2446–2454. 10.1093/nar/gkz030 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
83. Sarker SA, Sultana S, Reuteler G, Moine D, Descombes P, Charton F, et al. Oral Phage Therapy of Acute Bacterial Diarrhea With Two Coliphage Preparations: A Randomized Trial in Children From Bangladesh. EBioMedicine. 2016;4: 124–137. 10.1016/j.ebiom.2015.12.023 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
84. Brüssow H. Hurdles for Phage Therapy to Become a Reality—An Editorial Comment. Viruses. 2019:557 10.3390/v11060557 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
85. Kaper JB, Nataro JP, Mobley HL. Pathogenic Escherichia coli. Nat Rev Microbiol. 2004;2: 123–140. 10.1038/nrmicro818 [Abstract] [CrossRef] [Google Scholar]
86. Croxen MA, Brett Finlay B. Molecular mechanisms of Escherichia coli pathogenicity. Nature Reviews Microbiology. 2010:26–38. 10.1038/nrmicro2265 [Abstract] [CrossRef] [Google Scholar]
87. Smith H. Virulence determinants of Escherichia coli: present knowledge and questions. Canadian Journal of Microbiology. 1992:747–752. 10.1139/m92-121 [Abstract] [CrossRef] [Google Scholar]
88. Jeong H, Barbe V, Lee CH, Vallenet D, Yu DS, Choi S-H, et al. Genome sequences of Escherichia coli B strains REL606 and BL21(DE3). J Mol Biol. 2009;394: 644–652. 10.1016/j.jmb.2009.09.052 [Abstract] [CrossRef] [Google Scholar]
89. Daegelen P, Studier FW, Lenski RE, Cure S, Kim JF. Tracing ancestors and relatives of Escherichia coli B, and the derivation of B strains REL606 and BL21(DE3). J Mol Biol. 2009;394: 634–643. 10.1016/j.jmb.2009.09.022 [Abstract] [CrossRef] [Google Scholar]
90. Studier FW, Daegelen P, Lenski RE, Maslov S, Kim JF. Understanding the differences between genome sequences of Escherichia coli B strains REL606 and BL21(DE3) and comparison of the E. coli B and K-12 genomes. J Mol Biol. 2009;394: 653–680. 10.1016/j.jmb.2009.09.021 [Abstract] [CrossRef] [Google Scholar]
91. Wright A, McConnell M, Kanegasaki S. Lipopolysaccharide as a Bacteriophage Receptor. Virus Receptors. 1980:27–57. 10.1007/978-94-011-6918-9_3 [CrossRef] [Google Scholar]
92. Schwartz M. Interaction of Phages with their Receptor Proteins. Virus Receptors. 1980:59–94. 10.1007/978-94-011-6918-9_4 [CrossRef] [Google Scholar]
93. Lindberg AA. Bacteriophage receptors. Annu Rev Microbiol. 1973;27: 205–241. 10.1146/annurev.mi.27.100173.001225 [Abstract] [CrossRef] [Google Scholar]
94. Calendar R. The Bacteriophages. Oxford University Press on Demand; 2006. [Google Scholar]
95. Lenski RE. Dynamics of Interactions between Bacteria and Virulent Bacteriophage. Advances in Microbial Ecology. 1988:1–44. 10.1007/978-1-4684-5409-3_1 [CrossRef] [Google Scholar]
96. Bertani G. Lysogeny at mid-twentieth century: P1, P2, and other experimental systems. J Bacteriol. 2004;186: 595–600. 10.1128/jb.186.3.595-600.2004 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
97. Kortright KE, Chan BK, Koff JL, Turner PE. Phage Therapy: A Renewed Approach to Combat Antibiotic-Resistant Bacteria. Cell Host Microbe. 2019;25: 219–232. 10.1016/j.chom.2019.01.014 [Abstract] [CrossRef] [Google Scholar]
98. Bertozzi Silva J, Storms Z, Sauvageau D. Host receptors for bacteriophage adsorption. FEMS Microbiol Lett. 2016;363 10.1093/femsle/fnw002 [Abstract] [CrossRef] [Google Scholar]
99. Heller KJ. Molecular interaction between bacteriophage and the gram-negative cell envelope. Archives of Microbiology. 1992:235–248. 10.1007/bf00245239 [Abstract] [CrossRef] [Google Scholar]
100. Kiino DR, Rothman-Denes LB. Genetic analysis of bacteriophage N4 adsorption. J Bacteriol. 1989;171: 4595–4602. 10.1128/jb.171.9.4595-4602.1989 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
101. Letarov AV, Kulikov EE. Adsorption of Bacteriophages on Bacterial Cells. Biochemistry. 2017;82: 1632–1658. 10.1134/S0006297917130053 [Abstract] [CrossRef] [Google Scholar]
102. Raya RR, Oot RA, Moore-Maley B, Wieland S, Callaway TR, Kutter EM, et al. Naturally resident and exogenously applied T4-like and T5-like bacteriophages can reduce Escherichia coli O157:H7 levels in sheep guts. Bacteriophage. 2011;1: 15–24. 10.4161/bact.1.1.14175 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
103. Kutter E, Gachechiladze K, Poglazov A, Marusich E, Shneider M, Aronsson P, et al. Evolution of T4-related phages. Virus Genes. 1995;11: 285–297. 10.1007/BF01728666 [Abstract] [CrossRef] [Google Scholar]
104. Franklin NC. Mutation in gal U gene of E. coli blocks phage P1 infection. Virology. 1969:189–191. 10.1016/0042-6822(69)90144-5 [Abstract] [CrossRef] [Google Scholar]
105. Baba T, Ara T, Hasegawa M, Takai Y, Okumura Y, Baba M, et al. Construction of Escherichia coli K-12 in-frame, single-gene knockout mutants: the Keio collection. Mol Syst Biol. 2006;2: 2006.0008. [Europe PMC free article] [Abstract] [Google Scholar]
106. Wall E, Majdalani N, Gottesman S. The Complex Rcs Regulatory Cascade. Annu Rev Microbiol. 2018;72: 111–139. 10.1146/annurev-micro-090817-062640 [Abstract] [CrossRef] [Google Scholar]
107. Chamberlin M. Isolation and characterization of prototrophic mutants of Escherichia coli unable to support the intracellular growth of T7. J Virol. 1974;14: 509–516. 10.1128/JVI.14.3.509-516.1974 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
108. Friedman DI, Olson ER, Georgopoulos C, Tilly K, Herskowitz I, Banuett F. Interactions of bacteriophage and host macromolecules in the growth of bacteriophage lambda. Microbiol Rev. 1984;48: 299–325. [Europe PMC free article] [Abstract] [Google Scholar]
109. Friedman DI. Interaction between bacteriophage λ and its Escherichia coli host. Current Opinion in Genetics & Development. 1992:727–738. 10.1016/s0959-437x(05)80133-9 [Abstract] [CrossRef] [Google Scholar]
110. Polissi A, Goffin L, Georgopoulos C. The Escherichia coli heat shock response and bacteriophage lambda development. FEMS Microbiol Rev. 1995;17: 159–169. 10.1111/j.1574-6976.1995.tb00198.x [Abstract] [CrossRef] [Google Scholar]
111. Karow M, Raina S, Georgopoulos C, Fayet O. Complex phenotypes of null mutations in the htr genes, whose products are essential for Escherichia coli growth at elevated temperatures. Res Microbiol. 1991;142: 289–294. 10.1016/0923-2508(91)90043-a [Abstract] [CrossRef] [Google Scholar]
112. Pagnout C, Sohm B, Razafitianamaharavo A, Caillet C, Offroy M, Leduc M, et al. Pleiotropic effects of rfa-gene mutations on Escherichia coli envelope properties. Sci Rep. 2019;9: 9696 10.1038/s41598-019-46100-3 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
113. Klein G, Raina S. Regulated Assembly of LPS, Its Structural Alterations and Cellular Response to LPS Defects. Int J Mol Sci. 2019;20 10.3390/ijms20020356 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
114. Mitchell AM, Silhavy TJ. Envelope stress responses: balancing damage repair and toxicity. Nat Rev Microbiol. 2019;17: 417–428. 10.1038/s41579-019-0199-0 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
115. Meaden S, Paszkiewicz K, Koskella B. The cost of phage resistance in a plant pathogenic bacterium is context‐dependent. Evolution. 2015:1321–1328. 10.1111/evo.12652 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
116. Díaz-Muñoz SL, Koskella B. Bacteria–Phage Interactions in Natural Environments. Advances in Applied Microbiology. 2014:135–183. 10.1016/B978-0-12-800259-9.00004-4 [Abstract] [CrossRef] [Google Scholar]
117. Mirzaei MK, Maurice CF. Ménage à trois in the human gut: interactions between host, bacteria and phages. Nature Reviews Microbiology. 2017:397–408. 10.1038/nrmicro.2017.30 [Abstract] [CrossRef] [Google Scholar]
118. Padfield D, Castledine M, Buckling A. Temperature-dependent changes to host-parasite interactions alter the thermal performance of a bacterial host. ISME J. 2020;14: 389–398. 10.1038/s41396-019-0526-5 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
119. Alseth EO, Pursey E, Luján AM, McLeod I, Rollie C, Westra ER. Bacterial biodiversity drives the evolution of CRISPR-based phage resistance. Nature. 2019;574: 549–552. 10.1038/s41586-019-1662-9 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
120. Kortright KE, Chan BK, Turner PE. High-throughput discovery of phage receptors using transposon insertion sequencing of bacteria. Proc Natl Acad Sci U S A. 2020. 10.1073/pnas.2001888117 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
121. Shibuya I. Metabolic regulations and biological functions of phospholipids in Escherichia coli. Prog Lipid Res. 1992;31: 245–299. 10.1016/0163-7827(92)90010-g [Abstract] [CrossRef] [Google Scholar]
