Skip to main content

This is a preprint.

It has not yet been peer reviewed by a journal.

The National Library of Medicine is running a pilot to include preprints that result from research funded by NIH in PMC and PubMed.

bioRxiv logoLink to bioRxiv
[Preprint]. 2025 Jun 9:2025.06.04.657922. [Version 2] doi: 10.1101/2025.06.04.657922

Analysis of Essential Genes in Clostridioides difficile by CRISPRi and Tn-seq

Maia E Alberts 1,3, Micaila P Kurtz 1, Ute Müh 1, Jonathon P Bernardi 1, Kevin W Bollinger 1, Horia A Dobrila 1,4, Leonard Duncan, Hannah M Laster 1, Andres J Orea 1,5, Anthony G Pannullo 1,6, Juan G Rivera-Rosado 1, Facundo V Torres 1,7, Craig D Ellermeier 1,2,*, David S Weiss 1,2,*
PMCID: PMC12157513  PMID: 40502013

Abstract

Essential genes are interesting in their own right and as potential antibiotic targets. To date, only one report has identified essential genes on a genome-wide scale in Clostridioides difficile, a problematic pathogen for which treatment options are limited. That foundational study used large-scale transposon mutagenesis to identify 404 protein-encoding genes as likely to be essential for vegetative growth of the epidemic strain R20291. Here, we revisit the essential genes of strain R20291 using a combination of CRISPR interference (CRISPRi) and transposon-sequencing (Tn-seq). First, we targeted 181 of the 404 putatively essential genes with CRISPRi. We confirmed essentiality for >90% of the targeted genes and observed morphological defects for >80% of them. Second, we conducted a new Tn-seq analysis, which identified 346 genes as essential, of which 283 are in common with the previous report and might be considered a provisional essential gene set that minimizes false positives. We compare the list of essential genes to those of other bacteria, especially Bacillus subtilis, highlighting some noteworthy differences. Finally, we used fusions to red fluorescent protein (RFP) to identify 18 putative new cell division proteins, three of which are conserved in Bacillota but of largely unknown function. Collectively, our findings provide new tools and insights that advance our understanding of C. difficile.

Keywords: CRISPRi, Tn-seq

INTRODUCTION

Clostridioides difficile infections (CDI) kill close to 13,000 people a year in the United States (1). Treating CDI is challenging because the antibiotics effective against C. difficile also impact the normal intestinal microbiota needed to keep C. difficile in check (24). There is a need for improved antibiotics that inhibit C. difficile more selectively. Most clinically useful antibiotics target proteins or pathways that are essential for viability, so a deeper understanding of the essential genes in C. difficile might provide foundational knowledge to guide antibiotic development. Essential genes are also interesting in their own right, as they provide insights into the most fundamental aspects of bacterial physiology.

Transposon sequencing (Tn-seq) identifies essential genes on a genome-wide scale based on the absence of insertions following saturation transposon mutagenesis (5, 6). However, several caveats must be kept in mind when interpreting the output of a Tn-seq experiment. For instance, insertion mutants that are viable but grow slowly will be lost from the mutant pool during outgrowth, so some apparently essential genes can be deleted. This caveat underscores the fact that binary categorization of genes as essential or nonessential is useful but an oversimplification. Tn-seq might also erroneously classify non-essential genes as essential due to polarity onto bona fide essential genes or because the random nature of Tn insertions means genes might be missed for stochastic reasons. Finally, Tn-seq does not provide insight into the actual function of essential genes because the phenotypic defects of the corresponding insertion mutants are not observed. Despite these caveats and limitations, Tn-seq is a powerful tool for prioritizing genes to investigate by more laborious methods.

CRISPR interference (CRISPRi) is a complementary approach for genome-wide interrogation of essential genes in bacteria (712). CRISPRi uses a single guide RNA (sgRNA) to direct a catalytically inactive Cas9 protein (dCas9) to a gene of interest, thereby repressing transcription (13). As the organism continues to grow and divide it becomes depleted of the targeted protein, potentially revealing phenotypic changes that precede cell death. Thus, CRISPRi provides functional information that Tn-seq cannot. However, CRISPRi shares with Tn-seq the problem of polarity, which has to be taken into consideration when interpreting phenotypes.

In 2015 Dembek et al. used Tn-seq to identify 404 protein-encoding genes as essential for vegetative growth in C. difficile strain R20291 on BHI media (14). As expected, most of these genes encode proteins involved in core biological processes and cell surface biogenesis, but some are of unknown function or not expected to be essential. Here, we revisit the essential genes of strain R20291 using a combination of CRISPRi and Tn-seq. First, we targeted 181 of the 404 putatively essential genes with CRISPRi to vet essentiality and identify terminal phenotypes. We confirmed essentiality for >90% of the targeted genes and observed morphological defects for >80% of them. Second, we conducted a new and more thorough Tn-seq analysis to identify genes essential for vegetative growth on TY media. We classified 346 protein-coding genes as essential, of which 283 (~80%) were also essential in the previous study. Finally, we conducted a microscopy-based screen to identify potential cell division proteins. We discuss our findings in light of what is known about essential genes and cell division in other bacteria, particularly Bacillus subtilis.

RESULTS AND DISCUSSION

A library for CRISPRi knockdown of 181 putative essential genes

Our C. difficile CRISPRi plasmid has been described (15). It expresses dCas9 from a xylose-inducible promoter (Pxyl) and a sgRNA from a constitutively-active glutamate dehydrogenase promoter (Pgdh). Constructing a knockdown library involved several steps: selecting the genes to be targeted, designing the sgRNAs, cloning those sgRNAs into the CRISPRi plasmid, and moving the finished plasmids from E. coli into C. difficile by conjugation. Because conjugation efficiencies are low, plasmids have to be moved from E. coli into C. difficile one-by-one. This step imposes a bottleneck that made it impractical to target all 404 essential genes identified previously. We therefore trimmed the gene list by excluding all transposon and phage-related genes (because these are not part of the core genome), most genes for tRNA synthetases and ribosomal proteins (to limit redundancy), and most genes for small proteins, defined here as fewer than 80 amino acids (240 nucleotides). Short genes are small targets for Tn insertion, so a disproportionate fraction are likely to be false positives. At this point we were left with 252 genes. Because CRISPRi is polar (7, 13, 16, 17), there is little to be gained by targeting multiple genes in an operon, so in most cases we targeted only one gene per transcription unit as annotated in BioCyc (v28.5, release Dec 2025) (18).

In the end, we selected a total of 181 putatively essential genes for CRISPRi knockdown (Table S1). We constructed a library of individual sgRNA clones, using two sgRNAs per gene for a total of 362 CRISPRi plasmids (Table S2). As negative controls, we constructed 20 CRISPRi plasmids with scrambled sgRNAs that do not target anywhere in the R20291 genome (Table S2). Plasmids were confirmed by sequencing across the Pgdh::sgRNA element in E.coli and after conjugation into C. difficile. Of the genes targeted for knockdown, 86 have an essential ortholog in Bacillus subtilis, 62 have a non-essential ortholog in B. subtilis, and 33 have no B. subtilis ortholog, including four hypothetical genes. However, the number of genes of unknown function is larger than four because many of the non-hypotheticals have homology to domains with such broadly or ill-defined functions that it is not obvious what these genes do or why they would be essential (e.g., “glycosyltransferase,” “two-component response regulator,” or “DUF1846”). Considering that 110 of the targeted genes are predicted to be in operons with other apparently essential genes, our study encompasses 281 of the 404 genes identified as essential by Tn mutagenesis, close to 70% of the total (14).

Essentiality determined by CRISPRi knock-down largely agrees with Tn-seq data

The entire CRISPRi library was screened for viability defects by conducting spot titer assays on TY plates containing thiamphenicol at 10 μg/ml (hereafter TY-Thi10) and 1% xylose. Control plates lacked xylose. Using a 10-fold viability defect or small colony phenotype with at least one sgRNA as the cut-off, 167 of the 181 genes (92%) were confirmed as essential by CRISPRi, while 14 were not essential (Fig. 1A; Table S1). Similar results were obtained with both sgRNAs for 174 of the 181 genes tested (Table S1). None of the 20 non-targeting control sgRNAs caused a growth defect, indicating off-target effects are rare. We conclude that the vast majority of the genes Dembek et al. identified as essential by Tn-seq are also essential by CRISPRi (14).

Fig. 1. Summary of gene essentiality determined by CRISPRi and Tn-seq.

Fig. 1.

(A) CRISPRi-induced viability defects were determined from spot titer assays on TY-thiamphenicol with 1% xylose. Viability defects were scored as strong (≥ 1000-fold), moderate (≥100-fold), weak (≥10-fold or full viability but colonies were small), or none (full viability, normal colony size). In cases where the two sgRNAs produced different results, the stronger viability defect was used. (B) Comparison of Tn-seq datasets. Of the 346 genes determined to be essential in our study, 283 were essential in the Dembek et al. set, and 10 were called ambiguous. Conversely, of the 404 Dembek et al. essential genes, 283 were essential in our dataset, 74 were non-essential, 12 were not found owing to use of different genome annotations, and 35 were ambiguous (Here, ambiguous combines three categories from Supplemental Table 3: unclear (11 genes), short (6 genes), unclear/non-essential (18 genes)). (C) Viability defects in CRISPRi correlate with likelihood a gene will be scored as essential by Tn-seq. Viability defects are from Table S1. Tn-seq calls come from Table S3.

Terminal phenotypes due to CRISPRi knockdown of genes of known function

To look for morphological abnormalities that might facilitate provisional assignment of essential genes to functional pathways, cells were scraped from the last culture dilution that grew on the 1% xylose plates and examined by phase-contrast microscopy. As the project progressed, we added staining with FM4–64 to visualize the cytoplasmic membrane and Hoechst 33342 to visualize DNA. The morphological defects associated with CRISPRi silencing of all 181 genes are listed in Table S1.

CRISPRi knockdown of genes of known function often provoked expected morphological defects, such as filamentation in the case of cell division genes and aberrant nucleoid staining in the case of DNA replication genes (Fig. 2, Fig. S1, Table 1, Table S1). Also as expected, knockdown of DNA replication genes sometimes resulted in filamentation, presumably due to induction of the SOS response (19, 20). However, we also observed morphological defects that were not expected and are difficult to rationalize. For instance, knockdown of rpoB (β subunit of RNA polymerase) or era (GTPase involved in ribosome assembly) caused severe filamentation, while knockdown of guaA (synthesis of guanosine ribonucleotides) caused a mild chaining phenotype. To address whether the unexpected morphological abnormalities are an artifact of working with cells scraped from plates, we reexamined the filamentation phenotype of four non-division genes in broth about six doublings after inducing CRISPRi: dnaH, rpoB, prfB and tilS. We observed elongated cells in each case (Fig. S2). Thus, at least for this phenotype and these four genes, morphologies determined using plates are reliable.

Fig. 2. Morphology of CRISPRi strains with sgRNAs targeting genes in select functional pathways.

Fig. 2.

Left: pathway. Middle: Predicted transcription unit. Targeted genes are boxed and indicated above the operon diagrams. Numbers are R20291 locus tags. Genes are color coded to indicate essentiality based on Tn-seq calls in Table S3. Blue: essential. Light blue: ambiguous. White: non-essential. Operon structure is not to scale. Right: Morphological changes based on phase contrast and fluorescence micrographs of cells scraped from viability plates. Membranes were stained with FM4–64 and DNA was stained with Hoechst 33342. Size bars are 5 μm. The control strain expressed an sgRNA that does not target anywhere in the genome. Micrographs are representative of at least two experiments. Figure S1 shows microscopy of more genes.

Table 1.

CRISPRi phenotypes of functional pathways

Pathway Genesa Phenotypeb
Cell division ftsZ, maf, minC, minD, whiA, zapA Filaments, few septa, lysis, misshapen cells (swollen or bent), phase-bright cells, chaining, mostly normal chromosome morphology.
DNA replication dnaD, dnaE, dnaF, dnaG, dnaL, dnaX, holA, priA, ssb Filamentous, few septa, condensed chromosomes and regions devoid of DNA.
Fatty acid and phospholipid biosynthesis accB, acpS, cdsA, fabF, fabG2, fabK, fabZ, fapR, pgsA, plsC, plsY, yqhY Highly variable, including: mostly normal morphology, short cells, elongated cells, misshapen cells (swollen or bent), phase-bright cells.
Nucleotides guaA, guaB, nrdD, pyrH, thyA, tmk Mostly normal morphology, chaining, filamentous cells with areas of condensed chromosomes or devoid of DNA.
Peptidoglycan biosynthesis rodA, pbp1, pbp2, murJ2 pbp1: filamentous cells with few septa; pbp2 or rodA: short, swollen cells, often phase-bright, chaining. murJ2: misshapen cells (swollen or bent).
Peptidoglycan precursor biosynthesis ddl, glmU, mraY, murB, murE, murF, murG, murI Short cells, phase-bright cells, misshapen cells (swollen or bent), mild elongation, lysis.
Protein synthesis argS, cpgA, efp, era, fusA, infC, obg, prfA, prfB, rbgA, rimM, rlmL, rnrY, rplC, rplT, rpsM, serS1, smpB, thrS, tilS, tsf, tufA, tufB Normal morphology, elongation, misshapen cells (swollen or bent), chaining, condensed chromosomes.
Teichoic acid biosynthesis cdr_2657, cdr_2663, cdr_2665, gtaB, murJ1, pgm2, rkpK, tuaA, tuaG Chains of short, swollen cells, sometimes phase-bright, sometimes mild elongation.
Transcription rpoA, rpoB rpoA: misshapen cells, a few modestly elongated; rpoB: long aseptate filaments, chromosome morphology normal.
a

Depletion phenotypes for genes in bold are shown in Figure 2 and Figure S1.

b

Phenotypes reported encompass the range observed across the genes listed. The phenotypic defects often differed for different genes from the same functional pathway. Major phenotypes caused by repression of each gene are listed in Table S1.

