banner
Maison / Nouvelles / Pseudomonas aeruginosa : pathogenèse, facteurs de virulence, résistance aux antibiotiques, interaction avec l'hôte, avancées technologiques et thérapies émergentes
Nouvelles

Pseudomonas aeruginosa : pathogenèse, facteurs de virulence, résistance aux antibiotiques, interaction avec l'hôte, avancées technologiques et thérapies émergentes

Sep 24, 2023Sep 24, 2023

Signal Transduction and Targeted Therapy volume 7, Article number: 199 (2022) Cite this article

38k Accesses

47 Citations

22 Altmetric

Metrics details

Pseudomonas aeruginosa (P. aeruginosa) is a Gram-negative opportunistic pathogen that infects patients with cystic fibrosis, burn wounds, immunodeficiency, chronic obstructive pulmonary disorder (COPD), cancer, and severe infection requiring ventilation, such as COVID-19. P. aeruginosa is also a widely-used model bacterium for all biological areas. In addition to continued, intense efforts in understanding bacterial pathogenesis of P. aeruginosa including virulence factors (LPS, quorum sensing, two-component systems, 6 type secretion systems, outer membrane vesicles (OMVs), CRISPR-Cas and their regulation), rapid progress has been made in further studying host-pathogen interaction, particularly host immune networks involving autophagy, inflammasome, non-coding RNAs, cGAS, etc. Furthermore, numerous technologic advances, such as bioinformatics, metabolomics, scRNA-seq, nanoparticles, drug screening, and phage therapy, have been used to improve our understanding of P. aeruginosa pathogenesis and host defense. Nevertheless, much remains to be uncovered about interactions between P. aeruginosa and host immune responses, including mechanisms of drug resistance by known or unannotated bacterial virulence factors as well as mammalian cell signaling pathways. The widespread use of antibiotics and the slow development of effective antimicrobials present daunting challenges and necessitate new theoretical and practical platforms to screen and develop mechanism-tested novel drugs to treat intractable infections, especially those caused by multi-drug resistance strains. Benefited from has advancing in research tools and technology, dissecting this pathogen's feature has entered into molecular and mechanistic details as well as dynamic and holistic views. Herein, we comprehensively review the progress and discuss the current status of P. aeruginosa biophysical traits, behaviors, virulence factors, invasive regulators, and host defense patterns against its infection, which point out new directions for future investigation and add to the design of novel and/or alternative therapeutics to combat this clinically significant pathogen.

Pseudomonas aeruginosa is a multi-drug resistance (MDR) opportunistic pathogens, causing acute or chronic infection in immunocompromised individuals with chronic obstructive pulmonary disease (COPD), cystic fibrosis, cancer, traumas, burns, sepsis, and ventilator-associated pneumonia (VAP) including those caused by COVID-19.1,2,3 P. aeruginosa in biofilm states may survive in a hypoxic atmosphere or other extremely harsh environments.4,5 In addition, treatments of P. aeruginosa infection are extremely difficult due to its rapid mutations and adaptation to gain resistance to antibiotics.6 Furthermore, P. aeruginosa is also one of the top-listed pathogens causing hospital-acquired infections, which are widely found in medical devices (ventilation) because they tend to thrive on wet surfaces.7 Importantly, P. aeruginosa is one of the MDR ESKAPE pathogens, which stand for pathogens Enterococcus faecium, Staphylococcus aureus, Klebsiella pneumoniae, Acinetobacter baumannii, P. aeruginosa, and Enterobacter. P. aeruginosa with arbapenem-resistance is listed among the "critical" group of pathogens by WHO, which urgently need novel antibiotics in the clinics.8

Epidemiological studies have shown that nearly 700,000 people died of the antibiotic resistance bacterial infections each year. P. aeruginosa that was isolated from European populations with a combined resistance was 12.9%.9 Hospital-acquired infection caused by P. aeruginosa continues to produce resistance to conventionally effective antibiotics becoming a main healthcare problem.10 The 2016 EPINE survey found that Escherichia coli and P. aeruginosa are the most common cause of hospital-acquired infections in Spain, P. aeruginosa accounting for 10.5% of clinically isolated bacteria infections.11 The 2011–2012 HCAIs report that P. aeruginosa caused almost nosocomial 9% of infections, which is the fourth commonest pathogen of the European hospitals.12 Similarly, 7.1% of HCAI are caused by P. aeruginosa in the United States.13 In addition, the 2016 European Center for Disease Prevention and Control (ECDC) epidemiological reported that P. aeruginosa causes a variety of ICU-hospital-acquired infections, including pneumonia flares, urinary tract infections, and bloodstream infections.9,10,14,15 Furthermore, data from China Antimicrobial Surveillance Network (CHINET) (http://www.chinets.com/) identified 301,917 clinically isolated pathogenic strains and found P. aeruginosa was the fourth nosocomial infections after Escherichia coli, Klebsiella pneumoniae, and Staphylococcus aureus, accounting for 7.96%. Altogether, P. aeruginosa is not a local, but a global major threat to human health.

The aforementioned statistics necessitate the identification drug targets and development of new treatments and effective vaccines for P. aeruginosa to improve human health. However, both efforts have met huge difficulty due to the surging cases with MDR strains. This article broadly reviews the recent progress in P. aeruginosa research towards the regulatory and functional mechanisms of virulence factors, gene expression regulators, secretion systems, quorum sensing, and antibiotic resistance, as well as host-pathogen interaction, new technologic advances, and therapeutic development (Fig. 1).

Schema of P. aeruginosa pathogenesis. P. aeruginosa can be traced everywhere including hospital environments and cause serious infection of almost any organ. LPS induces TLR-4-dependent and -independent inflammatory responses in the lung after bacterial infection, epithelial cells secrete cytokines and chemokines, thereby recruiting and activating innate immune cells and adaptive immune cells. The recruitment of neutrophils is a sign of inflammatory response activation. Although the activation of neutrophils is critical for host defense, excessively activated immune cell infiltration will cause severe tissue damage and aggravate bacterial infections.478 Therefore, studying the balance between the virulence factors secreted by bacteria and corresponding host immunity is important for the treatment of infections

P. aeruginosa is able to adapt to the adverse environment in hosts by secreting a variety of virulence factors, which contribute to successful infection and causing disease.16 First, lipopolysaccharide (LPS) is an important surface structural component to protect the external leaflet and posion host cells and the endotoxicity of the lipid A in LPS enable tissue damage, attachment, and recognition by host receptors.17 LPS may be related to antibiotic tolerance and biofilm formation.18 Second, out membrane proteins (OMPs) contributes to nutrient exchange, adhesion, and antibiotic resistance.19 In addition, drug resistance caused by the formation of biofilms is associated with the flagellum, pili, and other adhesins.20 Fourth, there are six types of secretion systems, including flagella (T6SS-associated), pili (T4SS), and multi-toxin components type 3 secretion system (T3SS), which function at colonisation of the host, adhesion, swimming, and swarming responding chemotactic signaling. Finally, exopolysaccharides, such as alginate, Psl and Pel, may help facilitate the formation of biofilms, while impairing bacterial clearance.20

In terms of toxins, T3SS is a complex system and may severely impede host defense via injection of cytotoxins including ExoU, ExoT, ExoS and ExoY, which affect the intracellular environment, especially blocking phagocytosis and bacterial clearance. Exotoxin A (ETA) can inhibit host protein synthesis through ADP ribosylation activity.21,22 Pyocyanin is also toxic to hosts to cause disease severity, damage host tissue, and impair organs’ function.23 In addition, LasA and LasB elastases, alkaline protease (AprA) LipC lipases, phospholipase C, and esterase A enzymes comprise a large group of lytic enzymes that modulate other virulence factors.24 Moreover, rhamnolipids-mediated lung surfactant degrading and tight junction destroying can directly injure trachea or lung epithelial cells.25 Furthermore, siderophores (pyoverdine and pyochelin) as iron uptake systems help in bacterial survival in iron-depleted environment to augment virulence strength.26 Finally, antioxidant enzymes, such as catalases (KatA, KatB, and KatE), alkyl hydroperoxide reductases, and superoxide dismutases, neutralize activity of reactive oxygen species (ROS) in phagocyte environments to avoid clearance.27

LPS, an important classical structural component of the outer membrane (OM) of most Gram-negative bacteria, is a known potent agonist that elicits robust innate inflammatory immunity, its distal end may be capped with O antigen, a long polysaccharide that can range from a few to hundreds of sugars in length, which is critical for bacterial physiology and pathogenesis.28 At the early stage, scientists were interested in developing vaccines to prevent infection by focusing on LPS, which were later proven highly difficult due to the various serotypes and inefficacious outcomes.28,29,30

Pathogen-associated molecular patterns (PAMPs), as small molecular motifs conserved in a class of microorganisms, can be sensed by toll-like receptors (TLRs) and other pattern recognition receptors (PRRs) to activate innate immune responses, which effectively protect the host from infection.31 LPS, as the prototypical PAMP, can be recognized by multiple host receptors, including TLRs, PRRs, and nucleotide-binding oligomerization domain-like (NOD-like) receptors.32 The LPS-PRRs/NOD-like molecules activate inflammasome to produce proinflammatory cytokines,33,34 activating TNF-α and IL-1β, two of the most eminent inflammatory cytokines.35 Furthermore, LPS amongst five major Gram-negative bacteria have the ability to induce the production of NF-κB and proinflammatory IL-8 in a TLR4-dependent manner, suggesting that the pathogenesis of bacterial enhancement of chronic inflammatory diseases may be related to its serotype-specific LPS response.36

Apparently, LPS exhibits a crucial role in regulating the interaction of bacteria with their host, is the main cause of tissue degeneration and chronic damage. LPS induces respiratory tract infections by regulating epithelial-mesenchymal transition (EMT)-mediated airway remodeling.37 The mutations of LPS can result in attenuated virulence.38,39 Caffeine alleviates the excessive inflammatory response caused by P. aeruginosa infection by inhibiting the activation of LPS-mediated TLR4/MyD88/NF-κB/miR-301b signaling pathway, and improves lung tissue injury.40 Notably, LPS mutations confer bacteria gain tolerance to phage infection.41 Taken together, in addition to the direct interaction with the host PRRs receptors, LPS may use its unique molecular features to adjust bacterial pathogenesis and damage host immune defense, ultimately benefiting the fitness and invasive strength.

OMVs are bacterial components that can be released spontaneously to the environment during growth by many Gram-negative bacteria. Bacterial-derived OMVs have been characterized as a novel secretion mechanism that can deliver a variety of bacterial proteins and lipids into host cells without direct contact with host cells.42,43,44 OMVs can package and enrich a wide variety of proteins and nucleic acids, including lipoproteins, periplasmic proteins (E. coli cytolysin A, enterotoxigenic E. coli heat-labile enterotoxin, and Actinobacillus actinomycetemcomitans leukotoxin), plasmid containing chromosomal DNA fragments, phage DNA, virulence factors (LPS, alkaline phosphatase, phospholipase C, β-lactamase, and Cif et al.45,46 P. aeruginosa secretion of OMVs have been implicated in many virulence-associated behaviors, including the acquisition of drug resistance, the regulation of bacterial density and host immune escape.47,48,49,50,51 Mechanistically, P. aeruginosa secretes OMVs to deliver virulence factors and sRNAs into lung epithelial cells through the diffusion the mucus layer.44,47,52,53,54,55 Some studies also illustrate that OMVs could lead to an increased hydrophobicity of cell surface, resulting in enhanced ability to form biofilms.56 OMVs is controlled by quorum-sensing systems, which enable bacteria to colonize and immune escape.54,57 Interestingly, OMVs are naturally immunogenic and self-adjuvation, making them have potential to be developed as antibacterial vaccine, such as OMV vaccine for Neisseria meningitidis.58,59,60 Therefore, OMVs are not only an important functional constitute, but also a potential biotechnological engineering carrier for vaccination or drug delivery. More details about virulence factors and their associated treatment strategies in P. aeruginosa are listed in Table 1 (45, 82–100).

Gram-negative pathogens cause various types of nosocomial infections, and secreted virulence factors often mediate their interactions with the host.61 Bacteria can modulate the host immune response through the secretion system for secreting virulence factors into host cells, which facilitates immune escape and enable bacterial colonization.62 Currently, 6 types of secretion systems (T1SS to T6SS) have been identified in P. aeruginosa. Based on the secretion routes of transport proteins, the secretion systems are divided into two major classes, one-step secretion system (T1SS, T3SS, T4SS, and T6SS) and two-step secretion system (T2SS and T5SS). The one-step secretion system directly secretes proteins from bacterial cytosol to the surface, while the two-step secretion system requires a brief periplasmic stay of the secreted proteins on the export way and then releases the proteins to outside environments of the bacterium (Fig. 2).

