Figures
Abstract
The Gram-negative pathogen, Acinetobacter baumannii, poses a serious threat to human health due to its role in nosocomial infections that are resistant to treatment with current antibiotics. Despite this, our understanding of fundamental A. baumannii biology remains limited, as many essential genes have not been experimentally characterized. These essential genes are critical for bacterial survival and, thus, represent promising targets for drug discovery. Here, we systematically probe the function of essential genes by screening a CRISPR interference knockdown library against a diverse panel of chemical inhibitors, including antibiotics. We find that most essential genes show chemical-gene interactions, allowing insights into both inhibitor and gene function. For instance, knockdown of lipooligosaccharide (LOS) transport genes increased sensitivity to a broad range of chemicals. Cells with defective LOS transport showed cell envelope hyper-permeability that was dependent on continued LOS synthesis. Using phenotypes across our chemical-gene interaction dataset, we constructed an essential gene network linking poorly understood genes to well-characterized genes in cell division and other processes. Finally, our phenotype-structure analysis identified structurally related antibiotics with distinct cellular impacts and suggested potential targets for underexplored inhibitors. This study advances our understanding of essential gene and inhibitor function, providing a valuable resource for mechanistic studies, therapeutic strategies, and future key targets for antibiotic development.
Author summary
Acinetobacter baumannii is a bacterium that can cause multidrug-resistant infections, particularly in hospitals and in vulnerable patients. However, we still don’t fully understand many of the genes that help it survive and grow. In this study, we used a genetic tool called CRISPR interference to study genes essential for A. baumannii survival and how they interact with different antibiotics and other chemicals. We found that most of these genes affect how the bacteria respond to antibiotic exposure, giving us new insights into both bacterial survival and how these antibiotics work against A. baumannii. Additionally, we identified links between poorly understood genes and well-known processes like cell division. Finally, we developed and utilized a method to examine how the structure of antibiotics influences their effects on bacterial cells. Taken together, this study advances our understanding of A. baumannii biology and paves the way for further genetic studies, new treatment strategies, and antibiotic development.
Citation: Tran JS, Ward RD, Iruegas-López R, Ebersberger I, Peters JM (2025) Chemical genomics informs antibiotic and essential gene function in Acinetobacter baumannii . PLoS Genet 21(3): e1011642. https://doi.org/10.1371/journal.pgen.1011642
Editor: Jan-Willem Veening,, University of Lausanne: Universite de Lausanne, SWITZERLAND
Received: December 31, 2024; Accepted: March 3, 2025; Published: March 28, 2025
Copyright: © 2025 Tran et al. This is an open access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.
Data Availability: The datasets and computer code produced in this study are available in the following databases: • Raw sequencing data: NCBI BioProject (PRJNA1190454) • Scripts: Github (https://github.com/jasonpeterslab)
Funding: o This work (JMP lab) is supported by the National Institutes of Health (https://www.nih.gov/grants-funding) under award numbers K22AI137122 and R35GM150487. JST is supported by the Biotechnology Training Program (NIH 5T32GM135066; https://btp.wisc.edu/), SciMED Graduate Research Scholars (https://scimedgrs.wisc.edu/), and a National Science Foundation GRFP (https://www.nsfgrfp.org/). RDW is supported by the Predoctoral Training Program in Genetics (NIH 5T32GM007133) and SciMED Graduate Research Scholars. The work in IE lab was supported by a grant by the German Research Foundation (DFG; https://www.dfg.de/) in the scope of the Research Group FOR2251 “Adaptation and persistence of A. baumannii.” Grant ID EB-947 285-2/2 to IE. IE was further supported by the Research Funding Program Landes-Offensive zur Entwicklung Wissenschaftlich-ökonomischer Exzellenz (LOEWE; https://wissenschaft.hessen.de/forschen/landesprogramm-loewe) of the State of Hessen, Research Center for Translational Biodiversity Genomics (TBG). The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.
Competing interests: I have read the journal's policy and the authors of this manuscript have the following competing interests: JMP has filed for patents related to Mobile-CRISPRi technology and bacterial promoters.
Introduction
Systems biology provides a robust framework for deciphering the complex networks that govern cellular functions. This approach is especially pertinent in the study of infectious diseases and antibiotic resistance as it offers sophisticated tools to analyze the multifaceted interactions between pathogens and their environments, including their responses to therapeutic interventions. The Gram-negative, hospital-acquired pathogen Acinetobacter baumannii is categorized as an ‘urgent threat’ due to certain clinical strains having developed resistance to all known therapeutics and its ability to persist on surfaces that are typically adverse to cellular life, such as stainless steel [1,2]. Despite the critical dangers posed by A. baumannii, our understanding of how the fundamental elements of its biology interact with antibiotics or other inhibitors remains limited. Essential genes, which are vital for survival even under optimal growth conditions, are promising targets for drug discovery. Essential genes in A. baumannii have been cataloged using methods like transposon sequencing (Tn-seq), which identifies genes with low or nonexistent insertion frequencies as critical for survival [3,4]. However, lethality caused by knocking out essential genes limits more detailed studies of gene function and interactions. CRISPR interference (CRISPRi), which allows for the knockdown of gene expression without eliminating gene function [5,6], provides a solution for studying the function of essential genes [7]. CRISPRi uses a deactivated Cas9 protein, dCas9, directed by a single guide RNA (sgRNA) to specifically target genes for knockdown. This method has been used successfully in diverse bacteria including A. baumannii, non-baumannii Acinetobacter species, and other Gram-negative pathogens [8–13].
Chemical genomics, which combines gene perturbation libraries with chemical treatments, offers a strategy to explore gene function as changes in relative fitness of mutant or knockdown cells in response to chemical stresses [14,15]. This approach has been applied at the genome-scale in bacteria, facilitating the observation of subtle phenotypic responses and the identification of genetic connections from Tn-seq or deletion libraries [3,16,17] as well as in CRISPRi libraries [18–21]. Our previous work involved developing and characterizing a CRISPRi library in A. baumannii, which identified key vulnerabilities among essential genes and demonstrated how this library could be utilized to elucidate antibiotic-gene interactions [20].
In this study, we took a chemical-genomics approach that provided systems-level insights into gene and antibiotic function in A. baumannii. We screened our CRISPRi library against a set of 45 chemical stressors, allowing us to identify pathways that cause broad sensitivity to these agents when perturbed. Additionally, we created an essential gene network that informed the function of poorly characterized genes that are unique to or highly divergent in A. baumannii. Finally, we integrated phenotypic and chemoinformatic datasets to identify possible target pathways for inhibitors and show distinctions in the physiological impacts of structurally similar drugs. In doing so, this work advances our understanding of the cellular contributions of A. baumannii essential genes in the context of antibiotic stress.
Results
The vast majority of A. baumannii essential genes show significant chemical-gene interactions
To identify phenotypes for essential gene knockdown strains in the context of chemical treatment (i.e., chemical-gene interactions) in A. baumannii, we conducted a large-scale chemical genomics screen using our established inducible CRISPRi library in A. baumannii strain ATCC19606 (19606 throughout) [20]. This library consists of pooled CRISPRi strains with guides targeting 406 putatively essential genes derived from Tn-seq data in other strains [3,4], along with 1000 non-targeting control sgRNAs. For each targeted gene, we cloned four sgRNA spacers that exactly match the target sequence (perfect-match) and ten mismatch spacers with single-base variations from the target sequence that allow for titration of knockdown [22]. We performed competition fitness assays by inducing CRISPRi knockdown and adding chemicals at sublethal concentrations. These chemical concentrations ensured sufficient cell viability to preserve library diversity and representation, allowing us to measure phenotypic responses to treatment (Fig 1A). We screened the library against a diverse collection of chemical compounds, including clinically relevant antibiotics, heavy metals, and compounds with unknown mechanisms of action (Fig 1B). Knockdown strain abundance was analyzed by amplification and sequencing of the sgRNA spacer regions. We set quality control metrics—replicate correlation and sample complexity cutoffs—to remove samples with excessive noise or selective pressure (e.g., lethal chemical concentrations) that would skew calculations (S1A and S1B Fig; see Material and Methods for details). We verified that chemical solvent impacts were negligible as library composition for DMSO-only and mock treatment samples were highly correlated (r > 0.95) (S1C Fig).
(A) Schematic depicting experimental setup for chemical genomics. (B) Proportional area chart of reported cellular targets of screen chemicals. (C) Histogram showing significant chemical-gene scores (|median log2 fold change| ≥ 1 and Stouffer’s p < 0.05, calculated from perfect-match guides), across chemical conditions in our screen. Darker bars to the right of the dotted line represent at least one significant chemical interaction in this screen. (D) Stacked bar chart showing the sum of significant positive (pink) and negative (blue) CG scores across screen chemicals for each gene in the library.
To determine chemical-gene phenotypes from the pooled screens, we calculated chemical-gene (CG) scores for each gene as the median log2 fold change (medL2FC) of perfect guides with chemical treatment compared to induction alone. Because our A. baumannii CRISPRi system achieves only ~20-fold knockdown after 14 generations [20], perfect match guides cause only a partial knockdown. Positive CG scores indicate gene knockdowns with improved growth under chemical conditions relative to the untreated knockdown, whereas negative CG scores indicate chemical sensitivity or reduced growth. 93% (378/406) of the genes we investigated exhibited at least one significant CG score (medL2FC ≥ |1|, p <0.05) upon knockdown, with a median of 14 significant chemical interactions per gene (Fig 1C). Most of these scores were skewed toward reduced growth, with ~73% (3895/5345) of significant CG scores being negative (Fig 1D). This is consistent with previous studies suggesting essential genes are more frequently connected by negative interactions, where two perturbations (e.g., a chemical stressor and a knockdown) cause further loss of fitness than expected based the effects of the individual perturbations [23]. Overall, this dataset provides thousands of unique chemical-gene phenotypes and serves as a detailed resource for A. baumannii essential gene responses to chemical stress.