Because morphological defects were only loosely associated with the function of well-studied genes, we conclude that CRISPRi is not sufficient for assigning genes of unknown function to physiological pathways. We are not the first to report unanticipated complexity among terminal phenotypes in a CRISPRi screen. For example, CRISPRi knockdown of the RNA polymerase gene rpoC and the phospholipid synthesis genes psd and plsB caused filamentation in E. coli (8). In addition, knockdown of multiple genes with no direct role in envelope biogenesis caused morphological defects in B. subtilis (7). These reports contrast with the narrower spectrum of morphological defects induced by antibiotics that target specific pathways (2123). Antibiotics might be less subject to secondary effects because cells are visualized at early times after exposure and polarity is not an issue.

Terminal phenotypes due to CRISPRi knockdown of genes of unknown function

Our CRISPRi library targeted 11 genes that could not be assigned to a functional category and were confirmed as essential in our own Tn-seq analysis as will be described below. CRISPRi caused a viability defect in nine cases, often accompanied by abnormal morphologies (Table 2). Examples include cdr20291_0481 and cdr20291_0828 (elongation), the cdr20291_1053–1057 cluster (short, swollen, phase-bright cells and chaining), cdr20291_1124 (chaining and many misshapen phase-bright cells) and cdr20291_2526 (a few misshapen cells). The phenotype resulting from knockdown of cdr20291_1124 could be due to reverse polarity onto the upstream gene alaS, which encodes an alanyl-tRNA synthetase. These genes warrant further investigation.

Table 2.

Essential genes not assigned to a physiological pathwaya

Locus tag Annotation Size CRISPRi viability defect CRISPRi terminal morphology
CDR20291_0351 Phosphoesterase 230 a.a.b Weak Normal
CDR20291_0481 Sugar isomerase/endonuclease 251 a.a. Weak Elongated
CDR20291_0828 DUF1846 domain 501 a.a. Strong Elongated
CDR20291_1053 Pyrophosphokinase 373 a.a. Strong Chaining, short cells, swollen cells, phase-bright
CDR20291_1054 Putative exported protein 291 a.a. Strong See CDR20291_1053
CDR20291_1055 Family 2 glycosyl transferase 230 a.a. Strong See CDR20291_1053
CDR20291_1056 Glycosyl transferase family protein 274 a.a. Strong See CDR20291_1053
CDR20291_1057 DUF3866 domain 355 a.a. Not targeted
CDR20291_1124 Putative membrane protein 723 a.a. Moderate Chaining, curved cells, phase-bright
CDR20291_1171 UvrD/REP type DNA helicase 593 a.a. None Normal
CDR20291_1418B None 113 a.a. Not targeted Not done
CDR20291_2521 PDZ, Radical SAM and DUF512 domains 466 a.a. Not targeted Not done
CDR20291_2526 Two-component response regulator 230 a.a. Moderate Mostly normal, a few curved
CDR20291_2569 Putative calcium-chelating exported protein 308 a.a. None Normal
CDR20291_3525 Conserved hypothetical protein 61 a.a. Not targeted Not done
a

These proteins were classified as essential in our Tn-seq and either essential or ambiguous by Dembek et al., 2015. CDR20219_3519 and CDR20219_3520 are omitted because their essentiality is likely due to polarity onto dnaC and/or rplI.

b

a.a., amino acids

Rationale for decision to conduct Tn-seq

As noted above, our CRISPRi screen was based on the only published genome-wide analysis of gene essentiality in C. difficile. That study made essentiality calls based on a single pool of mutants containing ~77,000 unique Tn insertions plated on BHI media (14). We reasoned that an independently-derived list of essential genes based on multiple biological replicates and larger insertion libraries would serve as a useful resource to the C. difficile community. We also thought a new Tn-seq dataset might serve as a “tie-breaker” for the 14 putatively essential genes that did not appear to be essential by CRISPRi, i.e., failure to recover insertions in those genes would suggest our sgRNAs were ineffective, while recovery of insertions would suggest the genes are non-essential and were missed in the previous study for stochastic reasons.

Generation of Tn insertion libraries and identification of essential genes

We used the same R20291 strain and mariner-based transposon as in the previous study (14). Mariner is a good choice for C. difficile because it inserts at TA dinucleotides and the genome G+C content is 29% (24, 25). However, our experimental design differed from Dembek et al. in three noteworthy respects: (i) we used TY media, (ii) we constructed three independent insertion libraries, (iii) and we determined insertion profiles at both an early and a late timepoint because gradual loss of slow-growing mutants from the pools influences perceptions of gene essentiality. Our early timepoint consisted of primary insertion libraries recovered directly from selection plates after ~18 hours of incubation. For a later timepoint, libraries were sub-cultured in duplicate into TY and harvested after seven generations of outgrowth.

Insertion profiles were analyzed using TRANSIT2 and the C. difficile R20291 reference genome NC_013316.1 (26, 27). Depending on the experimental replicate, insertions were identified in 117,217 to 204,061 of the 502,945 unique TA dinucleotides in the R20291 genome (Table 3). A total of 289,505 TA sites sustained at least one Tn insertion across the three libraries. TRANSIT2 makes essentiality calls by comparing the observed frequency of Tn insertions to the availability of potential TA insertion sites. Genes are classified as essential (E or EB, depending on the model for statistical analysis), not essential (NE), or unclear (U) (28). Genes with too few TA sites for statistical analysis are designated S. After inspecting the output from TRANSIT2, we manually reclassified eleven NE or U genes as essential, giving them the designation Ei for “essential by inspection.” Ten of these genes had a large number of TA sites but very few insertions. An example is the tRNA-synthetase valS (CDR20291_3114), with insertions in only four of the possible 266 TA dinucleotides after outgrowth (Table S3A). For comparison, TRANSIT2 scored the cell division gene ftsZ as essential even though there were insertions in three out of 110 TA sites. All ten genes that we moved to Ei based on few insertions are considered essential in C. difficile and B. subtilis (14, 29). The final Ei gene, murJ2 (CDR20291_3335), had a large number of insertions but almost all of these were at the 3’ end of the gene (Fig. 3A). murJ2 was previously classified as essential in C. difficile by Dembek et al. but its ortholog is not essential in B. subtilis due to functional redundancy (29, 30).

Table 3.

Number of unique transposon insertions from experimental replicatesa

Replicate Unique TA sites hit Fraction of total TA sites
Library A 178,325 0.355
Library B 168,519 0.335
Library C 143,213 0.285
Combined Libraries A-C 289,505 0.576
Outgrowth A1 135,217 0.269
Outgrowth A2 117,217 0.233
Outgrowth B1 127,947 0.254
Outgrowth B2 135,056 0.269
Outgrowth C1 167,894 0.334
Outgrowth C2 204,061 0.406
a

Total TA sites = 502,945

Fig. 3. Essentiality follow-up.

Fig. 3.

(A) Transposon insertion profile for murJ2. Vertical lines represent mapped insertion sites and are scaled to indicate the number of sequence reads mapping to that site. Although murJ2 sustained numerous insertions, ~80% were in the last 10% of the gene, suggesting the non-essentiality call by TRANSIT2 is incorrect. (B) Transposon insertion profile for polA indicating that only the N-terminal domain is essential. Read frequency was scaled to 5 to highlight the absence of reads in the N-terminal domain. The average number of reads per polA site with at least one read was 173. (C) Spot titer assays of CRISPRi strains targeting genes in the atp operon. Serial dilutions of overnight cultures were spotted on TY-Thi10 plates with 1% xylose. Plates were imaged after incubation at 37˚C for ~18h. Silencing atpB and atpD resulted in small colonies, while growth after silencing atpI and atpF was comparable to the negative control. (D) Pre-depletion of ATP synthase proteins impairs growth. Starter cultures were grown overnight in TY-Thi10 without (left) or with (right) 1% xylose, then subcultured into TY-Thi10 with 1% xylose and growth was followed by measuring optical density at 600 nm. To prolong growth, cultures in the left panel were back-diluted at 7h. (E) Zone of inhibition assays reveal CRISPRi knockdown of the atp operon increases sensitivity to CCCP. Plates were imaged after incubation at 37˚C for ~18h. Novobiocin (Novo) served as a control. Guides in panels C-E were: atpI (5531), atpB (5583), atpF (5581), atpD (5579) or a negative control that does not target anywhere in the gemone.

Of the 3673 annotated protein-coding genes in R20291, 346 were scored as essential for vegetative growth in the initial libraries and/or after outgrowth (Table S3A). We grouped these genes into functional categories similar to those used in previous studies of B. subtilis and S. aureus (Table 4; Table S3B) (31, 32). As expected, over half are involved in DNA metabolism (25 genes), RNA metabolism (24 genes), protein synthesis (113 genes) or cell envelope biogenesis (76 genes). Also as expected, the majority of C. difficile’s essential genes are conserved; BioCyc assigned a B. subtilis ortholog for 272 of the 346 genes, of which 169 are essential (Table S3A, B) (29).

Table 4.

Tn-seq essential genes by category

Category Count
Cell envelope 76
 Cell division 7
 Cell shape 4
 Diaminopimelate biosynthesis 7
 Fatty acid biosynthesis 7
 Isoprenoid biosynthesis 7
 Peptidoglycan biosynthesis 5
 Peptidoglycan precursor biosynthesis 13
 Phospholipid biosynthesis 6
 Regulation 3
 S-layer 2
 Teichoic acid biosynthesis 14
 Other 1
Cofactors 24
 CoA 6
 Fe-S cluster 1
 Folate 7
 Heme 1
 NAD 4
 Riboflavin 1
 SAM 1
 Thiamine 3
DNA metabolism 25
 DNA packaging and segregation 4
 DNA recombination and repair 4
 DNA replication 16
 Other 1
 Category Count
Metabolism 15
 Amino acid biosynthesis 3
 Glycolysis 2
 Pentose phosphate pathway 1
 Phosphate metabolism 1
 Other 8
Nucleotides 11
 dNTP biosynthesis 2
 Purine biosynthesis 3
 Pyrimidine biosynthesis 4
 Regulatory nucleotides 2
Other/unknown 58
 Phage-related 4
 Sporulation 2
 Transporter 6
 Transposon-related 17
 Other 1
 Unknown 28
Protein synthesis 113
 Protein degradation 4
 Protein folding 2
 Protein modification 2
 Protein translocation 7
 Ribosomal proteins 52
 Ribosome biogenesis 12
 Translation factors 10
 tRNA synthetases 24
RNA metabolism 24
 Basic transcription machinery 5
 Regulation of RNA synthesis 4
 RNA processing and degradation 6
 tRNA modification 9
Total 346

Comparison of our Tn-seq data to Dembek et al. and to CRISPRi

There is good overall agreement between our Tn-seq essentiality calls and those made previously. Of the 346 genes identified as essential in our experiments, 283 (82%) were also essential for Dembek et al. (Fig. 1B; Table S3A). As expected based on the larger size of our insertion libraries, we scored fewer genes as essential, 346 versus 404 (Fig. 1B). Our list of essential genes includes 53 considered nonessential by Dembek et al., while those investigators identified 121 essential genes that did not make our cutoffs. Of these, 12 encode proteins that are not annotated in the genome sequence we used and were thus invisible to our analysis. An additional 6 were scored as S and 29 as U or U/NE. That leaves 74 bona fide discrepancies, genes that were essential for Dembek et al. but not essential for us. Several of these differences will be discussed below, but the most likely explanations have to do with statistical cut-offs that factor into essentiality calls and the stochastic nature of Tn mutagenesis.

There is also good overall agreement between our CRISPRi and Tn-seq data sets. Of the 141 genes for which CRISPRi elicited a strong or moderate viability defect, 129 (~90%) scored as essential in our Tn-seq (Fig. 1C; Table S3A). Conversely, only 4 out of 14 genes (~30%) that appeared to be nonessential by CRISPRi nevertheless scored as essential in our Tn-seq. These four genes are an uncharacterized DNA helicase (CDR20291_1171), a sporulation-associated phosphatase (ptpB), an acetyl-CoA thiolase (thlA2), and a putative exported Ca2+-chelating protein (ykwD). None of these has an essential ortholog in B. subtilis. Two labs have constructed null mutants of ptpB, indicating it is not essential (33, 34). One study reported a growth defect (34), which might explain why ptpB appears to be essential by Tn-seq.