Protein secretion systems in P. aeruginosa. The secretion systems are divided into two major classes, one-step secretion system (T1SS, T3SS, T6SS) and two-step secretion system (T2SS, T5SS). One-step secretion system exoproteins are directly absorbed into the cytoplasm through their cognate secretion mechanism. In contrast, the exoproteins of two-step secretion system are first exported to the periplasm through the Sec or Tat system, and then crossing outer membranes through specific secretion mechanisms

Two different T1SS types in P. aeruginosa elucidated, Apr secretion system and Has secretion system.63 The Apr secretion system consists of three major components: AprD (ATP-binding cassette transporter, ABC transporter), AprE (adaptor), AprF (outer membrane factor, OMF), and secretes two proteins: AprA (alkaline protease), and AprX (protein with unknown function).64 T1SS is found in a variety of bacteria (P. aeruginosa Salmonella enterica, Neisseria meningitidis, and E. coli).65 T1SS major transport proteins (such as proteases and lipases). The substrate protein containing a C-terminal uncleaved secretion signal were recognized by the ABC transporter, were directly transferred across bacterial inner and outer membranes in a one-step process.66 The Has secretion system is composed of HasD (ABC transporter), HasE (adaptor), HasF, and OMF.67 Has secretion system participates in iron regulation by secreting an extracellular haem-binding protein (hasAp).68 Thus far, data relating to T1SS is very limited and its function in pathogenesis and significance for bacterium physiology and fitness are largely unknown, requiring further elucidation in order to know whether it has potential important functions.

The T3SS of P. aeruginosa, playing a key role in virulence like quorum sensing, was first discovered in 199669 and is one of the most extensively-studied secreted toxins with increasing evidence for its important virulent effects.70 The T3SS regulon comprises five distinct operons, including the pscN to pscU, the exsD-pscB to pscL operons, the popN-pcr1-pcr2-pcr3-pcr4-pcrD-pcrR operon, the pcrGVH-popBD operon and finally the exsCEBA operon.71,72 The five distinct operons play important roles in the biogenesis and translocation mechanisms of type III secretions. Structurally, the T3SS, similar to a molecular syringe, comprises five components: the needle complex, the translocation apparatus, the regulatory proteins, the effector proteins, and the chaperones.73 T3SS secretes virulence effectors (ExoS, ExoT, ExoY, and ExoU) into the eukaryotic host cells to disrupt intracellular signaling and ultimately causing cell death.74,75,76 Many bacterial factors regulate T3SS genes of P. aeruginosa. The MgtC and OprF of PAO1 regulate T3SS and ExoS-induced host macrophage damage.77 The T3SS is positively regulated by PsrA,78 HigB,79 Vfr,80 and DeaD,81 but negatively regulated by MexT,82 AlgZR, GacAS/LadS/RetS,83,84,85,86,87,88 and MgtE.89 No only for P. aeruginosa, the T3SS is a highly important secretion mechanism for Gram-negative bacterial invasion factors, which may also facilitate the bacterial evasion from the host immune responses to establish invasion, colonization, replication, and spread.

In P. aeruginosa, T6SS is a newly identified powerful system with diverse and vital functions in virulence, bacterial interaction, and competition with the environmental microorganisms.90 Initially, the genome of P. aeruginosa was thought to constitute three gene clusters called Hcp Secretion Island (HSI) encoding T6SS components, which are later renamed H1-T6SS to H3-T6SS,91,92 with ~15–20 genes for each of them.93 In addition, the apparatus of T6SS, consisting of 13 core components, is divided into a baseplate-like structure, a sheathed inner tube assembled from the baseplate-like structure and a trans-membrane complex.94 The assembled T6SS appears to be an inverted phage tail, with the Hcp (hemolysin coregulated protein)/Vgr (valine-glycine repeat protein) complex forming the distal end of the cell-puncturing device.95 The sheath transits the effectors into targeted cells by a contraction-based mechanism.96 Furthermore, ClpV, an AAA+ family ATPase of T6SS, also provides the energy necessary to drive the secretory apparatus.97

Mechanically, the secretion process by T6SS needs other elements; for example, the H2-T6SS machinery to deliver the novel antibacterial toxin Tle3 requiring a cytoplasmic adaptor Tla3.98 The GacAS/Rsm regulates T6SS (H1-T6SS and H3-T6SS) by activating RsmY/Z to inhibit the binding activity of RsmA/RsmN to fha1/tssA1.99 H3-T6SS secrete TseF to facilitate the import of the PQS-Fe3+ complex into cells by incorporating it into OMVs with Pseudomonas quinolone signal (PQS).100 Interestingly, it was found that the quorum sensing (QS) system plays an important role in the expression of this secretion system. In P. aeruginosa, QS differentially regulates three loci of encoding T6SS (HSI-I, HSI-II, and HSI-III) by LasR and MvfR.101 The QS systems regulator of Las and Rhl controlls the expression of H2-T6SS in PAO1 strains102 and the QS regulator of MvfR directly modulates the expression of multiple proteins, including virulence factors and other regulators in PA14,103 respectively.

The H1-T6SS-dependent substrates have a broad research foundation, while only little is known about functional roles of H2-T6SSs and H3-T6SSs due to the limited substrates available for research.104 With collaboration efforts, we characterize H2-T6SS-dependent secretomes that are related to copper (Cu2+)-binding effector azurin (Azu).104 Azu and H2-T6SS are both inhibited by CueR while induced by low levels of Cu2+. Furthermore, our studies reveal an Azu-interacting partner OprC, which is a Cu2+-specific TonB-dependent outer membrane transporter and is also modulated by CueR. The Azu-OprC-mediated Cu2+ transport network may contribute to P. aeruginosa virulence. Our follow-up studies indeed illuminate that oprC dampens host response in cells and mice to Pseudomonas infection by potently enhancing Quorum-Sensing-associated virulence.105

In addition, we recently describe a function for H2-T6SS of P. aeruginosa for specific delivery of AmpDh3 (a paralogous zinc protease) to the periplasm of a prey bacterium upon contact.106 AmpDh3 can hydrolyze peptidoglycan located on the cell wall of the prey bacterium to induce prey cell death, which serves as evolutionary advantage for P. aeruginosa in a competitive environment.

In spite of the relative short time since discovery of T6SS, the progress in understanding the potent virulence pathway is fascinating and fast-moving, which may define many unknown functions that can be attributed to T6SS’ virulence and indicate new ways for treating patients who suffer the most severe disease and difficult to treat.107

T3SS and T6SS are indicated to regulate host and bacterial responses, including host cell apoptosis, inflammatory response, colonization, motility, biofilm formation, and bacterial competition/interaction108,109 (Fig. 3). Interaction regulation and inter-conversion of T6SS and T3SS may be especially helpful for coping with complex environmental pressures.110 The switch between T6SS and T3SS is directly regulated by the RNA-binding protein RtcB controlling colonization, establishment, and pathogenicity in P. aeruginosa.111 YbeY regulates T3SS and T6SS secretion systems and biofilm formation by controlling RetS.112 The function of T3SS is regulated by various regulators, including four main regulators genes (exsA, exsC, exsD, and exsE), which is involved in the transcription activation of the aforementioned classical effectors (exoS, exoT, exoU, and exoY).72,113 ExoS is a 48.3 kDa protein containing 453 amino acid length. It has been reported early that ExoS participates in host cell apoptosis via its GAP region or ADP-ribosyltransferase (ADPr) activity.114,115 Furthermore, the ExoS possesses ADPRT activity, which induces P. aeruginosa-afflicted host cell apoptosis by targeting a variety of Ras proteins.116 ExoU is the longest P. aeruginosa effector containing 687 amino acids (73.9 kDa).117 ExoU is the most acutely cytotoxic among the four effector proteins because it can induce rapid cell death and is considered to be the main driver of the cytotoxic phenotype.118 ExoU dysregulates the host's innate inflammatory response by poisoning and killing immune cells, including macrophages, neutrophils, epithelial cells, and endothelial cells, allowing bacteria to persist, proliferate, and spread, and ultimately leading to sepsis, Alzheimer's disease, acute respiratory distress syndrome, etc.114 Mechanistically, ExoU transiently represses Capase 1 and NLRC4 inflammasome activation, inhibiting the release of IL-1β, IL-18, and proinflammatory DAMPs, and thereby suppressing the host immune response.119,120 In addition, ExoU can activate AP-1 transcription factors to increase IL-8 production and induce tissue-damaging inflammatory by JNK/MAPK) pathway.121 ExoT containing 457 amino acids (48.5 kDa) has GAP and ADPRT activities, and can induce host cell apoptosis by targeting Crk proteins phosphorylation of p38β and JNK induces apoptosis, which subsequently interferes with integrin-mediated survival signaling via destroying the stability of focal adhesion sites.122 Notably, recent studies have shown that P. aeruginosa ExoT induces G1 cell cycle arrest in melanoma cells, suggesting its potential for regulating the cell cycle.123 ExoY is a 378 amino acid protein with a molecular size of 41.7 kDa, detected in 80–100% of P. aeruginosa.124 ExoY plays a direct role in immune escape by inhibiting TAK1 activation, which is a key factor in the TGF-inducible pathway that directly modulates immune responses, contributing to P. aeruginosa survival and infection severity.125 In addition, ExoY regulates host inflammatory responses by delaying activation of NF-κB and caspase-1.126

Mechanisms of T3SS and T6SS in regulating bacterial pathogenesis and host responses in P. aeruginosa. LPS is recognized by TLRs (TLR1/2 or TLR4/9) and then activates T3SS and T6SS.479 T3SS and and T6SS represent a critical network in regulating bacterial behaviors (growth, biofilm formation, and competition) and host defense (host cell apoptosis, inflammatory response, colonization, and motility). T6SS and T3SS interaction and inter-conversion are regulated by RtcB and YbeY. ExoS/ExoU induce P. aeruginosa-afflicted host cell apoptosis and colonization by targeting JNKS signal pathway. ExoY/ExoT reduces inflammasome activity through inhibition of bacterial motility to dampen NF-κB and caspase-1 activation. T6SS is a powerful antibacterial weapon that can be injected through multiple effectors to compete with other bacteria and allows P. aeruginosa colonization and biofilm formation108,109

The T6SS components and their effectors are diverse and complex beyond bacterial-cell targeting. T6SS systems have been detected in ~200 Gram-negative bacteria, including P. aeruginosa.109 To compete for survival in the living environment, H1-T6SS kills other bacteria by injecting Tse2 effector molecules into other target bacteria possessing antibacterial activity and providing advantages for P. aeruginosa growth. In addition, to protect itself from Tse2 toxins, P. aeruginosa also produces the antitoxin Tsi2.127 Similarly, H1-Tse1 and Tse3 are injected into the periplasm of other bacteria to hydrolyze peptidoglycan, which can be counteracted by periplasmic immune proteins Tse1 and Tse3.128 Tse4, Tse5, Tse6, and Tse7 that also show antibacterial activity are associated with homologous immunity.129 The phospholipase D enzymes PldA and PldB of Tle5 were injected to other bacteria by H2-T6SS and H3-T6SS to exert antibacterial activity.130 Further, VgrG2b is injected into epithelial cells by H2-T6SS, in which it targets the γ-tubulin ring complex component (γ-TuRC) and promotes the recruitment of PI3K at the apical membrane.131,132 Moreover, PldA/B targets the host PI3K (phosphoinositide 3-kinase)/Akt pathway to remodel the PIP3 (phosphatidylinositol-3,4,5-triphosphate) and actin in the apical membrane, which is essential for successful colonization of the host by P. aeruginosa.133 T6SS-related genes hcp1/hcp3 and lasR were significantly higher in the strong biofilm formation (SBF) compared to nonbiofilm formation (NBF) groups, which contributes to the biofilm production by P. aeruginosa.90 Collectively, T6SS is a powerful antibacterial weapon that can be injected with many different effectors to compete with other bacteria and allow P. aeruginosa colonization and biofilm formation.

Different from the One-Step secretion, Two-step secretion requires a brief periplasmic phase of the secreted proteins on the export route before being exported to the outside of the cell through general export pathways, which plays an important role in the transport of periplasmic and outer membrane proteins.