Lipooligosaccharide transport inhibition enhances drug susceptibility through increased membrane permeability
To identify A. baumannii pathways crucial for chemical resistance, we ranked genes based on significant negative CG scores (medL2FC ≤-1, p <0.05) across conditions and performed functional enrichment using the STRING database [24]. We found that knockdowns of the lipooligosaccharide transport (Lpt) system were significantly enriched (GO:0015920, enrichment score = 8.95, FDR = 1.04e-05), with negative CG scores for 70% of conditions in our screen (Fig 2B and 2C). Lpt is a highly conserved pathway required for lipopolysaccharide (LPS)/lipooligosaccharide (LOS) export to the outer leaflet of the outer membrane in Gram-negative bacteria. A. baumannii uses Lpt to transport LOS, which is comprised of lipid A and core oligosaccharide moieties found in LPS but lacks O-polysaccharide repeats (Fig 2A) [25]. Intriguingly, while LPS/LOS is essential in most Gram-negative bacteria, many strains of A. baumannii do not require LOS for survival, including the strain used for our library (19606) [26]. Indeed, knockdowns of lpx genes responsible for LOS synthesis showed fewer phenotypes than knockdowns of lpt needed for LOS transport (S2A Fig), suggesting that LOS transport unexpectedly plays a more crucial role in A. baumannii’s chemical sensitivity than LOS synthesis itself.
(A) Graphical depiction of Lpt machinery and LOS trafficking to the outer membrane in A. baumannii. (B) Histogram of significant negative chemical interactions (medL2FC ≤ -1 and Stouffer’s p < 0.05) across conditions. Bold line indicates location of lpt genes. (C) Heatmap displaying CG scores for lpt knockdown cluster. Y-axis clustering was conducted using the Ward method and Canberra distance across all library knockdowns. (D) Tigecycline and azithromycin MIC test strip assays for the lptA knockdown, lpxC knockdown, and control strain containing a nontargeting guide. Plates supplemented with 1mM IPTG; approximate MICs reported below images (μg/mL). (E) Ethidium bromide permeability assay for lptA knockdown and nontargeting control in 19606 or LOS- backgrounds after induction; increased fluorescence over time indicates membrane permeability. Ribbons represent standard deviation (n=4).
We validated our screen results for individual knockdowns outside of the pooled context by performing quantitative minimum inhibitory concentration (MIC) assays using antibiotic strips. This revealed that an lptA knockdown was sensitized to several clinically relevant antibiotics (Fig 2D). In contrast, an lpxC knockdown was not sensitized (Figs 2D and S2B) although loss of LOS has been shown to increase antibiotic sensitivity [27]. Both the lpxC and lptA knockdowns remained sensitized to colistin (S2B Fig)—a cationic lipopeptide that disrupts membranes by targeting the lipid A moiety of LOS [28] —suggesting that these knockdowns do not completely eliminate LOS production or transport and knockdown strains still maintain some level of LOS in the outer membrane that can be targeted by colistin [29]. We therefore introduced CRISPRi either targeting lptA or containing a non-targeting control sgRNA into a mutant strain background that does not produce LOS (i.e., LOS-), 19606 lpxC(S106R) [27]. In contrast to the 19606 lptA and lpxC knockdowns, LOS- strains were fully resistant to colistin (S2B Fig). Additionally, the LOS- strains showed considerable growth defects not exhibited by the 19606 knockdown strains (S2D Fig). Yet, the lptA knockdown strain was as similarly sensitized to levofloxacin as the LOS- strains (S2B Fig). This suggested levofloxacin, an antibiotic targeting gyrase/DNA synthesis in the cell cytoplasm, could penetrate the cell envelope more effectively in both the lptA and the LOS- strains.
These findings along with the sensitivity of lpt knockdowns to chemically diverse inhibitors pointed to a permeability defect. Therefore, we performed an ethidium bromide (EtBr) fluorescence assay to test for the knockdown’s effect on membrane permeability; in this assay, cells are briefly exposed to EtBr, which fluoresces after passing through the outer and inner membranes and binding to DNA. More permeable membranes will have higher fluorescence readings, and as expected, the LOS- strains showed increased permeability compared to the 19606 nontargeting control. Surprisingly though, the 19606 lptA knockdown was substantially more permeable than the LOS- and LOS- lptA knockdown strains (Figs 2E and S2E). This indicates that Lpt deficiencies cause a further increase in membrane permeability when LOS is being synthesized, possibly due to LOS accumulating in and destabilizing the inner membrane. Taken together, these findings further underscore the critical role of the LOS transport system in A. baumannii in maintaining membrane integrity and mediating resistance to chemicals.
Taxon-specific determinants of chemical susceptibility and resistance
Essential genes that are highly divergent in, or exclusive to, A. baumannii may contribute to its unique drug resistance and could help guide the development of targeted therapeutics. To identify these genes, we determined the phylogenetic profiles of the genes represented in our library across a broad collection species covering the known diversity of Gammaproteobacteria [30]. We identified 15 candidate genes that were highly responsive to chemicals when knocked down but were rare in Gammaproteobacteria outside of the order to which A. baumannii belongs (Figs 3A and S4). These genes exhibited significant CG scores (medL2FC ≥ |1|, p <0.05) for at least 10% of chemicals tested and likely play substantial roles in responding to chemical stressors.
(A) Dot plot depicting library chemical-gene scores and ortholog presence ratio (fraction of isolates possessing at least one ortholog out of the total number of analyzed isolates) across (A) representative non-Moraxellales Gammaproteobacteria or (B) non-A. baumannii Acinetobacter species. Dots in blue represent genes with <38% presence across representative groups and significant chemical-gene interactions in >10% of screen conditions. (C) Heatmap depicting CG scores for knockdowns of hits for genes rare outside A. baumannii.
Among others, this revealed that orthologs of A. baumannii lptC, which is involved in lipooligosaccharide transport, are considerably rare among Gammaproteobacteria (Figs 3A and S5). This suggests A. baumannii lptC is more divergent, in contrast to other lpt genes. Additional underexplored cell envelope genes—GO593_05760 (advA), GO593_07665 (blhA), and GO593_11915 (smc2), GO593_03530 (gtrOC1), GO593_06865 (wzx),—suggest distinct and under-characterized components of the A. baumannii cell envelope may play key roles in chemical response.
We then increased our resolution to pinpoint genes specific to A. baumannii compared to other Acinetobacter species. Three genes stood out as more responsive to chemicals upon knockdown: gtrOC1, wzx, and GO593_15125 (Fig 3B). Both gtrOC1 and wzx are involved in cell surface processes; wzx encodes a predicted polysaccharide flippase within a capsule biosynthesis locus [31], while gtrOC1 is part of a locus involved in LOS synthesis [32]. Orthologs to these genes are only sporadically found in Gammaproteobacteria outside A. baumannii (Figs 3A,3B,S3, and S4). For wzx, one explanation for this patchy occurrence of orthologs is the diversity of polysaccharides that the encoded protein transports [33,34]. Similarly, gtrOC1 has been implicated in a novel mechanism for core oligosaccharide assembly in A. baumannii [32]. Hierarchical clustering of gene knockdowns based on CG score correlations using Ward’s method confirmed association of gtrOC1 with LOS synthesis genes but revealed an unexpected clustering of wzx with thiD, a thiamin synthesis gene, for unknown reasons (Fig 3C).
GO593_15125 shows a similar ortholog distribution (Figs 3A, 3B, and S4) but encodes for an Arc-family repressor. In our previous work, we showed knockdown of a different Arc-family repressor caused significant fitness loss, likely due to its role in repressing toxic prophage genes [20]. GO593_15125 could play a similar role in regulating mobile genetic elements (MGEs) as it is linked to a predicted transposase and an antiphage defense island [35,36]). Additionally, the knockdown is most sensitive to levofloxacin, a DNA-damaging antibiotic (Fig 3C), and MGEs often modulate expression in response to DNA damage [37,38] and occasionally trigger DNA damage responses themselves [39]. Furthermore, this gene clusters with nusG, encoding a transcription elongation/termination factor (Fig 3C). NusG in E. coli induces the DNA damage response and rac prophage excision when depleted [40] and the essentiality of nusG is tied to suppression of toxic genes in rac [41], suggesting that GO593_15125 and nusG are both involved in prophage gene repression in A. baumannii.
These findings identify key chemically responsive genes in A. baumannii, which are less common in other Gammaproteobacteria. Their potential unique roles in bacterial survival provide a strong foundation for further investigation into their contribution to drug resistance.
An essential gene network in A. baumannii identifies novel gene connections
We next sought to identify further functional connections for these taxon-specific, chemically responsive genes. Using the CG scores, we calculated all pairwise correlations between genes then defined a significance threshold by generating an empirical null distribution through random permutation of the underlying data [7,16]. We used only correlations passing the significance threshold to construct a high-confidence essential gene network. Most connections in this network were between genes in different operons (Fig 4A), suggesting true functional relationships dominated over artifacts of CRISPRi polarity. As expected, genes in well-conserved pathways like translation and cell division showed numerous connections (Fig 4A). However, oxidative phosphorylation complexes were disconnected (Figs 4A and S6A), suggesting distinct functionalities outside of the electron transport chain (e.g., succinate dehydrogenase also functions in the TCA cycle).
Essential gene network in A. baumannii. Nodes (genes) are connected by edges (phenotypic correlations, r>0.76). Solid lines represent genes across different operons; dotted lines represent genes within the same operon. Subnetworks depict (B) DNA polymerase and replication-related genes, (C) signal-recognition, and (D) cell division and cell wall synthesis. (E) Quantification of cell length of knockdowns from microscopy of cell wall gene knockdowns (N=200-400). (F) Representative microscopy images. Ruler is 5 microns.
Connections in this network between poorly understood, taxon-specific genes and well-characterized, conserved genes point to shared functions. Lpt genes formed a subnetwork with gtrOC1, supporting the gene’s role in LOS synthesis (S6B Fig). In comparison, wzx was located in a subnetwork with genes involved in thiamin synthesis, ferredoxins, iron-sulfur cluster biogenesis, isoprenoid synthesis, cytochrome bo3 oxidase, and another monovalent cation/H+ antiporter, hinting at broader connections between aerobic metabolism and isoprenoid lipid-linked glycans operated on by Wzx [42] (S6A Fig).