DNA metabolism

Some DNA replication proteins have different names in B. subtilis and E. coli. Where there are conflicts, we adopted the names used in B. subtilis, which in some cases differ from the names used in BioCyc. We identified 16 widely conserved DNA replication genes as essential in C. difficile. All but pcrA and polA were previously classified as essential in C. difficile, and all but polA is essential in B. subtilis (14, 29). Interestingly, polA is domain essential in C. difficile—Tn insertions were recovered in the C-terminal 3’ to 5’ exonuclease and DNA polymerase domains but not in the N-terminal 5’ to 3’ exonuclease domain, which removes Okazaki fragments (Fig. 3B). Similar restricted essentiality of the polA 5’ to 3’ exonuclease domain has been reported in Streptococcus and Haemophilus (35, 36). Organisms like B. subtilis in which the entire polA gene is dispensable have an RNAse H that can remove Okazaki fragments (37).

Interestingly, C. difficile lacks dnaB (38). DnaB is an essential protein in B. subtilis, where it works together with DnaD and DnaI to load the replicative helicase DnaC onto oriC DNA (39). DnaB and DnaD are structurally related. It has been proposed that in C. difficile the DnaD ortholog (CDR20291_3512) fulfills the functions of both DnaB and DnaD (38).

C. difficile has four essential DNA packaging and segregation genes, all of which are also essential in B. subtilis. In addition, there are three essential DNA recombination and repair genes, none of which are essential in B. subtilis.

LexA, which represses genes involved in the SOS response, is required for viability in C. difficile but not in B. subtilis. A C. difficile lexA Clostron insertion mutant has been described and grows poorly, so its apparent essentiality by Tn-seq may be due to slow growth rather than lack of viability per se (19). However, the strong viability defect we observed upon CRISPRi knockdown of lexA (Table S1) raises the possibility that the reported mutant retains partial function or acquired a suppressor.

RNA metabolism

As expected, the core subunits and major sigma factor (σ70) of RNA polymerase are all essential. Surprisingly, the omega subunit (rpoZ) is also essential according to Tn-seq, even though it is not essential in B. subtilis, S. aureus or E. coli (29, 40, 41). The apparent essentiality of rpoZ is likely to be an artifact of polarity because it is predicted to be co-transcribed with three widely conserved essential genes: dapF, gmk, coaBC. The elongation factor greA and three termination/anti-termination factors (nusA, nusG and rho) are essential. Of these, only nusA is essential in B. subtilis (29, 42). In C. difficile rho mutations have been reported, including an early frameshift, but the gene could not be deleted, possibly because the mutant is too sick (43).

Fifteen genes for enzymes that modify RNA were essential in our analysis, of which twelve were essential or ambiguous for Dembek et al., but only eight are considered essential in B. subtilis. Most of these genes encode proteins needed to generate mature tRNAs or rRNAs from precursor transcripts.

Protein synthesis

There are 55 annotated ribosomal proteins in BioCyc (Dec 18, 2024), 52 of which scored as essential by our Tn-seq. Most of these were also identified as essential by Dembek et al. and are essential in B. subtilis. Instances of ribosomal protein genes scored as essential in C. difficile but as non-essential in B. subtilis could reflect polarity. Five widely conserved small GTPases involved in ribosome assembly are essential, as are ten translation factors, including smpB, which encodes a component of the SsrA tagging complex that rescues stalled ribosomes by trans-translation (44). We confirmed essentiality of smpB by CRISPRi (Table S1). The essentiality of smpB is unlikely to be an artifact of polarity because it is not predicted to be co-transcribed with any other genes. SmpB is essential in S. aureus (32, 45) but not in E. coli, Streptococcus sanguinis, or B. subtilis (29, 46, 47). An interesting omission from the list of essential translation factors is elongation factor Tu (EF-Tu), which is essential in B. subtilis (29). This difference can be explained by the presence of two EF-Tu genes in C. difficile, tufA and tufB, which are 100% identical at the DNA level. Simultaneous knockdown of tufA and tufB with CRISPRi caused a strong viability defect, demonstrating EF-Tu is indeed required for viability (Table S1).

We identified 24 essential tRNA synthetases, all of which are also essential according to Dembek et al. There are several noteworthy differences in comparison to B. subtilis. First, synthetases for asparagine (asnS), threonine (thrS) and tyrosine (tyrS) are essential in C. difficile but not B. subtilis, which has alternative routes for generating the corresponding charged tRNAs (4850). Second, although glnS is essential in C. difficile, this gene does not exist in B. subtilis or most other gram-positive bacteria, which generate Gln-tRNAGln by a different route. Namely, C. difficile charges tRNAGln directly with glutamine, as in E. coli, while most Gram-positive bacteria generate glutaminyl-tRNAGln by (mis)charging tRNAGln with glutamate, which is then amidated to glutamine (51, 52). Lastly, C. difficile has two annotated genes for ligating proline to tRNApro, the essential gene proS1 (CDR20291_0038) and the non-essential gene proS2 (CDR20291_0039). According to RNA-sequencing, both are expressed during vegetative growth (53). B. subtilis has only a single proS gene, which is essential and more similar to C. difficile proS1 than proS2.

Five proteases appear to be important for viability in C. difficile: clpX, htrA, lon, prp and the M16 family protease cdr20291_1161. Of these, only prp is essential in B. subtilis. Prp is a cysteine protease needed to remove an N-terminal extension from ribosomal protein L27 (54). The apparent essentiality of lon and cdr20291_1161 in C. difficile are likely to be artifacts of polarity onto engB and dapG, respectively. ClpX is a component of the ClpXP protease complex, one of the major housekeeping proteases in bacteria (55). C. difficile has only one clpX gene but two genes for ClpP, which might explain why clpX is essential but clpP1 and clpP2 are not. HtrA proteases are involved in protein quality control (56). TRANSIT2 scored htrA as essential despite a high number of Tn insertions (67 out of 127 TA sites) and this gene was not essential for Dembek et al.

In bacteria, protein synthesis begins with N-formyl methionine (fMet). Peptide deformylase (def) and methionine aminopeptidase (map) are essential enzymes that work sequentially to remove the formyl group from about 90% of proteins and the initiating methionine from about half of proteins. E. coli has only one def and one map gene, both of which are essential (57). C. difficile has two predicted map genes and two predicted def genes. Of these, only map1 is essential by Tn-seq. This situation is reminiscent of B. subtilis, which also has two def and two map genes. The def genes are functionally redundant and at least one must be present for viability (58, 59). The essentiality of the map genes in B. subtilis is less clear. One study found mapA is essential but mapB is not (60), while another found neither is individually essential (29).

Bacteria have a plethora of systems for exporting proteins out of the cytoplasm, of which the three most important are the General Secretion (Sec) system, the Twin Arginine Translocation (Tat) system, and the Signal Recognition Particle (SRP) system (61). There is no Tat system in C. difficile, but the genes for the Sec and SRP systems are present and essential. The Sec system uses an ATPase named SecA to power export of proteins through a membrane channel composed of SecEYG. Interestingly, C. difficile has two secA paralogs, which handle different protein substrates and are both essential (62). The SRP system works together with SecEYG to integrate proteins into the cytoplasmic membrane. Three genes associated with the SRP system (ffh, ftsY and srpM) were scored as essential, although the apparent essentiality of srpM might result from polarity onto ffh; srpM is not essential in B. subtilis.

Cell envelope

Numerous genes involved in membrane biogenesis are essential in C. difficile. An unexpected exception is the accBCDA gene cluster for synthesis of malonyl-CoA, the substrate for fatty acid synthesis. This result is difficult to explain and probably incorrect because the acc cluster is essential according to Dembek et al. and we confirmed essentiality by CRISPRi (Table S1). Moreover, acc genes are also essential in B. subtilis (29). Nevertheless, the acc cluster sustained numerous Tn insertions in our study (e.g., ten of the 48 TA sites in accB, the first gene in the operon). We identified three membrane biogenesis genes that are essential in C. difficile but not in B. subtilis: fabH, yqhY, and gpsA. The fabH discrepancy can be explained by the presence of two fabH genes in B. subtilis (63). B. subtilis ΔyqhY mutants are not stable (64), implying yqhY is quasi essential in that organism. Regarding gpsA, although Koo et al. reported it is dispensable in B. subtilis (29), an earlier study found it is essential (65), which agrees with what we see in C. difficile.

C. difficile synthesizes isoprenoids via the methylerythritol (MEP) pathway (66). Accordingly, dxr and ispDEFGH were all essential by Tn-seq. Isoprenoids are essential in bacterial because they are precursors for quinones and carrier lipids such as undecaprenyl phosphate (Und-P) required for synthesis of peptidoglycan and teichoic acids (67). C. difficile lacks quinones (68) so the essentiality of the MEP pathway presumably reflects the importance of Und-P. Consistent with this inference, the predicted undecaprenyl pyrophosphate synthase UppS1 is essential, although that conclusion comes with a caveat because insertions in uppS1 are probably polar onto the essential phospholipid biosynthesis gene cdsA (69). Interestingly, C. difficile has a non-essential uppS paralog called uppS2 that might be involved in synthesis of the wall teichoic acid PS-II (70). UppS2 is not essential by Tn-seq, and RNA-sequencing implies expression of uppS2 is ~60-fold lower in vegetative cells compared to uppS1 (53).

The C. difficile cell has a unique proteinaceous surface-layer (S-layer) and a unique wall teichoic acid, PS-II, whose structure is very different from the wall teichoic acids of other Gram-positive bacteria (71). Both the S-layer and PS-II are essential by Tn-seq, although the existence of (unhealthy) null mutants of slpA indicates the S-layer is not strictly required for viability (72, 73). Multiple studies point to essentiality of PS-II (70, 74, 75). Whether PS-II is essential because it plays a critical role in cell envelope integrity or because disruption of the PS-II gene cluster depletes the pool of Und-P needed for peptidoglycan synthesis remains to be determined (76, 77).

The universal precursor for peptidoglycan synthesis is lipid II, a disaccharide-pentapeptide attached to Und-P (78). As expected, many lipid II genes are essential, including six dap genes for biosynthesis of lysine and diaminopimelic acid, and nine mur genes for various steps in lipid II assembly. Lipid II is transported across the cytoplasmic membrane by flippases, of which there are two known families, MurJ and Amj (30, 79). BLAST searches indicate C. difficile lacks Amj but has two MurJ orthologs, both of which are essential. MurJ1 is part of the PS-II gene cluster and proposed to transport a lipid-linked precursor for PS-II synthesis (74), which leaves MurJ2 as the likely lipid II flippase for peptidoglycan synthesis. Some non-essential proteins distantly related to MurJ can be identified using HHPred and could also potentially transport lipid II (80, 81). Further work is needed to establish the functions of the two clear MurJ paralogs and rule out the presence of alternative or additional lipid II transporters (30, 82, 83).

The final steps of peptidoglycan synthesis involve incorporation of new disaccharide-pentapeptide subunits into the existing wall by sequential glycosyltransferase (GTase) and transpeptidase (TPase) reactions (84, 85). These reactions are catalyzed by two types of penicillin-binding proteins (PBPs) (86). Class A PBPs (aPBPs) are bifunctional enzymes with both a GTase domain and a TPase domain, while class B PBPs (bPBPs) have a TPase domain and form a complex with a SEDS-family GTase (8789). C. difficile encodes one aPBP (PBP1), three bPBPs (PBP2, PBP3, and SpoVD), and two SEDS proteins (RodA and SpoVE). Of these proteins, we confirmed by Tn-seq that PBP1, PBP2 and RodA are essential for vegetative growth (14). Although spoVE was also classified as essential, it sustained Tn insertions in about half the available TA sites (Table S3A) and the gene has been deleted previously (90). In confirmation and extension of previous reports (15, 91), CRISPRi knockdown of PBP1 caused filamentation, while CRISPRi knockdown of PBP2 and RodA resulted in formation of short, swollen, phase-bright cells, with some chaining (Fig. S1). These morphologies implicate PBP1 in cell division and PBP2 in elongation, respectively. We also examined red fluorescent protein (RFP) fusions to the PBPs and observed that both localize to division sites (Fig. 4). Septal localization of PBP1 has been reported by Shen’s group, who showed it is the primary synthase for septal peptidoglycan (91). Septal localization of PBP2 suggests the RodA/PBP2 complex might also contribute to cell division, as further suggested by the mild chaining phenotypes caused by CRISPRi knockdown. Both RFP-PBP1 and RFP-PBP2 exhibited some fluorescence along the cell cylinder, which could indicate they contribute to elongation, especially in the case of PBP2. However, localization to the cell cylinder is not diagnostic of a function in elongation because this is the default location of divisome proteins when they are not at the septum. Finally, it should be noted that non-canonical 3–3 crosslinks made by L,D-transpeptidases (LDTs) are essential for vegetative growth in C. difficile, but none of the five LDTs in the C. difficile genome is individually essential owing to functional redundancy (92).

Fig. 4.

Fig. 4.

Representative fluorescence micrographs of fixed cells that produced the indicated proteins fused to RFP. Percentages indicate the fraction of cells scored positive for septal localization (n ≥ 202 cells). Space bar = 10 μm.