The function of T2SS is one of the less characterized secretion systems and is thought to secrete folded proteins from the periplasm.134 Two different pathways exist in T2SS: the general secretory (Sec) and twin-arginine translocation (Tat).135 The secreted proteins are first transited through the inner membrane, stays briefly in the periplasm and then secreted into the extracellular environment.136,137 The Sec pathway consists of a protein targeting component, a motor protein, and a membrane integrated conducting channel called SecYEG translocase, the secreted proteins with a SecB-specific signal sequence might be guided to the periplasm or the extracellular environment.138 The twin-arginine translocation (Tat) pathway of Gram-negative consists of TatA and TatB, which can decide whether the secreted is retained in the periplasm or translocated to the extracellular space with a twin-arginine motif.139 Functionally, T2SS participates in the secretion of guanylate cyclase ExoA, proteases lasA/B and multiple other factors and many of which have emerged as potential therapeutic targets.140,141

The T5SS of P. aeruginosa is composed of five subtypes (type Va to Ve) and exports the proteins across the inner membrane via the Sec pathway.142 The proteins of the V-type secretion system are often referred to as autotransporters (ATs). Typically, the T5SS consists of only one polypeptide chain with a β-barrel translocator domain in the membrane, and an extracellular passenger or effector region.143 Under the regulation of the Bam complex (β-barrel assembly mechanism) and TAM complex (translocation and assembly module), outer membrane proteins fold to form a β-barrel conformation and insert into the outer membrane.144 T5SS secretes a variety of proteins, including EstA, LepB, and LepA. EstA has esterase activity and is involved in rhamnolipid production and biofilm formation.143

The type Vb secretion system comprises two distinct polypeptide chains encoded in one operon, therefore, it is also known as the dual-partner secretion system (TPSS) containing the passenger domain (TpsA) and the β domain (TpsB).145 TpsA has a TPS secretion motif and a functional/catalytic domain and the TpsB is a 16-chain OM-integrating β-barrel protein with two periplasmic POTRA (polypeptide transport-associated) domains.146 The Vc-type secretion system forms a highly intertwined trimeric structure and is therefore also known as a trimeric autotransporter adhesin (TAA).147 The C-terminal β-barrel domain of the Type Vd consists of 16 β-strands, similar to the β-barrel of TpsB proteins.145 The Ve-type ATs share obvious similarities with Va-type ATs, the main difference is that Ve-type ATs have an inverted domain order, with the β-barrel at the N-terminus and the passenger at the C-terminus.148 Despite these preliminary studies, there is much more to be learnt regarding T5SS for the physiology and virulence in P. aeruginosa.

T4SS is a multisubunit cell-envelope-spanning structure that can transfer protein and nucleoprotein complexes across membranes, which is related to horizontal gene transfer-mediated antibiotic resistance, adaptation, evolution, and virulence.149 T4SS in bacteria is divided into the IVA-type secretion system represented by Agrobacterium tumefaciens VirB/VirD4 and the IVB-type secretion existed as Legionella pneumophila Dot/Icm system. Those distinct from the above are classified as "Other T4SS", which contain genomic islands pKLC102 and PAPI in P. aeruginosa and are less characterized than other two types.149 P. aeruginosa T4SS comprises abundant major pilin subunits, PilA, and low abundance minor pilins FimU and PilVWXE. Both major pilin and minor pilins are processed by the pre-prepilin peptidase, PilD to exert function. T4SS assembly system is evolutionarily associated with T2SS contributing to the process of assembly and disassembly of pili. Minor pilins can impact assembly, retraction, extension, and adhesion.150

Again, the chief function of T4SS is horizontal genetic transfer (HGT) between different microorganisms and potentially relating to pathogenesis.151 For the genetic island containing T4SS, there is a discrepancy: whether a conjugation mechanism exists but this is likely related to differences between strains.152 Nevertheless, there are relatively limited studies of this system compared to other secretion systems such as T3SS. We will further discuss the potential role of T4SS in host response context, initiating/activating inflammasome independent on nucleotide oligomerization domain (NOD)-like receptor (NLR) family CARD (C-terminal caspase recruitment domain) domain-containing protein 4 (NLRC4),153 in addition, the virulence factors and their associated secretion systems in P. aeruginosa were summarized in Table 2.

The regulation of all these virulence factors is cell density-dependent through release of autoinducers of critical quorum sensing (QS) (e.g., Las, Rhl, Pqs, and Iqs), a mass communication system.154 QS may help large population fitness by a hierarchical signal pattern to survive in fierce host environments and thrive, leading to persistent infection in individuals with cystic fibrosis, which cannot be completely cured even with tremendous progress in drug development, drastically improved medcare systems and living conditions.155 Hence, QS systems along with some other critical virulence factors, such as six types of secretion systems (of toxic molecules), two-component systems (TCSs), have become an intense interest in mechanistic understanding of this bacterium.

QS describes a method that is widely utilized by bacteria for cell–cell mass communication. Both Gram-negative and -positive bacteria detect the local population density by sensing chemical signals and coordinate gene expression and group-beneficial behaviors.156,157 Bacteria produce autoinducer or quoromone as diffusion signaling molecules and release into the environment for communication. Once the population reaching a threshold, the autoinducers activate their cognate receptors to directly or indirectly induce gene expression.158 Over the past two decades, QS has been extensively studied as a potential target for ‘antivirulence agents’, which may be harnessed to counteract bacterial virulence via a noncytotoxic mechanism as alternatives of traditional antibiotics.159,160

Another essential function for QS in P. aeruginosa is to regulate the production of multiple virulence factors, such as extracellular proteases,161 iron chelators,162 efflux pump expression,163 biofilm development,164 swarming motility,165 and the response to host immunity.166 As a model organism, P. aeruginosa serves as one of the most suitable bacteria to study the fundamental mechanisms of QS signaling regulating virulence and search for chemical agents to block the QS system.167,168

Bacteria communicate with each other through the QS, which acts as a global regulatory system by directly or indirectly controlling the expression of genes, has been the central point of research.157 For example, scientists have made significant efforts in understanding the interactions between all the four QS systems and also how environmental cues may affect gene expression and function of the QS. Two canonical N-acyl L-homoserine lactone (AHL) based (Las and Rhl) and two 2-alkyl-4 quinolones (AQ) based (Pqs and its precursor Hhq) signaling systems.70 These systems connect and coregulate each other. Rhl and Pqs were positively regulated by the Las system, while Rhl represses Pqs and Pqs augments Rhl.169 For example, in response to various nutritional and environmental stimuli, the regulatory relationship between Rhl and Pqs systems can change independently of Las.170

The activation of QS genes generally requires a large number of regulatory factors to control receptor expression or function, and/or to coregulate some QS-controlled target genes since the QS systems are functional diverse, organizational complex, and consisting of a spectrum of key regulators (including rpoS, vqsR, mvfR, rhlR, rsaL et al.).171 RpoS indirectly plays a subtle role in activating lasR and rhlR expression and modulating ~40% QS-controlled gene expression during the stationary phase.172 VqsR controls the production of AHL signaling molecules and virulence factors by inhibiting the LuxR-type regulator QscR, which represses lasI expression to regulate the timing of QS activation.173 VqsM positively regulates the QS systems by controlling several relevant QS regulators ranging from QS to antibiotic resistance, and P. aeruginosa pathogenicity.174 MvaT and its homolog MvaU control the magnitude and timing of QS-dependent gene expression,175 which have a massive impact on all three QS systems by directly regulating lasR, lasI, rsaL, pprB, mvfR, algT/U, mexR, and rpoS.176 RsaL binds to the lasI promoter and prevents LasR-mediated activation, regulates las signaling,177,178 and modulates the activity of PqsH and a recently identified regulator, CdpR, which are required for PQS synthesis.179 AmpR activates QS-regulated genes to positively influence acute virulence, while negatively regulating biofilm formation.180 CdpR negatively modulates bacterial virulence by impacting the expression of pqsH, which is positively regulated by LasI and VqsM along with QscR and RsaL.181 We also recently found that BfmR/S and/or its variants modulate the rhl QS system in P. aeruginosa.182 Crc regulates rhl QS by promoting Hfq-mediated suppression of lon gene expression.183 More recently, our laboratory delineated that AnvM is a critical regulator of virulence in P. aeruginosa by directly interacting with the QS regulator MvfR and anaerobic regulator Anr.184 The aforementioned discoveries have highlighted the rapid progress in understanding diverse, heterogenous regulatory mechanisms of QS coordinated by seemingly a large number of unprecedented factors, which may finally characterize powerful, versatile regulatory proteins or systems to be applied to better control the notorious pathogen.

Several factors (such as QscR and QteE) have been identified to regulate the activation threshold of quorum-regulated genes, which control QS activation timing through additional homeostatic mechanisms.174 In addition to the QscR mentioned above, QteE also blocks QS-regulated genes’ expression by preventing LasR and RhlR accumulation and blocking Rhl-mediated signaling.185 Moreover, QslA is found to interact with LasR.186

Based on extensive studies to date, we may presume that QS may be one of the most important regulatory systems in Pseudomonas contributing substantially to bacterial physiology, adaptation, and pathogenesis. Although, various studies have tested the ideas of targeting QS for potential therapy of bacterial infection, the effect of using QS-associated approach for treatment is unsatisfactory.187 In particular, there were only limited reports of in vivo treatments targeting QS.188 Hence, it is necessary to further study the fundamental role of QS in bacterial pathogenesis and identify new anti-virulence targets and approaches that would help develop urgently needed medicines for treating refractory infection in clinics for QS bearing bacteria.

The two-component systems (TCSs) are ubiquitous, complex signaling regulators that play vital roles in bacterial survival, metabolism, and virulence.189 As a versatile opportunistic pathogen, P. aeruginosa virulence network is tightly controlled by a growing number of TCSs.190 In general, a TCS pair of genes consists of a membrane-bound sensory histidine kinase (HK) and a cytoplasmic response regulator (RR). In P. aeruginosa, 64 HKs and 72 RRs have been identified.191 More than 50% of the TCSs and their corresponding HKs are linked to virulence and/or antibiotic resistance of P. aeruginosa.192 For instance, CzcR/CzcS is implicated in carbapenem-resistance;193 KinB/AlgB is involved in the alginate synthesis and virulence;194 and GacS/GacA is essential for pathogenicity.195 Remarkably, an attenuation in virulence behavior can be achieved by blocking TCS signaling. Goswami et al. reported that inhibition of HKs, especially Rilu-4 and 12, significantly reduced the production of virulence factors and toxins, and severely impacted the motility behavior of PA14.195

Our recent studies discovered a new copper-responsive TCS called DsbRS in P. aeruginosa, in which DsbS (sensor of histidine kinase) and DsbR (cognate response regulator) modulate gene transcription for disulfide bond formation (Dsb). DsbS (phosphatase) targets DsbR to interfere with the transcription of Dsb genes and help the bacterium cope with copper stress.196 Intriguingly, transcription factors can also regulate the behaviors of bacteria to adapt host environments; for instance, imidazole-4-acetic acid (ImAA) and its receptor HinK are recently implicated in the response of P. aeruginosa to histamine.197 These findings help understand the communication between P. aeruginosa and hosts to adjust bacterial health. We have summarized TCSs and their roles in controlling the key virulence factors in P. aeruginosa (Table 3).

The number of TCSs related to the virulence of P. aeruginosa has recently grown considerably, which may be highlighted by the multi-kinase networks containing multiple sensor kinases through crosstalk (network) to impact virulence.192 The networks include: (1) GacS network governs the switch between acute and chronic infections;198,199,200,201 (2) Roc network and Rcs/Pvr network control cup fimbriae production;202,203,204 (3) A complex network of five TCSs (PmrB, PhoQ, ColS, ParS, and CprS) regulates the aminoarabinose modification of lipopolysaccharide;205,206,207 (4) Wsp chemosensory pathway modulates biofilm formation and chronic infection;208,209 and (5) Chp/FimS/AlgR network involves biofilm formation and virulence.210,211 Although TCSs may control many functions, some of them may have unknown roles that require further studying. TCSs play a significant role in controlling either P. aeruginosa virulence or virulence-related behaviors (such as biofilm formation and antibiotic resistance). The functions of TCSs may contribute to the clinical significance exemplified by a more recently identified BfmRS TCS.182 Indeed, the plasticity of TCSs mediated by spontaneous mutations of BfmRS in patients has been assessed, and the data suggest that mutation-induced activation of BfmRS may be related to host adaptation by P. aeruginosa in chronic infections.212 Despite increasing identification of novel TCSs, several critical questions remain to be answered in future investigation: how multi-kinase networks process and integrate signals, which of the kinases plays a dominant role in multi-kinase networks, and how these TCSs interact with other systems, quorum sensing and secondary messenger signals to confer pathogenicity (e.g., cAMP and c-di-GMP).