122. Dalebroux ZD, Edrozo MB, Pfuetzner RA, Ressl S, Kulasekara BR, Blanc M-P, et al. Delivery of cardiolipins to the Salmonella outer membrane is necessary for survival within host tissues and virulence. Cell Host Microbe. 2015;17: 441–451. 10.1016/j.chom.2015.03.003 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
123. Nepper JF, Lin YC, Weibel DB. Rcs phosphorelay activation in cardiolipin-deficient Escherichia coli reduces biofilm formation. J Bacteriol. 2019: 201(9): e00804–18. 10.1128/JB.00804-18 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
124. Zhou Y, Filter JJ, Court DL, Gottesman ME, Friedman DI. Requirement for NusG for transcription antitermination in vivo by the lambda N protein. J Bacteriol. 2002;184: 3416–3418. 10.1128/jb.184.12.3416-3418.2002 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
125. Sullivan SL, Gottesman ME. Requirement for E. coli NusG protein in factor-dependent transcription termination. Cell. 1992;68: 989–994. 10.1016/0092-8674(92)90041-a [Abstract] [CrossRef] [Google Scholar]
126. Henthorn KS. Multilevel regulation of gene expression in lambda and related bacteriophages [thesis]. University of Michigan; 1994. http://hdl.handle.net/2027.42/103954.
127. Bertani B, Ruiz N. Function and Biogenesis of Lipopolysaccharides. EcoSal Plus. 2018;8 10.1128/ecosalplus.ESP-0001-2018 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
128. Konovalova A, Kahne DE, Silhavy TJ. Outer Membrane Biogenesis. Annu Rev Microbiol. 2017;71: 539–556. 10.1146/annurev-micro-090816-093754 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
129. Potts AH, Vakulskas CA, Pannuri A, Yakhnin H, Babitzke P, Romeo T. Global role of the bacterial post-transcriptional regulator CsrA revealed by integrated transcriptomics. Nat Commun. 2017;8: 1596 10.1038/s41467-017-01613-1 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
130. Romeo T, Babitzke P. Global Regulation by CsrA and Its RNA Antagonists. Microbiol Spectr. 2018;6 10.1128/microbiolspec.RWR-0009-2017 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
131. Mojardín L, Salas M. Global Transcriptional Analysis of Virus-Host Interactions between Phage ϕ29 and Bacillus subtilis. Journal of Virology. 2016:9293–9304. 10.1128/JVI.01245-16 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
132. Chevallereau A, Blasdel BG, De Smet J, Monot M, Zimmermann M, Kogadeeva M, et al. Next-Generation “-omics” Approaches Reveal a Massive Alteration of Host RNA Metabolism during Bacteriophage Infection of Pseudomonas aeruginosa. PLoS Genet. 2016;12: e1006134 10.1371/journal.pgen.1006134 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
133. Hantke K. Major outer membrane proteins of E. coli K12 serve as receptors for the phages T2 (protein Ia) and 434 (protein Ib). Molecular and General Genetics MGG. 1978:131–135. 10.1007/bf00267377 [Abstract] [CrossRef] [Google Scholar]
134. Grigonyte AM, Harrison C, MacDonald PR, Montero-Blay A, Tridgett M, Duncan J, et al. Comparison of CRISPR and Marker-Based Methods for the Engineering of Phage T7. Viruses. 2020;12 10.3390/v12020193 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
135. Prelich G. Gene overexpression: uses, mechanisms, and interpretation. Genetics. 2012;190: 841–854. 10.1534/genetics.111.136911 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
136. Veitia RA, Bottani S, Birchler JA. Gene dosage effects: nonlinearities, genetic interactions, and dosage compensation. Trends Genet. 2013;29: 385–393. [Abstract] [Google Scholar]
137. van Leeuwen J, Pons C, Boone C, Andrews BJ. Mechanisms of suppression: The wiring of genetic resilience. Bioessays. 2017;39 10.1002/bies.201700042 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
138. Holzmayer TA, Pestov DG, Roninson IB. Isolation of dominant negative mutants and inhibitory antisense RNA sequences by expression selection of random DNA fragments. Nucleic Acids Res. 1992;20: 711–717. 10.1093/nar/20.4.711 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
139. Sopko R, Huang D, Preston N, Chua G, Papp B, Kafadar K, et al. Mapping pathways and phenotypes by systematic gene overexpression. Mol Cell. 2006;21: 319–330. 10.1016/j.molcel.2005.12.011 [Abstract] [CrossRef] [Google Scholar]
140. Cardarelli L, Maxwell KL, Davidson AR. Assembly mechanism is the key determinant of the dosage sensitivity of a phage structural protein. Proc Natl Acad Sci U S A. 2011;108: 10168–10173. 10.1073/pnas.1100759108 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
141. Plumbridge J. Control of the expression of the manXYZ operon in Escherichia coli: Mlc is a negative regulator of the mannose PTS. Mol Microbiol. 1998;27: 369–380. 10.1046/j.1365-2958.1998.00685.x [Abstract] [CrossRef] [Google Scholar]
142. Decker K, Plumbridge J, Boos W. Negative transcriptional regulation of a positive regulator: the expression of malT, encoding the transcriptional activator of the maltose regulon of Escherichia coli, is negatively controlled by Mlc. Mol Microbiol. 1998;27: 381–390. 10.1046/j.1365-2958.1998.00694.x [Abstract] [CrossRef] [Google Scholar]
143. Joly N, Danot O, Schlegel A, Boos W, Richet E. The Aes protein directly controls the activity of MalT, the central transcriptional activator of the Escherichia coli maltose regulon. J Biol Chem. 2002;277: 16606–16613. 10.1074/jbc.M200991200 [Abstract] [CrossRef] [Google Scholar]
144. Lee SJ, Boos W, Bouché JP, Plumbridge J. Signal transduction between a membrane-bound transporter, PtsG, and a soluble transcription factor, Mlc, of Escherichia coli. EMBO J. 2000;19: 5353–5361. 10.1093/emboj/19.20.5353 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
145. Schlegel A, Danot O, Richet E, Ferenci T, Boos W. The N Terminus of the Escherichia coli Transcription Activator MalT Is the Domain of Interaction with MalY. Journal of Bacteriology. 2002:3069–3077. 10.1128/jb.184.11.3069-3077.2002 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
146. Reidl J, Boos W. The malX malY operon of Escherichia coli encodes a novel enzyme II of the phosphotransferase system recognizing glucose and maltose and an enzyme abolishing the endogenous induction of the maltose system. J Bacteriol. 1991;173: 4862–4876. 10.1128/jb.173.15.4862-4876.1991 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
147. Lengsfeld C, Schönert S, Dippel R, Boos W. Glucose- and glucokinase-controlled mal gene expression in Escherichia coli. J Bacteriol. 2009;191: 701–712. 10.1128/JB.00767-08 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
148. Hong JS, Smith GR, Ames BN. Adenosine 3’:5’-cyclic monophosphate concentration in the bacterial host regulates the viral decision between lysogeny and lysis. Proc Natl Acad Sci U S A. 1971;68: 2258–2262. 10.1073/pnas.68.9.2258 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
149. Grodzicker T, Arditti RR, Eisen H. Establishment of repression by lambdoid phage in catabolite activator protein and adenylate cyclase mutants of Escherichia coli. Proc Natl Acad Sci U S A. 1972;69: 366–370. 10.1073/pnas.69.2.366 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
150. Jenal U, Reinders A, Lori C. Cyclic di-GMP: second messenger extraordinaire. Nat Rev Microbiol. 2017;15: 271–284. 10.1038/nrmicro.2016.190 [Abstract] [CrossRef] [Google Scholar]
151. Römling U, Galperin MY, Gomelsky M. Cyclic di-GMP: the first 25 years of a universal bacterial second messenger. Microbiol Mol Biol Rev. 2013;77: 1–52. 10.1128/MMBR.00043-12 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
152. Hengge R. Principles of c-di-GMP signalling in bacteria. Nat Rev Microbiol. 2009;7: 263–273. 10.1038/nrmicro2109 [Abstract] [CrossRef] [Google Scholar]
153. Hall BG. The rtn gene of Proteus vulgaris is actually from Escherichia coli. Journal of Bacteriology. 1997:2433–2434. 10.1128/jb.179.7.2433-2434.1997 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
154. Chae KS, Yoo OJ. Cloning of the lambda resistant genes from Brevibacterium albidum and Proteus vulgaris into Escherichia coli. Biochem Biophys Res Commun. 1986;140: 1101–1105. 10.1016/0006-291x(86)90748-5 [Abstract] [CrossRef] [Google Scholar]
155. Norioka S, Ramakrishnan G, Ikenaka K, Inouye M. Interaction of a transcriptional activator, OmpR, with reciprocally osmoregulated genes, ompF and ompC, of Escherichia coli. J Biol Chem. 1986;261: 17113–17119. [Abstract] [Google Scholar]
156. Yoshida T, Qin L, Egger LA, Inouye M. Transcription regulation of ompF and ompC by a single transcription factor, OmpR. J Biol Chem. 2006;281: 17114–17123. 10.1074/jbc.M602112200 [Abstract] [CrossRef] [Google Scholar]