Our Tn-seq identified two cell envelope-related regulatory loci as essential: walRK and ddlR. These regulators were also essential for Dembek et al. walRK is a two-component system known to be essential for cell wall homeostasis and viability in numerous Bacillota, including C. difficile (53, 93). DdlR is essential for peptidoglycan synthesis because it activates expression of the D-alanyl-D-alanine ligase ddl (94).

Cell shape and division

In rod-shaped bacteria, the essential peptidoglycan synthases work in the context of loosely defined complexes known as the elongasome and the divisome (84, 85). The C. difficile elongasome appears to comprise the RodA/PBP2 bipartite peptidoglycan synthase and four Mre proteins (MreB1, MreB2, MreC and MreD). All of these are essential by Tn-seq and CRISPRi, although this inference will need to be revisited with non-polar deletions. CRISPRi knockdown implicates these genes primarily in elongation, because the predominant terminal morphologies include short, swollen cells with some chaining (Fig. S1, Table S1).

Among canonical divisome proteins, only ftsZ and its assembly factors sepF and zapA are essential in C. difficile. Neither sepF nor zapA is essential in B. subtilis (9597). The greater importance of sepF and zapA in C. difficile might be due to the absence of an ftsA ortholog (98). As noted above, the primary septal peptidoglycan synthase is the class A enzyme PBP1 (91). Consistent with that inference, CRISPRi against pbp1 induces filamentation, however additional morphological defects such as bending and chaining suggest PBP1 might contribute to elongation as well (15). Curiously, the division site placement genes minCDE are essential in C. difficile. This result might be an artifact of polarity onto the essential SEDS gene rodA, because the Min system is not essential in B. subtilis or most other bacteria (99). Tn-seq identified maf as essential. Maf is a nucleotide pyrophosphatase whose overproduction causes filamentation in both B. subtilis and E. coli, but Maf is not essential in either organism (100102). The DNA-binding protein WhiA was essential for Dembek et al. and we observed a weak viability defect and modest cell elongation by CRISPRi, but whiA is not essential in our Tn-seq experiments. WhiA is conserved in monoderms and essential in Mycobacterium tuberculosis but not Streptomyces or B. subtilis, where it has been linked to cell division and chromosome segregation (103107).

Use of RFP fusions to identify new divisome proteins

We have a long-standing interest in bacterial cell division, so we extended our studies to include a screen for divisome proteins (108112). Using CRISPRi knockdown to identify divisome proteins by screening for a filamentous phenotype comes with two major caveats—polarity onto a bona fide division gene will generate false positives and depletion of non-essential divisome proteins might not cause cells to become longer than normal. A more direct approach is to use fluorescent tags to screen for proteins that localize to the division site. Here the major caveat is that the tag might interfere with proper localization. We used BLAST searches to identify homologs of known morphogenesis proteins, which were fused to a codon-optimized red fluorescent protein (RFP) and produced from a plasmid under control of the xylose-inducible promoter, Pxyl (15). Some of these proteins are encoded in (predicted) operons with proteins of unknown function, so we constructed RFP fusions to several of these as well. Although septal localization is strong evidence for a role in cell division, lack of septal localization is uninformative because we did not test whether our RFP fusions are functional. We screened a total of 25 proteins, of which 18 localized and are discussed below (Fig. 4). The seven that did not localize are MreB1, MreB2, FtsL, FtsB, SpoVE, CDR_3330, and CDR_2504.

Seven enzymes for peptidoglycan synthesis exhibited convincing midcell localization, including the two essential PBPs (PBP1 and PBP2), one essential SEDS protein (RodA), one non-essential monofunctional glycosyltransferase related to PBPs (Mgt), and three non-essential LDTs (Ldt1, Ldt4 and Ldt5). Of these, PBP1 was already known to localize to sites of cell division (91), but septal localization of the remaining enzymes is new and suggests they too contribute to synthesis of septal peptidoglycan. Somewhat surprisingly, the canonical elongasome proteins MreC and MreD localized strongly to the midcell, even though our fusions to MreB1 and MreB2 did not. Mre proteins have been reported to localize transiently at or near the midcell in a few other bacteria (113116). Further work is warranted to investigate the role of the Mre proteins in C. difficile and the possibility that MreC and MreD localize independently of MreB, for which there is precedent from non-rod-shaped bacteria that have MreC and MreD but lack MreB (116, 117).

C. difficile orthologs of five widely-conserved divisome proteins localized to the midcell: FtsZ, FtsK, FtsQ, SepF, DivIVA, as did CDR_3331, a unique protein with limited structural similarity to both FtsL and FtsB, which in C. difficile are used for asymmetric division during sporulation (14, 91). Septal localization of C. difficile FtsZ has been reported previously (118). Septal localization of FtsQ is new but probably misleading because C. difficile ftsQ is a sporulation gene and not expressed during vegetative growth (14, 53, 90), whereas we produced RFP-FtsQ from Pxyl. Immediately downstream of ftsQ are two genes of unknown function, ylxW and ylxX, that according to RNA-sequencing are expressed in vegetative cells (53). YlxW and YlxX are encoded downstream of ftsQ in many Bacillota and have been proposed on this basis to play a role in envelope biogenesis (119). Our observation that these proteins localize to the midcell argues they are involved in cell division. Another novel divisome protein identified in our screen is YlmG, a small membrane protein encoded in the sepF operon of many Gram-positive bacteria and Cyanobacteria (98). Mutants of ylmG have been constructed in several organisms and exhibit thin septa, poor sporulation, and/or aberrant nucleoid compaction and segregation, depending on the species (98). In closing, and for completeness, we note that four additional proteins have been shown previously to localize to the division site in C. difficile: ZapA, MldA, MldB and MldC (112, 120). This brings total number of documented divisome proteins to 22.

Metabolism

For an insightful overview of energy metabolism in C. difficile, readers are referred to a review by Neumann-Schaal et al. (121). Briefly, C. difficile is an obligate anaerobe that generates energy through fermentation of sugars and amino acids, the latter by a process known as Stickland reactions (122, 123). There is no electron transport chain. Hence, the five genes that are essential for menaquinone biosynthesis in B. subtilis are not found in C. difficile’s genome. The TCA cycle is incomplete and is used to generate precursor metabolites rather than energy. Fermentation pathways generate ATP directly by substrate level phosphorylation but can also be used via electron bifurcation and the Rnf complex to generate a motive force across the cytoplasmic membrane (124, 125). Whether this is a proton or a sodium-ion motive force is not yet known; we will assume protons for simplicity, i.e., a PMF. C. difficile has an F0F1-type ATP synthase, which, depending on the needs of the organism, can consume the PMF to generate ATP or hydrolyze ATP to generate a PMF.

Few of the genes involved in these various pathways scored as essential by Tn-seq. Genes for the TCA cycle, acetate kinase, and the major Stickland reductases for glycine, proline and leucine are all non-essential, as are the genes for the RNF complex and three electron bifurcation complexes (etf genes). The essentiality of genes for glycolysis is less clear because eight of these were essential for Dembek et al. but only two (eno, tpiA) were essential in our experiments. Glycolysis might have more of a contribution on BHI, which contains glucose, than on TY. Differences in slow growth and statistical cutoffs that impact essentiality calls may also factor into the discrepancies. In support of this explanation, we observed a small colony phenotype when we used CRISPRi to knock down expression of four glycolysis genes (fba, gapB, pgi, and pfkA) that were essential for Dembek et al. but not in our Tn-seq (Table S1). A further point to keep in mind is that glycolysis genes could be more important for supplying precursor metabolites rather than energy in C. difficile.

A noteworthy discrepancy concerns the ten gene operon for the F-type ATPase. Dembek et al. scored nine of the genes as essential, but all ten were non-essential in our Tn-seq experiments. This gene cluster is too large to have escaped Tn insertions by chance. The most likely explanation for this discrepancy has to do with how slow growth affects perceptions of essentiality because we observed that CRISPRi knockdown of atpB and atpD resulted in a small colony phenotype (Fig. 3C). We also tested the effect of knockdowns in TY broth using one sgRNA that caused a small colony phenotype (atpD) and one that did not (atpF). Interestingly, both knockdowns caused a strong growth defect, but only if cultures were pre-grown overnight in 1% xylose to deplete the AtpD or AtpF proteins before sub-culturing (Fig. 3D). As an aside, we found that all four atp operon knockdowns were sensitized to subinhibitory concentrations of the uncoupler carbonyl cyanide m-chlorophenylhydrazone (CCCP, Fig. 3E), which hints at the potential for using our CRISPRi library to study drug targets in C. difficile (7, 12).

Three genes (hisC, ilvB and ilvC) involved in amino acid biosynthesis were identified as essential despite utilization of growth medium rich in tryptone. These genes were not essential for Dembek et al. Note, however, that there are seven essential lysine biosynthesis genes, which we categorized under cell envelope rather than metabolism owing to their role in synthesis of diaminopimelate for peptidoglycan. Finally, the global regulator CodY is essential by Tn-seq. CodY is widely conserved in Bacillota and senses GTP and branched chain amino acids to regulate gene expression in response to the energetic and nutritional needs of the cell. In C. difficile CodY represses hundreds of genes during exponential growth, and a codY null mutant grows poorly upon entry into stationary phase (126128), which likely explains the Tn-seq result.

Nucleotides and cofactors

We identified eleven genes essential for nucleotide biosynthesis. All eleven were also essential or ambiguous for Dembek et al. and five are essential in B. subtilis as well. One interesting difference is that an anaerobic ribonucleotide reductase encoded by nrdD and nrdG is essential in C. difficile, but these genes are not found in B. subtilis, which has instead an aerobic ribonucleotide reductase encoded by nrdE and nrdF that are not found in C. difficile (129). Two additional exceptions are guaA (GMP synthase) and thyA (thymidylate synthase), which are essential in C. difficile and S. aureus but not B. subtilis (14, 29, 32). Two genes for regulatory nucleotides appear to be essential in C. difficile, the cyclic-di-GMP phosphodiesterase yybT and the bifunctional (pp)pGpp synthase/hydrolase relA. Essentiality of relA was confirmed by CRISPRi (Table S1). The B. subtilis paralogs of these genes are not essential (29). Essentiality of yybT is likely to be an artifact of polarity onto rplI or dnaC, but the genes transcribed with relA are not essential. In C. difficile relA is called rsh and synthesizes exclusively pGpp (130, 131). As an aside, we note that cyclic-di-AMP is essential in C. difficile growing on rich media, but c-di-AMP synthases were not identified by Tn-seq because there are two of them, neither of which is individually essential (132).

Twenty-four genes are essential for synthesis of cofactors despite utilization of media containing tryptone and yeast extract. All but two of these were also essential or ambiguous for Dembek et al., and 14 have an essential ortholog in B. subtilis. Curiously, neither we nor Dembek et al. scored dihydrofolate reductase (dfrA) as essential. Dihydrofolate reductase is the target of several important antibiotics and essential in E. coli, B. subtilis, S. sanguinis, and S. aureus (29, 32, 46, 47).

Phage and Transposon-related genes

The C. difficile genome has a remarkably high content of mobile genetic elements (25, 133). Mobile genetic elements are not part of the core genome and thus should not be essential for viability. Nevertheless, twenty-one genes classified as essential appear to reside on a prophage or a transposon. Some of these might be false positives because only eight were also essential or ambiguous for Dembek et al. Even the eight genes classified as essential in both studies are likely due to indirect effects such as induction of a lytic prophage.

Transporters

Six genes for transporters were classified as essential in our Tn-seq, three of which were also essential for Dembek et al. and were confirmed by CRISPRi (Table S1). These encode a predicted Ktr potassium transporter and a predicted CorA-like divalent metal ion transporter. In B. subtilis, there are two Ktr systems, which are not essential but improve growth at high osmolarity (134).

Sporulation

Curiously, both we and Dembek et al. classified the sporulation-associated phosphatases ptpA and ptpB as ambiguous or essential for vegetative growth. Two labs have reported null mutants of these genes, so they are not formally essential (33, 135, 136). Loss of ptpA or ptpB enhances sporulation, which we confirmed using CRISPRi against ptpB (Table S1). We presume that ptp genes are essential by Tn-seq because enhanced sporulation reduces vegetative growth.

Genes of unknown function

Our Tn-seq analysis identified 28 putatively essential genes that could not be assigned to a functional pathway. None of these genes have an essential ortholog in B. subtilis, although in five cases BioCyc identified a non-essential ortholog. Eleven of these genes were not essential for Dembek et al. and in two cases (cdr20291_3519 and cdr20291_3520) essentiality is likely due to polarity onto rplI or dnaC. That leaves 15 genes that are essential or ambiguous in two independent Tn-seq studies and are therefore likely to be bona fide essential genes. As noted above in the discussion of our CRISPRi experiments, we silenced expression of eleven of these genes and observed a viability defect for nine of them, often accompanied by abnormal morphologies (Table 2, Table S1). The apparently essential genes of unknown function constitute a high value gene set from the perspectives of bacterial physiology and antibiotic development.