Small non-coding regulatory RNAs (sRNAs) range from 50 to 400 nucleotides in length.213 Since the discovery as part of a large group of transcriptional regulators in Escherichia coli in the 1960s, sRNAs are gradually recognized as key modulators to mount rapid responses when necessary and are encoded widespread in the bacterial genomes213,214,215 despite to different extent (the number of sRNAs in PAO1 is approximately twice higher than that in PA14). sRNAs play an essential role in bacterial pathogenicity and virulence mechanisms, such that participation in quorum-sensing regulation, ion metabolism, biofilm formation, stress responses, host cell invasion, and adaption to growth conditions.216,217,218 Hfq-dependent sRNAs are instrumental for modulating P. aeruginosa virulence.219 rsmY and rsmZ act as early responders to finely modulate bacterial cooperation under environmental stimuli to optimize population density.220

In P. aeruginosa, sRNAs can regulate bacterial metabolism instantly and precisely to establish successful infection in the hosts.218 A total of 573 sRNAs were detected in PAO1 and 233 sRNAs in PA14, indicating their quantity variation in different strains.201 Although 126 sRNAs are found in common to both strains, sRNAs can evolve rapidly, and many sRNAs exhibit strain‐specific or environmental-dependent expression.221 Interestingly, we recently revealed that sRNA PhrS may help generate CRISPR RNA (crRNA) for initiating bacterial immunity against bacteriophages, which is achieved through inhibiting Rho function on transcription-termination.222 Collectively, it would be intriguing to further understand the functions and underpinning mechanisms involving sRNA in this bacterium, which may identify novel pathway regulators and drug targets for controlling bacterial invasion.

The acquisition of drug resistance by P. aeruginosa depends primarily on multiple intrinsic and acquired antibiotic resistance mechanisms, including the biofilm-mediated formation of resistant and multi-drug-resistant persistent cells.223 Therefore, P. aeruginosa can quickly develop resistance to various antibiotics, including aminoglycosides, quinolones, and β-lactams.224

In the course of long-term evolution, bacteria have developed a variety of ancient genetic resistance mechanisms that have significant genetic plasticity against harmful antibiotic molecules, enabling them to respond to various environmental threats, including possible harm (e.g., antibiotics, chemical compounds, and antimicrobial peptides) to their survival. The acquisition of antibiotic resistance with P. aeruginosa is quite diverse, and some primary mechanisms including outer membrane permeability, efflux systems, and antibiotic-inactivating enzymes will be addressed below225 (Fig. 4).

Mechanisms of antimicrobial resistance in P. aeruginosa. Mechanisms of antimicrobial resistance in P. aeruginosa can be divided into intrinsic antibiotic resistance (① outer membrane permeability, ② efflux systems, and ④antibiotic-modifying enzymes or ⑤ antibiotic-inactivating enzymes), acquired antibiotic resistance (⑥ resistance by mutations and acquisition of resistance genes), and adaptive antibiotic resistance (③ biofilm-mediated resistance). Alteration of outer membrane protein porins decreases the penetration of drugs into cells by reducing membrane permeability. The efflux system directly pumps out drugs. Drug-hydrolyzing and modification enzymes render them inactive. Similarly, some enzymes cause target alterations so that the drug cannot bind its target, resulting in drug inactivity. Antibiotic resistance genes carried on plasmids can be acquired via horizontal gene transfer from the same or different bacterial species,225 quorum-sensing signaling molecules activate the formation of biofilms, which act as physical barriers and prevent antibiotics penetrating the cell

To treat P. aeruginosa infections, most antibiotics need to penetrate the cell membrane to reach the intracellular compartment to function.226 The outer membrane of P. aeruginosa can act as a specific hurdle inhibiting antibiotic penetration. The outer membrane of P. aeruginosa is chiefly composed of bilayer phospholipid molecules, LPS and porins embedded in phospholipids. The outer membrane is responsible for the specific and non-specific uptake of extracellular substances relying on different porin functions, including non-specific porins (OprF), specific porins (OprB, OprD, OprE, OprO, and OprP), gated porins (OprC and OprH), and efflux porins (OprM, OprN, and OprJ).227,228,229,230 P. aeruginosa manipulates different porins to limit antibiotic penetration and increase antibiotic resistance. OprF promotes the formation and attachment function of P. aeruginosa biofilm to protect the bacterium against antibiotics.231,232,233 Mutations of specific porins OprD involving conformational features cause carbapenem resistance, a serious challenge for medical treatment practices.234 Outer membrane protein H (OprH) of P. aeruginosa enhances the stability of the outer membrane through direct interaction with LPS to regulate antibiotic resistance.235,236 Efflux porins (OprM, OprN, and OprJ) contribute to the active efflux of several antibiotics, including tetracycline, norfloxacin, and β-lactam antibiotics.237 These findings demonstrate that the elegant effects and diverse mechanisms by which porins determine antibiotic resistance.

In a separate account, in recent years, OMVs secreted by P. aeruginosa are shown to be able to transport multiple virulence factors, like hemolytic phospholipase C, mRNA, DNA, β-lactamase, alkaline phosphatase, to the host cytoplasm, which may be a part of pathogenic mechanisms of antibiotic resistance. OMVs may be fused with the host plasma membrane through receptor-mediated endocytosis. While OMVs are detrimental to the host by delivering antibiotic resistance molecules or enzymes (β-lactamase), they have been harnessed as alternative delivery vehicles for transporting treatments or vaccines for various diseases including infection and cancer.47,238

The toxic compounds derived from metabolism or antimicrobials inside bacterial cells are harmful to bacterial survival, and require mechanisms to expel. The efflux pump is a main, conserved mechanism to remove antibiotics. The efflux pump may upregulate virulence genes (QS) for enhancing antibiotic resistance and maintaining bacterial homeostasis. Currently, five components are described in the efflux pump family: ATP-binding cassette (ABC) superfamily, major facilitator superfamily (MFS), and multidrug, toxic compound extrusion (MATE) family, resistance-nodulation-division (RND) family, and small multidrug resistance (SMR) family.239 Among all efflux pumps, the RND efflux pump family has the strongest correlation with antibiotic resistance. Twelve RND family efflux pumps are identified in P. aeruginosa: multidrug efflux AB-outer membrane porin M (MexAB-OprM), multidrug efflux CD-outer membrane porin J (MexCD-OprJ), multidrug efflux EF-outer membrane porin N (MexEF-OprN) and multidrug efflux XY-outer membrane porin M (Mex XY-OprM) can increase resistance to a variety of antibiotics through efflux.240 MexAB-OprM is crucial for developing carbapenem-resistant P. aeruginosa, which is a lingering clinical issue.237 Upregulation of MexCD-OprJ is closely associated with increased resistance of most clinical strains to ciprofloxacin, cefepime, and chloramphenicol.241 Quinolones MexEF-OprN are overproduced by the QS deficiency due to kynurenine extrusion.242 Furthermore, the resistance of P. aeruginosa to aminoglycosides, fluoroquinolones, and zwitterionic cephalosporins depends on the efflux function of MexXY-OprM.243 These findings suggest that despite some similarity in substrates, their affinity of efflux pumps may vary greatly, displaying different extent of anti-antimicrobial activities.

Multiple lines of evidence suggest that the chief mechanism for P. aeruginosa success in infection is highly related to its stubborn resistance to antibiotics or other therapeutics, which is regulated by the efflux pump. Hence, targeting this mechanism such as inhibiting the critical efflux pump—mexAB-oprM or enhancing the repressor—mexR, will likely reveal new strategies to overcome antibiotic resistance mechanisms in the bacterium and achieve the improved treatment efficacy.244

Antibiotics often contain chemical bonds (e.g., amides and esters), and bacteria can produce antibiotic-inactivating enzymes (hydrolase) to degrade or alter antibiotics, leading to antibiotic resistance.245 P. aeruginosa is highly resistant to various antibiotics including penicillin, cephalosporin, and aztreonam by producing extended-spectrum β-lactamases (ESBLs).246,247 In addition, the bacterium is also resistant to the cefazolin-tazobactam combination therapy via ESBL GES-6.248 Again, the ESBL of P. aeruginosa is thought to be the most significant mechanism in terms of countering antibiotics, which would be a major target for designing and developing the most potent antimicrobial drugs.

From enzymatic angles, P. aeruginosa can modify the amino groups and glycosidic groups of aminoglycoside antibiotics to produce antibiotic resistance. The currently known aminoglycoside modifying enzymes utilize three common mechanisms: aminoglycoside phosphotransferase (APH), aminoglycoside acetyltransferase (AAC), and aminoglycoside nucleotide transferase (ANT).249,250 These enzymes trigger the resistance of different antibiotics through various mechanisms, providing powerful resistant activities towards different types of antibiotics. APH can inactivate streptomycin by transferring the phosphate group to the 3′-hydroxyl group of aminoglycosides.251,252 AAC may cause gentamicin resistance by transferring the acetyl group to the amino group at the 3′ and 6′ positions of the aminoglycoside.253 ANT confers P. aeruginosa resistance to amikacin by transferring adenosine groups to the amino or hydroxyl groups of these aminoglycosides.254 The diverse enzyme list is growing (currently more than 50 enzymes), which helps in bacterial success in the anti-drug battle with humans.

Bacteria can acquire antibiotic resistance through mutations or horizontal gene transfer.255,256 Mutations of OprD in P. aeruginosa confers resistance to carbapenems234 and mutations of DNA gyrase (GyrA) causes resistance to quinolone antibiotics.257 Importantly, mutations in the β-lactamase gene ampC causes a significant increase in resistance to cephalosporins.258 There are already a host of enzymes in this bacterium may counter antibiotics while it continues to gain new resistance factors, which is debatably the biggest challenge for drug industry and scientific research. As bacteria can conveniently obtain antibiotic resistance genes from the same or different bacterial species through horizontal gene transfer, despite challenging targeting this mechanism may be a niche to search for better treatments.259 A typical example is that P. aeruginosa may obtain aminoglycoside and β-lactam resistance genes through horizontal gene transfer from the environment or other microbes at an unpredictable, fast pace,223,260 it may be highly difficult for scientists and clinicians to design new tools in impeding this natural robust mechanism in this bacterium.

When facing the diverse environmental stimuli, bacteria obtain adaptive resistance to increase the resistance to antibiotics through transient changes in gene and/or protein expression by a spectrum of approaches.261 In P. aeruginosa, the formation of biofilms is the most typical strategy to acquire adaptive antibiotic resistance.262,263 P. aeruginosa can produce biofilms to enhance pathogenicity. Meanwhile, P. aeruginosa infection can also cope with antibiotic treatment by forming persistent cells or persisters, thereby preventing the synthesis of antibiotic targets.223,262,263 Persistent molecules from the persisters can maintain vitality and refill biofilms; once antibiotics are not present, bacteria will resume growth and cause chronic infections.264

Altogether, P. aeruginosa can become exceedingly resistant to antibiotics through myriads of mechanisms, and currently we probably only know a tip of iceberg regarding these bacterial behaviors. It necessitates accelerated research efforts to fully understand the detailed mechanisms by which bacteria constantly grow antibiotic resistance, providing insight into the design and development of innovative and efficious drugs to overcome drug resistance.

The opportunistic pathogen P. aeruginosa exists almost everywhere and in any environmental conditions. In immunosuppressed people, there is extreme susceptibility to P. aeruginosa infection, developing either acute or chronic phenotypes. As the first line of host defense systems, the innate immune system plays a vital role in battling with P. aeruginosa via multiple mechanisms, such as phagocytosis and inflammatory responses. Several types of host systems, such as pattern recognition receptors (PRRs), plasma membrane signals, intracellular enzymes, and secreted cytokines/chemokines participate in inflammatory response against the bacterial infections. Although a well-balanced inflammatory response is required for restraining P. aeruginosa invasion, overzealous inflammation is associated with rapid disease progression, tissue injury, and even death. Some host molecules including cytosolic protein annexin A2 (AnxA2), autophagy-related protein 7 (ATG7), NLRC4, as well as non-coding RNAs (lncRNAs and microRNAs), are also implicated in P. aeruginosa-induced inflammation and/or other aspects of host defense mechanisms,265,266,267 and understanding of the mechanisms of inflammatory response is just beginning to be unfolded.

TLRs are highly conserved transmembrane PRRs, containing leucine-rich repeats and Toll/interleukin-1 receptor (TIR) homolog domains, which rapidly recognize invading microorganisms and elicit innate immunity and inflammatory response upon bacterial invasion.268 TLRs play vital roles in initiating innate immunity for eradicating invading pathogens.269 Upon the sensing of pathogen-associated molecular patterns (PAMPs), TLRs activate NF-κB and AP-1 to mediate inflammatory response signal pathways.270 Correspondingly, TLRs are capable of recognizing different pathogen-associated molecular patterns found in P. aeruginosa. LPS is a major component of the outer membrane of P. aeruginosa, responsible for maintaining bacterial membrane structure and initiating immune response as a major antigenic factor. Research shows that a large amount of hex-acylated lipid A from LPS is isolated from infected patients, which is a strong TLR4 agonist, capable of activating TLR4-dependent inflammatory response.271,272 In addition, several TLR4 co-receptors MD-2 and CD14 are indicated necessary for TLR4 recognition of LPS and TLR4 activation.273 Airway epithelial cells and macrophages, both expressed TLR4 in the cell membrane, serve as the first line of host defense for the initial contact to P. aeruginosa. As TLR4 is the essential trigger of host immunity against P. aeruginosa infection, its deficient mice show higher bacterial burdens and severe lung injury under infection.274 Nevertheless, the TLR4 pathway does not function alone and is very complex, which may be impacted or coregulated by a number of signaling systems including AnxA2,275 autophagy,276,277 inflammasome,278 cGAS,279,280 ion channel proteins,281 DNA repair proteins,282 miRNAs,40,266,283 and lncRNA,229 to name a few.