157. Mizuno T, Mizushima S. Signal transduction and gene regulation through the phosphorylation of two regulatory components: the molecular basis for the osmotic regulation of the porin genes. Mol Microbiol. 1990;4: 1077–1082. 10.1111/j.1365-2958.1990.tb00681.x [Abstract] [CrossRef] [Google Scholar]
158. Andersen J, Forst SA, Zhao K, Inouye M, Delihas N. The function of micF RNA. micF RNA is a major factor in the thermal regulation of OmpF protein in Escherichia coli. J Biol Chem. 1989;264: 17961–17970. [Abstract] [Google Scholar]
159. Deighan P, Free A, Dorman CJ. A role for the Escherichia coli H-NS-like protein StpA in OmpF porin expression through modulation of micF RNA stability. Mol Microbiol. 2000;38: 126–139. 10.1046/j.1365-2958.2000.02120.x [Abstract] [CrossRef] [Google Scholar]
160. Kao C, Snyder L. The lit gene product which blocks bacteriophage T4 late gene expression is a membrane protein encoded by a cryptic DNA element, e14. J Bacteriol. 1988;170: 2056–2062. 10.1128/jb.170.5.2056-2062.1988 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
161. Kitagawa M, Ara T, Arifuzzaman M, Ioka-Nakamichi T, Inamoto E, Toyonaga H, et al. Complete set of ORF clones of Escherichia coli ASKA library (a complete set of E. coli K-12 ORF archive): unique resources for biological research. DNA Res. 2005;12: 291–299. 10.1093/dnares/dsi012 [Abstract] [CrossRef] [Google Scholar]
162. Cano DA, Domínguez-Bernal G, Tierrez A, Garcia-Del Portillo F, Casadesús J. Regulation of capsule synthesis and cell motility in Salmonella enterica by the essential gene igaA. Genetics. 2002;162: 1513–1523. [Europe PMC free article] [Abstract] [Google Scholar]
163. Cho S-H, Szewczyk J, Pesavento C, Zietek M, Banzhaf M, Roszczenko P, et al. Detecting envelope stress by monitoring β-barrel assembly. Cell. 2014;159: 1652–1664. 10.1016/j.cell.2014.11.045 [Abstract] [CrossRef] [Google Scholar]
164. Stevenson G, Andrianopoulos K, Hobbs M, Reeves PR. Organization of the Escherichia coli K-12 gene cluster responsible for production of the extracellular polysaccharide colanic acid. Journal of bacteriology. 1996:4885–4893. 10.1128/jb.178.16.4885-4893.1996 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
165. Hussein NA, Cho S-H, Laloux G, Siam R, Collet J-F. Distinct domains of Escherichia coli IgaA connect envelope stress sensing and down-regulation of the Rcs phosphorelay across subcellular compartments. PLoS Genet. 2018;14: e1007398 10.1371/journal.pgen.1007398 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
166. Majdalani N, Gottesman S. The Rcs phosphorelay: a complex signal transduction system. Annu Rev Microbiol. 2005;59: 379–405. 10.1146/annurev.micro.59.050405.101230 [Abstract] [CrossRef] [Google Scholar]
167. Chaudhry W, Lee E, Worthy A, Weiss Z, Grabowicz M, Vega N, et al. Mucoidy, a general mechanism for maintaining lytic phage in populations of bacteria. FEMS Microbiology Ecology. 2020. 10.1093/femsec/fiaa162 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
168. Scanlan PD, Buckling A. Co-evolution with lytic phage selects for the mucoid phenotype of Pseudomonas fluorescens SBW25. ISME J. 2012;6: 1148–1158. 10.1038/ismej.2011.174 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
169. Martin DR. Mucoid Variation in Pseudomonas AeruginosaInduced by the Action of Phage. Journal of Medical Microbiology. 1973:111–118. 10.1099/00222615-6-1-111 [Abstract] [CrossRef] [Google Scholar]
170. Mizoguchi K, Morita M, Fischer CR, Yoichi M, Tanji Y, Unno H. Coevolution of bacteriophage PP01 and Escherichia coli O157:H7 in continuous culture. Appl Environ Microbiol. 2003;69: 170–176. 10.1128/aem.69.1.170-176.2003 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
171. Keseler IM, Mackie A, Santos-Zavaleta A, Billington R, Bonavides-Martínez C, Caspi R, et al. The EcoCyc database: reflecting new knowledge about Escherichia coli K-12. Nucleic Acids Res. 2017;45: D543–D550. 10.1093/nar/gkw1003 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
172. Reinders A, Hee C-S, Ozaki S, Mazur A, Boehm A, Schirmer T, et al. Expression and Genetic Activation of Cyclic Di-GMP-Specific Phosphodiesterases in Escherichia coli. J Bacteriol. 2016;198: 448–462. 10.1128/JB.00604-15 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
173. Hengge R, Galperin MY, Ghigo J-M, Gomelsky M, Green J, Hughes KT, et al. Systematic Nomenclature for GGDEF and EAL Domain-Containing Cyclic Di-GMP Turnover Proteins of Escherichia coli. J Bacteriol. 2016;198: 7–11. 10.1128/JB.00424-15 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
174. Povolotsky TL, Hengge R. Genome-Based Comparison of Cyclic Di-GMP Signaling in Pathogenic and Commensal Escherichia coli Strains. J Bacteriol. 2016;198: 111–126. 10.1128/JB.00520-15 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
175. Park JH, Lee KH, Kim TY, Lee SY. Metabolic engineering of Escherichia coli for the production of L-valine based on transcriptome analysis and in silico gene knockout simulation. Proc Natl Acad Sci U S A. 2007;104: 7797–7802. 10.1073/pnas.0702609104 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
176. Lehnen D, Blumer C, Polen T, Wackwitz B, Wendisch VF, Unden G. LrhA as a new transcriptional key regulator of flagella, motility and chemotaxis genes in Escherichia coli. Mol Microbiol. 2002;45: 521–532. 10.1046/j.1365-2958.2002.03032.x [Abstract] [CrossRef] [Google Scholar]
177. Göpel Y, Khan MA, Görke B. Ménage à trois: post-transcriptional control of the key enzyme for cell envelope synthesis by a base-pairing small RNA, an RNase adaptor protein, and a small RNA mimic. RNA Biol. 2014;11: 433–442. 10.4161/rna.28301 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
178. McPartland J, Rothman-Denes LB. The tail sheath of bacteriophage N4 interacts with the Escherichia coli receptor. J Bacteriol. 2009;191: 525–532. 10.1128/JB.01423-08 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
179. Washizaki A, Yonesaki T, Otsuka Y. Characterization of the interactions between Escherichia coli receptors, LPS and OmpC, and bacteriophage T4 long tail fibers. Microbiologyopen. 2016;5: 1003–1015. 10.1002/mbo3.384 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
180. Hattman S, Fukasawa T. HOST-INDUCED MODIFICATION OF T-EVEN PHAGES DUE TO DEFECTIVE GLUCOSYLATION OF THEIR DNA. Proc Natl Acad Sci U S A. 1963;50: 297–300. 10.1073/pnas.50.2.297 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
181. Prehm P, Jann B, Jann K, Schmidt G, Stirm S. On a bacteriophage T3 and T4 receptor region within the cell wall lipopolysaccharide of escherichia coli B. J Mol Biol. 1976;101: 277–281. 10.1016/0022-2836(76)90377-6 [Abstract] [CrossRef] [Google Scholar]
182. Lenski RE. Two-step resistance by Escherichia coli B to bacteriophage T2. Genetics. 1984;107: 1–7. [Europe PMC free article] [Abstract] [Google Scholar]
183. Boos W, Shuman H. Maltose/maltodextrin system of Escherichia coli: transport, metabolism, and regulation. Microbiol Mol Biol Rev. 1998;62: 204–229. [Europe PMC free article] [Abstract] [Google Scholar]
184. Plumbridge J. Control of the expression of the manXYZ operon in Escherichia coli: Mlc is a negative regulator of the mannose PTS. Mol Microbiol. 1998;27: 369–380. 10.1046/j.1365-2958.1998.00685.x [Abstract] [CrossRef] [Google Scholar]
185. Decker K, Plumbridge J, Boos W. Negative transcriptional regulation of a positive regulator: the expression of malT, encoding the transcriptional activator of the maltose regulon of Escherichia coli, is negatively controlled by Mlc. Mol Microbiol. 1998;27: 381–390. 10.1046/j.1365-2958.1998.00694.x [Abstract] [CrossRef] [Google Scholar]
186. Lee SJ, Boos W, Bouché JP, Plumbridge J. Signal transduction between a membrane-bound transporter, PtsG, and a soluble transcription factor, Mlc, of Escherichia coli. EMBO J. 2000;19: 5353–5361. 10.1093/emboj/19.20.5353 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
187. Lee SJ, Boos W, Bouché JP, Plumbridge J. Signal transduction between a membrane-bound transporter, PtsG, and a soluble transcription factor, Mlc, of Escherichia coli. EMBO J. 2000;19: 5353–5361. 10.1093/emboj/19.20.5353 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
188. Danese PN, Pratt LA, Kolter R. Exopolysaccharide production is required for development of Escherichia coli K-12 biofilm architecture. J Bacteriol. 2000;182: 3593–3596. 10.1128/jb.182.12.3593-3596.2000 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