One of the candidate genes restricted to Moraxellales, GO593_00140, was linked with multiple DNA replication genes and an unannotated gene, GO593_11615 (Fig 4B). Knockdowns of GO593_00140 and GO593_11615 showed significant, negative CG scores in levofloxacin, which targets replication (S7A and S7B Fig). Despite poor conservation (Figs 2A and S4), structural prediction of GO593_00140 using Phyre2 showed a domain similar to that of E. coli HolD, the psi subunit of the bacterial DNA polymerase III holoenzyme (48% of the sequence modelled with 96.8% confidence despite only 12% sequence ID) (S7C Fig). A previous hidden Markov model (HMM)-based homolog search identified similar putative psi subunits in other Acinetobacter species [43], and our work provides phenotypic evidence to further support that this gene is involved in DNA replication in A. baumannii. GO593_11615 likely encodes a phage-associated transcriptional regulator, as the gene is situated within an intact prophage predicted by PHASTEST [44]. This protein has domain architecture similar to LexA, a conserved repressor of DNA repair pathways that senses DNA damage through an interaction with RecA-DNA filaments that triggers self-cleavage and subsequent de-repression of genes involved in the SOS response [45] (S7D Fig). Depletion of LexA is known to induce excision of some prophages and subsequent host lysis [37]. Interestingly, A. baumannii lacks lexA [46], suggesting this prophage encodes its own LexA-like protein as a host-independent strategy to detect DNA damage.
Other candidate genes, GO593_10640 and GO593_18720, grouped with the conserved signal recognition particle (SRP) genes ftsY and ffh as well as ybgC, a Tol-Pal-associated thioesterase linked to membrane integrity [47] (Fig 4C). GO593_10640 shows structural similarities to ion transporters and significant chemical responses consistent with ftsY and ffh knockdowns (S8A and S8B Fig). GO593_18720, however, showed CG scores that were smaller in magnitude, suggesting a less critical or defined role (S8A Fig). These genes may be directly or indirectly related to membrane protein localization and membrane integrity, though SRP mechanisms remain understudied in A. baumannii.
Our network analysis revealed a cell wall and division subnetwork (Fig 4D), containing three genes we identified as rare in other Gammaproteobacteria: GO593_05760 (advA), GO593_07665 (blhA), and GO593_11915 (smc2, or SMC_prok_B). These genes correlate with numerous well-characterized, highly conserved division genes, corroborating initial data from other studies suggesting roles for advA, blhA, and smc2 in cell division [3,48]. This subnetwork also featured another gene of unknown function, GO593_07290, which encodes for a SPOR domain-containing protein, annotated here as spor. While SPOR domains are widely conserved and are known to interact with cell wall peptidoglycan (PG) [49], functions of SPOR domain-containing proteins vary considerably. Indeed, one of the most extensively characterized SPOR domain proteins, FtsN from E. coli, does not require the SPOR domain for its essential function [50]. The short periplasmic region required for FtsN function, EFtsN [51], does not appear to be present at the amino acid level in GO593_07290, suggesting a functionally distinct role in cell division (S7E Fig).
To further support these genes’ roles in cell division, we generated individual CRISPRi knockdowns for blhA, smc2, and spor and examined cell shape with microscopy. Knockdown of advA has previously been shown to cause filamentation, a hallmark of impaired division [12]. Similarly, blhA, smc2, and spor knockdowns produced elongated or chained cells (Fig 4E and 4F), corroborating prior findings from a blhA transposon mutant [48]. We next attempted to determine the localization pattern of putative division genes using trans expression of fluorescent protein fusions in A. baumannii strain ATCC17978, due to complications caused by phase switching in 19606 Similar to what had been previously shown with AdvA [3], a fusion of superfolder GFP to the N-terminus of BlhA (sfGFP::BlhA) showed localization to the septum, or the site of division, suggesting that BlhA is a bona-fide member of the divisome (S9A Fig). In contrast, the SPOR domain protein localized to both the cell periphery and septum (S9B Fig), consistent with PG binding, but suggests a more generalized cell wall synthesis function. Interestingly, expression of the Smc2::sfGFP hybrid resulted in cell elongation and division defects, with the fluorescently tagged Smc2 protein forming elongated aggregates in the cells (S9C Fig). The impaired division phenotypes were verified with overexpression of Smc2 alone, suggesting precise protein levels are needed for proper cell division (S9D Fig). This trans expression additionally allowed us to disentangle the role of smc2 from polar knockdown effects on downstream gene zapA, a known division gene. As an important caveat, we note that our fusion constructs may not be expressed at the same level as their native counterparts and, therefore, may show aberrant localization. Nonetheless, we contend that these localizations are at least possible. Overall, our network reveals functional connections and underlying phenotypes linking crucial pathways in A. baumannii, such as cell division, and provides a basis for further targeted experiments for mechanistic analysis.
Structure-informed clustering of chemical perturbations reveals insights into inhibitor function
We next aimed to develop an approach for characterizing inhibitor function in A. baumannii that leveraged our chemical genomics dataset. To broadly define inhibitor impacts on cell physiology, we analyzed chemical-gene interactions at the pathway level rather than the individual gene level. This approach evaluated the L2FCs of all guides, including mismatches, associated with a STRING functional pathway (“pathway effect”) [52]. Pathway level analysis provided increased statistical robustness (many more guides were measured at the pathway level than the gene level) at the cost of single gene resolution; this allowed us to detect many more significant pathways. We then clustered chemicals by pathway-level interactions, showing the relationship between chemicals solely based on the cellular pathways they impacted in our screen (Fig 5A). Inhibitors from the same antibiotic classes, such as quinolones (ciprofloxacin and levofloxacin) or aminoglycosides (amikacin, apramycin, and tobramycin), clustered as expected, validating our method.
(A) Dendrogram depicting hierarchical clustering (Ward method) of screen chemicals by similarities of effects across essential gene pathways. (B) Dendrogram displaying hierarchical clustering (Ward method) of chemical molecular fingerprints (ECFP6) by Tanimoto coefficient. Asterisks represent conditions with multiple chemical structures, in which a single structure was used (see S4 Table). (C) Sina plots depicting differential guide counts per million (CPMs) for protein translocation in carvacrol and thymol, or (D) for heme-copper terminal oxidase in tetracycline-class antibiotics, compared to corresponding no chemical controls. Chemical-gene interactions are described above each graph—up (positive), down (negative), or no change—with FDR values. Guides are weighted in this calculation by predicted efficacy, with perfect guides weighted at 100. Non-significant comparisons are shown with grey backgrounds.
To more systematically compare our results to inhibitors outside of established antibiotic classes, we performed molecular fingerprinting—a chemoinformatic approach that numerically represents unique chemical structure and properties of molecules as binary “fingerprints” (extended-connectivity fingerprints) [53,54]—and measured similarity between chemicals using the Tanimoto coefficient [55,56]. Clustering based on distances derived from Tanimoto coefficients verified that antibiotics with similar core structures (beta-lactams, fluoroquinolones, aminoglycosides, etc.) group together (Fig 5B). Comparing the pathway impact clustering with the structure clustering of these chemicals allowed us to investigate (a) consistent chemical classes with a not-yet-established mechanism of action, (b) structurally similar chemicals that have distinct effects on A. baumannii, and (c) structurally dissimilar chemicals that similarly affect A. baumannii physiology.
Carvacrol and thymol are chemicals that cluster tightly by both pathway effect and structure (Figs 5A,5B, and S10). These compounds exhibit antibacterial activity but lack a defined mechanism of action [57], though they have been proposed as antimicrobial agents or additives [58]. Our pathway analysis revealed that genes encoding NADH dehydrogenase (nuo genes) and protein transport complexes (ftsY, ffh, and sec genes) knockdowns were significantly sensitized while ubiquinone (ubi genes) knockdowns were more resistant (Figs 5C,S11A,S11B,S11C, and S11D). In particular, ftsY, encoding the SRP receptor protein, was a major negative outlier in both carvacrol and thymol (S11C and S11D Fig). These pathways influence membrane homeostasis and proton motive force (PMF), supporting previous studies stating that membrane disruption and depolarization underlie carvacrol and thymol antibacterial activity [59]. However, carvacrol and thymol cluster more distantly in their pathway effects from other known membrane disruptors, like chlorhexidine and EDTA [60,61] (Fig 5A). This suggests that the mechanisms of action for these phenolic monoterpenes are unique, potentially due to more specific effects on SRP-mediated protein transport.
Identification of structurally similar chemicals with differing effects on A. baumannii pathways could reveal how relatively small chemical modifications alter their physiological impact. To explore this, we scored how well chemical pairs clustered in pathway effect or in structure and calculated the differences (S12 Fig). We found that the pathways affected by the tetracycline class antibiotics minocycline, tetracycline, and tigecycline diverged despite their strong structural similarity (S13–S15 Figs). Tetracycline class antibiotics target the ribosome, but pathway effects and CG scores were similar for minocycline, tetracycline, and tigecycline treated cells (S14 Fig). The most distinct pathway effects across the tetracycline class occurred in genes encoding the cytochrome bo3 oxidase, as cyo knockdown strains were severely depleted in tigecycline, but not significantly affected in minocycline and tetracycline (Fig 5D). We observed more subtle, but statistically significant differences in LOS transport and ATP synthesis pathways (S13 Fig), with Lpt encoding genes depleted from cells treated with minocycline and tetracycline but not tigecycline and ATP synthase gene knockdowns enriched in tetracycline and tigecycline but not minocycline. Although the precise mechanisms underlying these interactions are unknown, small changes in structure may affect permeability; for instance, RND-type efflux pumps that are sensitive to the cellular energy state are able to efflux tigecycline and tetracycline but not minocycline [62]. Tigecycline has a unique structural modification (a glycylcycline side chain) that may impact permeability [63] or could contribute to its clustering with other chemicals that affect oxidative phosphorylation and oxidative stress (e.g., phenazine and myricetin [64–67]). Additional work is required to validate and investigate the biological underpinnings of our tetracycline class results.
An important potential caveat our structure-informed clustering approach is that non-bioequivalent doses across chemical inhibitors could drive differential clustering and obscure the relationship between biological targets and pathway effects. One such likely example of this phenomenon in our data set is the carbapenem class antibiotic meropenem, which fails to cluster with other carbapenems we screened (doripenem and imipenem; S15 Fig). CG score (S15A Fig) and pathway analyses (S15B Fig) show a muted effect for meropenem versus other carbapenems, even on the known target pathway. Despite our efforts to normalize growth inhibition across chemicals, the sigmoid nature of dose-response curves can make it challenging to achieve precise bioequivalence in the context of partial inhibition. In the case of meropenem, it is likely that we simply failed to add a sufficient concentration to achieve bioequivalent inhibition. However, we note that our screening quality control steps, such as monitoring the diversity of non-targeting control guides as internal controls (see methods), likely guard against scenarios in which an excessive inhibition has been applied.