Apart from TLR4, TLR2 has also been reported as an LPS recognizer, which is capable of recognizing P. aeruginosa‐derived lipoproteins and T3SS effector ExoS.269 In addition, TLR5 is responsible for detecting P. aeruginosa flagellin, which can induce specific inflammasome.284 TLR9 is responsible for detecting P. aeruginosa unmethylated bacterial CpG DNA intracellularly.285 Interestingly, TLR9 is also activated by P. aeruginosa infection-induced mtDNA release (Wang Biao et al., under revision, Immunology).

Although paradoxical at times, TLRs are arguably the single most important PPRs for initiating the initial recognition to mount a robust immune defense against both bacteria and viruses. It is imperative that scientists be focused on verifying the roles of played out by TLRs using a comprehensive approach or systems biology in an unbiased manner in animals as well as human subjects,277 which may provide insight into the disease pathogenesis and suggesting new therapeutic development.

The inflammasome is a multiprotein complex, which is attributed to the production and release of inflammatory cytokines, IL-1β and IL-18.286,287 Recent work reveals that inflammasome is involved in pyroptosis dependent on the cleavage of Gasdermin D, which contributed to the formation of plasma membrane pores, and in turn promoting the release of inflammatory cytokines and pyroptosis.288,289 Typically, inflammasome consists of cytosolic PRRs, ASC (an apoptosis-associated speck-like protein containing a CARD), and caspase 1. Inflammasome PRRs are responsible for detecting exogenous PAMPs like TLRs, which are also essential for monitoring P. aeruginosa invasion and activate host inflammatory response for promoting the clearance of P. aeruginosa.269,290 Remarkably, TLRs are involved in the priming of inflammasome activation by promoting the transcription of inflammasome-related genes.290 NLRC4 and AIM2 (absent in melanoma 2) are well characterized among numerous inflammasome PRRs linked to infection of P. aeruginosa.

NLRC4 inflammasome is shown capable of recognizing needle and inner rod (PrgJ) T3SS proteins, independently of T3SS exotoxins, or intracellular flagellin utilizing different murine NAIP as adaptors.291,292,293,294 Unexpectively, only one factor, human NAIP (hNAIP), has been found, homologous to murine NAIP5, responsible for sensing P. aeruginosa T3SS inner-rod protein PscI and needle protein PscF.295 Apart from T3SS, the T4SS pilin is also capable of activating inflammasome independent of NLRC4 and ASC.153 P. aeruginosa-induced mitochondrial dysfunction also promotes the assembly and activation of NLRC4 via T3SS.296 Mitochondrial ROS and release of mitochondrial DNA are key to NLRC4 activation under P. aeruginosa infection, which is also dependent on autophagy.296 Removal of damaged mitochondria blocks the activation of NLRC4.296 However, AIM2 appears to be dispensable for recognizing and promoting P. aeruginosa infection-induced inflammation in most cases.297

Although several P. aeruginosa components are implicated in activating inflammasome-related immune defense, little is known about how the bacterium evades immune response after inflammasome activation.298 Recent research broadens our knowledge that P. aeruginosa takes advantage of bacterial QS-dependent secretant, including proteases, pyocyanin, and 3-oxo-C12-HSL, to inhibit the activation of NLRC4 or NLR family, pyrin domain containing 3 (NLRP3) inflammasomes.298 A further study supports that pyocyanin specifically inhibits activation of the NLRP3 inflammasome (but not NLRC4 and AIM2) through induced excessive oxidation,299 contrary to the positive role of oxidation in NLRP3 activation.300 Hence, the potential role of oxidation in P. aeruginosa infection and inflammasome activation requires further study. In addition, K+ efflux is necessary for the activation of inflammasome against P. aeruginosa infection.301

To date, the role and underlying mechanisms of the inflammasome in host defense against P. aeruginosa remain largely unclear and sometimes paradoxical: some researches show that inflammasome activation enhances host defense to clear P. aeruginosa,302,303 whereas others present opposite results that inflammasome activation triggered severe host tissue damage, which may impact disease progression and mortality.304,305 One explanation for such a mixed response is that infection may involve different pathways dependent on the bacterial strains, model systems, environments, and experimental conditions. Another caveat is that most experiments have not been performed at holistic and/or spatiotemporal levels to evaluate the dynamics, rather a single cell type, specific location, and one or two timepoints. Therefore, enhanced or advanced approaches are needed to further clarify the role and underlying regulatory mechanisms of inflammasome activation in P. aeruginosa infection.277

cGAS is a recently discovered nucleic acid sensor that is initially identified to recognize viruses.306 Typically, activation of cGAS contributes to the induction of inflammasome as a means against viral or intracellular bacterial infection.307 However, the latest research shows that cGAS downstream effector STING may also play an anti-inflammatory role under extracellular bacterial P. aeruginosa infection by inhibiting NF-κB activity.308 More recently, we have found that cGAS may be involved in an uncoupled protein response (UPR) during P. aeruginosa infection. Mechanistically, our studies uncover cGAS as a novel nucleic acids’ sensor to initiate immune responses against P. aeruginosa infection through a canonical pathway involving STING and IRF-1,280,309 suggesting that cGAS pathways may be a critical target for control of this bacterium. It is a new and important function for cGAS since previous reports were primarily focused on viral sensing and intracellular bacterial sensing.

CRISPR-Cas (clustered regularly interspaced short palindromic repeats-CRISPR-associated) is widely existing in bacteria and archaea providing protection against genetic intruders (plasmids), phages (phage) and other parasites.310 To date, two classes, six types of CRISPR-CRISPR-associated (Cas) systems, based on the characteristic of Cas proteins, are present in bacteria.311,312,313 Class I CRISPR-Cas systems (types II, V, and VI systems) rely on multiple CRISPR-Cas protein effector complexes, while Class II CRISPR-Cas systems (types I, III, and IV systems) are dependent on a single CRISPR-Cas effector protein.314,315,316 CRISPR-Cas I-C, type I-E and type I-F systems have been found in many clinically isolated P. aeruginosa.317,318 It is known that CRISPR-Cas systems provide adaptive immunity against the invasion of bacteriophages or plasmids.319 However, the role of CRISPR-Cas systems is far beyond the adaptive immunity of bacteria based on the current reports.320 Our research showed that type I CRISPR-Cas targeted the QS regulator LasR to inhibit toll-like receptor 4 recognition, thereby evading mammalian host immunity, suggesting that the CRISPR system is linked to host immunity modulation by targeting their own genes, potentially evading host defense mechanisms.321 Type I CRISPR-Cas systems may elicit inflammasome activation in mammalian hosts by regulating autophagy in P. aeruginosa.278 To ensure maximum CRISPR-Cas function, P. aeruginosa PA14 utilized quorum sensing to activate the expression of cas genes to promote CRISPR adaptation when bacteria are at high risk of phage invasion.322 The adaptability and virulence of P. aeruginosa can be regulated by CRISPR-Cas adaptive immunity based on the biological complexity of microbial communities in natural environments.323 The CRISPR-Cas system actively maintains the virulence of P. aeruginosa by limiting virus-like accessory genomic elements.324 Increased expression of phage-related genes initiates CRISPR-mediated biofilm-specific death of P. aeruginosa.325 CRISPR-Cas systems may help mediate antibiotic resistance to multiple membrane irritants, including enhancing the integrity of the envelope in pathogen Francisella novicida.326 Consequently, various corresponding anti-CRISPR (Acr) proteins have evolved by phages to inhibit bacterial CRISPR systems.315,327,328,329,330,331,332,333 Our previous research identified a series of type VI-A anti-CRISPRs (acrVIA1-7) genes that block the activities of Type VI-A CRISPR-Cas13a system,333 and designed Type III CRISPR endonuclease antivirals for coronaviruses (TEAR-CoV) as an experimental therapeutic to combat SARS-CoV-2 infection,334 suggesting that exploiting the mechanism of CRISPR-mediated adaptive immunity has great potential for treating bacterial and viral infections. However, it is not clear whether CRISPR-Cas systems can regulate antibiotic resistance in P. aeruginosa, which may be studied in future (Fig. 5).

CRISPR-mediated adaptive immunity. Type I-C, type I-E, and type I-F CRISPR-Cas systems have been identified in P. aeruginosa. Type I CRISPR-Cas targeted endogenous LasR gene to decrease TLR4 expression and TLR4-mediated host inflammatory responses. Similarly, type I CRISPR-Cas systems elicited inflammasome activation by promoting mitochondrial-mediated autophagy. Ultimately, CRISPR-mediated adaptive immunity helps P. aeruginosa evade mammalian host immunity

P. aeruginosa have strong ability to develop natural and acquired drug resistance through various mechanisms, including the production of antibiotic inactivating enzymes or antibiotic modifying enzymes, inhibiting the penetration of the drug through the cell wall, changing the target of antibacterial action, and limiting the drug to reach its target and adaptation to the adverse environment.223 Due to the increasing difficulty in treating antibiotic resistance strain infection, the development of the anti-P. aeruginosa therapy depends on targeting the resistance mechanism. Research on the resistance mechanism of P. aeruginosa has been an urgent topic for decades since antibiotic resistance has escalated exceedingly. Even with the intense interest, development of new antibiotics and other therapeutic strategies for P. aeruginosa infections is at a painstakingly slow pace due to the variability and complexity of drug resistance, as well as the lack of a deep understanding of the pathogenic mechanisms for P. aeruginosa. Designing effective therapeutic approaches (including phage therapy, immunotherapy, gene editing therapy, antimicrobial peptides, and vaccine therapies) to counteracting P. aeruginosa invasion has been an increasing urgency, requiring consorted efforts (Fig. 6).

Novel therapeutics for P. aeruginosa. Multidrug-resistant P. aeruginosa poses a major challenge to traditional antibiotics therapeutics, which have limited efficacy and cause serious side effects. Phage therapy, immunotherapy, gene editing therapy, antimicrobial peptides, and vaccine therapies have become the most promising strategy and garnered great expectations to overcome multidrug-resistant bacterial infections. A full-scale network of regulatory understanding of P. aeruginosa virulence is expected to be unveiled, thus, we will be in a much better position for rationale drug design to control Pseudomonas aeruginosa infections

With an alarming rise in pathogens with resistance to existing drugs, a number of new antibiotics have recently entered the antibiotic development pipeline; however, the hope for patients and clinicians is rapidly dwindling once a new antibiotic resistance strain emerges. Hence, we should invest unconventionally research efforts in searching for new treatments of MDR bacteria. Recently developed antimicrobials are discussed below.

The biological activity of substituted guanidines was known in the mid-1930s when a series of guanidines and metformin compounds were found to possess bactericidal and disinfectant properties. Subsequently, many guanidine derivatives were studied for therapeutic purposes. Chin et al., recently reported a polyguanidine polycarbonates with strong antimicrobial activities through a distinctive mechanism that does not intensify drug resistance in multi-drug resistance (MDR) superbugs including methicillin-resistant S. aureus, P. aeruginosa, A. baumannii, K. pneumoniae,335 suggesting that the polyguanidine compounds may be potential antibacterial candidates.

In addition, novel antimicrobial peptides are promising to become the next generation of antimicrobials. The sequences of amino acids with antibiotic nature can be found in insects, soil bacteria, amphibians, plants, and even mammals. With the diversity of species and antimicrobial mechanisms, peptides render a niche as an alternative treatment for the MDR bacteria.336 A variety of new antibiotic formulations for treating P. aeruginosa infection have been tested in the clinic despite mostly at preliminary stages. Plazomicin is a novel semi-synthetic parenteral aminoglycoside that inhibits bacterial protein synthesis, and is shown to inhibit P. aeruginosa growth.337 Plazomicin is assessed in two Phase III randomized controlled trials, with an EPIC trial compared with meropenem in Complicated Urinary Tract Infection (cUTI). Plazomicin demonstrated a composite cure 81.7% (P) vs. 70.1% (M), a difference of 11.6%.337 Similarly, another trial CARE with plazomicin-based or colistin-based combinations to treat infection by carbapenem-resistant Enterobacteriaceae (CRE) shows that therapy with plazomicin-based combinations reduced mortality and complications vs. colistin-based combinations (23.5% vs 50%). This latter (CARE) trial seems more effective than the former (Plazomicin) and is achieved by enhanced sustained microbiological eradication.