189. Eydallin G, Morán-Zorzano MT, Muñoz FJ, Baroja-Fernández E, Montero M, Alonso-Casajús N, et al. An Escherichia coli mutant producing a truncated inactive form of GlgC synthesizes glycogen: further evidences for the occurrence of various important sources of ADPglucose in enterobacteria. FEBS Lett. 2007;581: 4417–4422. 10.1016/j.febslet.2007.08.016 [Abstract] [CrossRef] [Google Scholar]
190. Baker CS, Morozov I, Suzuki K, Romeo T, Babitzke P. CsrA regulates glycogen biosynthesis by preventing translation of glgC in Escherichia coli. Mol Microbiol. 2002;44: 1599–1610. 10.1046/j.1365-2958.2002.02982.x [Abstract] [CrossRef] [Google Scholar]
191. Romeo T, Gong M, Liu MY, Brun-Zinkernagel AM. Identification and molecular characterization of csrA, a pleiotropic gene from Escherichia coli that affects glycogen biosynthesis, gluconeogenesis, cell size, and surface properties. J Bacteriol. 1993;175: 4744–4755. 10.1128/jb.175.15.4744-4755.1993 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
192. Wang X, Dubey AK, Suzuki K, Baker CS, Babitzke P, Romeo T. CsrA post-transcriptionally represses pgaABCD, responsible for synthesis of a biofilm polysaccharide adhesin of Escherichia coli. Mol Microbiol. 2005;56: 1648–1663. 10.1111/j.1365-2958.2005.04648.x [Abstract] [CrossRef] [Google Scholar]
193. Jackson DW, Suzuki K, Oakford L, Simecka JW, Hart ME, Romeo T. Biofilm formation and dispersal under the influence of the global regulator CsrA of Escherichia coli. J Bacteriol. 2002;184: 290–301. 10.1128/jb.184.1.290-301.2002 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
194. Palaniyandi S, Mitra A, Herren CD, Lockatell CV, Johnson DE, Zhu X, et al. BarA-UvrY two-component system regulates virulence of uropathogenic E. coli CFT073. PLoS ONE. 2012;7: e31348 10.1371/journal.pone.0031348 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
195. Klein G, Raina S. Regulated Control of the Assembly and Diversity of LPS by Noncoding sRNAs. Biomed Res Int. 2015;2015: 153561 10.1155/2015/153561 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
196. Murata A, Odaka M, Mukuno S. The Bacteriophage-Inactivating Effect of Basic Amino Acids; Arginine, Histidine, and Lysine. Agric Biol Chem. 1974;38: 477–478. [Google Scholar]
197. Lau CKY, Krewulak KD, Vogel HJ. Bacterial ferrous iron transport: the Feo system. FEMS Microbiol Rev. 2016;40: 273–298. 10.1093/femsre/fuv049 [Abstract] [CrossRef] [Google Scholar]
198. Kammler M, Schön C, Hantke K. Characterization of the ferrous iron uptake system of Escherichia coli. J Bacteriol. 1993;175: 6212–6219. 10.1128/jb.175.19.6212-6219.1993 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
199. Hantke K. Regulation of ferric iron transport in Escherichia coli K12: isolation of a constitutive mutant. Mol Gen Genet. 1981;182: 288–292. 10.1007/BF00269672 [Abstract] [CrossRef] [Google Scholar]
200. Carpenter BM, Whitmire JM, Merrell DS. This Is Not Your Mother’s Repressor: the Complex Role of Fur in Pathogenesis. Infection and Immunity. 2009. pp. 2590–2601. 10.1128/IAI.00116-09 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
201. Braun V. FhuA (TonA), the career of a protein. J Bacteriol. 2009;191: 3431–3436. 10.1128/JB.00106-09 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
202. Heinrichs DE, Yethon JA, Whitfield C. Molecular basis for structural diversity in the core regions of the lipopolysaccharides of Escherichia coli and Salmonella enterica. Mol Microbiol. 1998;30: 221–232. 10.1046/j.1365-2958.1998.01063.x [Abstract] [CrossRef] [Google Scholar]
203. Ornellas EP, Stocker BA. Relation of lipopolysaccharide character to P1 sensitivity in Salmonella typhimurium. Virology. 1974;60: 491–502. 10.1016/0042-6822(74)90343-2 [Abstract] [CrossRef] [Google Scholar]
204. Ho TD, Waldor MK. Enterohemorrhagic Escherichia coli O157:H7 gal mutants are sensitive to bacteriophage P1 and defective in intestinal colonization. Infect Immun. 2007;75: 1661–1666. 10.1128/IAI.01342-06 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
205. Raetz CRH, Whitfield C. Lipopolysaccharide Endotoxins. Annual Review of Biochemistry. 2002. pp. 635–700. 10.1146/annurev.biochem.71.110601.135414 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
206. Nikaido H. Molecular basis of bacterial outer membrane permeability revisited. Microbiol Mol Biol Rev. 2003;67: 593–656. 10.1128/mmbr.67.4.593-656.2003 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
207. Thibault D, Jensen PA, Wood S, Qabar C, Clark S, Shainheit MG, et al. Droplet Tn-Seq combines microfluidics with Tn-Seq for identifying complex single-cell phenotypes. Nat Commun. 2019;10: 5729 10.1038/s41467-019-13719-9 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
208. Hernandez CA, Koskella B. Phage resistance evolution in vitro is not reflective of in vivo outcome in a plant-bacteria-phage system. Evolution. 2019;73: 2461–2475. 10.1111/evo.13833 [Abstract] [CrossRef] [Google Scholar]
209. Qimron U, Kulczyk AW, Hamdan SM, Tabor S, Richardson CC. Inadequate inhibition of host RNA polymerase restricts T7 bacteriophage growth on hosts overexpressing udk. Mol Microbiol. 2008;67: 448–457. 10.1111/j.1365-2958.2007.06058.x [Abstract] [CrossRef] [Google Scholar]
210. Liu H, Price MN, Waters RJ, Ray J, Carlson HK, Lamson JS, et al. Magic Pools: Parallel Assessment of Transposon Delivery Vectors in Bacteria. mSystems. 2018;3 10.1128/mSystems.00143-17 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
211. Nora LC, Westmann CA, Guazzaroni M-E, Siddaiah C, Gupta VK, Silva-Rocha R. Recent advances in plasmid-based tools for establishing novel microbial chassis. Biotechnol Adv. 2019;37: 107433 10.1016/j.biotechadv.2019.107433 [Abstract] [CrossRef] [Google Scholar]
212. Oechslin F. Resistance Development to Bacteriophages Occurring during Bacteriophage Therapy. Viruses. 2018;10 10.3390/v10070351 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
213. Abedon ST. Bacteriophage Ecology: Population Growth, Evolution, and Impact of Bacterial Viruses. Cambridge University Press; 2008. [Google Scholar]
214. Díaz-Muñoz SL, Koskella B. Bacteria-phage interactions in natural environments. Adv Appl Microbiol. 2014;89: 135–183. 10.1016/B978-0-12-800259-9.00004-4 [Abstract] [CrossRef] [Google Scholar]
215. Oechslin F, Piccardi P, Mancini S, Gabard J, Moreillon P, Entenza JM, et al. Synergistic Interaction Between Phage Therapy and Antibiotics Clears Pseudomonas Aeruginosa Infection in Endocarditis and Reduces Virulence. J Infect Dis. 2017;215: 703–712. 10.1093/infdis/jiw632 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
216. Hung C-H, Kuo C-F, Wang C-H, Wu C-M, Tsao N. Experimental phage therapy in treating Klebsiella pneumoniae-mediated liver abscesses and bacteremia in mice. Antimicrob Agents Chemother. 2011;55: 1358–1365. 10.1128/AAC.01123-10 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
217. Taylor VL, Fitzpatrick AD, Islam Z, Maxwell KL. The Diverse Impacts of Phage Morons on Bacterial Fitness and Virulence. Adv Virus Res. 2019;103: 1–31. 10.1016/bs.aivir.2018.08.001 [Abstract] [CrossRef] [Google Scholar]
218. Koskella B, Brockhurst MA. Bacteria-phage coevolution as a driver of ecological and evolutionary processes in microbial communities. FEMS Microbiol Rev. 2014;38: 916–931. 10.1111/1574-6976.12072 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
219. Castillo D, Rørbo N, Jørgensen J, Lange J, Tan D, Kalatzis PG, et al. Phage defense mechanisms and their genomic and phenotypic implications in the fish pathogen Vibrio anguillarum. FEMS Microbiol Ecol. 2019;95 10.1093/femsec/fiz004 [Abstract] [CrossRef] [Google Scholar]
220. Wang X, Wei Z, Yang K, Wang J, Jousset A, Xu Y, et al. Phage combination therapies for bacterial wilt disease in tomato. Nat Biotechnol. 2019;37: 1513–1520. 10.1038/s41587-019-0328-3 [Abstract] [CrossRef] [Google Scholar]
221. McCallin S, Oechslin F. Bacterial Resistance to Phage and Its Impact on Clinical Therapy In: Phage Therapy: A Practical Approach. Springer; Nature Switzerland AG; 2019. pp. 59–88. [Google Scholar]
222. Pires DP, Cleto S, Sillankorva S, Azeredo J, Lu TK. Genetically Engineered Phages: a Review of Advances over the Last Decade. Microbiol Mol Biol Rev. 2016;80: 523–543. 10.1128/MMBR.00069-15 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
223. Kilcher S, Loessner MJ. Engineering Bacteriophages as Versatile Biologics. Trends Microbiol. 2019;27: 355–367. 10.1016/j.tim.2018.09.006 [Abstract] [CrossRef] [Google Scholar]
224. Ando H, Lemire S, Pires DP, Lu TK. Engineering Modular Viral Scaffolds for Targeted Bacterial Population Editing. Cell Syst. 2015;1: 187–196. 10.1016/j.cels.2015.08.013 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