Conclusions

In summary, we identified 346 protein-encoding genes that by Tn mutagenesis are essential for vegetative growth of C. difficile strain R20291 on TY media. Of these, 283 were also identified as essential by Tn mutagenesis in a previous study (14) and 169 have an essential ortholog in B. subtilis (29). Overall, these results are broadly consistent with studies of gene essentiality in model organisms such as E. coli, B. subtilis and S. aureus (29, 32, 45, 46, 57). The 283 C. difficile genes identified as essential in two independent Tn mutagenesis studies can be regarded as a consensus “essentialome” that minimizes false positives. Most of these genes play key roles in foundational cellular processes such as DNA replication, transcription, translation and cell envelope biogenesis. But the consensus essentialome also includes 15 genes that could not be assigned to any functional pathway (Table 2, Table S3A, B). These genes might be targets for antibiotics that kill C. difficile without decimating the healthy microbiota needed to keep C. difficile in check.

We also used CRISPRi knockdown to investigate 181 genes that had been identified as essential in a previous Tn-seq analysis (14). Our goals were to vet essentiality and screen for morphological defects that would facilitate assigning genes of unknown function to physiological pathways. Our CRISPRi platform used a plasmid that expresses dCas9 from a xylose-inducible promoter (Pxyl) and an sgRNA from a strong constitutive promoter (Pgdh) (15). CRISPRi resulted in reduced plating efficiencies and/or small colony phenotypes on TY-xylose plates for 167 of the 181 genes targeted, a very high confirmation rate of 92%. The 14 genes for which no viability defect was observed could be false positives from the previous report or genes for which our sgRNAs were ineffective. Of these genes, ten sustained insertions in our Tn-seq experiments, so we infer they are non-essential. Four did not sustain Tn insertions and are therefore likely to be essential genes that were poorly repressed by our sgRNAs. Importantly, no growth defects were observed using 20 control sgRNAs that did not target anywhere in the genome, indicating off-target effects are rare.

Microscopy of surviving cells scraped from the TY-xylose plates revealed most knockdowns resulted in morphological abnormalities (151 out of 181 genes, 83%). Disappointingly, however, the utility of these defects for making functional assignments was limited by the observation that repressing genes of known function often resulted in non-intuitive defects. For example, repressing RNA polymerase gene rpoB resulted in severe filamentation suggestive of a cell division defect, while repressing the nucleotide biosynthesis gene guaA caused a chaining phenotype suggestive of a daughter cell separation defect. Non-intuitive phenotypes have also been reported in other CRISPRi screens (7, 8).

The findings and resources presented here should help guide future studies of C. difficile. First, our results can be used to prioritize genes for more rigorous but labor-intensive investigation using depletion strains with in-frame deletions (137). The 15 apparently essential genes that could not be assigned to a functional pathway seem like a good place to start. Second, our CRISPRi library can be leveraged to investigate antibiotic sensitivities (7, 12, 138), which might illuminate gene function and reveal vulnerabilities that can be exploited to improve treatment of C. difficile infections. Third, the identification of 18 proteins that localize to the midcell raises new questions related to C. difficile morphogenesis. For example, septal localization of the canonical elongation proteins MreC and MreD suggests they contribute to cell division and/or C. difficile elongates by inserting new peptidoglycan near the midcell. In addition, our discovery that YlmG, YlxW and YlxX localize to the division site provides the most direct evidence to date that these conserved but enigmatic proteins play a role in cell division.

METHODS

Strains, media, and growth conditions.

Most bacterial strains used in this study are listed in Table S4. Strains and plasmids constructed for the CRISPRi library are summarized separately in Table S2. C. difficile strains were derived from R20291 (139). C. difficile was routinely grown in tryptone-yeast extract (TY) medium, supplemented as needed with thiamphenicol at 10 μg/ml (TY-Thi10). TY medium consisted of 3% tryptone, 2% yeast extract, and 2% agar (for plates). Brain heart infusion (BHI) media was prepared per manufacturer’s (DIFCO) instructions. C. difficile strains were maintained at 37°C in an anaerobic chamber (Coy Laboratory Products) in an atmosphere of 2% H2, 5% CO2, and 93% N2. Escherichia coli strains were grown in LB medium at 37°C with chloramphenicol at 10 μg/ml and/or ampicillin at 100 μg/ml as needed. LB medium contained 1% tryptone, 0.5% yeast extract, 0.5% NaCl, and 1.5% agar (for plates). OD600 measurements were made with the WPA Biowave CO8000 tube reader in the anaerobic chamber.

Plasmid and strain construction.

Plasmids are listed in Table S5 and were constructed with HiFi DNA Assembly from New England Biolabs (Ipswich, MA). Oligonucleotide primers (Table S6) were synthesized by Integrated DNA Technologies (Coralville, IA). CRISPRi plasmids were constructed as described in (15). Regions constructed by PCR were verified by DNA sequencing. Plasmids were propagated in E. coli HB101/pRK24 and conjugated into C. difficile R20291 according to (53). Final R20291 CRISPRi strains were verified by PCR amplifying and sequencing the guide region. Details relevant to other plasmid construction are provided in Table S5.

CRISPRi screen.

Overnight cultures grown in TY-Thi10 were serially diluted 10-fold in TY, and 5 µL spotted on TY-Thi10 and TY-Thi10 1% (w/v) xylose plates. Plates were incubated at 37°C overnight and imaged the following morning (~18 h). Cells were scraped from select spots (usually the last spot with growth) and resuspended in 50 µL TY. Cell suspensions were supplemented with 5 µg/mL FM4–64 (red fluorescent membrane stain, Thermo Scientific) and 15 µg/mL Hoechst 33342 (blue fluorescent DNA stain, Invitrogen) and imaged by phase-contrast and fluorescence microscopy.

Protein localization.

R20291 harboring plasmids that expressed RFP-tagged proteins under xylose control were grown in TY-Thi10 overnight, subcultured into TY-Thi10 with 0.1% or 1% xylose, grown to an OD600 of about 0.6, and fixed with 4% buffered paraformaldehyde as described (53, 120, 140). Fixed cells were photographed under phase-contrast and (red) fluorescence. Septal localization was scored manually by inspecting cells for the presence of a fluorescent band near the midcell. MicrobeJ was used to keep track of cells scored positive or negative for septal localization (141).

Microscopy.

Cells were immobilized using thin agarose pads (1% w/v agarose). Phase-contrast micrographs were recorded on an Olympus BX60 microscope equipped with a 100× UPlanApo objective (numerical aperture, 1.35). Micrographs were captured with a Hamamatsu Orca Flash 4.0 V2+ complementary metal oxide semiconductor (CMOS) camera. Excitation light was generated with an X-Cite XYLIS LED light source. Red fluorescence was detected with the Chroma filter set 49008 (538 to 582 nm excitation filter, 587 nm dichroic mirror, and a 590 to 667 nm emission filter). Blue fluorescence was detected with the Olympus filter set U-MWU (330–385 nm excitation filter, 400 nm dichroic mirror and a 420 nm barrier emission filter).

Transposon library construction.

Plasmid pRPF215 is a quasi-suicide plasmid that harbors the Himar1 mariner transposase gene under control of Ptet (14). The gene for TetR does not have a terminator and transcription reads through into the origin of replication, presumably disrupting plasmid replication. Addition of anhydrotetracycline therefore both induces the transposase and causes plasmid loss. A single colony of R20291/pRPF215 was used to inoculate a 2 mL overnight culture in TY-Thi10. Twenty independent overnight cultures were grown for each transposon library construction. After overnight growth, each was then sub-cultured 1:50 into 2 mL TY and grown to an OD600 of 0.3. From each subculture an aliquot was removed and spread on to two large (15 cm diameter) plates of TY agar with 80 µg/mL lincomycin (RPI) and 100 ng/mL anhydrotetracycline (Sigma), for a total of 40 plates. We used higher concentrations of lincomycin than originally published (14) because we found 80 µg/mL lincomycin decreased the number of false positives. The amount of subculture to plate was experimentally determined to give roughly 5000–8000 colonies. Typically, we used 220 µL of subculture diluted with TY to 600 µL, a volume suitable for spreading evenly on a large plate. A dilution series of one subculture was also plated on TY to calculate plating efficiency. Selection plates typically grew one colony for every 500 plated (i.e., an efficiency of about 2 × 10−3). Plates were incubated for 20 hours at 37˚C. Cells were then scraped off the plates with 5 mL TY each, pooled, amended to 10% DMSO, aliquoted and stored at −80˚C. This material was referred to as the primary transposon library. Suspensions of the primary libraries typically had an OD600 of about 6. The concentration of viable cells was quantitated by plating aliquots on TY plates and was typically around 3 × 108 CFU/mL. Three independent libraries were constructed on different days.

Tn-seq sample preparation.

DNA samples were prepared directly from 1 mL of primary library or from 10 mL culture that had been grown for an additional 7 doublings in TY. To avoid creating a bottleneck, 10 mL TY was inoculated with 2.2 × 107 CFU. There are 502,945 possible TA insertion sites in the R20291 chromosome, thus cultures were started with a ratio of about 45 CFU per TA site. DNA libraries for Illumina sequencing were prepared based on modifications of Karash et al. (142). Briefly, regions adjacent to any transposon insertion were amplified by single primer extension. The resulting products were extended with a cytosine-tail, which then allowed further amplification by PCR. The upstream primer recognizes the transposon sequence, incorporates the P5 sequence for Illumina sequencing and a sample-specific barcode; the downstream primer recognizes the C-tail and incorporates the P7 sequence.

Genomic DNA was prepared using the Monarch Genomic DNA purification kit from NEB, using the protocol for Gram positive bacteria. A maximum of 2 × 109 cells were pelleted. Lysis was facilitated through the addition of 0.5 mg hen egg white lysozyme (Boehringer Mannheim) and 20 U mutanolysin (Sigma), and DNA was eluted in 35 µL with a typical yield of 200 ng/µL. Linear extension PCR was performed on 100 ng DNA in 50 µL with Taq polymerase (NEB) and primer Tn-ermB-2 (anneal: 30 s at 55˚C, extend 30 s at 68˚C, 50 cycles). The resulting product was spin-column purified (Zymo Research Clean & Concentrator kit) and eluted in 12 µL. A C-tail was added by extending with terminal transferase (NEB) in a 20 µL reaction, using 1.25 mM dCTP (NEB) and 50 µM ddCTP (MilliporeSigma/Roche). The product was again spin-column purified and eluted in 10 µL. Final PCR amplification used 1 µL of C-tailed DNA in a 35 µL reaction mixture, Taq polymerase and primers P7–16G and P5-Tn-Px (x: variable barcode; anneal: 30 s at 62˚C, extend 30 s at 68˚C, 35 cycles). The resulting product was separated on a 1.5% agarose gel in Tris Acetate EDTA buffer (TAE). Fragments of 300–500 base pair length were excised, purified with the Zymo Research Gel DNA recovery kit, and eluted in 10 µL. DNA concentration was quantitated with the Qubit dsDNA assay and was typically around 5 ng/µL. Four samples with distinct barcodes were combined and submitted for sequencing (Illumina HiSeq X, 150-bp PE reads) with Admera Health Biopharma Services (South Plainfield, NJ). Samples were spiked with 5% PhiX DNA to improve data quality.

Sequencing data processing.

Raw sequencing files were first trimmed with Trimmomatic to eliminate poor quality reads (143). The first four bases before the barcodes were then removed using Trim Sequences and the resulting files were de-multiplexed using the Barcode splitter, both on Galaxy (144). Reads were aligned to the reference genome of R20291 (NC_013316.1 or ASM2710v1) using the Burrows-Wheeler Aligner (BWA) provided in TRANSIT (27). Finally, the resulting Wig files were compared in TRANSIT2 which evaluates gene essentiality both by Gumbel analysis and binomial analysis (145). The former makes essentiality calls based on insertion gaps, i.e. consecutive TA sites lacking transposon insertions, using the Gumbel distribution (146). The latter calls essentiality for small genes lacking insertions which can be difficult to detect by the more conservative Gumbel algorithm (28). Essentiality calls are either “E” when identified by Gumbel or “EB” when identified by the Binomial analysis. Table S3 lists genes that were called essential in primary insertion libraries using cells scraped from plates, or after an additional 7 generations of growth. The library dataset was generated from three independently constructed transposon libraries. The outgrowth dataset was generated from two independent growth cultures from each of the three independent libraries. We present both the separate data output as well as a combined essentiality call (Table S3). The latter was further hand-edited by including 11 genes (indicated as “Ei” for “essential by inspection”) that appeared to have mistakenly called non-essential by TRANSIT2. Ten of these genes had very few insertions despite numerous possible TA sites, while the eleventh had a large number of insertions but mostly at the 3’ end of the gene.

Supplementary Material

Supplement 1
media-1.pdf (614.2KB, pdf)
Supplement 2
media-2.xlsx (36.9KB, xlsx)
Supplement 3
media-3.xlsx (36.4KB, xlsx)
Supplement 4
media-4.xlsx (423KB, xlsx)
Supplement 5
media-5.pdf (201.1KB, pdf)

IMPORTANCE.

Clostridioides difficile is an opportunistic pathogen for which better antibiotics are sorely needed. Most antibiotics target pathways that are essential for viability. Here we use saturation transposon mutagenesis and gene silencing with CRISPR interference to identify and characterize genes required for growth on laboratory media. Comparison to the model organism B. subtilis reveals many similarities and a few striking differences that warrant further study and may include opportunities for developing antibiotics that kill C. difficile without decimating the healthy microbiota needed to keep C. difficile in check.