Discussion
Essential cellular processes are the current and likely future targets of mono- and poly-therapies against antibiotic resistant pathogens. Despite the biological and clinical importance of essential genes, few studies have comprehensively characterized their functions, especially in Gram-negative pathogens. This work advances our understanding of gene and antibiotic function in A. baumannii through large-scale, chemical-genomic profiling of an essential gene knockdown library. The breadth and diversity of our chemical perturbations provided functional insights that could not be obtained with a smaller scope. We identified >5000 significant phenotypes for essential genes (Fig 1) and found that specific pathways, such as LOS transport, were highly enriched for sensitive phenotypes (Fig 2), representing compelling targets for combination therapies. Our evolutionary analysis revealed inhibitor-responsive essential genes that are unique to or are highly divergent in A. baumannii (Fig 3), providing possible targets for narrow-spectrum therapeutics. Our essential gene network showed robust connections between known and poorly characterized essential genes (Fig 4), hinting at new functions and corroborating prior work. Finally, our structure-function analysis of chemical-gene phenotypes uncovered a surprising discordance among the phenotypic impacts of antibiotics within the same class (Fig 5), highlighting how genes outside of the direct target can alter antibiotic efficacy. Moreover, this study provides a rich dataset of essential gene phenotypes to the A. baumannii and larger Gram-negative pathogen research communities, serving as a resource and framework for furthering our understanding of essential gene function.
Our findings linking Lpt function to antibiotic resistance accentuate the potential of targeting LOS trafficking for combination therapies, especially in the context of recent pharmaceutical advances.
Our observation that lpt knockdown strains show a hyper-permeable phenotype that is dependent on LOS production suggests that increased permeability is caused by both loss of LOS in the outer membrane and retention of LOS in the inner membrane, although additional work is needed to show this definitively. Other studies have suggested that inhibiting LPS transport in other Gram-negatives results in toxic accumulation of LOS intermediates that cause membrane defects and antibiotic susceptibility [68–70]; these hypotheses are not mutually exclusive with our own. Moreover, the increased permeability of lpt knockdowns is consistent with classic E. coli lptD (imp4213 [71]) mutants that are broadly permeable [72] and are often used in drug screening efforts [73]. Recent studies have characterized novel Lpt-targeting macrocyclic peptides [74,75] that selectively disrupt LOS transport in Acinetobacter species. Interestingly, our conservation analysis revealed that LptC is far more divergent among Gammaproteobacteria than other proteins in the Lpt complex, which may be related to the fact that these Lpt-targeting macrocyclic peptides are inactive against other Gram-negatives. Taken together, these data highlight the potential for recently described LOS transport inhibitors [70,74,75] to be effective in conjunction with other antibiotics as a combination therapy against A. baumannii infections.
Our work also raises intriguing questions regarding the functions of essential genes common in A. baumannii yet rare across other Gammaproteobacteria. Our essential gene network revealed functional connections between many of these poorly characterized and well-studied ones, generating hypotheses that can be validated in future mechanistic studies. For instance, GO593_10640 knockdown strains show strong phenotypic similarity to knockdowns of SRP components. Whether GO593_10640 is required for SRP activity or is a key SRP substrate involved in antibiotic resistance is unknown. GO593_10640 contains a predicted signal sequence, and, thus, is a potential SRP substrate. SRP is poorly characterized in A. baumannii, although recent work has highlighted its potential as an antimicrobial target [76]. Advances in protein structural modeling have caused a paradigm shift in contemporary biology [77]; synthesizing structural predictions with systematic phenotyping is becoming a core strategy to elucidate gene functions. Using this approach, we uncovered GO593_00140 as a possible holD analog in A. baumannii. Although GO593_00140 is highly divergent from E. coli HolD at the primary sequence level, it contains similar predicted folds and a GO593_00140 knockdown strain has overlapping chemical phenotypes to knockdowns of other DNA polymerase III (DNAP) subunits, corroborating previous in silico predictions [43]. Additionally, we identified a possible LexA analog (GO593_11615) inside a predicted prophage. Although A. baumannii has a distinct DNA damage response that does not rely on LexA [46], knockdown strains of this analog display phenotypes that match disruption of DNAP activity. These findings open the possibility that mobile genetic elements could serve as a resource to better understand the DNA damage response or identify proteins that are particularly toxic to A. baumannii. Finally, our ortholog and network analyses suggest that A. baumannii contains may unique or divergent genes associated with cell division. Synthesizing our work with prior studies [3,48] supports the hypothesis that A. baumannii contains a unique divisome with AdvA and BlhA localizing to the division septum and additional proteins possibly participating in division either directly (SPOR) or indirectly (Smc2). The precise functions of these players and their order of assembly into the divisome [78] has yet to be elucidated.
Our structure-function studies combining chemoinformatics, bioinformatics, and phenotype data provided insights into the functions of known and poorly characterized inhibitors. Our results with carvacrol and thymol, two inhibitors of unknown mechanism, suggest that perturbance of membrane associated complexes such as NADH dehydrogenase complex I and the SRP underly their toxicity to A. baumannii. This is broadly consistent with previous studies that have noted that thymol disrupts PMF [59] or generates reactive oxygen species through the Fenton reaction [79]. The direct target remains unknown; indeed, multiple targets are possible. We further show that antibiotics within the same class can have divergent effects on A. baumannii physiology. Among tetracycline class antibiotics, only minocycline was unaffected by ATP synthase knockdowns and only tigecycline showed increased efficacy against cytochrome bo3 knockdowns. Despite tigecycline being a more recently synthesized tetracycline derivative, it is often less effective against A. baumannii than minocycline due to efflux [62]. Efflux is dependent on the cellular energy state [80], which may link the cytochrome bo3 oxidase to reduced tigecycline efficacy. Previous work suggests that tigecycline is more prone to inactivation by dissolved oxygen in media than other tetracyclines [81] and is known to have anti-cancer activities due to induction of mitochondrial oxidative stress [82]. Cytochrome bo3 oxidase, as the main terminal oxidase in A. baumannii, has an important role in reducing oxygen that could impact tigecycline stability. It is important to note that these genes are essential in A. baumannii because of its obligate aerobic lifestyle, and knockdowns of ATP synthase or cytochrome bo3 oxidase may be causing complex compensation or stress responses that affect observed phenotypes. Our identification of distinct physiological impacts of tetracyclines underscores the utility of our chemical-genomics approach to identify mechanistic differences between closely related compounds.
Overall, this study enhances our understanding of essential genes in A. baumannii, highlighting unique aspects of its biology and establishing a useful resource for future studies. A limitation of our approach is dependence on CRISPRi, which causes polar effects on downstream genes in operons and can have off-target effects, although our experimental design and analysis strategy mitigate some of the risks of toxic or otherwise off-target guides. Additionally, much of our computational analysis relies on similarity to known proteins and functional predictions, particularly from the STRING database [24]. A. baumannii has a poorly characterized proteome compared to model bacteria; despite this, our approach still successfully identified many expected gene connections, likely due to the high level of conservation in essential gene sequences and functions. Future studies could integrate our dataset or network analyses to enhance the breadth and coverage of predicted genetic pathways. While this screen focuses on essential genes and a curate set of chemicals known to cause growth defects in A. baumannii, our systems approach can be readily applied to full knockdown libraries or small molecule screens in other pathogens.
Methods
Bacterial strains and growth
All strains, plasmids, and oligos used in this study are listed in S1–S3 Tables. Strains were grown in EZ Rich Defined Media (Teknova) supplemented 40 mM sodium succinate as the carbon source (AbRDM) or AbRDM + 1.5% agar at 37°C unless otherwise noted. Selective antibiotics were used when necessary: for E. coli, 100 µg/mL ampicillin (amp) or carbenicillin (carb), 15 µg/mL gentamicin (gent), or 30 µg/mL kanamycin (kan); and for A. baumannii, 10 µg/mL polymyxin B (pmb), 150 µg/mL gentamicin (gent), 60 µg/mL kanamycin (kan). Diaminopimelic acid (DAP) was added at 300 µM to support growth of E. coli dap- donor strains. IPTG (isopropyl b-D-1-thiogalactopyranoside) (0 to 1 mM) was added where indicated in the figures or figure legends. Strains were preserved in 15% glycerol at -80°C.
General molecular biology techniques and plasmid construction
Plasmids, oligos, and other construction details are listed in S2 and S3 Tables. Oligonucleotides were synthesized by Integrated DNA Technologies (Coralville, IA). Plasmid DNA was purified using the GeneJet plasmid miniprep kit (K0503; Thermo Scientific); genomic DNA was purified using the GeneJet genomic DNA purification kit (K0721; Thermo Scientific). DNA fragments were amplified with Q5 DNA polymerase (M0491; New England Biolabs (NEB)) or OneTaq DNA Polymerase (NEB). All enzymes for restriction digests were from NEB. DNA fragments were spin-purified using DNA Clean & Concentrator kit (D4004; Zymo Research) after digestion or amplification. Ligations performed using T4 DNA ligase (M0202; NEB). Plasmids were assembled from restriction enzyme-linearized or PCR-amplified vector and PCR products or synthetic DNA fragments using the NEBuilder Hifi DNA assembly kit (E2621; NEB). Plasmids were transformed into electrocompetent E. coli cells using a Bio-Rad Gene Pulser Xcell on the EC1 setting. Sanger DNA sequencing was performed by Functional Biosciences (Madison, WI); full plasmid sequencing was performed by Plasmidsaurus.
Library growth experiment
The A. baumannii essential gene CRISPRi library (sJMP2949) was diluted 1000-fold from frozen stock (OD600=15) in 50 mL AbRDM (starting OD600= 0.015) and incubated in 250 mL flasks at 37°C with shaking until mid-log (OD600=0.2; ~3.5 hrs)(t0). This culture was then diluted to OD600 =0.02 in 4 mL media with 1mM IPTG and antibiotics (antibiotic stocks diluted 1:100 into media and used immediately; see S4 Table) in 14 mL snapcap culture tubes (Falcon 352051) and incubated with shaking at 37°C. At 18 hours, saturated cultures were serially diluted back to OD600=0.02 into fresh tubes containing the same media and incubated for another 18 hours at 37°C with shaking, for a total of ~14 doublings. Cells were pelleted from 1 mL culture in duplicate before (t0) and after growth in antibiotics.