225. Chan BK, Abedon ST, Loc-Carrillo C. Phage cocktails and the future of phage therapy. Future Microbiol. 2013;8: 769–783. 10.2217/fmb.13.47 [Abstract] [CrossRef] [Google Scholar]
226. Yehl K, Lemire S, Yang AC, Ando H, Mimee M, Torres MDT, et al. Engineering Phage Host-Range and Suppressing Bacterial Resistance through Phage Tail Fiber Mutagenesis. Cell. 2019;179: 459–469.e9. 10.1016/j.cell.2019.09.015 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
227. Yosef I, Goren MG, Globus R, Molshanski-Mor S, Qimron U. Extending the Host Range of Bacteriophage Particles for DNA Transduction. Mol Cell. 2017;66: 721–728.e3. 10.1016/j.molcel.2017.04.025 [Abstract] [CrossRef] [Google Scholar]
228. Yosef I, Manor M, Kiro R, Qimron U. Temperate and lytic bacteriophages programmed to sensitize and kill antibiotic-resistant bacteria. Proc Natl Acad Sci U S A. 2015;112: 7267–7272. 10.1073/pnas.1500107112 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
229. Wright RCT, Friman V-P, Smith MCM, Brockhurst MA. Resistance Evolution against Phage Combinations Depends on the Timing and Order of Exposure. MBio. 2019;10 10.1128/mBio.01652-19 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
230. Kilcher S, Studer P, Muessner C, Klumpp J, Loessner MJ. Cross-genus rebooting of custom-made, synthetic bacteriophage genomes in L-form bacteria. Proc Natl Acad Sci U S A. 2018;115: 567–572. 10.1073/pnas.1714658115 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
231. Kutter E, Sulakvelidze A. Bacteriophages: Biology and Applications. CRC Press; 2004. [Google Scholar]
232. Dunne M, Rupf B, Tala M, Qabrati X, Ernst P, Shen Y, et al. Reprogramming Bacteriophage Host Range through Structure-Guided Design of Chimeric Receptor Binding Proteins. Cell Rep. 2019;29: 1336–1350.e4. 10.1016/j.celrep.2019.09.062 [Abstract] [CrossRef] [Google Scholar]
233. Nobrega FL, Costa AR, Kluskens LD, Azeredo J. Revisiting phage therapy: new applications for old resources. Trends in Microbiology. 2015:185–191. 10.1016/j.tim.2015.01.006 [Abstract] [CrossRef] [Google Scholar]
234. Sturino JM, Klaenhammer TR. Engineered bacteriophage-defence systems in bioprocessing. Nat Rev Microbiol. 2006;4: 395–404. 10.1038/nrmicro1393 [Abstract] [CrossRef] [Google Scholar]
235. Mahony J, Ainsworth S, Stockdale S, van Sinderen D. Phages of lactic acid bacteria: The role of genetics in understanding phage-host interactions and their co-evolutionary processes. Virology. 2012:143–150. 10.1016/j.virol.2012.10.008 [Abstract] [CrossRef] [Google Scholar]
236. Ma NJ, Isaacs FJ. Genomic Recoding Broadly Obstructs the Propagation of Horizontally Transferred Genetic Elements. Cell Syst. 2016;3: 199–207. 10.1016/j.cels.2016.06.009 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
237. Lajoie MJ, Rovner AJ, Goodman DB, Aerni H-R, Haimovich AD, Kuznetsov G, et al. Genomically recoded organisms expand biological functions. Science. 2013;342: 357–360. 10.1126/science.1241459 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
238. Spears PA, Mitsu Suyemoto M, Hamrick TS, Wolf RL, Havell EA, Orndorff PE. In VitroProperties of a Listeria monocytogenes Bacteriophage-Resistant Mutant Predict Its Efficacy as a Live Oral Vaccine Strain. Infection and Immunity. 2011:5001–5009. 10.1128/IAI.05700-11 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
239. Capparelli R, Nocerino N, Lanzetta R, Silipo A, Amoresano A, Giangrande C, et al. Bacteriophage-resistant Staphylococcus aureus mutant confers broad immunity against staphylococcal infection in mice. PLoS ONE. 2010;5: e11720 10.1371/journal.pone.0011720 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
240. Capparelli R, Nocerino N, Iannaccone M, Ercolini D, Parlato M, Chiara M, et al. Bacteriophage therapy of Salmonella enterica: a fresh appraisal of bacteriophage therapy. J Infect Dis. 2010;201: 52–61. 10.1086/648478 [Abstract] [CrossRef] [Google Scholar]
241. Dąbrowska K, Abedon ST. Pharmacologically Aware Phage Therapy: Pharmacodynamic and Pharmacokinetic Obstacles to Phage Antibacterial Action in Animal and Human Bodies. Microbiol Mol Biol Rev. 2019;83 10.1128/MMBR.00012-19 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
242. Adler BA, Zhong C, Liu H, Kutter E, Deutschbauer AM, Mutalik VK, et al. Systematic Discovery of Salmonella Phage-Host Interactions via High-Throughput Genome-Wide Screens. bioRxiv [Preprint]. 2020. [cited 2020 Mar 10] 10.1101/2020.04.27.058388 [CrossRef] [Google Scholar]
243. Mangalea MR, Duerkop BA. Fitness Trade-Offs Resulting from Bacteriophage Resistance Potentiate Synergistic Antibacterial Strategies. Infection and Immunity. 2020. 10.1128/iai.00926-19 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
244. Torres-Barceló C, Hochberg ME. Evolutionary Rationale for Phages as Complements of Antibiotics. Trends Microbiol. 2016;24: 249–256. 10.1016/j.tim.2015.12.011 [Abstract] [CrossRef] [Google Scholar]
245. Tagliaferri TL, Jansen M, Horz H-P. Fighting Pathogenic Bacteria on Two Fronts: Phages and Antibiotics as Combined Strategy. Front Cell Infect Microbiol. 2019;9: 22 10.3389/fcimb.2019.00022 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
246. Abedon ST. Phage-Antibiotic Combination Treatments: Antagonistic Impacts of Antibiotics on the Pharmacodynamics of Phage Therapy? Antibiotics (Basel). 2019;8 10.3390/antibiotics8040182 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
247. Ausubel FM, Brent R, Kingston RE, Moore DD, Seidman JG, Smith JA, et al. Short Protocols in Molecular Biology: A Compendium of Methods from Current Protocols in Molecular Biology. John Wiley & Sons; 1995. [Google Scholar]
248. Grenier F, Matteau D, Baby V, Rodrigue S. Complete Genome Sequence of Escherichia coli BW25113. Genome Announc. 2014;2 10.1128/genomeA.01038-14 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
249. Ptashne M. A Genetic Switch: Phage Lambda Revisited. CSHL Press; 2004. [Google Scholar]
250. Ikeda H, Tomizawa J-I. Transducing fragments in generalized transduction by phage P1 I. Molecular origin of the fragments. Journal of Molecular Biology. 1966:226 10.1016/s0022-2836(66)80067-0 [Abstract] [CrossRef] [Google Scholar]
251. Price MN, Ray J, Iavarone AT, Carlson HK, Ryan EM, Malmstrom RR, et al. Oxidative Pathways of Deoxyribose and Deoxyribonate Catabolism. mSystems. 2019;4 10.1128/mSystems.00297-18 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
252. McCarthy DJ, Chen Y, Smyth GK. Differential expression analysis of multifactor RNA-Seq experiments with respect to biological variation. Nucleic Acids Res. 2012;40: 4288–4297. 10.1093/nar/gks042 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
253. Robinson MD, McCarthy DJ, Smyth GK. edgeR: a Bioconductor package for differential expression analysis of digital gene expression data. Bioinformatics. 2010;26: 139–140. 10.1093/bioinformatics/btp616 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
254. Datta S, Costantino N, Court DL. A set of recombineering plasmids for gram-negative bacteria. Gene. 2006;379: 109–115. 10.1016/j.gene.2006.04.018 [Abstract] [CrossRef] [Google Scholar]
255. Arkin AP, Cottingham RW, Henry CS, Harris NL, Stevens RL, Maslov S, et al. KBase: The United States Department of Energy Systems Biology Knowledgebase. Nat Biotechnol. 2018;36: 566–569. 10.1038/nbt.4163 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
256. Bolger AM, Lohse M, Usadel B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics. 2014;30: 2114–2120. 10.1093/bioinformatics/btu170 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
257. Kim D, Paggi JM, Park C, Bennett C, Salzberg SL. Graph-based genome alignment and genotyping with HISAT2 and HISAT-genotype. Nat Biotechnol. 2019;37: 907–915. 10.1038/s41587-019-0201-4 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
258. Pertea M, Pertea GM, Antonescu CM, Chang T-C, Mendell JT, Salzberg SL. StringTie enables improved reconstruction of a transcriptome from RNA-seq reads. Nat Biotechnol. 2015;33: 290–295. 10.1038/nbt.3122 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
259. Love MI, Huber W, Anders S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 2014;15: 550 10.1186/s13059-014-0550-8 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
260. Bak G, Lee J, Suk S, Kim D, Young Lee J, Kim K-S, et al. Identification of novel sRNAs involved in biofilm formation, motility, and fimbriae formation in Escherichia coli. Sci Rep. 2015;5: 15287 10.1038/srep15287 [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
2020 Oct; 18(10): e3000877.
Published online 2020 Oct 13. 10.1371/journal.pbio.3000877.r001