Acknowledgements

This work was supported by Public Health Service Grants R21 AI159071 (D.S.W) and R01 AI155492 (C.D.E. and D.S.W). A.J.O. and F.V.T. were supported by NSF REU DBI-1852070. H.M.L. and J.G.R.-R. were supported by NSF REU DBI-2244169. We thank members of the Ellermeier and Weiss laboratories for helpful discussions, John Cronan for information on gpsA and Erin Purcell for information on relA.

REFERENCES

  • 1.CDC. 2019. Antibiotic Resistance Threats in the United States. Centers for Disease Control and Prevention, Atlanta, GA. [Google Scholar]
  • 2.Miller AC, Arakkal AT, Sewell DK, Segre AM, Tholany J, Polgreen PM, Group CDCM-H. 2023. Comparison of Different Antibiotics and the Risk for Community-Associated Clostridioides difficile Infection: A Case-Control Study. Open Forum Infect Dis 10:ofad413. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Theriot CM, Koenigsknecht MJ, Carlson PE Jr., Hatton GE, Nelson AM, Li B, Huffnagle GB, J ZL, Young VB. 2014. Antibiotic-induced shifts in the mouse gut microbiome and metabolome increase susceptibility to Clostridium difficile infection. Nat Commun 5:3114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Cole SA, Stahl TJ. 2015. Persistent and Recurrent Clostridium difficile Colitis. Clin Colon Rectal Surg 28:65–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Cain AK, Barquist L, Goodman AL, Paulsen IT, Parkhill J, van Opijnen T. 2020. A decade of advances in transposon-insertion sequencing. Nat Rev Genet 21:526–540. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.van Opijnen T, Camilli A. 2013. Transposon insertion sequencing: a new tool for systems-level analysis of microorganisms. Nat Rev Microbiol 11:435–42. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Peters JM, Colavin A, Shi H, Czarny TL, Larson MH, Wong S, Hawkins JS, Lu CHS, Koo BM, Marta E, Shiver AL, Whitehead EH, Weissman JS, Brown ED, Qi LS, Huang KC, Gross CA. 2016. A Comprehensive, CRISPR-based Functional Analysis of Essential Genes in Bacteria. Cell 165:1493–1506. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Silvis MR, Rajendram M, Shi H, Osadnik H, Gray AN, Cesar S, Peters JM, Hearne CC, Kumar P, Todor H, Huang KC, Gross CA. 2021. Morphological and Transcriptional Responses to CRISPRi Knockdown of Essential Genes in Escherichia coli. mBio 12:e0256121. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Liu X, Gallay C, Kjos M, Domenech A, Slager J, van Kessel SP, Knoops K, Sorg RA, Zhang JR, Veening JW. 2017. High-throughput CRISPRi phenotyping identifies new essential genes in Streptococcus pneumoniae. Mol Syst Biol 13:931. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Shields RC, Walker AR, Maricic N, Chakraborty B, Underhill SAM, Burne RA. 2020. Repurposing the Streptococcus mutans CRISPR-Cas9 System to Understand Essential Gene Function. PLoS Pathog 16:e1008344. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.de Wet TJ, Winkler KR, Mhlanga M, Mizrahi V, Warner DF. 2020. Arrayed CRISPRi and quantitative imaging describe the morphotypic landscape of essential mycobacterial genes. Elife 9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Ward RD, Tran JS, Banta AB, Bacon EE, Rose WE, Peters JM. 2024. Essential gene knockdowns reveal genetic vulnerabilities and antibiotic sensitivities in Acinetobacter baumannii. mBio 15:e0205123. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Qi LS, Larson MH, Gilbert LA, Doudna JA, Weissman JS, Arkin AP, Lim WA. 2013. Repurposing CRISPR as an RNA-guided platform for sequence-specific control of gene expression. Cell 152:1173–83. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Dembek M, Barquist L, Boinett CJ, Cain AK, Mayho M, Lawley TD, Fairweather NF, Fagan RP. 2015. High-throughput analysis of gene essentiality and sporulation in Clostridium difficile. mBio 6:e02383. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Müh U, Pannullo AG, Weiss DS, Ellermeier CD. 2019. A Xylose-Inducible Expression System and a CRISPR Interference Plasmid for Targeted Knockdown of Gene Expression in Clostridioides difficile. J Bacteriol 201:e00711–18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Rousset F, Cui L, Siouve E, Becavin C, Depardieu F, Bikard D. 2018. Genome-wide CRISPR-dCas9 screens in E. coli identify essential genes and phage host factors. PLoS Genet 14:e1007749. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Wang T, Guan C, Guo J, Liu B, Wu Y, Xie Z, Zhang C, Xing XH. 2018. Pooled CRISPR interference screening enables genome-scale functional genomics study in bacteria with superior performance. Nat Commun 9:2475. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Karp PD, Billington R, Caspi R, Fulcher CA, Latendresse M, Kothari A, Keseler IM, Krummenacker M, Midford PE, Ong Q, Ong WK, Paley SM, Subhraveti P. 2019. The BioCyc collection of microbial genomes and metabolic pathways. Brief Bioinform 20:1085–1093. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Walter BM, Cartman ST, Minton NP, Butala M, Rupnik M. 2015. The SOS Response Master Regulator LexA Is Associated with Sporulation, Motility and Biofilm Formation in Clostridium difficile. PLoS One 10:e0144763. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Kawai Y, Moriya S, Ogasawara N. 2003. Identification of a protein, YneA, responsible for cell division suppression during the SOS response in Bacillus subtilis. Mol Microbiol 47:1113–22. [DOI] [PubMed] [Google Scholar]
  • 21.Nonejuie P, Burkart M, Pogliano K, Pogliano J. 2013. Bacterial cytological profiling rapidly identifies the cellular pathways targeted by antibacterial molecules. Proc Natl Acad Sci U S A 110:16169–74. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Htoo HH, Brumage L, Chaikeeratisak V, Tsunemoto H, Sugie J, Tribuddharat C, Pogliano J, Nonejuie P. 2019. Bacterial Cytological Profiling as a Tool To Study Mechanisms of Action of Antibiotics That Are Active against Acinetobacter baumannii. Antimicrob Agents Chemother 63. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Quach D, Sharp M, Ahmed S, Ames L, Bhagwat A, Deshpande A, Parish T, Pogliano J, Sugie J. 2025. Deep learning-driven bacterial cytological profiling to determine antimicrobial mechanisms in Mycobacterium tuberculosis. Proc Natl Acad Sci U S A 122:e2419813122. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Lampe DJ, Grant TE, Robertson HM. 1998. Factors affecting transposition of the Himar1 mariner transposon in vitro. Genetics 149:179–87. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Sebaihia M, Wren BW, Mullany P, Fairweather NF, Minton N, Stabler R, Thomson NR, Roberts AP, Cerdeno-Tarraga AM, Wang H, Holden MT, Wright A, Churcher C, Quail MA, Baker S, Bason N, Brooks K, Chillingworth T, Cronin A, Davis P, Dowd L, Fraser A, Feltwell T, Hance Z, Holroyd S, Jagels K, Moule S, Mungall K, Price C, Rabbinowitsch E, Sharp S, Simmonds M, Stevens K, Unwin L, Whithead S, Dupuy B, Dougan G, Barrell B, Parkhill J. 2006. The multidrug-resistant human pathogen Clostridium difficile has a highly mobile, mosaic genome. Nat Genet 38:779–86. [DOI] [PubMed] [Google Scholar]
  • 26.He M, Sebaihia M, Lawley TD, Stabler RA, Dawson LF, Martin MJ, Holt KE, Seth-Smith HM, Quail MA, Rance R, Brooks K, Churcher C, Harris D, Bentley SD, Burrows C, Clark L, Corton C, Murray V, Rose G, Thurston S, van Tonder A, Walker D, Wren BW, Dougan G, Parkhill J. 2010. Evolutionary dynamics of Clostridium difficile over short and long time scales. Proc Natl Acad Sci U S A 107:7527–32. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.DeJesus MA, Ambadipudi C, Baker R, Sassetti C, Ioerger TR. 2015. TRANSIT--A Software Tool for Himar1 TnSeq Analysis. PLoS Comput Biol 11:e1004401. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Choudhery S, Brown AJ, Akusobi C, Rubin EJ, Sassetti CM, Ioerger TR. 2021. Modeling Site-Specific Nucleotide Biases Affecting Himar1 Transposon Insertion Frequencies in TnSeq Data Sets. mSystems 6:e0087621. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Koo BM, Kritikos G, Farelli JD, Todor H, Tong K, Kimsey H, Wapinski I, Galardini M, Cabal A, Peters JM, Hachmann AB, Rudner DZ, Allen KN, Typas A, Gross CA. 2017. Construction and Analysis of Two Genome-Scale Deletion Libraries for Bacillus subtilis. Cell Syst 4:291–305 e7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Meeske AJ, Sham LT, Kimsey H, Koo BM, Gross CA, Bernhardt TG, Rudner DZ. 2015. MurJ and a novel lipid II flippase are required for cell wall biogenesis in Bacillus subtilis. Proc Natl Acad Sci U S A 112:6437–42. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Kobayashi K, Ehrlich SD, Albertini A, Amati G, Andersen KK, Arnaud M, Asai K, Ashikaga S, Aymerich S, Bessieres P, Boland F, Brignell SC, Bron S, Bunai K, Chapuis J, Christiansen LC, Danchin A, Debarbouille M, Dervyn E, Deuerling E, Devine K, Devine SK, Dreesen O, Errington J, Fillinger S, Foster SJ, Fujita Y, Galizzi A, Gardan R, Eschevins C, Fukushima T, Haga K, Harwood CR, Hecker M, Hosoya D, Hullo MF, Kakeshita H, Karamata D, Kasahara Y, Kawamura F, Koga K, Koski P, Kuwana R, Imamura D, Ishimaru M, Ishikawa S, Ishio I, Le Coq D, Masson A, Mauel C, et al. 2003. Essential Bacillus subtilis genes. Proc Natl Acad Sci U S A 100:4678–83. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Chaudhuri RR, Allen AG, Owen PJ, Shalom G, Stone K, Harrison M, Burgis TA, Lockyer M, Garcia-Lara J, Foster SJ, Pleasance SJ, Peters SE, Maskell DJ, Charles IG. 2009. Comprehensive identification of essential Staphylococcus aureus genes using Transposon-Mediated Differential Hybridisation (TMDH). BMC Genomics 10:291. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Underwood S, Guan S, Vijayasubhash V, Baines SD, Graham L, Lewis RJ, Wilcox MH, Stephenson K. 2009. Characterization of the sporulation initiation pathway of Clostridium difficile and its role in toxin production. J Bacteriol 191:7296–305. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Edwards AN, McBride SM. 2017. Determination of the in vitro Sporulation Frequency of Clostridium difficile. Bio Protoc 7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Diaz A, Lacks SA, Lopez P. 1992. The 5’ to 3’ exonuclease activity of DNA polymerase I is essential for Streptococcus pneumoniae. Mol Microbiol 6:3009–19. [DOI] [PubMed] [Google Scholar]
  • 36.Bayliss CD, Sweetman WA, Moxon ER. 2005. Destabilization of tetranucleotide repeats in Haemophilus influenzae mutants lacking RnaseHI or the Klenow domain of PolI. Nucleic Acids Res 33:400–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Fukushima S, Itaya M, Kato H, Ogasawara N, Yoshikawa H. 2007. Reassessment of the in vivo functions of DNA polymerase I and RNase H in bacterial cell growth. J Bacteriol 189:8575–83. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.van Eijk E, Paschalis V, Green M, Friggen AH, Larson MA, Spriggs K, Briggs GS, Soultanas P, Smits WK. 2016. Primase is required for helicase activity and helicase alters the specificity of primase in the enteropathogen Clostridium difficile. Open Biol 6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Smits WK, Goranov AI, Grossman AD. 2010. Ordered association of helicase loader proteins with the Bacillus subtilis origin of replication in vivo. Mol Microbiol 75:452–61. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Weiss A, Moore BD, Tremblay MHJ, Chaput D, Kremer A, Shaw LN. 2017. The omega Subunit Governs RNA Polymerase Stability and Transcriptional Specificity in Staphylococcus aureus. J Bacteriol 199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Gentry DR, Burgess RR. 1989. rpoZ, encoding the omega subunit of Escherichia coli RNA polymerase, is in the same operon as spoT. J Bacteriol 171:1271–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Jayasinghe OT, Mandell ZF, Yakhnin AV, Kashlev M, Babitzke P. 2022. Transcriptome-Wide Effects of NusA on RNA Polymerase Pausing in Bacillus subtilis. J Bacteriol 204:e0053421. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Trzilova D, Anjuwon-Foster BR, Torres Rivera D, Tamayo R. 2020. Rho factor mediates flagellum and toxin phase variation and impacts virulence in Clostridioides difficile. PLoS Pathog :e1008708. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Karzai AW, Roche ED, Sauer RT. 2000. The SsrA-SmpB system for protein tagging, directed degradation and ribosome rescue. Nat Struct Biol 7:449–55. [DOI] [PubMed] [Google Scholar]