Eravacycline is a novel fluorocycline that is evaluated for antimicrobial activity against anaerobic Gram-negative and Gram-positive bacteria. Eravacycline demonstrates potent broad-spectrum activity against 90% of bacterial isolates having a concentration range of ≤0.008 to 2 μg/ml.338 Eravacycline is more effective when an expression of tetracycline-specific efflux and ribosomal protection mechanisms is present. Eravacycline is effective against multi-drug-resistant bacteria towards other antibiotic classes. Eravacycline has shown broad-spectrum antibacterial activity when put unique chemical modifications at C-9 and C-7 of the tetracycline core.338

Baxdela, also known as delafloxacin meglumine, is a broad-spectrum anionic fluoroquinolone and has distinct chemical structures that increase potency in acidic environments. Delafloxacin is known for inhibiting DNA replication and repairs targeting DNA gyrase and topoisomerase IV. Currently, Baxdela is under a Phase III clinical trial for therapy of community-acquired pneumonia.308 Zidebactam (WCK 5222) is a dual mechanism antibiotic in a development phase, which is involved in binding Gram-negative PBP2 to exert β-lactamase inhibition. As an active antibiotic against Enterobacteriacease producing CTX-M-15, Zidebactam has some effects against Enterobacteriaceae and P. aeruginosa.339 Although we only reviewed a small portion of ongoing development of new therapeutics for P. aeruginosa with some hope, development of effective therapeutics to counter the growing drug resistance is still challenging.

Nanoparticles have being tested in a number of therapeutic applications, such as drug delivery, gene delivery, and vaccine delivery in addition to direct killing of bacteria by its membrane piercing ability and induction of ROS. Since the CRISPR/Cas system can target to self-genes within a bacterium, delivery of CRISPR-Cas targeting a vital gene (for survival) by nanoparticles may directly kill the bacterium. Use of nanoparticles for delivering vaccines can also be achieved but preparation of optimal particles that can readily control protein loading and folding remains challenging.340 We will focus on the studies of vaccines and nanoparticles. There are several types of nanoparticles: nanoantibiotics liposomes, polymers, hydrogels, metallic particles, magnetic structures, carbon-based materials, and mesoporous with silica and each has its advantages and disadvantages. Hence, a multivalent nanoparticle strategy is considered a superior means and is used in vaccine delivery with P. aeruginosa, specifically to target T6SS. T6SS can divide into two distinct forms: phage tail-like structure and transmembrane complex, which can be exploited to create multivalent nanoparticle delivery of vaccine antigens. For the delivery of T6SS structural device, many proteins in a single nanoparticle may be delivered simultaneously.

The drawback of using nanoparticle structures is the likelihood that host immune responses may be activated against the delivery system, which can prohibit the repeated delivery.340 Future experiments need to be conducted to control antigen exposure to generate a uniform particle. The advantages of T6SS delivery system can be adopted for other gene targets with different nanoparticles. The advancement of nanotechnology has allowed a variety of nanostructures as a strategy to minimize the undesirable characteristic and natural and synthetic AMPs. Reports have indicated that in nanostructures, peptides present lower cytotoxicity and better efficiency of desired targets. Hence, nanostructures may help produce biomolecules and implementation of vaccines and drugs against extracellular degradation and target treatment. Using two approaches for encapsulating AMPs, nanotechnology can use an indirect approach where a passive delivery occurs involving a conventional nano-delivery system. Direct delivery is involved in an active target for modifying the surface chemistry of the nano-carrier with ligands. When comparing the two approaches, both have their rpos and cons, using the passive system involving the encapsulated peptide results in the best fitting solution.336

As biofilms are a major cause of drug resistance, nanoparticles targeting biofilms are also a promising strategy. Zhang et al. found that magnetite hybrid nanocomplexes can penetrate and disrupt bacterial biofilms by breaking through the biofilm matrix barrier.341 Others also demonstrate that silver nanoparticles (metallic particles) combined with polymyxin B can target P. aeruginosa biofilms with significant effects.342 Biofilms are also an important factor to aggravate cystic fibrosis disease. Hence, exotoxin A (ETA) is encapsulated into poly-lactic-co-glycolic acid (PLGA) nanoparticles, forming the ETA-PLGA nanoparticles, which may function as a continuous immunogen to trigger cellular immunity.343 Therefore, these nanoparticles may be one of the potential vaccine candidates to penetrate biofilms. To date, many nanoparticles-mediated compounds are in clinical development and several of them have been approved for human therapy, especially cancer therapy, hepatic fibrosis, virus infection, fungal infections, hypercholesterolemia, and pneumonia.344,345,346 In particular, the recently approved lipid nanoparticle (LNP) mRNA vaccines for COVID-19 was an acclaimed success in nanoparticle research for contribution to novel vaccines to combat diseases.347

Due to the lack of effective therapy for drug-resistant superbugs, phage therapy has been a viable alternative for bacterial treatment, especially for patients who have been non-responsive to conventional antibiotics. There are two types of bacteriophages, lytic and temperate, only lytic phages are utilized for clinical therapy. Lytic phages adhere to the host cell surface and inject DNA into host cells. Progeny phages release and migrate towards infection sites through targeting specific bacteria. For P. aeruginosa, flagella vaccines have been used together with phage or other therapies and showed the remarkable therapeutic results. However, P. aeruginosa displays heterogeneity and varying clinical outcomes with acute infections in later phases, where bacteria start losing their flagella and pili. A vaccine based on flagella will direct toward motile bacteria instead of established/colonized biofilms-forming strains.348 Advanced techniques have been developed for formulating phage cocktails to improve efficacy in vitro. For example, Forti et al. made a cocktail with six phages: E215, E217, PAK_P1, PYO2, DEV, and PAK_P4 and showed some effects against clinical P. aeruginosa strains.349 A number of studies showed that phage cocktails possess enhanced efficiency in killing P. aeruginosa compared to single phage therapy.350,351 As strains from patients are distinct, personalized phages combined with antibiotics have also been applied for effective and safe therapy in clinics. An excellent such application was reported by Forti et al., who constructed phage cocktails based on the genomic information and host range.349 Clinical trials of phage therapy have shown viable and sustained efficacy in therapy. Marked reduction in bacterial colonies with a single dose of phage therapy has been reported in small studies.352 Previous studies have primarily been on patients with burns. Phages were applied on the wounds, but the minimal effect was observed.353 With P. aeruginosa being a ubiquitous pathogen that infects lungs and many other organs, further research will investigate how phage therapy would help control chronic CF or COPD diseases. Phage neutralizing antibodies that are induced by the body's immune system and the safety concern about exotoxins during phage preparations are among the biggest challenge of phage therapy.

Another limitation is that bacteria can become resistant rapidly to phage therapy due in part to CRISPR-Cas systems, which are a bacterial adaptive immunity system that can degrade invading DNA in a high specificity. Research has been conducted with a CRISPR-Cas system that targets endogenous genes to alter host defense.324 To conquer the bacterial immune response for promoting phage replication, phages should gain access to do the job by evolving an anti-CRISPR system (Acrs). With this mechanism, future goals are to use natural anti-CRISPR or synthetic proteins that can counteract the CRISPR systems.319 A protein called AcrIF was shown to be a potent inhibitor of the CRISPR system in P. aeruginosa. This protein is locked with the Cys complex of CRISPR and diminished the phages’ ability to bind to DNA.354 Further research may be conducted using the same mechanism for bacteria to become resistant to CRISPR. Creating a synthetic phage with desired characteristics and genomic content would be a creative approach.355 Recently, mutations in phage host-range-determining regions (HRDRs) to cope with resistance mutation of bacteria have been implemented as an effective method to prevent Escherichia coli infection.41 Engineered bacteriophages BPsΔ33HTH_HRM10 and D29_HRMGD40 cocktail cured a cystic fibrosis patient with a disseminated drug-resistant Mycobacterium abscessus, which shows that the therapeutics of engineered phages to break the evolutionary constraints, holding great potential to create the next-generation of antibacterial antimicrobials.356,357 In spite of recent progress, the strategy for phage therapy is highly difficult. Through creative design, basic research and clinical testing, phage therapy can be rejuvenated against bacterial infections. Hence, it is promising to combine the phage therapy and anti-CRISPR approach for development of clinically feasible antibacterials.333

Genome editing and nucleic acid-based antibiotics, such as single-stranded DNA (ssDNA), double-stranded DNA (dsDNA), plasmid DNA, and ssRNA358 have emerged as the new types of antimicrobials. For gene editing, there is no doubt that editing with CRISPR systems is of high promise. Since scientists first used Cas9 to edit genomes in mice and human cells in 2013,359 the technology has been thought to have the greatest potential to cure some of the deadly diseases that humans cannot control today, including against bacteria-plagued diseases (i.e., CF), cancer and more importantly hereditary disorders. Although there is no therapy in clinics yet, the advance in off-target control and improved editing efficiency will translate gene therapy from bench to bedside in the foreseeable future. Both the gene editing and nucleic acid-based antibiotics need to be delivered into the host and cross spontaneously bacterial cell walls and membranes.

The major strategies for gene delivery are viral and non-viral vector systems. Non-viral strategy includes electroporation (direct injection, in vitro), lipids, and polymers. Research has also shown that LPN (Liposome polymer nucleic acid) is a versatile platform for efficiently delivering diverse nucleic acids to Gram-negative bacteria.358 Intracellular delivery of LPN can be directed against essential genes, resulting in bacterial growth inhibition or death.360 For viral delivery strategy, lentivirus, adenovirus-associated virus (AAV), and adenovirus vectors are often used.361 The lentiviral vector can efficiently integrate a foreign gene into the host chromosome, thereby achieving the effect of persistently expressing the sequence of interest. For some cells that are difficult to transfect, such as primary cells, stem cells, undifferentiated cells, etc., the use of lentiviral vectors has shown versatility to greatly improve the transduction efficiency of the target gene or the target shRNA. Therefore, in vitro and in vivo experiments, lentivirus has become one of the commonly used forms of expression for expressing foreign genes or exogenous shRNAs, and is gaining increasing attention.362 AAV's function mechanism is very similar to lentiviruses. Compared to the lentiviruses, AAV shows better safety, a wide range of host cells, and low immunogenicity, so it has become the most promising gene transfer vector, especially for the CRISPR system.363,364 The AAV-based gene delivery may bear high transduction efficiency for delivering nucleic acid-based antibiotics, leading to better therapeutic effects.

Nip in the bud, vaccines have been a viable approach for preventing millions of people from suffering devastating pandemic infections. Vaccine formulations can be efficiently targeted by antigen-presenting cells to generate both humoral and cellular immune responses.

Only can a few of vaccines that have being tested pass the hurdles of clinical trials to inoculate human populations. A major component of the bacterium targeted by vaccines are LPS and OMP with high rates of variations and mutations, which present challenges in vaccine designs. Vaccines based on these components have shown effects against homologous strains, but no effect against different serotypes. Two strategies have been adopted with LPS-vaccines: the first was to use lipid droplets for carrying LPS components into the host; and the second was stripping the lipid portion off LPS to create serotype-specific protection, which is considered the best option.365 Proteins or genetic vaccines appear to be the best candidates to vaccinate for P. aeruginosa. Emerging strains of P. aeruginosa are a key model system for investigating T4P structure and function, which may provide insight into vaccine design. Research indicates a serious nosocomial pathogen with facile growth requirements or contribution to motility.366 Flagellin and flagella have high immunogenicity, therefore, they are potential targets for vaccination against P. aeruginosa.367 The use of flagella is not only for motility purpose, but also for assisting attachment to the respiratory epithelium and activating TLR5. P. aeruginosa flagellin is a major unit of flagellar filament that is required to target protective antibodies and virulence. Flagella vaccines seem to help target acute infections and provide substantial relief of disease by preventing bacterial invasions and spreads. Klebsiella surface O polysaccharides (OPS) are protective antigens against infection in animals. The development of combined K. pneumoniae and P. aeruginosa glycoconjugate vaccine of four common Klebsiella OPS types may improve the control of human infections. OPS is chemically associated with two flagellin types, FlaA and FlaB.368 Monoclonal antibodies as passive immunity have shown protection of susceptible individuals to delay or prevent initial infections, especially those with cystic fibrosis.369 Increasing vaccines targeting iron chelation, lectin inhibitors, QS inhibition, have been developed. Finally, OMVs are being tested in clinical settings either vaccines or therapeutic carriers.58 Taken together, despite the strong interest in research and urgent need in public health, vaccines against P. aeruginosa are still unavailable for clinical application.