Sample preparation and sequencing
Genomic DNA was extracted from cell pellets and resuspended in a final volume of 200 μL. The sgRNA-encoding region was amplified using Q5 DNA polymerase (NEB) in a 100 μL reaction with ~100 ng gDNA and primers oJMP697 and oJMP698 according to manufacturer’s protocol, using a BioRad C1000 thermocycler: 30s at 98°C, followed by 16 cycles of 15s at 98°C, 15s at 65°C, 15s at 72°C, then 72°C for 5 min. PCR products were spin-purified and eluted in a final volume of 20 μL.
Samples were sequenced by the UW-Madison Biotech Center Next Gen DNA Sequencing Core facility. PCR products were amplified with nested primers containing i5 and i7 indexes and Illumina TruSeq adapters, cleaned, quantified, pooled, and run on a Novaseq 6000 (150 bp paired-end reads).
Sequencing read quality control and processing
Guides were counted using a custom Python script designed to minimize noise and accurately count guides with overlapping targets. This script samples sequencing reads to identify barcode diversity, orientations, and offsets, and determines flanking sequences for correct barcode identification. This is particularly useful when guides differ by only one or two bases (e.g., mismatch guides) and correctly identifies unintended guides present in the library due to oligo synthesis errors. Only counts for library barcodes were used in downstream analyses here.
Coefficients of determination (R²) between samples were calculated from barcode counts per million (CPM), and conditions with replicates where R² < 0.5 were removed to ensure replicate agreement. Samples were also excluded if population diversity (Nb) was less than 10,000 for nontargeting control guides.
Log2 fold changes and confidence values were computed using edgeR (S5 Table). Gene-level values were calculated by taking the median guide-level log2 fold change for perfect match guides; confidence was determined by computing Stouffer’s p-value using guide-level FDR values (poolr R package) (S6 Table). Functional enrichment analyses were conducted using STRING-db version 11.5 with a custom A. baumannii ATCC 19606 proteome (Organism ID: STRG0060QIE).
Ortholog analyses
Genome assemblies of Gammaproteobacteria were collected from the NCBI Reference Sequence database (RefSeq release 213). The dataset comprises 2792 isolates distributed among 474 genera and 18 taxonomic orders.
Essential genes represented in the library were pulled from the genome assembly for A. baumannii ATCC 19606 (GCA_009759685.1). Orthology assignment was computed with the phylogenetic profiling tool fDOG v.0.1.26 [30], compiling a maximum of 25 core-orthologs for each gene using a taxonomic distance minimum of genus and maximum of class. Visualizations were performed using PhyloProfile [83].
The ortholog presence ratio for each gene was measured within a taxonomic group by calculating the fraction of isolates possessing at least one ortholog out of the total number of analyzed isolates. We evaluated these ratios at the genus level with non-ACB (Acinetobacter calcoaceticus-baumannii complex) isolates, at the order level with non-Acinetobacter Moraxellales, and at the class level with non-Moraxellales Gammaproteobacteria. Orthologs and ortholog presence ratios are listed in S7 Table with the full genome dataset and assemblies in S8 Table.
Creation of knockdown strains
As previously described [20,84], sgRNA-encoding sequences were cloned between the BsaI restriction sites of MCi plasmid pJMP2776. Two oligonucleotides encoding an sgRNA were designed to overlap such that when annealed, their ends were complementary to the BsaI-cut ends on the vector.
Next, the Mobile-CRISPRi system was transferred to the attTn7 site on the chromosome of A. baumannii by quad-parental conjugation of three donor strains - one with a mobilizable plasmid (pTn7C1) encoding Tn7 transposase, another with a conjugal helper plasmid (pEVS104), and a third with a mobilizable plasmid containing a Tn7 transposon encoding the CRISPRi system - and the recipient strain A. baumannii ATCC 19606 or ATCC 17978. Briefly, 100 µL of culture of donor and recipient strains were added to 600 µL LB, pelleted at ~8,000 × g, washed twice with LB prior to depositing cells on a nitrocellulose filter (Millipore HAWP02500) on an LB plate, and incubated at 37°C, ~4 h. Cells were removed from the filter by vortexing in 500 μL PBS and selected on LB-gent plates at 37°C.
MIC test strip assays
500 μL of culture (OD=0.1, or 0.5 for LOS- strains) was spread on Mueller-Hinton agar (MHA) plates and allowed to dry for 30 minutes. A sterile MIC test strip was added to the plate and incubated overnight at 37°C. The MIC was determined as the value where the zone of inhibition meets the strip.
Spot assays
Strains were grown on LB agar overnight. Cell scrapes were resuspended from plates into 500 μL PBS, and normalized to an OD600 of ~1.0. 10 μL of ten-fold dilutions of these resuspensions were then spotted on AbRDM and AbRDM +1mM IPTG agar plates.
Membrane permeability assay
ATCC 19606 and LOS- strains containing sgRNAs targeting lptA or a non-targeting control (sJMP4539 and 4324, respectively) were grown on LB agar or LB agar containing 1mM IPTG overnight. Cell scrapes were resuspended from plates into 500 μL PBS and normalized to an OD600 of 0.3. Ethidium bromide (EtBr; Bio-Rad #1610433) was added to cell suspensions at a final concentration of 10 ug/mL, and OD600 and fluorescence at 545nm excitation/600 emission were read in a plate reader (Tecan Infinite 200 Pro M Plex). Readings occurred over 100 cycles (~75 min).
Chemical comparison analyses
After edgeR was used to import, organize, filter, and normalize counts, the limma package with voom method was used to perform gene set testing using functional groups predicted by STRING-db (custom proteome; Organism ID: STRG0060QIE). Chemical effects on pathways were scored for direction and significance, then transformed into a correlation matrix. Chemicals were clustered using hierarchical clustering with the Ward method, and a dendrogram visualized the clustering based on pathway impacts.
SMILES strings for each chemical were parsed using the rcdk package in R to generate molecular fingerprints. For combination or mixed-structure antibiotics, a representative structure was selected (e.g., trimethoprim for trimethoprim-sulfamethoxazole or polymyxin E1 for colistin). Tanimoto coefficients were calculated to measure chemical similarity and generate a similarity matrix, which was then clustered using the Ward method to produce a chemical structure dendrogram.
To assess the relationships between chemical pairs, each dendrogram was iteratively cut at all k-values (1-45), and pairs were scored if they clustered together. The final association scores are the sums across all k-values. Differences in association scores between pathway-effect and structure were plotted to highlight disparities. Additionally, the sums of association scores for pathway-effect and structure were plotted to highlight similarities.
Protein structural prediction
Structures of GO593_10640 and GO593_00140 (with chi subunit ortholog) were predicted from amino acid sequence using Alphafold2 [77,85] and AlphaFold-Multimer [86]. Predicted structures were visualized using PyMOL (Schrödinger).
Cell division protein localization and microscopy
Expression vector plasmids were transferred by conjugation into A. baumannii ATCC 17978; briefly, 100 µL of culture of donor and recipient strains were mixed and deposited on a nitrocellulose filter (Millipore HAWP02500) on an LB plate, and incubated at 37°C, ~2 h. Cells were removed from the filter by vortexing in 500 μL PBS and selected on LB-kan plates at 37°C.
Cultures were grown from a starting OD600~0.01 to mid-log at 37°C for ~3 hours in AbRDM + 1mM IPTG (or 500 uM IPTG for Smc2 and Smc2-sfGFP overexpression strains, due to toxicity). 1 mL of culture was centrifuged (8000xg for 2 min) and resuspended in equal volume PBS. Cells were fixed with paraformaldehyde (final concentration 5%) and quenched with equal volume 2.5M glycine. 10 μL of each sample were spotted on a glass slide for microscopy.
Bacteria were imaged with a Nikon Ti-E inverted microscope with an Orca Fusion BT digital CMOS camera (Hamamatsu) using NIS-Elements. Fluorescence images were collected using Prior Lumen 200 metal halide light source and a FITC-specific filter set. Cell length analysis was done with MicrobeJ [87].
Supporting information
S1 Fig. Replicate agreement and sample diversity as quality metrics.
(A) Correlation of library spacer counts per million (CPM) across the two biological replicates for samples containing polymyxin B at concentrations of 0.5 ug/mL (left) or 1 ug/mL (right). Count density represented by contour map; right graph shows over-depletion of guides. (B) Population diversity (Nb) for polymyxin B replicates. Samples with nontargeting guide complexity below the cutoff (dotted line) were excluded from analyses. (C) Heatmap showing correlations of spacer CPMs between solvent controls. All solvent control samples show r≥0.97.
https://doi.org/10.1371/journal.pgen.1011642.s001
(PDF)
S2 Fig. Growth and permeability assays for lpt and lpx knockdowns.
(A) Box plot of significant negative CG scores for lpt or lpx genes. Asterisk denotes p < 0.05 (Student’s t-test). (B) Colistin and levofloxacin MIC test strip assays for lptA or lpxC knockdown, LOS-, or nontargeting guide control strains. Plates supplemented with 1mM IPTG. Approximate MICs indicated below images. (C) Heatmap of CG scores of lpx knockdowns across conditions, showing muted phenotypes compared to lpt knockdowns. (D) 10 uL spots of ten-fold serial dilutions on plates with and without induction (N=3 biological replicates). The lptA knockdown, but not the lpxC knockdown, exhibits a minor growth defect upon induction. LOS- strains show major growth defects. (E) Ethidium bromide permeability assay for lptA knockdown and nontargeting control in 19606 or LOS- backgrounds without induction; increased fluorescence over time indicates membrane permeability. Ribbons represent standard deviation (N=4 biological replicates). Without induction, the strain containing the lptA guide behaves similarly to a nontargeting guide control.
https://doi.org/10.1371/journal.pgen.1011642.s002
(PDF)
S3 Fig. Chemical-gene scores and conservation in Moraxellales.