  • 45.Santiago M, Matano LM, Moussa SH, Gilmore MS, Walker S, Meredith TC. 2015. A new platform for ultra-high density Staphylococcus aureus transposon libraries. BMC Genomics 16:252. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Yamamoto N, Nakahigashi K, Nakamichi T, Yoshino M, Takai Y, Touda Y, Furubayashi A, Kinjyo S, Dose H, Hasegawa M, Datsenko KA, Nakayashiki T, Tomita M, Wanner BL, Mori H. 2009. Update on the Keio collection of Escherichia coli single-gene deletion mutants. Mol Syst Biol 5:335. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Xu P, Ge X, Chen L, Wang X, Dou Y, Xu JZ, Patel JR, Stone V, Trinh M, Evans K, Kitten T, Bonchev D, Buck GA. 2011. Genome-wide essential gene identification in Streptococcus sanguinis. Sci Rep 1:125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Nair N, Raff H, Islam MT, Feen M, Garofalo DM, Sheppard K. 2016. The Bacillus subtilis and Bacillus halodurans Aspartyl-tRNA Synthetases Retain Recognition of tRNA(Asn). J Mol Biol 428:618–630. [DOI] [PubMed] [Google Scholar]
  • 49.Putzer H, Gendron N, Grunberg-Manago M. 1992. Co-ordinate expression of the two threonyl-tRNA synthetase genes in Bacillus subtilis: control by transcriptional antitermination involving a conserved regulatory sequence. EMBO J 11:3117–27. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Williams-Wagner RN, Grundy FJ, Raina M, Ibba M, Henkin TM. 2015. The Bacillus subtilis tyrZ gene encodes a highly selective tyrosyl-tRNA synthetase and is regulated by a MarR regulator and T box riboswitch. J Bacteriol 197:1624–31. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Strauch MA, Zalkin H, Aronson AI. 1988. Characterization of the glutamyl-tRNA(Gln)-to-glutaminyl-tRNA(Gln) amidotransferase reaction of Bacillus subtilis. J Bacteriol 170:916–20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Curnow AW, Hong K, Yuan R, Kim S, Martins O, Winkler W, Henkin TM, Soll D. 1997. Glu-tRNAGln amidotransferase: a novel heterotrimeric enzyme required for correct decoding of glutamine codons during translation. Proc Natl Acad Sci U S A 94:11819–26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Müh U, Ellermeier CD, Weiss DS. 2022. The WalRK Two-Component System Is Essential for Proper Cell Envelope Biogenesis in Clostridioides difficile. J Bacteriol 204:e0012122. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Wall EA, Caufield JH, Lyons CE, Manning KA, Dokland T, Christie GE. 2015. Specific N-terminal cleavage of ribosomal protein L27 in Staphylococcus aureus and related bacteria. Mol Microbiol 95:258–69. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Sauer RT, Baker TA. 2011. AAA+ proteases: ATP-fueled machines of protein destruction. Annu Rev Biochem 80:587–612. [DOI] [PubMed] [Google Scholar]
  • 56.Clausen T, Kaiser M, Huber R, Ehrmann M. 2011. HTRA proteases: regulated proteolysis in protein quality control. Nat Rev Mol Cell Biol 12:152–62. [DOI] [PubMed] [Google Scholar]
  • 57.Baba T, Ara T, Hasegawa M, Takai Y, Okumura Y, Baba M, Datsenko KA, Tomita M, Wanner BL, Mori H. 2006. Construction of Escherichia coli K-12 in-frame, single-gene knockout mutants: the Keio collection. Mol Syst Biol 2:2006 0008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Haas M, Beyer D, Gahlmann R, Freiberg C. 2001. YkrB is the main peptide deformylase in Bacillus subtilis, a eubacterium containing two functional peptide deformylases. Microbiology (Reading) 147:1783–1791. [DOI] [PubMed] [Google Scholar]
  • 59.Cai Y, Chandrangsu P, Gaballa A, Helmann JD. 2017. Lack of formylated methionyl-tRNA has pleiotropic effects on Bacillus subtilis. Microbiology (Reading) 163:185–196. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.You C, Lu H, Sekowska A, Fang G, Wang Y, Gilles AM, Danchin A. 2005. The two authentic methionine aminopeptidase genes are differentially expressed in Bacillus subtilis. BMC Microbiol 5:57. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Green ER, Mecsas J. 2016. Bacterial Secretion Systems: An Overview. Microbiol Spectr 4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Fagan RP, Fairweather NF. 2011. Clostridium difficile has two parallel and essential Sec secretion systems. J Biol Chem 286:27483–93. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Choi KH, Heath RJ, Rock CO. 2000. beta-ketoacyl-acyl carrier protein synthase III (FabH) is a determining factor in branched-chain fatty acid biosynthesis. J Bacteriol 182:365–70. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Todter D, Gunka K, Stulke J. 2017. The Highly Conserved Asp23 Family Protein YqhY Plays a Role in Lipid Biosynthesis in Bacillus subtilis. Front Microbiol 8:883. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Morbidoni HR, de Mendoza D, Cronan JE, Jr. 1995. Synthesis of sn-glycerol 3-phosphate, a key precursor of membrane lipids, in Bacillus subtilis. J Bacteriol 177:5899–905. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Heuston S, Begley M, Gahan CGM, Hill C. 2012. Isoprenoid biosynthesis in bacterial pathogens. Microbiology (Reading) 158:1389–1401. [DOI] [PubMed] [Google Scholar]
  • 67.Manat G, Roure S, Auger R, Bouhss A, Barreteau H, Mengin-Lecreulx D, Touze T. 2014. Deciphering the metabolism of undecaprenyl-phosphate: the bacterial cell-wall unit carrier at the membrane frontier. Microb Drug Resist 20:199–214. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Monot M, Boursaux-Eude C, Thibonnier M, Vallenet D, Moszer I, Medigue C, Martin-Verstraete I, Dupuy B. 2011. Reannotation of the genome sequence of Clostridium difficile strain 630. J Med Microbiol 60:1193–1199. [DOI] [PubMed] [Google Scholar]
  • 69.Zbylicki BR, Murphy CE, Petsche JA, Müh U, Dobrila HA, Ho TD, Daum MN, Pannullo AG, Weiss DS, Ellermeier CD. 2024. Identification of Clostridioides difficile mutants with increased daptomycin resistance. J Bacteriol 206:e0036823. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Willing SE, Candela T, Shaw HA, Seager Z, Mesnage S, Fagan RP, Fairweather NF. 2015. Clostridium difficile surface proteins are anchored to the cell wall using CWB2 motifs that recognise the anionic polymer PSII. Mol Microbiol 96:596–608. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Kirk JA, Banerji O, Fagan RP. 2017. Characteristics of the Clostridium difficile cell envelope and its importance in therapeutics. Microb Biotechnol 10:76–90. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Kirk JA, Gebhart D, Buckley AM, Lok S, Scholl D, Douce GR, Govoni GR, Fagan RP. 2017. New class of precision antimicrobials redefines role of Clostridium difficile S-layer in virulence and viability. Sci Transl Med 9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Wang S, Courreges MC, Xu L, Gurung B, Berryman M, Gu T. 2024. Revealing roles of S-layer protein (SlpA) in Clostridioides difficile pathogenicity by generating the first slpA gene deletion mutant. Microbiol Spectr 12:e0400523. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Chu M, Mallozzi MJ, Roxas BP, Bertolo L, Monteiro MA, Agellon A, Viswanathan VK, Vedantam G. 2016. A Clostridium difficile Cell Wall Glycopolymer Locus Influences Bacterial Shape, Polysaccharide Production and Virulence. PLoS Pathog 12:e1005946. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Malet-Villemagne J, Yucheng L, Evanno L, Denis-Quanquin S, Hugonnet JE, Arthur M, Janoir C, Candela T. 2023. Polysaccharide II Surface Anchoring, the Achilles’ Heel of Clostridioides difficile. Microbiol Spectr 11:e0422722. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Jorgenson MA, Young KD. 2016. Interrupting Biosynthesis of O Antigen or the Lipopolysaccharide Core Produces Morphological Defects in Escherichia coli by Sequestering Undecaprenyl Phosphate. J Bacteriol 198:3070–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.D’Elia MA, Millar KE, Beveridge TJ, Brown ED. 2006. Wall teichoic acid polymers are dispensable for cell viability in Bacillus subtilis. J Bacteriol 188:8313–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Bouhss A, Trunkfield AE, Bugg TD, Mengin-Lecreulx D. 2008. The biosynthesis of peptidoglycan lipid-linked intermediates. FEMS Microbiol Rev 32:208–33. [DOI] [PubMed] [Google Scholar]
  • 79.Sham LT, Butler EK, Lebar MD, Kahne D, Bernhardt TG, Ruiz N. 2014. Bacterial cell wall. MurJ is the flippase of lipid-linked precursors for peptidoglycan biogenesis. Science 345:220–2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Zimmermann L, Stephens A, Nam SZ, Rau D, Kubler J, Lozajic M, Gabler F, Söding J, Lupas AN, Alva V. 2018. A Completely Reimplemented MPI Bioinformatics Toolkit with a New HHpred Server at its Core. J Mol Biol 430:2237–2243. [DOI] [PubMed] [Google Scholar]
  • 81.Elhenawy W, Davis RM, Fero J, Salama NR, Felman MF, Ruiz N. 2016. The O-Antigen Flippase Wzk Can Substitute for MurJ in Peptidoglycan Synthesis in Helicobacter pylori and Escherichia coli. PLoS One 11:e0161587. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Englehart K, Dworkin J. 2025. Bacillus subtilis MurJ and Amj Lipid II flippases are not essential for growth. J Bacteriol doi: 10.1128/jb.00078-25:e0007825. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Mohammadi T, van Dam V, Sijbrandi R, Vernet T, Zapun A, Bouhss A, Diepeveen-de Bruin M, Nguyen-Disteche M, de Kruijff B, Breukink E. 2011. Identification of FtsW as a transporter of lipid-linked cell wall precursors across the membrane. EMBO J 30:1425–32. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Rohs PDA, Bernhardt TG. 2021. Growth and Division of the Peptidoglycan Matrix. Annu Rev Microbiol 75:315–336. [DOI] [PubMed] [Google Scholar]
  • 85.Egan AJF, Errington J, Vollmer W. 2020. Regulation of peptidoglycan synthesis and remodelling. Nat Rev Microbiol 18:446–460. [DOI] [PubMed] [Google Scholar]
  • 86.Sauvage E, Kerff F, Terrak M, Ayala JA, Charlier P. 2008. The penicillin-binding proteins: structure and role in peptidoglycan biosynthesis. FEMS Microbiol Rev 32:234–58. [DOI] [PubMed] [Google Scholar]
  • 87.Meeske AJ, Riley EP, Robins WP, Uehara T, Mekalanos JJ, Kahne D, Walker S, Kruse AC, Bernhardt TG, Rudner DZ. 2016. SEDS proteins are a widespread family of bacterial cell wall polymerases. Nature 537:634–638. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Sjodt M, Rohs PDA, Gilman MSA, Erlandson SC, Zheng S, Green AG, Brock KP, Taguchi A, Kahne D, Walker S, Marks DS, Rudner DZ, Bernhardt TG, Kruse AC. 2020. Structural coordination of polymerization and crosslinking by a SEDS-bPBP peptidoglycan synthase complex. Nat Microbiol 5:813–820. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Taguchi A, Welsh MA, Marmont LS, Lee W, Sjodt M, Kruse AC, Kahne D, Bernhardt TG, Walker S. 2019. FtsW is a peptidoglycan polymerase that is functional only in complex with its cognate penicillin-binding protein. Nat Microbiol 4:587–594. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Shrestha S, Dressler JM, McNellis ME, Shen A. 2024. Penicillin-binding proteins exhibit catalytic redundancy during asymmetric cell division in Clostridioides difficile. bioRxiv doi: 10.1101/2024.09.26.615255. [DOI] [Google Scholar]
  • 91.Shrestha S, Taib N, Gribaldo S, Shen A. 2023. Diversification of division mechanisms in endospore-forming bacteria revealed by analyses of peptidoglycan synthesis in Clostridioides difficile. Nat Commun 14:7975. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Bollinger KW, Müh U, Ocius KL, Apostolos AJ, Pires MM, Helm RF, Popham DL, Weiss DS, Ellermeier CD. 2024. Identification of a family of peptidoglycan transpeptidases reveals that Clostridioides difficile requires noncanonical cross-links for viability. Proc Natl Acad Sci U S A 121:e2408540121. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Takada H, Yoshikawa H. 2018. Essentiality and function of WalK/WalR two-component system: the past, present, and future of research. Biosci Biotechnol Biochem 82:741–751. [DOI] [PubMed] [Google Scholar]
  • 94.Bouillaut L, Newton W, Sonenshein AL, Belitsky BR. 2019. DdlR, an essential transcriptional regulator of peptidoglycan biosynthesis in Clostridioides difficile. Mol Microbiol 112:1453–1470. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Gueiros-Filho FJ, Losick R. 2002. A widely conserved bacterial cell division protein that promotes assembly of the tubulin-like protein FtsZ. Genes Dev 16:2544–56. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Hamoen LW, Meile JC, de Jong W, Noirot P, Errington J. 2006. SepF, a novel FtsZ-interacting protein required for a late step in cell division. Mol Microbiol 59:989–99. [DOI] [PubMed] [Google Scholar]