In the past several decades, although human beings have witnessed great scientific advances in basic research of pathogenesis and host defense as well as preventative and therapeutic development with P. aeruginosa, our understanding of this bacterium is still limited, impeding the hunt for novel effective therapeutic in the clinical settings. There are a large number of questions requiring clarification. (1) How do drug-resistant invasive strains emerge in the context with abuse of antibiotics or normal evolution? (2) How many other virulence factors are not discovered and characterized but critical for bacterial pathogenesis and evolution, especially those with remarkable virulence? (3) How the elements of the bacterium trigger and how the bacterium evade the activation, phagocytosis, and clearance by the human immune system? (4) What are new mechanisms responsible for increasing resistance in antimicrobial therapy? (5) How to better identify new host factors for improved immunity against this bacterium? (6) Can the scientific community establish a systematic and all-inclusive guideline to facilitate the discovery and revelation of novel antimicrobials, immune modulators, and other disease-modifying therapeutics?

As an opportunistic pathogen, P. aeruginosa has a complex regulatory system that is closely connected and mutually regulated to cope with the harsh external and internal environment, which causes substantial morbidity, debilitating diseases, shortened life span, and high mortality in humans (Fig. 7). In this comprehensive review and other articles, scientists have discussed the virulence factors of P. aeruginosa, such as LPS, adherence factors, elastase, secretion systems, and OMVs.116,153,212,366,369,370,371,372,373,374,375,376,377,378,379,380,381,382,383,384,385,386,387,388,389,390,391,392,393,394,395,396,397,398,399,400,401,402,403,404,405,406,407,408,409,410,411,412,413,414,415,416,417,418,419,420,421,422,423,424,425,426,427,428,429,430,431,432,433,434,435,436,437,438,439,440,441,442,443,444,445,446,447,448,449,450,451,452,453,454,455 We also introduced the recent progress in new antibiotic formulations and compounds, phage therapy strategy, vaccine approaches, nanoparticle fabrication as well as gene editing and nucleic acid-based antibiotics. Furthermore, we have included a large set of immune responses from hosts, including cell types, innate and adaptive immunity, and emerging advances in immunological research.

Interlinked and mechanistic regulatory network in P. aeruginosa. Interactions between various regulatory systems of P. aeruginosa are linked to regulate adaptation, survival, and resistance to multiple antibiotics, enabling P. aeruginosa to survive environmental stresses. QS systems primarily comprise Las, Rhl, and Pqs and are driven by autoinducer signaling molecules N-(3-oxododecanoyl)-L-homoserine lactone (3-oxo-C12-HSL) N-butanoyl-L-homoserine lactone (C4-HSL), and quinolone signal (PQS) activated through the interaction of transcription factors LasR/RhlR/Pqs and 3-oxo-C12-HSL/C4-HSL/PQS. When signaling molecules reach a putative threshold concentration in the cell environment, QS regulates toxicity expression by regulating two-component systems (TCSs). TCSs regulatory systems, consisting of sensor kinase and response regulator pairs, play roles in bacterial adaptation by regulating the expression of a variety of extracellular enzymes, virulence factors, and QS molecules. GacS/GacA TCS is regulated by sensor kinases RetS (positive regulation) and LadS (negative regulation). Transcription of non-coding regulatory RNAs of the RsmY/Z depends on the activation of GacS/GacA to activate RsmA and regulate the T3SS-mediated virulence secretion and biofilm formation. Biofilms are encapsulated in a self-generating extracellular polymer (EPS) matrix for species survival in surprising alterations of living conditions, like temperature fluctuations and nutrient availability, especially the antibiotic threat. Likewise, the secretory system activates the host immune response through virulence factors. TLRs play a key role in innate immunity. TLR1, 2, 4, 5, 6, and 9 are reported for recognizing P. aeruginosa infection and mediating inflammatory response signal pathways. T3SS and bacterial QS-dependent secretants have roles in modulating NLRP3 and NLRC4 inflammasome activation under P. aeruginosa challenge

Our paper is a rarely extensive review that covers both bacterial pathogenesis and host defense in a great depth, serving as an irreplaceable reference for both students, doctors, and scientists who want to better understand P. aeruginosa. This comprehensive and analytic summary of current literature may enrich our knowledge in the balance between P. aeruginosa invasion and host responses. Despite the vast progress made over the years, a number of questions ranging from basic to clinical and applied aspects remain to be answered and further increased research efforts are still needed to study P. aeruginosa, which will improve our design to more effectively combat the infection caused by emerging drug-resistance strains.

Rossi, E. et al. Pseudomonas aeruginosa adaptation and evolution in patients with cystic fibrosis. Nat. Rev. Microbiol. 19, 331–342 (2021).

Article CAS PubMed Google Scholar

Jurado-Martin, I., Sainz-Mejias, M. & McClean, S. Pseudomonas aeruginosa: an audacious pathogen with an adaptable arsenal of virulence factors. Int. J. Mol. Sci. 22, 1–35 (2021).

Cendra, M. D. M. & Torrents, E. Pseudomonas aeruginosa biofilms and their partners in crime. Biotechnol. Adv. 49, 107734 (2021).

Article CAS PubMed Google Scholar

Sinha, M. et al. Pseudomonas aeruginosa theft biofilm require host lipids of cutaneous wound. Ann. Surg. 5252, 1–23 (2021).

Tang, P. et al. BNT162b2 and mRNA-1273 COVID-19 vaccine effectiveness against the SARS-CoV-2 Delta variant in Qatar. Nat. Med. 27, 2136–2143 (2021).

Article CAS PubMed Google Scholar

Blomquist, K. C. & Nix, D. E. A critical evaluation of newer beta-lactam antibiotics for treatment of Pseudomonas aeruginosa infections. Ann. Pharmacother. 55, 1010–1024 (2021).

Article CAS PubMed Google Scholar

Jangra, V., Sharma, N. & Chhillar, A. K. Therapeutic approaches for combating Pseudomonas aeruginosa Infections. Microbes Infect. 24, 104950 (2022).

Daikos, G. L. et al. Review of ceftazidime-avibactam for the treatment of infections caused by Pseudomonas aeruginosa. Antibiotics. 10, 1–24 (2021).

Botelho, J., Grosso, F. & Peixe, L. Antibiotic resistance in Pseudomonas aeruginosa—mechanisms, epidemiology and evolution. Drug Resist. Updat. 44, 100640 (2019).

Haque, M., Sartelli, M., McKimm, J. & Abu Bakar, M. Health care-associated infections—an overview. Infect. Drug Resist. 11, 2321–2333 (2018).

Article PubMed PubMed Central Google Scholar

Lopez-Calleja, A. I. et al. Antimicrobial activity of ceftolozane-tazobactam against multidrug-resistant and extensively drug-resistant Pseudomonas aeruginosa clinical isolates from a Spanish hospital. Rev. Esp. Quimioter. 32, 68–72 (2019).

CAS PubMed PubMed Central Google Scholar

European Centre for Disease Prevention and Control publishes Annual epidemiological report 2011. Euro. Surveill. 16, 20012 (2011).

McCarthy, K. L. & Paterson, D. L. Increased risk of death with recurrent Pseudomonas aeruginosa bacteremia. Diagn. Microbiol. Infect. Dis. 88, 152–157 (2017).

Article CAS PubMed Google Scholar

Feng, W. et al. Epidemiology and resistance characteristics of Pseudomonas aeruginosa isolates from the respiratory department of a hospital in China. J. Glob. Antimicrob. Resist. 8, 142–147 (2017).

Article PubMed Google Scholar

Xin, X. F., Kvitko, B. & He, S. Y. Pseudomonas syringae: what it takes to be a pathogen. Nat. Rev. Microbiol. 16, 316–328 (2018).

Article CAS PubMed PubMed Central Google Scholar

Vidaillac, C. & Chotirmall, S. H. Pseudomonas aeruginosa in bronchiectasis: infection, inflammation, and therapies. Expert Rev. Respir. Med. 15, 649–662 (2021).

Article CAS PubMed Google Scholar

Park, W. S. et al. Benzyl isothiocyanate attenuates inflammasome activation in Pseudomonas aeruginosa LPS-stimulated THP-1 cells and exerts regulation through the MAPKs/NF-kappaB pathway. Int. J. Mol. Sci. 23, 1–10 (2022).

Chambers, J. R., Cherny, K. E. & Sauer, K. Susceptibility of Pseudomonas aeruginosa dispersed cells to antimicrobial agents is dependent on the dispersion cue and class of the antimicrobial agent used. Antimicrob. Agents Chemother. 61, 1–18 (2017).

Sabnis, A. et al. Colistin kills bacteria by targeting lipopolysaccharide in the cytoplasmic membrane. Elife. 10, 1–26 (2021).

Ozer, E. et al. An inside look at a biofilm: Pseudomonas aeruginosa flagella biotracking. Sci. Adv. 7, 1–15 (2021).

Yang, J. J., Tsuei, K. C. & Shen, E. P. The role of Type III secretion system in the pathogenesis of Pseudomonas aeruginosa microbial keratitis. Tzu. Chi. Med. J. 34, 8–14 (2022).

Karash, S., Nordell, R., Ozer, E. A. & Yahr, T. L. Genome sequences of two Pseudomonas aeruginosa isolates with defects in type III secretion system gene expression from a chronic ankle wound infection. Microbiol. Spectr. 9, e0034021 (2021).

Article PubMed Google Scholar

Alatraktchi, F. A., Svendsen, W. E. & Molin, S. Electrochemical detection of pyocyanin as a biomarker for Pseudomonas aeruginosa: a focused review. Sensors. 20, 1–15 (2020).

Chadha, J., Harjai, K. & Chhibber, S. Revisiting the virulence hallmarks of Pseudomonas aeruginosa: a chronicle through the perspective of quorum sensing. Environ. Microbiol. 00, 1–27 (2021).

Zhao, F., Wang, Q., Zhang, Y. & Lei, L. Anaerobic biosynthesis of rhamnolipids by Pseudomonas aeruginosa: performance, mechanism and its application potential for enhanced oil recovery. Micro. Cell Fact. 20, 103 (2021).

Article CAS Google Scholar

Perraud, Q. et al. Opportunistic use of catecholamine neurotransmitters as siderophores to access iron by Pseudomonas aeruginosa. Environ. Microbiol. 24, 878–893 (2022).

Article CAS PubMed Google Scholar

Dar, H. H. et al. A new thiol-independent mechanism of epithelial host defense against Pseudomonas aeruginosa: iNOS/NO(*) sabotage of theft-ferroptosis. Redox. Biol. 45, 102045 (2021).

Maldonado, R. F., Sa-Correia, I. & Valvano, M. A. Lipopolysaccharide modification in Gram-negative bacteria during chronic infection. FEMS Microbiol. Rev. 40, 480–493 (2016).

Article CAS PubMed PubMed Central Google Scholar

Cross, A. S. Anti-endotoxin vaccines: back to the future. Virulence. 5, 219–225 (2014).

Goldberg, J. B. & Pler, G. B. Pseudomonas aeruginosa lipopolysaccharides and pathogenesis. Trends Microbiol. 4, 490–494 (1996).

Article CAS PubMed Google Scholar

Amarante-Mendes, G. P. et al. Pattern recognition receptors and the host cell death molecular machinery. Front. Immunol. 9, 2379 (2018).

Article PubMed PubMed Central CAS Google Scholar

Mogensen, T. H. Pathogen recognition and inflammatory signaling in innate immune defenses. Clin. Microbiol. Rev. 22, 240–273 (2009).

Article CAS PubMed PubMed Central Google Scholar

Li, D. & Wu, M. Pattern recognition receptors in health and diseases. Signal Transduct. Target Ther. 6, 291 (2021).

Article CAS PubMed PubMed Central Google Scholar

Nadatani, Y. et al. NOD-like receptor protein 3 inflammasome priming and activation in Barrett's epithelial cells. Cell Mol. Gastroenterol. Hepatol. 2, 439–453 (2016).

Article PubMed PubMed Central Google Scholar

Sebastian-Valverde, M. & Pasinetti, G. M. The NLRP3 inflammasome as a critical actor in the inflammaging process. Cells. 9, 1–28 (2020).

Stephens, M. & von der Weid, P. Y. Lipopolysaccharides modulate intestinal epithelial permeability and inflammation in a species-specific manner. Gut Microbes. 11, 421–432 (2020).

Liu, W., Sun, T. & Wang, Y. Integrin alphavbeta6 mediates epithelial-mesenchymal transition in human bronchial epithelial cells induced by lipopolysaccharides of Pseudomonas aeruginosa via TGF-beta1-Smad2/3 signaling pathway. Folia Microbiol. 65, 329–338 (2020).