Decision Letter 0

Roland G Roberts, Senior Editor

16 Apr 2020

Dear Dr Mutalik,

Thank you for submitting your manuscript entitled "High-throughput mapping of the phage resistance landscape in E. coli" for consideration as a Research Article by PLOS Biology.

Your manuscript has now been evaluated by the PLOS Biology editorial staff, as well as by an academic editor with relevant expertise, and I'm writing to let you know that we would like to send your submission out for external peer review.

However, before we can send your manuscript to reviewers, we need you to complete your submission by providing the metadata that is required for full assessment. To this end, please login to Editorial Manager where you will find the paper in the 'Submissions Needing Revisions' folder on your homepage. Please click 'Revise Submission' from the Action Links and complete all additional questions in the submission questionnaire.

Please re-submit your manuscript within two working days, i.e. by Apr 20 2020 11:59PM.

Login to Editorial Manager here: https://www.editorialmanager.com/pbiology

During resubmission, you will be invited to opt-in to posting your pre-review manuscript as a bioRxiv preprint. Visit http://journals.plos.org/plosbiology/s/preprints for full details. If you consent to posting your current manuscript as a preprint, please upload a single Preprint PDF when you re-submit.

Once your full submission is complete, your paper will undergo a series of checks in preparation for peer review. Once your manuscript has passed all checks it will be sent out for review.

Given the disruptions resulting from the ongoing COVID-19 pandemic, please expect delays in the editorial process. We apologise in advance for any inconvenience caused and will do our best to minimize impact as far as possible.

Feel free to email us at gro.solp@ygoloibsolp if you have any queries relating to your submission.

Kind regards,

Roli Roberts

Roland G Roberts, PhD,

Senior Editor

PLOS Biology

    2020 Oct; 18(10): e3000877.
    Published online 2020 Oct 13. 10.1371/journal.pbio.3000877.r002

    Decision Letter 1

    Roland G Roberts, Senior Editor

    24 May 2020

    Dear Dr Mutalik,

    Thank you very much for submitting your manuscript "High-throughput mapping of the phage resistance landscape in E. coli" for consideration as a Research Article at PLOS Biology. Your manuscript has been evaluated by the PLOS Biology editors, an Academic Editor with relevant expertise, and by three independent reviewers.

    You'll see that while the reviewers are broadly positive about your study, they each raise a few concerns that should be addressed. The Academic Editor asked me to emphasise that reviewer #1 asks for a few important clarifications (including why affected RM mechanisms don't show more clear resistance phenotypes), and reviewers #2 and #3 request several additional essential pieces of information (e.g. quality of the GoF/LoF libraries, MOI and host/phage densities in the assays); we also anticipate that the requests by reviewer #3 to reduce the text and improve readability will involve significant work (his recommendations seem sensible, especially given our wide readership).

    In light of the reviews (below), we are pleased to offer you the opportunity to address the comments from the reviewers in a revised version that we anticipate should not take you very long. We will then assess your revised manuscript and your response to the reviewers' comments and we may consult the reviewers again.

    We expect to receive your revised manuscript within 1 month.

    Please email us (gro.solp@ygoloibsolp) if you have any questions or concerns, or would like to request an extension. At this stage, your manuscript remains formally under active consideration at our journal; please notify us by email if you do not intend to submit a revision so that we may end consideration of the manuscript at PLOS Biology.

    **IMPORTANT - SUBMITTING YOUR REVISION**

    Your revisions should address the specific points made by each reviewer. Please submit the following files along with your revised manuscript:

    1. A 'Response to Reviewers' file - this should detail your responses to the editorial requests, present a point-by-point response to all of the reviewers' comments, and indicate the changes made to the manuscript.

    *NOTE: In your point by point response to the reviewers, please provide the full context of each review. Do not selectively quote paragraphs or sentences to reply to. The entire set of reviewer comments should be present in full and each specific point should be responded to individually.

    You should also cite any additional relevant literature that has been published since the original submission and mention any additional citations in your response.

    2. In addition to a clean copy of the manuscript, please also upload a 'track-changes' version of your manuscript that specifies the edits made. This should be uploaded as a "Related" file type.

    *Resubmission Checklist*

    When you are ready to resubmit your revised manuscript, please refer to this resubmission checklist: https://plos.io/Biology_Checklist

    To submit a revised version of your manuscript, please go to https://www.editorialmanager.com/pbiology/ and log in as an Author. Click the link labelled 'Submissions Needing Revision' where you will find your submission record.

    Please make sure to read the following important policies and guidelines while preparing your revision:

    *Published Peer Review*

    Please note while forming your response, if your article is accepted, you may have the opportunity to make the peer review history publicly available. The record will include editor decision letters (with reviews) and your responses to reviewer comments. If eligible, we will contact you to opt in or out. Please see here for more details:

    https://blogs.plos.org/plos/2019/05/plos-journals-now-open-for-published-peer-review/

    *PLOS Data Policy*

    Please note that as a condition of publication PLOS' data policy (http://journals.plos.org/plosbiology/s/data-availability) requires that you make available all data used to draw the conclusions arrived at in your manuscript. If you have not already done so, you must include any data used in your manuscript either in appropriate repositories, within the body of the manuscript, or as supporting information (N.B. this includes any numerical values that were used to generate graphs, histograms etc.). For an example see here: http://www.plosbiology.org/article/info%3Adoi%2F10.1371%2Fjournal.pbio.1001908#s5

    *Blot and Gel Data Policy*

    We require the original, uncropped and minimally adjusted images supporting all blot and gel results reported in an article's figures or Supporting Information files. We will require these files before a manuscript can be accepted so please prepare them now, if you have not already uploaded them. Please carefully read our guidelines for how to prepare and upload this data: https://journals.plos.org/plosbiology/s/figures#loc-blot-and-gel-reporting-requirements

    *Protocols deposition*

    To enhance the reproducibility of your results, we recommend that if applicable you deposit your laboratory protocols in protocols.io, where a protocol can be assigned its own identifier (DOI) such that it can be cited independently in the future. For instructions see: https://journals.plos.org/plosbiology/s/submission-guidelines#loc-materials-and-methods

    Thank you again for your submission to our journal. We hope that our editorial process has been constructive thus far, and we welcome your feedback at any time. Please don't hesitate to contact us if you have any questions or comments.

    Sincerely,

    Roli Roberts

    Roland G Roberts, PhD,

    Senior Editor

    PLOS Biology

    *****************************************************

    REVIEWERS' COMMENTS:

    Reviewer #1:

    Overview: The resistance strategies used by E. coli to prevent bacteriophage infection have previously been characterized for several phage:bacteria pairs by using labor-intensive methods that are difficult to scale. To develop an unbiased and comprehensive screen compatible with many bacterial and phage strains, Mutalik and colleagues utilized three previously published E. coli libraries, two for loss of function (RbTn-Seq and CRISPRi) and one for gain of function (DubSeq). With these libraries in two strains of E. coli and 14 diverse phages across three families, the authors were able to validate many existing phage receptors and uncover unexpected pathways involved in phage infection, including colonic acid synthesis and cyclic-di-GMP. Impressively, many hits were validated through plaque assays, and RNA-sequencing was performed to further determine mechanism for previously undescribed hits.

    All in all, this paper is an incredible amount of work, that I believe will be of interest to a broad readership as it opens up a lot of avenues for future research and adaptation to other phage pairs. Before I can fully recommend for publication, I only have a few concerns to address:

    Major points:

    * The mention of resistance mechanisms throughout the introduction primes the reader to think about classic bacteria defense mechanisms such as Restriction/Modification, Methylation, CRISPR etc. Besides CRISPR, which is not expressed in these strains, these are all likely categories of genes that would appear as hits in the DubSeq, overexpression experiments. Why do the authors think that restriction or methylation enzymes are not strong hits? Along these lines, the conclusion section would benefit from a further discussion into the conditions under which a bacterium would actively overexpress the colanic acid synthesis pathway as a phage-defense mechanism.

    * Most of the phage strains in the study are lytic, as suggested by the choice of using lambda cI857 and p1vir. However, P2 and 186 are known to be lysogenic phages. Whether lytic phages were targeted in this study, and how the lysogenic nature of P2 and 186 could have affected the results is not addressed. Could certain bacterial mutants bias lysogeny of those phages leading to "resistance" as indicated by growth of those lysogens? Also, why is phage P2 excluded from the K-12 DubSeq study?