Dot plot depicting significant CG scores and ortholog presence ratio (fraction of isolates possessing at least one ortholog out of the total number of analyzed isolates) across representative non-Acinetobacter Moraxellales for library essential genes. Dots in blue represent genes with <38% presence across representative groups and significant chemical-gene interactions in >10% of screen conditions.
https://doi.org/10.1371/journal.pgen.1011642.s003
(PDF)
S4 Fig. Phyloprofile of A. baumannii candidate genes rare in other Gammaproteobacteria.
Phylogenetic profile of candidate genes showing ortholog presence ratios (fraction of isolates possessing at least one ortholog out of the total number of analyzed isolates; circle sizes) for selected A. baumannii genes across Gammaproteobacteria orders. The color encodes the median feature architecture similarity (FAS score) between the protein in 19606 and its orthologs within an order. The dot color gradient (FAS_F; blue to orange) captures architecture differences using the 19606 protein as reference. The cell color gradient (FAS_B; white to pink) captures architecture differences using the ortholog as reference. The score decreases if features in reference are missing in the respective orthologs.
https://doi.org/10.1371/journal.pgen.1011642.s004
(PDF)
S5 Fig. Phyloprofile of lpt genes.
Phylogenetic profile of candidate genes showing ortholog presence ratios (fraction of isolates possessing at least one ortholog out of the total number of analyzed isolates; circle sizes) for selected lpt genes across Gammaproteobacteria orders. The color encodes the median feature architecture similarity (FAS score) between the protein in 19606 and its orthologs within an order. The dot color gradient (FAS_F; blue to orange) captures architecture differences using the 19606 protein as reference. The cell color gradient (FAS_B; white to pink) captures architecture differences using the ortholog as reference. The score decreases if features in reference are missing in the respective orthologs.
https://doi.org/10.1371/journal.pgen.1011642.s005
(PDF)
S6 Fig. Essential gene subnetworks.
Sections from the essential gene network for (A) cytochrome bo3 oxidase (cyo), ion transporter (phaAB, mnhC1, phaD, mrpF), and related genes or (B) lpt-associated genes, including gtrOC1. Colors indicate STRING functional groups; solid or dotted lines indicate genes between or within operons, respectively.
https://doi.org/10.1371/journal.pgen.1011642.s006
(PDF)
S7 Fig. Phenotypes and structural analyses of uncharacterized A. baumannii genes.
Heatmaps of CG scores for clusters containing (A) GO593_00140 or (B) GO593_11615. Y-axis clustering was conducted using the Ward method and Canberra distance across all library knockdowns. (C) Heterodimer structural prediction (Alphafold2) for GO593_00140 and the chi subunit of DNA Polymerase III in A. baumannii (left) and the solved structure for the E. coli psi-chi subunit heterodimer (right). GO593_00140 has a lengthy unstructured N-terminal end absent in E. coli HolD. (D) Domain architecture map from InterProScan analysis, comparing E. coli LexA to GO593_11615. Both have a DNA-binding domain and S24 peptidase/LexA domain but differ in overall protein length and specific type of DNA-binding domain. (E) Domain architectures for E. coli FtsN and A. baumannii GO593_07290. GO593_07290 lacks similarity to the short EFtsN region required for function (47). BLAST-aligned amino acid sequences (24% identity) are depicted with dotted lines.
https://doi.org/10.1371/journal.pgen.1011642.s007
(PDF)
S8 Fig. Functional phenotypes for network genes of unknown function.
(A) Heatmaps display medL2FC for the signal-recognition cluster containing ftsY, ffh, and GO593_10640. GO593_18720 does not cluster in the heatmap with these genes due to muted phenotypes but is connected in the network. (B) GO593_10640 is a predicted transmembrane protein similar to a transporter. Predicted structure (Alphafold) and orientation in the membrane (TMHMM) are shown.
https://doi.org/10.1371/journal.pgen.1011642.s008
(PDF)
S9 Fig. Microscopy of fluorescently tagged cell division proteins.
3 representative images for each construct are shown with corresponding empty vector controls. (A) sfGFP::BlhA localizes at sites of division at the mid-cell at either existing or forming septa. (B) Fluorescently tagged SPOR-domain protein shows localization at the cell membrane and at septa for dividing cells. (C) Smc2::sfGFP expression leads to elongated cells, with fluorescently tagged proteins aggregating inconsistently within the cells. (D) Expression of Smc2 from a replicative plasmid produces elongated cells. All bars are 5 μm; GFP fluorescence/phase contrast channels merged in composite images for all except (D).
https://doi.org/10.1371/journal.pgen.1011642.s009
(PDF)
S10 Fig. Chemical pairs similar in pathway effect and structure
. Heatmap shows the normalized sum of pathway effect and structure association scores for chemical pairs. Axes are identical. Dark blue represents chemical pairs clustered tightly in both structure or pathway effect; white represents chemicals not clustered in either structure or pathway effect. Stars represent compounds or treatments with multiple chemical structures, where one was selected for structural comparisons (see S4 Table).
https://doi.org/10.1371/journal.pgen.1011642.s010
(PDF)
S11 Fig. Pathways impacted by carvacrol and thymol
. Sina plots depict differential guide CPMs for (A) NADH dehydrogenase and (B) ubiquinone pathways in carvacrol and thymol. Chemical-gene interactions are described above each graph—up (positive), down (negative), or no change—with FDR values. Guides are weighted in this calculation by predicted efficacy, with perfect guides at 100. Volcano plots show relative fitness and FDRs at the gene-level in (C) carvacrol or (D) thymol. Genes in pathways of interest are highlighted, and chemical structures are depicted.
https://doi.org/10.1371/journal.pgen.1011642.s011
(PDF)
S12 Fig. Chemical pairs opposed in pathway effect and structure
. Heatmap shows the normalized differences between pathway effect and structure association scores for chemical pairs. Axes are identical. Green represents chemical pairs clustered tightly by structure but not pathway effect; purple represents chemicals clustered by pathway effect but not by structure. Stars represent compounds or treatments with multiple chemical structures, where one was selected for structural comparisons (see S4 Table).
https://doi.org/10.1371/journal.pgen.1011642.s012
(PDF)
S13 Fig. Lpt and ATP synthesis pathway effects in tetracycline-class antibiotics.
Sina plots with boxplots of differential guide CPMs for listed STRING identifiers in tetracycline-class antibiotics: (A) lipopolysaccharide (lipooligosaccharide in A. baumannii) transport or (B) ATP synthesis pathways. Guides are weighted by predicted efficacy, with perfect guides at 100. Chemical interactions compared to no chemical control (up, down, or no change) and FDRs are listed. Non-significant comparisons are shown with grey backgrounds. (C) Chemical structures of tetracycline, minocycline, and tigecycline.
https://doi.org/10.1371/journal.pgen.1011642.s013
(PDF)
S14 Fig. Target pathway effects for tetracycline-class antibiotics.
(A) Sina plots depict differential guide CPMs for ribosome genes (STRING identifier CL:113) in minocycline, tetracycline, or tigecycline compared to no chemical control. Chemical-gene interactions are described above each graph—up (positive), down (negative), or no change—with FDR values. Guides are weighted in this calculation by predicted efficacy, with perfect guides at 100. (B) Heatmap shows relative fitness scores(median log2 fold changes) of ribosomal gene knockdowns of the mock treatment control compared to T0. Ribosome knockdowns all have substantial loss of fitness without additional chemical treatment. (C) Volcano plots of library gene CG scores (median log2 fold change compared to mock treatment) in minocycline, tetracycline, and tigecycline. Genes in the ribosome group are colored green. Lines depict cutoffs for significance (|median log2 fold change| ≥1, Stouffer’s p < 0.05).
https://doi.org/10.1371/journal.pgen.1011642.s014
(PDF)
S15 Fig. Ineffective chemical concentrations lack significant pathway effects.
(A) Volcano plots of genes in meropenem, imipenem, or doripenem compared to no drug control. Genes with significant CG scores are shown in black. Meropenem has very few significant CG scores, suggesting insufficient chemical dosage. (B) Sina plots with boxplots of differential guide CPMs for target pathway in doripenem, imipenem, or meropenem. Chemical interactions compared to no chemical control (up, down, or no change) and FDRs are listed. Non-significant comparison for meropenem is shown with grey background. Doripenem and imipenem are significantly impacted.
https://doi.org/10.1371/journal.pgen.1011642.s015
(PDF)
S4 Table. Table of chemical conditions and chemical information.
https://doi.org/10.1371/journal.pgen.1011642.s019
(XLSX)
S5 Table. Guide-level (CG score) log2FCs and confidence values.
https://doi.org/10.1371/journal.pgen.1011642.s020
(XLSX)
S6 Table. Gene-level (CG score) log2FCs and confidence values.
https://doi.org/10.1371/journal.pgen.1011642.s021
(XLSX)
S7 Table. Ortholog presence ratios for library genes.
https://doi.org/10.1371/journal.pgen.1011642.s022
(XLSX)
S8 Table. Genome dataset for ortholog analysis.
https://doi.org/10.1371/journal.pgen.1011642.s023
(XLSX)
Acknowledgments
We thank the UW-Madison Biotech Center for their assistance with sequencing. Thanks to Stephen Trent and Brent Simpson for their gift of the LOS- strain, and special thanks to Rachel Salemi and Alex Goetsch for their assistance and training with microscopy. We thank Warren Rose and Tim Bugni for helpful comments.
References
- 1. Acinetobacter in Healthcare Settings | HAI | CDC. 2019. [cited 11 Jul 2023. ]. Available: https://www.cdc.gov/hai/organisms/acinetobacter.html
- 2. Wong D, Nielsen TB, Bonomo RA, Pantapalangkoor P, Luna B, Spellberg B. Clinical and pathophysiological overview of acinetobacter infections: a century of challenges. Clin Microbiol Rev. 2017;30(1):409–47. pmid:27974412
- 3. Geisinger E, Mortman NJ, Dai Y, Cokol M, Syal S, Farinha A, et al. Antibiotic susceptibility signatures identify potential antimicrobial targets in the Acinetobacter baumannii cell envelope. Nat Commun. 2020;11(1).