  • 97.Ishikawa S, Kawai Y, Hiramatsu K, Kuwano M, Ogasawara N. 2006. A new FtsZ-interacting protein, YlmF, complements the activity of FtsA during progression of cell division in Bacillus subtilis. Mol Microbiol 60:1364–80. [DOI] [PubMed] [Google Scholar]
  • 98.White ML, Eswara PJ. 2021. ylm Has More than a (Z Anchor) Ring to It! J Bacteriol 203. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Lee S, Price CW. 1993. The minCD locus of Bacillus subtilis lacks the minE determinant that provides topological specificity to cell division. Mol Microbiol 7:601–10. [DOI] [PubMed] [Google Scholar]
  • 100.Briley K Jr., Prepiak P, Dias MJ, Hahn J, Dubnau D. 2011. Maf acts downstream of ComGA to arrest cell division in competent cells of B. subtilis. Mol Microbiol 81:23–39. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Butler YX, Abhayawardhane Y, Stewart GC. 1993. Amplification of the Bacillus subtilis maf gene results in arrested septum formation. J Bacteriol 175:3139–45. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Jin J, Wu R, Zhu J, Yang S, Lei Z, Wang N, Singh VK, Zheng J, Jia Z. 2015. Insights into the cellular function of YhdE, a nucleotide pyrophosphatase from Escherichia coli. PLoS One 10:e0117823. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Bosch B, DeJesus MA, Poulton NC, Zhang W, Engelhart CA, Zaveri A, Lavalette S, Ruecker N, Trujillo C, Wallach JB, Li S, Ehrt S, Chait BT, Schnappinger D, Rock JM. 2021. Genome-wide gene expression tuning reveals diverse vulnerabilities of M. tuberculosis. Cell 184:4579–4592 e24. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.DeJesus MA, Gerrick ER, Xu W, Park SW, Long JE, Boutte CC, Rubin EJ, Schnappinger D, Ehrt S, Fortune SM, Sassetti CM, Ioerger TR. 2017. Comprehensive Essentiality Analysis of the Mycobacterium tuberculosis Genome via Saturating Transposon Mutagenesis. mBio 8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Bush MJ, Bibb MJ, Chandra G, Findlay KC, Buttner MJ. 2013. Genes required for aerial growth, cell division, and chromosome segregation are targets of WhiA before sporulation in Streptomyces venezuelae. mBio 4:e00684–13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Bohorquez LC, de Sousa J, Garcia-Garcia T, Dugar G, Wang B, Jonker MJ, Noirot-Gros MF, Lalk M, Hamoen LW. 2023. Metabolic and chromosomal changes in a Bacillus subtilis whiA mutant. Microbiol Spectr 11:e0179523. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Surdova K, Gamba P, Claessen D, Siersma T, Jonker MJ, Errington J, Hamoen LW. 2013. The conserved DNA-binding protein WhiA is involved in cell division in Bacillus subtilis. J Bacteriol 195:5450–60. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Weiss DS, Chen JC, Ghigo JM, Boyd D, Beckwith J. 1999. Localization of FtsI (PBP3) to the septal ring requires its membrane anchor, the Z ring, FtsA, FtsQ, and FtsL. J Bacteriol 181:508–20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Schmidt KL, Peterson ND, Kustusch RJ, Wissel MC, Graham B, Phillips GJ, Weiss DS. 2004. A predicted ABC transporter, FtsEX, is needed for cell division in Escherichia coli. J Bacteriol 186:785–93. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Arends SJ, Williams K, Scott RJ, Rolong S, Popham DL, Weiss DS. 2010. Discovery and characterization of three new Escherichia coli septal ring proteins that contain a SPOR domain: DamX, DedD, and RlpA. J Bacteriol 192:242–55. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Yahashiri A, Babor JT, Anwar AL, Bezy RP, Piette EW, Arends SJR, Muh U, Steffen MR, Cline JM, Stanek DN, Lister SD, Swanson SM, Weiss DS. 2020. DrpB (YedR) Is a Nonessential Cell Division Protein in Escherichia coli. J Bacteriol 202. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Ransom EM, Williams KB, Weiss DS, Ellermeier CD. 2014. Identification and characterization of a gene cluster required for proper rod shape, cell division, and pathogenesis in Clostridium difficile. J Bacteriol 196:2290–300. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Gitai Z, Dye N, Shapiro L. 2004. An actin-like gene can determine cell polarity in bacteria. Proc Natl Acad Sci U S A 101:8643–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Fenton AK, Gerdes K. 2013. Direct interaction of FtsZ and MreB is required for septum synthesis and cell division in Escherichia coli. EMBO J 32:1953–65. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Velazquez-Suarez C, Valladares A, Luque I, Herrero A. 2022. The Role of Mre Factors and Cell Division in Peptidoglycan Growth in the Multicellular Cyanobacterium Anabaena. mBio 13:e0116522. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Land AD, Winkler ME. 2011. The requirement for pneumococcal MreC and MreD is relieved by inactivation of the gene encoding PBP1a. J Bacteriol 193:4166–79. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Tavares AC, Fernandes PB, Carballido-Lopez R, Pinho MG. 2015. MreC and MreD Proteins Are Not Required for Growth of Staphylococcus aureus. PLoS One 10:e0140523. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Buckley AM, Jukes C, Candlish D, Irvine JJ, Spencer J, Fagan RP, Roe AJ, Christie JM, Fairweather NF, Douce GR. 2016. Lighting Up Clostridium Difficile: Reporting Gene Expression Using Fluorescent Lov Domains. Sci Rep 6:23463. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Megrian D, Taib N, Jaffe AL, Banfield JF, Gribaldo S. 2022. Ancient origin and constrained evolution of the division and cell wall gene cluster in Bacteria. Nat Microbiol 7:2114–2127. [DOI] [PubMed] [Google Scholar]
  • 120.Ransom EM, Ellermeier CD, Weiss DS. 2015. Use of mCherry Red fluorescent protein for studies of protein localization and gene expression in Clostridium difficile. Appl Environ Microbiol 81:1652–60. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Neumann-Schaal M, Jahn D, Schmidt-Hohagen K. 2019. Metabolism the Difficile Way: The Key to the Success of the Pathogen Clostridioides difficile. Front Microbiol 10:219. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Stickland LH. 1934. Studies in the metabolism of the strict anaerobes (genus Clostridium): The chemical reactions by which Cl. sporogenes obtains its energy. Biochem J 28:1746–59. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Johnstone MA, Self WT. 2022. d-Proline Reductase Underlies Proline-Dependent Growth of Clostridioides difficile. J Bacteriol 204:e0022922. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Buckel W, Thauer RK. 2018. Flavin-Based Electron Bifurcation, A New Mechanism of Biological Energy Coupling. Chem Rev 118:3862–3886. [DOI] [PubMed] [Google Scholar]
  • 125.Muller V, Chowdhury NP, Basen M. 2018. Electron Bifurcation: A Long-Hidden Energy-Coupling Mechanism. Annu Rev Microbiol 72:331–353. [DOI] [PubMed] [Google Scholar]
  • 126.Dineen SS, McBride SM, Sonenshein AL. 2010. Integration of metabolism and virulence by Clostridium difficile CodY. J Bacteriol 192:5350–62. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Dineen SS, Villapakkam AC, Nordman JT, Sonenshein AL. 2007. Repression of Clostridium difficile toxin gene expression by CodY. Mol Microbiol 66:206–19. [DOI] [PubMed] [Google Scholar]
  • 128.Daou N, Wang Y, Levdikov VM, Nandakumar M, Livny J, Bouillaut L, Blagova E, Zhang K, Belitsky BR, Rhee K, Wilkinson AJ, Sun X, Sonenshein AL. 2019. Impact of CodY protein on metabolism, sporulation and virulence in Clostridioides difficile ribotype 027. PLoS One 14:e0206896. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Hartig E, Hartmann A, Schatzle M, Albertini AM, Jahn D. 2006. The Bacillus subtilis nrdEF genes, encoding a class Ib ribonucleotide reductase, are essential for aerobic and anaerobic growth. Appl Environ Microbiol 72:5260–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Poudel A, Pokhrel A, Oludiran A, Coronado EJ, Alleyne K, Gilfus MM, Gurung RK, Adhikari SB, Purcell EB. 2022. Unique Features of Alarmone Metabolism in Clostridioides difficile. J Bacteriol 204:e0057521. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Pokhrel A, Poudel A, Castro KB, Celestine MJ, Oludiran A, Rinehold AJ, Resek AM, Mhanna MA, Purcell EB. 2020. The (p)ppGpp Synthetase RSH Mediates Stationary-Phase Onset and Antibiotic Stress Survival in Clostridioides difficile. J Bacteriol 202. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Oberkampf M, Hamiot A, Altamirano-Silva P, Belles-Sancho P, Tremblay YDN, DiBenedetto N, Seifert R, Soutourina O, Bry L, Dupuy B, Peltier J. 2022. c-di-AMP signaling is required for bile salt resistance, osmotolerance, and long-term host colonization by Clostridioides difficile. Sci Signal 15:eabn8171. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Brouwer MS, Roberts AP, Mullany P, Allan E. 2012. In silico analysis of sequenced strains of Clostridium difficile reveals a related set of conjugative transposons carrying a variety of accessory genes. Mob Genet Elements 2:8–12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Holtmann G, Bakker EP, Uozumi N, Bremer E. 2003. KtrAB and KtrCD: two K+ uptake systems in Bacillus subtilis and their role in adaptation to hypertonicity. J Bacteriol 185:1289–98. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Edwards AN, Wetzel D, DiCandia MA, McBride SM. 2022. Three Orphan Histidine Kinases Inhibit Clostridioides difficile Sporulation. J Bacteriol 204:e0010622. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Childress KO, Edwards AN, Nawrocki KL, Anderson SE, Woods EC, McBride SM. 2016. The Phosphotransfer Protein CD1492 Represses Sporulation Initiation in Clostridium difficile. Infect Immun 84:3434–3444. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Ng YK, Ehsaan M, Philip S, Collery MM, Janoir C, Collignon A, Cartman ST, Minton NP. 2013. Expanding the repertoire of gene tools for precise manipulation of the Clostridium difficile genome: allelic exchange using pyrE alleles. PLoS One 8:e56051. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Li S, Poulton NC, Chang JS, Azadian ZA, DeJesus MA, Ruecker N, Zimmerman MD, Eckartt KA, Bosch B, Engelhart CA, Sullivan DF, Gengenbacher M, Dartois VA, Schnappinger D, Rock JM. 2022. CRISPRi chemical genetics and comparative genomics identify genes mediating drug potency in Mycobacterium tuberculosis. Nat Microbiol 7:766–779. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Stabler RA, He M, Dawson L, Martin M, Valiente E, Corton C, Lawley TD, Sebaihia M, Quail MA, Rose G, Gerding DN, Gibert M, Popoff MR, Parkhill J, Dougan G, Wren BW. 2009. Comparative genome and phenotypic analysis of Clostridium difficile 027 strains provides insight into the evolution of a hypervirulent bacterium. Genome Biol 10:R102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Ransom EM, Weiss DS, Ellermeier CD. 2016. Use of mCherryOpt Fluorescent Protein in Clostridium difficile. Methods Mol Biol 1476:53–67. [DOI] [PubMed] [Google Scholar]
  • 141.Ducret A, Quardokus EM, Brun YV. 2016. MicrobeJ, a tool for high throughput bacterial cell detection and quantitative analysis. Nat Microbiol 1:16077. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Karash S, Jiang T, Samarth D, Chandrashekar R, Kwon YM. 2019. Preparation of Transposon Library and Tn-Seq Amplicon Library for Salmonella Typhimurium. Methods Mol Biol 2016:3–15. [DOI] [PubMed] [Google Scholar]
  • 143.Bolger AM, Lohse M, Usadel B. 2014. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics 30:2114–20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Blankenberg D, Gordon A, Von Kuster G, Coraor N, Taylor J, Nekrutenko A, Galaxy T. 2010. Manipulation of FASTQ data with Galaxy. Bioinformatics 26:1783–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Ioerger TR. 2022. Analysis of Gene Essentiality from TnSeq Data Using Transit. Methods Mol Biol 2377:391–421. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.DeJesus MA, Zhang YJ, Sassetti CM, Rubin EJ, Sacchettini JC, Ioerger TR. 2013. Bayesian analysis of gene essentiality based on sequencing of transposon insertion libraries. Bioinformatics 29:695–703. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplement 1
media-1.pdf (614.2KB, pdf)
Supplement 2
media-2.xlsx (36.9KB, xlsx)
Supplement 3
media-3.xlsx (36.4KB, xlsx)
Supplement 4
media-4.xlsx (423KB, xlsx)
Supplement 5
media-5.pdf (201.1KB, pdf)

Articles from bioRxiv are provided here courtesy of Cold Spring Harbor Laboratory Preprints

RESOURCES