    Minor Points:

    * Line 156 - "As described later, for comparative purposes, we also performed high-throughput genetic assays in the E. coli BL21 strain background…. Different serotypes of E. coli are also important pathogens with significant global threat and are crucial players in specific human-relevant ecologies [84-86], leading to the question of whether strain variation is also important in predicting the response to phage-mediated selection or whether the mechanisms are likely to be conserved."

    o This study does not address pathogenic strains of E. coli and gives an unclear motivation for the use of strain BL21. The organization of this paragraph guides the reader think that non-pathogenic and pathogenic strains will be compared, whereas this study characterizes two similar non-pathogenic strains. These strains may be expected to behave more similarly than a comparison with a pathogenic strain. Therefore, this paragraph would benefit from a stronger introduction to why strain BL21 is used.

    o The use of two novel shiga-toxin phages in this study is very intriguing and can link to the pathogen point made above. The conclusion could benefit from a discussion of how the genes responsible for preventing infection of these two pathogen associated phages differ from non-pathogenic phages of similar structure.

    * Line 178 - Figure 1.

    o The organization of Figure 1 appears like Myoviridae is used with RB-TnSeq, Podoviridae is used with CRISPRi and Siphoviridae is used with GOF. This figure could benefit from either moving the bacterial strains to "Technology" box and then having one arrow from technology to "Phage Strains" or removing the Bacterial Strain information.

    * Figure 2 - Where are the MOIs described for each phage? The methods section does not describe the MOIs used.

    * Figure 3 - In the CRISPRi screen, igaA is such a strong hit for all phages besides 186. In RbTn-Seq, however, igaA is a strong hit for 186. Why do the authors think that this result was not repeatable with CRISPRi? Does phage 186 contain a depolymerase to degrade the capsule?

    * Line 495 - "This is the first systematic analysis of how gene overexpression impacts phage resistance"

    o This is overstated given that Qimron 2006 use the entire ASKA collection to probe T7 resistance.

    * Lines 1010 and 1035 - Should this should be -80°C?

    * Supplementary Figure S5 - There are no 0.1mM IPTG pictures for T5 and T6 phages

    Reviewers #2:

    [identify themselves as Sankar Adhya and Shayla Hesse]

    Mutalik et al apply the array of specific and tractable genome-wide perturbations present in RB-TnSeq, Dub-Seq, and CRISPRi E. coli libraries to broadly probe the phage resistance landscape for multiple phage-bacteria pairs. Their work highlights the key advantages of such an approach: overcoming the limitations of complete LOF screens and venturing into the largely uncharted territory of GOF and intergenic screens - all while maintaining a relatively high degree of technical efficiency. Their work reveals novel insights on an age-old subject (coliphage-host interaction and evolution) and is well-presented in both written word and graphics. The degree to which non-phage-receptor- related resistance genes varied for the 14 different phages was particularly interesting, as well as the central role of Rcs signalling (igaA downregulation and rcsA upregulation) in modulating bacterial sensitivity to phage infection more generally.

    I recommend publication of the manuscript after authors satisfactorily respond to some comments listed below.

    Comments:

    A RB-TnSeq E.coli library seems fairly straightforward, but I'm sure that CRISPRi and Dub-Seq library generation are trickier propositions. One cannot easily assess the overall quality of sgRNAs (In terms of target affinity and avoiding off-target effects). Can you control gene dosage in Dub-Seq library generation? If not, what is the range of variability for different plasmids?

    Specific Comments

    Strain-specific differences between K-12 and BL21: Side-by-side comparison of Fig. 2a and 7 reveal some notable differences in RB-TnSeq results, while there were many more that were observed with the Dub-Seq data - do you think most of them are artifacts? What do you think are the most likely explanations?

    Dub-Seq: What is the average increase in gene copy number on a per cell basis?

    Figure 5: consider moving to Supplemental, not sure how much visualization of those raw data enhances the reader's understanding of the key points of the paper

    Page 22, Lines 850-855: The authors dutifully point out the potential for phage-resistant strains with significantly lower fitness to be outcompeted to the point of being obscured. It may also be worth pointing out that the relative fitness of these lab-generated strains is distinct from the likelihood that their cognate mutant will naturally evolve under phage selection pressure. (A more obvious limitation, yes, but also more consequential. And particularly relevant given the authors' claim in the abstract that an upshot to the application of their methods may be the generation of "datasets that allow predictive models of how phage-mediated selection will shape bacterial phenotype and evolution.")

    Page 30, Line 183-184: The igaA mutant map weblink returned an error message.

    Supplementary table S12: recommend addition of phage genome accession numbers/links for the non-canonical phages

    Reviewer #3:

    [identifies himself as Jeremy J Barr]

    The manuscript by Mutalik et. al., presents a tour de force in mapping the phage resistance landscape across two E. coli strains. Using a combination of LOF (Tn-Seq & CRISPRi) and GOF (DubSeq) techniques, they screened for host factors associated with phage resistance across 14 E. coli phages, many of which are well established model phages, along with novel and less-well characterised phages, demonstrating the broad applicability of the technique. Using these techniques, they confirm established phage-receptor mutants, while also uncovering novel resistance mechanisms, some of which were validated using conventional knock-out and complementation.

    The amount of work invested into this manuscript is both impressive and staggering, and this could have easily been split across multiple manuscripts. Yet the amount of data and analysis presented did make parts of the manuscript very detailed and difficult to read, with parts that could be reduced to improve readability. Overall, I have no major concerns with the experimental approach and analysis conducted and believe this is a very important paper to the field. Most of my comments address readability and accessibility of the manuscript.

    I had two major comments with the paper that I felt should be addressed.

    1) MOIs and both host and phage concentrations for these most experiments are not reported in text or methods. The authors simply state 'varying MOIs' and provide a qualitative scale for this (with exception of CRISPRi where MOI 1 was used). It is important that both the MOIs and concentrations used in each assay are report to confirm their validity. Please report all MOIs along with actual concentrations of phage and host used.

    2) I felt that the authors ignored the limitations of their approach, particularly the time and costs associated with establishing these libraries coupled with the limited phage that may target these library strains. Suggestion is to add an extra section to the discussion on this, and speculating how these limitations may be overcome in future developments of the technology.

    3) At points the manuscript was overly descriptive and long. The authors should consider reducing the amount of text in sections to improve readability and focus of each section, some specific examples are below.

    See below for line by line comments and suggestions:

    Lines 154-159 & 164-166 - Sections describing the libraries and hosts were repetitive, please combine points/sections

    Lines 171-172 - Please highlight in text the novel and STEC targeting phages

    Line 188 - term 'are' please change to past tense

    Lines 194-200 - the descriptions of positive fitness scores were confusingly written and long, I understand the rationale, but recommend simplifying this section

    Lines 228-229 - See comment above, please report both your MOIs and phage/host concentrations used.

    Line 236 - should read 'more than one'

    Lines 243-244 - This section reads like either the LPS or proteinaceous receptors are required for phage infection, this may be the case for your 14 phages studied, but not as a general rule - please rewrite to highlight the large diversity of potential phage receptors

    Figure 2 - Recommend adding darker border lines between phages as it is difficult to interpret inner parts of the graph. Also recommend adding something to the figure panel C to highlight that phages affected are colored purple.

    Lines 363 - briefly re-iterate what a high fitness score implies in this assay

    Lines 446 - Is not clear where this genomic DNA comes from, assuming this is taken from the same K-12 host, but this should be briefly explained here.

    Lines 453 - "67 genome wide GOF competition" - Can you clarify what this means? Does one genome-wide GOF assay encompass a single library containing the whole K12 genome of sheared 3kb DNA?

    Lines 457-458 - Again re-iterate what positive-growth implies in this experiment - multiple other points in manuscript this could be explained (lines 517,

    Lines 485-488 - Further explanation needed here, does the increased expression of PDEs lead to degradation of cyclic-di-GMP and thereby inhibition of biofilm formation? How does this confer broad resistance to phages?

    Lines 550-561 - reduce complexity here

    Lines 566-567 ¬- regarding Rcs pathway, useful to also state what this pathway activates (e.g., colonic acid)

    Lines 601-637 - In general this section was difficult to read and a lot of detail could be cut to focus on the main outcomes

    Section starting 638 - This section seems entirely focused on N4 phage infectivity - which is needed, but the heading doesnt convey this. please reword

    Figure 6 - Panel B: could briefly state the mechanism of igaA knock down and how it leads to colanic acid and EPS activation. Panel D: Similar comment, briefly explain mechanism of up and down regulation

    Lines 775-776 - include brief description on which host library was made from - assuming BL21

    Figure 7 - Why not highlight yellow boxes for confirmed phage receptors again?

      2020 Oct; 18(10): e3000877.
      Published online 2020 Oct 13. 10.1371/journal.pbio.3000877.r003

      Author response to Decision Letter 1

      21 Jul 2020

      Attachment

      Submitted filename:

        2020 Oct; 18(10): e3000877.
        Published online 2020 Oct 13. 10.1371/journal.pbio.3000877.r004

        Decision Letter 2

        Roland G Roberts, Senior Editor

        6 Aug 2020

        Dear Dr Mutalik,

        Thank you for submitting your revised Research Article entitled "High-throughput mapping of the phage resistance landscape in E. coli" for publication in PLOS Biology. I have now obtained advice from the original reviewers and have discussed their comments with the Academic Editor.