- 4. Gallagher LA, Ramage E, Weiss EJ, Radey M, Hayden HS, Held KG, et al. Resources for genetic and genomic analysis of emerging pathogen Acinetobacter baumannii. J Bacteriol. 2015;197(12):2027–35. pmid:25845845
- 5. Qi LS, Larson MH, Gilbert LA, Doudna JA, Weissman JS, Arkin AP, et al. Repurposing CRISPR as an RNA-guided platform for sequence-specific control of gene expression. Cell. 2013;152(5):1173–83. pmid:23452860
- 6. Bikard D, Jiang W, Samai P, Hochschild A, Zhang F, Marraffini LA. Programmable repression and activation of bacterial gene expression using an engineered CRISPR-Cas system. Nucleic Acids Res. 2013;41(15):7429–37. pmid:23761437
- 7. Peters JM, Colavin A, Shi H, Czarny TL, Larson MH, Wong S, et al. A comprehensive, CRISPR-based functional analysis of essential genes in bacteria. Cell. 2016;165(6):1493–506. pmid:27238023
- 8. Peters JM, Koo B-M, Patino R, Heussler GE, Hearne CC, Qu J, et al. Enabling genetic analysis of diverse bacteria with Mobile-CRISPRi. Nat Microbiol. 2019;4(2):244–50. pmid:30617347
- 9. Geyman L, Tanner M, Rosario-Melendez N, Peters J, Mandel MJ, van Kessel JC. Mobile-CRISPRi as a powerful tool for modulating Vibrio gene expression. bioRxiv. 2024:2024.01.17.575898. pmid:38293084
- 10. Zhou D, Huang G, Xu G, Xiang L, Huang S, Chen X, et al. CRISPRi-mediated gene suppression reveals putative reverse transcriptase gene PA0715 to be a global regulator of Pseudomonas aeruginosa. Infect Drug Resist. 2022;15:7577–99. pmid:36579125
- 11. Hogan AM, Rahman ASMZ, Lightly TJ, Cardona ST. A broad-host-range CRISPRi toolkit for silencing gene expression in burkholderia. ACS Synth Biol. 2019;8(10):2372–84. pmid:31491085
- 12. Bai J, Dai Y, Farinha A, Tang AY, Syal S, Vargas-Cuebas G, et al. Essential gene analysis in Acinetobacter baumannii by high-density transposon mutagenesis and CRISPR interference. J Bacteriol. 2021;203(12):e0056520. pmid:33782056
- 13. Luo J, Efimova E, Volke DC, Santala V, Santala S. Engineering cell morphology by CRISPR interference in Acinetobacter baylyi ADP1. Microb Biotechnol. 2022;15(11):2800–18. pmid:36005297
- 14. Zheng XFS, Chan T-F. Chemical genomics: a systematic approach in biological research and drug discovery. Curr Issues Mol Biol. 2002;4(2):33–43. pmid:11931568
- 15. Alaimo PJ, Shogren-Knaak MA, Shokat KM. Chemical genetic approaches for the elucidation of signaling pathways. Curr Opin Chem Biol. 2001;5(4):360–7. pmid:11470597
- 16. Nichols RJ, Sen S, Choo YJ, Beltrao P, Zietek M, Chaba R, et al. Phenotypic landscape of a bacterial cell. Cell. 2011;144(1):143–56. pmid:21185072
- 17. Cain AK, Barquist L, Goodman AL, Paulsen IT, Parkhill J, van Opijnen T. A decade of advances in transposon-insertion sequencing. Nat Rev Genet. 2020;21(9):526–40. pmid:32533119
- 18. Enright AL, Banta AB, Ward RD, Rivera Vazquez J, Felczak MM, Wolfe MB, et al. The genetics of aerotolerant growth in an alphaproteobacterium with a naturally reduced genome. mBio. 2023;14(6):e0148723. pmid:37905909
- 19. Li S, Poulton NC, Chang JS, Azadian ZA, DeJesus MA, Ruecker N, et al. CRISPRi chemical genetics and comparative genomics identify genes mediating drug potency in Mycobacterium tuberculosis. Nat Microbiol. 2022;7(6):766–79. pmid:35637331
- 20. Ward RD, Tran JS, Banta AB, Bacon EE, Rose WE, Peters JM. Essential gene knockdowns reveal genetic vulnerabilities and antibiotic sensitivities in Acinetobacter baumannii. mBio. 2024;15(2):e0205123. pmid:38126769
- 21. Rousset F, Cabezas-Caballero J, Piastra-Facon F, Fernández-Rodríguez J, Clermont O, Denamur E, et al. The impact of genetic diversity on gene essentiality within the Escherichia coli species. Nat Microbiol. 2021;6(3):301–12. pmid:33462433
- 22. Hawkins JS, Silvis MR, Koo B-M, Peters JM, Osadnik H, Jost M, et al. Mismatch-CRISPRi reveals the co-varying expression-fitness relationships of essential genes in Escherichia coli and Bacillus subtilis. Cell Syst. 2020;11(5):523-535.e9. pmid:33080209
- 23. Costanzo M, VanderSluis B, Koch EN, Baryshnikova A, Pons C, Tan G, et al. A global genetic interaction network maps a wiring diagram of cellular function. Science. 2016;353(6306):aaf1420. pmid:27708008
- 24. Szklarczyk D, Kirsch R, Koutrouli M, Nastou K, Mehryary F, Hachilif R, et al. The STRING database in 2023: protein-protein association networks and functional enrichment analyses for any sequenced genome of interest. Nucleic Acids Res. 2023;51(D1):D638–46. pmid:36370105
- 25. Preston A, Mandrell RE, Gibson BW, Apicella MA. The lipooligosaccharides of pathogenic Gram-negative bacteria. Crit Rev Microbiol. 1996;22(3):139–80. pmid:8894399
- 26. Boll JM, Crofts AA, Peters K, Cattoir V, Vollmer W, Davies BW, et al. A penicillin-binding protein inhibits selection of colistin-resistant, lipooligosaccharide-deficient Acinetobacter baumannii. Proc Natl Acad Sci U S A. 2016;113(41):E6228–37. pmid:27681618
- 27. Powers MJ, Trent MS. Phospholipid retention in the absence of asymmetry strengthens the outer membrane permeability barrier to last-resort antibiotics. Proc Natl Acad Sci U S A. 2018;115(36):E8518–27. pmid:30087182
- 28. Sabnis A, Hagart KL, Klöckner A, Becce M, Evans LE, Furniss RCD, et al. Colistin kills bacteria by targeting lipopolysaccharide in the cytoplasmic membrane. Elife. 2021;10:e65836. pmid:33821795
- 29. Moffatt JH, Harper M, Harrison P, Hale JDF, Vinogradov E, Seemann T, et al. Colistin resistance in Acinetobacter baumannii is mediated by complete loss of lipopolysaccharide production. Antimicrob Agents Chemother. 2010;54(12):4971–7. pmid:20855724
- 30. BIONF/fDOG. BIONF; 2024. Available from: https://github.com/BIONF/fDOGf
- 31. Kenyon JJ, Hall RM. Variation in the complex carbohydrate biosynthesis loci of Acinetobacter baumannii genomes. PLoS One. 2013;8(4):e62160. pmid:23614028
- 32. VanOtterloo LM, Macias LA, Powers MJ, Brodbelt JS, Trent MS. Characterization of Acinetobacter baumannii core oligosaccharide synthesis reveals novel aspects of lipooligosaccharide assembly. mBio. 2024;15(3):e0301323. pmid:38349180
- 33. Hong Y, Liu MA, Reeves PR. Progress in our understanding of Wzx flippase for translocation of bacterial membrane lipid-linked Oligosaccharide. J Bacteriol. 2017;200(1):e00154-17. pmid:28696276
- 34. Hong Y, Reeves PR. Diversity of o-antigen repeat unit structures can account for the substantial sequence variation of wzx translocases. J Bacteriol. 2014;196(9):1713–22. pmid:24532778
- 35. Vassallo CN, Doering CR, Littlehale ML, Teodoro GIC, Laub MT. A functional selection reveals previously undetected anti-phage defence systems in the E. coli pangenome. Nat Microbiol. 2022;7(10):1568–79. pmid:36123438
- 36. Tesson F, Hervé A, Mordret E, Touchon M, d’Humières C, Cury J, et al. Systematic and quantitative view of the antiviral arsenal of prokaryotes. Nat Commun. 2022;13(1):2561. pmid:35538097
- 37. Fornelos N, Browning DF, Butala M. The use and abuse of LexA by mobile genetic elements. Trends Microbiol. 2016;24(5):391–401. pmid:26970840
- 38. Auchtung JM, Lee CA, Monson RE, Lehman AP, Grossman AD. Regulation of a Bacillus subtilis mobile genetic element by intercellular signaling and the global DNA damage response. Proc Natl Acad Sci U S A. 2005;102(35):12554–9. pmid:16105942
- 39. Lane D, Cavaillé J, Chandler M. Induction of the SOS response by IS1 transposase. J Mol Biol. 1994;242(4):339–50. pmid:7932694
- 40. Yakhnin AV, Bubunenko M, Mandell ZF, Lubkowska L, Husher S, Babitzke P, et al. Robust regulation of transcription pausing in Escherichia coli by the ubiquitous elongation factor NusG. Proc Natl Acad Sci U S A. 2023;120:e2221114120.