        We're delighted to let you know that we're now editorially satisfied with your manuscript. However before we can formally accept your paper and consider it "in press", we also need to ensure that your article conforms to our guidelines. A member of our team will be in touch shortly with a set of requests. As we can't proceed until these requirements are met, your swift response will help prevent delays to publication. Please also make sure to address the Data Policy-related requests noted at the end of this email.

        *Copyediting*

        Upon acceptance of your article, your final files will be copyedited and typeset into the final PDF. While you will have an opportunity to review these files as proofs, PLOS will only permit corrections to spelling or significant scientific errors. Therefore, please take this final revision time to assess and make any remaining major changes to your manuscript.

        NOTE: If Supporting Information files are included with your article, note that these are not copyedited and will be published as they are submitted. Please ensure that these files are legible and of high quality (at least 300 dpi) in an easily accessible file format. For this reason, please be aware that any references listed in an SI file will not be indexed. For more information, see our Supporting Information guidelines:

        https://journals.plos.org/plosbiology/s/supporting-information

        *Published Peer Review History*

        Please note that you may have the opportunity to make the peer review history publicly available. The record will include editor decision letters (with reviews) and your responses to reviewer comments. If eligible, we will contact you to opt in or out. Please see here for more details:

        https://blogs.plos.org/plos/2019/05/plos-journals-now-open-for-published-peer-review/

        *Early Version*

        Please note that an uncorrected proof of your manuscript will be published online ahead of the final version, unless you opted out when submitting your manuscript. If, for any reason, you do not want an earlier version of your manuscript published online, uncheck the box. Should you, your institution's press office or the journal office choose to press release your paper, you will automatically be opted out of early publication. We ask that you notify us as soon as possible if you or your institution is planning to press release the article.

        *Protocols deposition*

        To enhance the reproducibility of your results, we recommend that if applicable you deposit your laboratory protocols in protocols.io, where a protocol can be assigned its own identifier (DOI) such that it can be cited independently in the future. For instructions see: https://journals.plos.org/plosbiology/s/submission-guidelines#loc-materials-and-methods

        *Submitting Your Revision*

        To submit your revision, please go to https://www.editorialmanager.com/pbiology/ and log in as an Author. Click the link labelled 'Submissions Needing Revision' to find your submission record. Your revised submission must include a cover letter, a Response to Reviewers file that provides a detailed response to the reviewers' comments (if applicable), and a track-changes file indicating any changes that you have made to the manuscript.

        Please do not hesitate to contact me should you have any questions.

        Sincerely,

        Roli Roberts

        Roland G Roberts, PhD,

        Senior Editor,

        gro.solp@streborr,

        PLOS Biology

        ------------------------------------------------------------------------

        DATA POLICY:

        You may be aware of the PLOS Data Policy, which requires that all data be made available without restriction: http://journals.plos.org/plosbiology/s/data-availability. For more information, please also see this editorial: http://dx.doi.org/10.1371/journal.pbio.1001797

        Many thanks for depositing your raw data in the SRA and Figshare. However, we also ask that all individual quantitative observations that underlie the data summarized in the figures and results of your paper be made available in one of the following forms:

        1) Supplementary files (e.g., excel). Please ensure that all data files are uploaded as 'Supporting Information' and are invariably referred to (in the manuscript, figure legends, and the Description field when uploading your files) using the following format verbatim: S1 Data, S2 Data, etc. Multiple panels of a single or even several figures can be included as multiple sheets in one excel file that is saved using exactly the following convention: S1_Data.xlsx (using an underscore).

        2) Deposition in a publicly available repository. Please also provide the accession code or a reviewer link so that we may view your data before publication.

        Regardless of the method selected, please ensure that you provide the individual numerical values that underlie the summary data displayed in the following figure panels as they are essential for readers to assess your analysis and to reproduce it: Figs 2A, 3ABC, 4A, 6ACE, S8. NOTE: the numerical data provided should include all replicates AND the way in which the plotted mean and errors were derived (it should not present only the mean/average values).

        Please also ensure that figure legends in your manuscript include information on where the underlying data can be found, and ensure your supplemental data file/s has a legend.

        Please ensure that your Data Statement in the submission system accurately describes where your data can be found.

        ------------------------------------------------------------------------

        BLOT AND GEL REPORTING REQUIREMENTS:

        For manuscripts submitted on or after 1st July 2019, we require the original, uncropped and minimally adjusted images supporting all blot and gel results reported in an article's figures or Supporting Information files. We will require these files before a manuscript can be accepted so please prepare and upload them now. Please carefully read our guidelines for how to prepare and upload this data: https://journals.plos.org/plosbiology/s/figures#loc-blot-and-gel-reporting-requirements

        ------------------------------------------------------------------------

        REVIEWERS' COMMENTS:

        Reviewer #1:

        The authors have adequately addressed each of my major and minor comments. I have no additional comments.

        Reviewer #2:

        [identifies herself as Shayla Hesse]

        Thank you for your clear, well-reasoned, and very thorough responses to each point raised by the reviewers. In our opinion, the revisions made to the manuscript adequately address the issues raised and significantly strengthen the paper.

        Reviewer #3:

        [identifies himself as Jeremy Barr]

        I would like to thank the authors for their detailed reviewer response and additional work on the revised manuscript. I think this new manuscript is much improved, including some excellent additions, particularly the critical review of the costings and benefits of the GOF & LOF approaches. The authors have addressed all my concerns and I have no further comments,

          2020 Oct; 18(10): e3000877.
          Published online 2020 Oct 13. 10.1371/journal.pbio.3000877.r005

          Author response to Decision Letter 2

          30 Aug 2020

          Attachment

          Submitted filename:

            2020 Oct; 18(10): e3000877.
            Published online 2020 Oct 13. 10.1371/journal.pbio.3000877.r006

            Decision Letter 3

            Roland G Roberts, Senior Editor

            8 Sep 2020

            Dear Dr Mutalik,

            On behalf of my colleagues and the Academic Editor, J. Arjan G. M. de Visser, I am pleased to inform you that we will be delighted to publish your Research Article in PLOS Biology.

            The files will now enter our production system. You will receive a copyedited version of the manuscript, along with your figures for a final review. You will be given two business days to review and approve the copyedit. Then, within a week, you will receive a PDF proof of your typeset article. You will have two days to review the PDF and make any final corrections. If there is a chance that you'll be unavailable during the copy editing/proof review period, please provide us with contact details of one of the other authors whom you nominate to handle these stages on your behalf. This will ensure that any requested corrections reach the production department in time for publication.

            Early Version

            The version of your manuscript submitted at the copyedit stage will be posted online ahead of the final proof version, unless you have already opted out of the process. The date of the early version will be your article's publication date. The final article will be published to the same URL, and all versions of the paper will be accessible to readers.

            PRESS

            We frequently collaborate with press offices. If your institution or institutions have a press office, please notify them about your upcoming paper at this point, to enable them to help maximise its impact. If the press office is planning to promote your findings, we would be grateful if they could coordinate with gro.solp@sserpygoloib. If you have not yet opted out of the early version process, we ask that you notify us immediately of any press plans so that we may do so on your behalf.

            We also ask that you take this opportunity to read our Embargo Policy regarding the discussion, promotion and media coverage of work that is yet to be published by PLOS. As your manuscript is not yet published, it is bound by the conditions of our Embargo Policy. Please be aware that this policy is in place both to ensure that any press coverage of your article is fully substantiated and to provide a direct link between such coverage and the published work. For full details of our Embargo Policy, please visit http://www.plos.org/about/media-inquiries/embargo-policy/.

            Thank you again for submitting your manuscript to PLOS Biology and for your support of Open Access publishing. Please do not hesitate to contact me if I can provide any assistance during the production process.

            Kind regards,

            Alice Musson

            Publishing Editor,

            PLOS Biology

            on behalf of

            Roland Roberts,

            Senior Editor

            PLOS Biology


              Articles from PLOS Biology are provided here courtesy of PLOS

              Citations & impact 


              Impact metrics

              Jump to Citations

              Citations of article over time

              Alternative metrics

              Altmetric item for https://www.altmetric.com/details/92317829
              Altmetric
              Discover the attention surrounding your research
              https://www.altmetric.com/details/92317829

              Smart citations by scite.ai
              Smart citations by scite.ai include citation statements extracted from the full text of the citing article. The number of the statements may be higher than the number of citations provided by EuropePMC if one paper cites another multiple times or lower if scite has not yet processed some of the citing articles.
              Explore citation contexts and check if this article has been supported or disputed.
              https://scite.ai/reports/10.1371/journal.pbio.3000877

              Supporting
              Mentioning
              Contrasting
              17
              98
              0

              Article citations


              Go to all (52) article citations

              Data 


              Data behind the article

              This data has been text mined from the article, or deposited into data resources.

              Similar Articles 


              To arrive at the top five similar articles we use a word-weighted algorithm to compare words from the Title and Abstract of each citation.


              Funding 


              Funders who supported this work.

              Innovative Genomics Institute

                U.S. Department of Energy (1)