- 41. Cardinale CJ, Washburn RS, Tadigotla VR, Brown LM, Gottesman ME, Nudler E. Termination factor Rho and its cofactors NusA and NusG silence foreign DNA in E. coli. Science. 2008;320(5878):935–8. pmid:18487194
- 42. Marolda CL, Tatar LD, Alaimo C, Aebi M, Valvano MA. Interplay of the Wzx translocase and the corresponding polymerase and chain length regulator proteins in the translocation and periplasmic assembly of lipopolysaccharide o antigen. J Bacteriol. 2006;188(14):5124–35. pmid:16816184
- 43. Robinson A, Brzoska AJ, Turner KM, Withers R, Harry EJ, Lewis PJ, et al. Essential biological processes of an emerging pathogen: DNA replication, transcription, and cell division in Acinetobacter spp. Microbiol Mol Biol Rev. 2010;74(2):273–97. pmid:20508250
- 44. Wishart DS, Han S, Saha S, Oler E, Peters H, Grant JR, et al. PHASTEST: faster than PHASTER, better than PHAST. Nucleic Acids Res. 2023;51(W1):W443–50. pmid:37194694
- 45. Cox MM. Regulation of bacterial RecA protein function. Crit Rev Biochem Mol Biol. 2007;42(1):41–63. pmid:17364684
- 46. Hare JM, Bradley JA, Lin C-L, Elam TJ. Diverse responses to UV light exposure in Acinetobacter include the capacity for DNA damage-induced mutagenesis in the opportunistic pathogens Acinetobacter baumannii and Acinetobacter ursingii. Microbiology (Reading). 2012;158(Pt 3):601–11. pmid:22117008
- 47. Hubloher JJ, Van der Sande L, Schaudinn C, Müller V, Averhoff B. The Tol-Pal system of Acinetobacter baumannii is important for cell morphology, antibiotic resistance and virulence. Int Microbiol. 2023;26(3):543–50. pmid:36648597
- 48. Knight D, Dimitrova DD, Rudin SD, Bonomo RA, Rather PN. Mutations decreasing intrinsic β-lactam resistance are linked to cell division in the nosocomial pathogen Acinetobacter baumannii. Antimicrob Agents Chemother. 2016;60(6):3751–8. pmid:27067318
- 49. Yahashiri A, Jorgenson MA, Weiss DS. The SPOR Domain, a widely conserved peptidoglycan binding domain that targets proteins to the site of cell division. J Bacteriol. 2017;199(14):e00118-17. pmid:28396350
- 50. Ursinus A, van den Ent F, Brechtel S, de Pedro M, Höltje J-V, Löwe J, et al. Murein (peptidoglycan) binding property of the essential cell division protein FtsN from Escherichia coli. J Bacteriol. 2004;186(20):6728–37. pmid:15466024
- 51. Gerding MA, Liu B, Bendezú FO, Hale CA, Bernhardt TG, de Boer PAJ. Self-enhanced accumulation of FtsN at Division Sites and Roles for Other Proteins with a SPOR domain (DamX, DedD, and RlpA) in Escherichia coli cell constriction. J Bacteriol. 2009;191(24):7383–401. pmid:19684127
- 52. Law CW, Alhamdoosh M, Su S, Dong X, Tian L, Smyth GK, et al. RNA-seq analysis is easy as 1-2-3 with limma, Glimma and edgeR. F1000Res. 2016;5:ISCB Comm J-1408. pmid:27441086
- 53. Rogers D, Hahn M. Extended-connectivity fingerprints. J Chem Inf Model. 2010;50(5):742–54. pmid:20426451
- 54. Guha R. Chemical informatics functionality inR. J Stat Soft. 2007;18(5).
- 55.
Tanimoto TT. IBM Internal Report. 1957.
- 56. Rogers DJ, Tanimoto TT. A computer program for classifying plants. Science. 1960;132(3434):1115–8. pmid:17790723
- 57. Valdivieso-Ugarte M, Gomez-Llorente C, Plaza-Díaz J, Gil Á. Antimicrobial, antioxidant, and immunomodulatory properties of essential oils: a systematic review. Nutrients. 2019;11(11):2786. pmid:31731683
- 58. Hajibonabi A, Yekani M, Sharifi S, Nahad JS, Dizaj SM, Memar MY. Antimicrobial activity of nanoformulations of carvacrol and thymol: New trend and applications. OpenNano. 2023;13:100170.
- 59. Xu J, Zhou F, Ji B-P, Pei R-S, Xu N. The antibacterial mechanism of carvacrol and thymol against Escherichia coli. Lett Appl Microbiol. 2008;47(3):174–9. pmid:19552781
- 60. Epand RF, Pollard JE, Wright JO, Savage PB, Epand RM. Depolarization, bacterial membrane composition, and the antimicrobial action of ceragenins. Antimicrob Agents Chemother. 2010;54(9):3708–13. pmid:20585129
- 61. Rzycki M, Drabik D, Szostak-Paluch K, Hanus-Lorenz B, Kraszewski S. Unraveling the mechanism of octenidine and chlorhexidine on membranes: does electrostatics matter? Biophys J. 2021;120:3392–408.
- 62. Lashinsky JN, Henig O, Pogue JM, Kaye KS. Minocycline for the treatment of multidrug and extensively drug-resistant A. baumannii: a review. Infect Dis Ther. 2017;6(2):199–211. pmid:28357705
- 63. Rose WE, Rybak MJ. Tigecycline: first of a new class of antimicrobial agents. Pharmacotherapy. 2006;26(8):1099–110. pmid:16863487
- 64. Stewart-Tull DE, Armstrong AV. The effect of 1-hydroxyphenazine and pyocyanin from Pseudomonas aeruginosa on mammalian cell respiration. J Med Microbiol. 1972;5(1):67–73. pmid:4623349
- 65. Bisschop A, Bergsma J, Konings WN. Site of interaction between phenazine methosulphate and the respiratory chain of Bacillus subtilis. Eur J Biochem. 1979;93(2):369–74. pmid:218814
- 66. Donadio G, Mensitieri F, Santoro V, Parisi V, Bellone ML, De Tommasi N, et al. Interactions with microbial proteins driving the antibacterial activity of flavonoids. Pharmaceutics. 2021;13(5):660. pmid:34062983
- 67. Taheri Y, Suleria HAR, Martins N, Sytar O, Beyatli A, Yeskaliyeva B, et al. Myricetin bioactive effects: moving from preclinical evidence to potential clinical applications. BMC Complement Med Ther. 2020;20(1):241. pmid:32738903
- 68. Zhang G, Baidin V, Pahil KS, Moison E, Tomasek D, Ramadoss NS, et al. Cell-based screen for discovering lipopolysaccharide biogenesis inhibitors. Proc Natl Acad Sci U S A. 2018;115(26):6834–9. pmid:29735709
- 69. Richie DL, Takeoka KT, Bojkovic J, Metzger LE 4th, Rath CM, Sawyer WS, et al. Toxic accumulation of LPS pathway intermediates underlies the requirement of LpxH for growth of Acinetobacter baumannii ATCC 19606. PLoS One. 2016;11(8):e0160918. pmid:27526195
- 70. Wang H, Ishchenko A, Skudlarek J, Shen P, Dzhekieva L, Painter RE, et al. Cerastecins inhibit membrane lipooligosaccharide transport in drug-resistant Acinetobacter baumannii. Nat Microbiol. 2024;9(5):1244–55. pmid:38649414
- 71. Sampson BA, Misra R, Benson SA. Identification and characterization of a new gene of Escherichia coli K-12 involved in outer membrane permeability. Genetics. 1989;122(3):491–501. pmid:2547691
- 72. Ruiz N, Falcone B, Kahne D, Silhavy TJ. Chemical conditionality: a genetic strategy to probe organelle assembly. Cell. 2005;121(2):307–17. pmid:15851036
- 73. Singh MP, Petersen PJ, Weiss WJ, Janso JE, Luckman SW, Lenoy EB, et al. Mannopeptimycins, new cyclic glycopeptide antibiotics produced by Streptomyces hygroscopicus LL-AC98: antibacterial and mechanistic activities. Antimicrob Agents Chemother. 2003;47(1):62–9. pmid:12499170
- 74. Zampaloni C, Mattei P, Bleicher K, Winther L, Thäte C, Bucher C, et al. A novel antibiotic class targeting the lipopolysaccharide transporter. Nature. 2024;625(7995):566–71. pmid:38172634
- 75. Pahil KS, Gilman MSA, Baidin V, Clairfeuille T, Mattei P, Bieniossek C, et al. A new antibiotic traps lipopolysaccharide in its intermembrane transporter. Nature. 2024;625(7995):572–7. pmid:38172635
- 76. Faoro C, Wilkinson-White L, Kwan AH, Ataide SF. Discovery of fragments that target key interactions in the signal recognition particle (SRP) as potential leads for a new class of antibiotics. PLoS One. 2018;13(7):e0200387. pmid:30044812
- 77. Jumper J, Evans R, Pritzel A, Green T, Figurnov M, Ronneberger O, et al. Highly accurate protein structure prediction with AlphaFold. Nature. 2021;596(7873):583–9. pmid:34265844
- 78. Goehring NW, Beckwith J. Diverse paths to midcell: assembly of the bacterial cell division machinery. Curr Biol. 2005;15(13):R514-26. pmid:16005287
- 79. Zhang J-Y, Meng X, Zhu X-L, Peng S-R, Li H-B, Mo H-Z, et al. Thymol induces fenton-reaction-dependent ferroptosis in vibrio parahemolyticus. J Agric Food Chem. 2024;72(25):14337–48. pmid:38867141
- 80. Amaral L, Martins A, Spengler G, Molnar J. Efflux pumps of Gram-negative bacteria: what they do, how they do it, with what and how to deal with them. Front Pharmacol. 2014;4:168. pmid:24427138
- 81. Hope R, Warner M, Mushtaq S, Ward ME, Parsons T, Livermore DM. Effect of medium type, age and aeration on the MICs of tigecycline and classical tetracyclines. J Antimicrob Chemother. 2005;56(6):1042–6. pmid:16286361
- 82. Xiong Y, Liu W, Huang Q, Wang J, Wang Y, Li H, et al. Tigecycline as a dual inhibitor of retinoblastoma and angiogenesis via inducing mitochondrial dysfunctions and oxidative damage. Sci Rep. 2018;8(1):11747. pmid:30082885
- 83. Tran N-V, Greshake Tzovaras B, Ebersberger I. PhyloProfile: dynamic visualization and exploration of multi-layered phylogenetic profiles. Bioinformatics. 2018;34(17):3041–3. pmid:29659708
- 84. Banta AB, Ward RD, Tran JS, Bacon EE, Peters JM. Programmable gene knockdown in diverse bacteria using Mobile-CRISPRi. Curr Protoc Microbiol. 2020;59(1):e130. pmid:33332762
- 85. Varadi M, Anyango S, Deshpande M, Nair S, Natassia C, Yordanova G, et al. AlphaFold protein structure database: massively expanding the structural coverage of protein-sequence space with high-accuracy models. Nucleic Acids Res. 2022;50(D1):D439–44. pmid:34791371
- 86. Evans R, O’Neill M, Pritzel A, Antropova N, Senior A, Green T, et al. Protein complex prediction with AlphaFold-Multimer. 2021.
- 87. Ducret A, Quardokus EM, Brun YV. MicrobeJ, a tool for high throughput bacterial cell detection and quantitative analysis. Nat Microbiol. 2016;1(7):16077. pmid:27572972