Article CAS Google Scholar

Ramjeet, M. et al. Truncation of the lipopolysaccharide outer core affects susceptibility to antimicrobial peptides and virulence of Actinobacillus pleuropneumoniae serotype 1. J. Biol. Chem. 280, 39104–39114 (2005).

Article CAS PubMed Google Scholar

Murray, G. L. et al. Mutations affecting Leptospira interrogans lipopolysaccharide attenuate virulence. Mol. Microbiol. 78, 701–709 (2010).

Article CAS PubMed Google Scholar

Li, X. et al. Pseudomonas aeruginosa infection augments inflammation through miR-301b repression of c-Myb-mediated immune activation and infiltration. Nat. Microbiol. 1, 16132 (2016).

Article CAS PubMed PubMed Central Google Scholar

Yehl, K. et al. Engineering phage host-range and suppressing bacterial resistance through phage tail fiber mutagenesis. Cell. 179, 459–469. e459 (2019).

Sartorio, M. G., Pardue, E. J., Feldman, M. F. & Haurat, M. F. Bacterial outer membrane vesicles: from discovery to applications. Annu Rev. Microbiol. 75, 609–630 (2021).

Article PubMed CAS Google Scholar

Kulp, A. & Kuehn, M. J. Biological functions and biogenesis of secreted bacterial outer membrane vesicles. Annu. Rev. Microbiol. 64, 163–184 (2010).

Article CAS PubMed PubMed Central Google Scholar

Bonnington, K. E. & Kuehn, M. J. Protein selection and export via outer membrane vesicles. Biochim. Biophys. Acta. 1843, 1612–1619 (2014).

Kim, J. Y. et al. Engineered bacterial outer membrane vesicles with enhanced functionality. J. Mol. Biol. 380, 51–66 (2008).

Article CAS PubMed PubMed Central Google Scholar

Jan, A. T. Outer membrane vesicles (OMVs) of Gram-negative bacteria: a perspective update. Front. Microbiol. 8, 1053 (2017).

Article PubMed PubMed Central Google Scholar

Bomberger, J. M. et al. Long-distance delivery of bacterial virulence factors by Pseudomonas aeruginosa outer membrane vesicles. PLoS Pathog. 5, e1000382 (2009).

Article PubMed PubMed Central CAS Google Scholar

Kadurugamuwa, J. L. & Beveridge, T. J. Virulence factors are released from Pseudomonas aeruginosa in association with membrane vesicles during normal growth and exposure to gentamicin: a novel mechanism of enzyme secretion. J. Bacteriol. 177, 3998–4008 (1995).

Article CAS PubMed PubMed Central Google Scholar

Kato, S., Kowashi, Y. & Demuth, D. R. Outer membrane-like vesicles secreted by Actinobacillus actinomycetemcomitans are enriched in leukotoxin. Micro. Pathog. 32, 1–13 (2002).

Article CAS Google Scholar

Wai, S. N. et al. Vesicle-mediated export and assembly of pore-forming oligomers of the enterobacterial ClyA cytotoxin. Cell. 115, 25–35 (2003).

Kesty, N. C. & Kuehn, M. J. Incorporation of heterologous outer membrane and periplasmic proteins into Escherichia coli outer membrane vesicles. J. Biol. Chem. 279, 2069–2076 (2004).

Article CAS PubMed Google Scholar

Bomberger, J. M. et al. A Pseudomonas aeruginosa toxin that hijacks the host ubiquitin proteolytic system. PLoS Pathog. 7, e1001325 (2011).

Article CAS PubMed PubMed Central Google Scholar

Bauman, S. J. & Kuehn, M. J. Purification of outer membrane vesicles from Pseudomonas aeruginosa and their activation of an IL-8 response. Microbes Infect. 8, 2400–2408 (2006).

Article CAS PubMed PubMed Central Google Scholar

Koeppen, K. et al. A novel mechanism of host-pathogen interaction through sRNA in bacterial outer membrane vesicles. PLoS Pathog. 12, e1005672 (2016).

Article PubMed PubMed Central CAS Google Scholar

Ballok, A. E. et al. Epoxide-mediated differential packaging of Cif and other virulence factors into outer membrane vesicles. J. Bacteriol. 196, 3633–3642 (2014).

Article PubMed PubMed Central CAS Google Scholar

Stanton, B. A. Effects of Pseudomonas aeruginosa on CFTR chloride secretion and the host immune response. Am. J. Physiol. Cell Physiol. 312, C357–C366 (2017).

Article PubMed PubMed Central Google Scholar

Furuyama, N. & Sircili, M. P. Outer membrane vesicles (OMVs) produced by Gram-negative bacteria: structure, functions, biogenesis, and vaccine application. Biomed. Res. Int. 2021, 1490732 (2021).

Article PubMed PubMed Central CAS Google Scholar

van der Pol, L., Stork, M. & van der Ley, P. Outer membrane vesicles as platform vaccine technology. Biotechnol. J. 10, 1689–1706 (2015).

Article PubMed PubMed Central CAS Google Scholar

Balhuizen, M. D., Veldhuizen, E. J. A. & Haagsman, H. P. Outer membrane vesicle induction and isolation for vaccine development. Front. Microbiol. 12, 629090 (2021).

Article PubMed PubMed Central Google Scholar

Martins, P. et al. Outer membrane vesicles from Neisseria Meningitidis (Proteossome) used for nanostructured Zika virus vaccine production. Sci. Rep. 8, 8290 (2018).

Article PubMed PubMed Central CAS Google Scholar

Rasko, D. A. & Sperandio, V. Anti-virulence strategies to combat bacteria-mediated disease. Nat. Rev. Drug Disco. 9, 117–128 (2010).

Article CAS Google Scholar

Sharma, A. K. et al. Bacterial virulence factors: secreted for survival. Indian J. Microbiol. 57, 1–10 (2017).

Article PubMed Google Scholar

Filloux, A. Protein secretion systems in Pseudomonas aeruginosa: an essay on diversity, evolution, and function. Front. Microbiol. 2, 155 (2011).

Article CAS PubMed PubMed Central Google Scholar

de Sousa, T. et al. Genomic and metabolic characteristics of the pathogenicity in Pseudomonas aeruginosa. Int. J. Mol. Sci. 22, 1–28 (2021).

Thomas, S., Holland, I. B. & Schmitt, L. The Type 1 secretion pathway—the hemolysin system and beyond. Biochim. Biophys. Acta. 1843, 1629–1641 (2014).

Smith, T. J., Sondermann, H. & O’Toole, G. A. Type 1 does the two-step: type 1 secretion substrates with a functional periplasmic intermediate. J. Bacteriol. 200, e00168–18 (2018).

Alav, I. et al. Structure, assembly, and function of tripartite efflux and type 1 secretion systems in Gram-negative bacteria. Chem. Rev. 121, 5479–5596 (2021).

Article CAS PubMed PubMed Central Google Scholar

Cassat, J. E. & Skaar, E. P. Iron in infection and immunity. Cell Host Microbe. 13, 509–519 (2013).

Yahr, T. L., Goranson, J. & Frank, D. W. Exoenzyme S of Pseudomonas aeruginosa is secreted by a type III pathway. Mol. Microbiol. 22, 991–1003 (1996).

Article CAS PubMed Google Scholar

Jimenez, P. N. et al. The multiple signaling systems regulating virulence in Pseudomonas aeruginosa. Microbiol. Mol. Biol. Rev. 76, 46–65 (2012).

Article CAS PubMed Google Scholar

Soscia, C. et al. Cross talk between type III secretion and flagellar assembly systems in Pseudomonas aeruginosa. J. Bacteriol. 189, 3124–3132 (2007).

Article CAS PubMed PubMed Central Google Scholar

Hauser, A. R. The type III secretion system of Pseudomonas aeruginosa: infection by injection. Nat. Rev. Microbiol. 7, 654–665 (2009).

Article CAS PubMed PubMed Central Google Scholar

Coburn, B., Sekirov, I. & Finlay, B. B. Type III secretion systems and disease. Clin. Microbiol. Rev. 20, 535–549 (2007).

Article CAS PubMed PubMed Central Google Scholar

Francis, M. S., Wolf-Watz, H. & Forsberg, A. Regulation of type III secretion systems. Curr. Opin. Microbiol. 5, 166–172 (2002).

Article CAS PubMed Google Scholar

Zhu, M. et al. Modulation of type III secretion system in Pseudomonas aeruginosa: involvement of the PA4857 gene product. Front. Microbiol. 7, 7 (2016).

CAS PubMed PubMed Central Google Scholar

Galle, M., Carpentier, I. & Beyaert, R. Structure and function of the Type III secretion system of Pseudomonas aeruginosa. Curr. Protein Pept. Sci. 13, 831–842 (2012).

Article CAS PubMed PubMed Central Google Scholar

Garai, P. et al. Killing from the inside: intracellular role of T3SS in the fate of Pseudomonas aeruginosa within macrophages revealed by mgtC and oprF mutants. PLoS Pathog. 15, e1007812 (2019).

Article CAS PubMed PubMed Central Google Scholar

Shen, D. K. et al. PsrA is a positive transcriptional regulator of the type III secretion system in Pseudomonas aeruginosa. Infect. Immun. 74, 1121–1129 (2006).

Article CAS PubMed PubMed Central Google Scholar

Li, M. et al. HigB of Pseudomonas aeruginosa enhances killing of phagocytes by up-regulating the type III secretion system in ciprofloxacin induced persister cells. Front. Cell Infect. Microbiol. 6, 125 (2016).

CAS PubMed PubMed Central Google Scholar

Marsden, A. E. et al. Vfr directly activates exsA transcription to regulate expression of the Pseudomonas aeruginosa Type III secretion system. J. Bacteriol. 198, 1442–1450 (2016).

Article CAS PubMed PubMed Central Google Scholar

Intile, P. J., Balzer, G. J., Wolfgang, M. C. & Yahr, T. L. The RNA helicase DeaD stimulates ExsA translation to promote expression of the Pseudomonas aeruginosa Type III secretion system. J. Bacteriol. 197, 2664–2674 (2015).

Article CAS PubMed Google Scholar

Jin, Y., Yang, H., Qiao, M. & Jin, S. MexT regulates the type III secretion system through MexS and PtrC in Pseudomonas aeruginosa. J. Bacteriol. 193, 399–410 (2011).

Article CAS PubMed Google Scholar

Goodman, A. L. et al. A signaling network reciprocally regulates genes associated with acute infection and chronic persistence in Pseudomonas aeruginosa. Dev. Cell. 7, 745–754 (2004).

Intile, P. J. et al. The AlgZR two-component system recalibrates the RsmAYZ posttranscriptional regulatory system to inhibit expression of the Pseudomonas aeruginosa type III secretion system. J. Bacteriol. 196, 357–366 (2014).

Article PubMed PubMed Central CAS Google Scholar

Brencic, A. et al. The GacS/GacA signal transduction system of Pseudomonas aeruginosa acts exclusively through its control over the transcription of the RsmY and RsmZ regulatory small RNAs. Mol. Microbiol. 73, 434–445 (2009).

Article CAS PubMed PubMed Central Google Scholar

Laskowski, M. A., Osborn, E. & Kazmierczak, B. I. A novel sensor kinase-response regulator hybrid regulates type III secretion and is required for virulence in Pseudomonas aeruginosa. Mol. Microbiol. 54, 1090–1103 (2004).

Article CAS PubMed PubMed Central Google Scholar

Ventre, I. et al. Multiple sensors control reciprocal expression of Pseudomonas aeruginosa regulatory RNA and virulence genes. Proc. Natl Acad. Sci. USA. 103, 171–176 (2006).

Goodman, A. L. et al. Direct interaction between sensor kinase proteins mediates acute and chronic disease phenotypes in a bacterial pathogen. Genes Dev. 23, 249–259 (2009).

Article CAS PubMed PubMed Central Google Scholar

Chakravarty, S. et al. Pseudomonas aeruginosa magnesium transporter MgtE inhibits type III secretion system gene expression by stimulating rsmYZ transcription. J. Bacteriol. 199, e00268–17 (2017).

Chen, L., Zou, Y., Kronfl, A. A. & Wu, Y. Type VI secretion system of Pseudomonas aeruginosa is associated with biofilm formation but not environmental adaptation. Microbiologyopen. 9, e991 (2020).

Vettiger, A. & Basler, M. Type VI secretion system substrates are transferred and reused among sister cells. Cell. 167, 99–110. e112 (2016).

Chen, L., Zou, Y., She, P. & Wu, Y. Composition, function, and regulation of T6SS in Pseudomonas aeruginosa. Microbiol. Res. 172, 19–25 (2015).

Article CAS PubMed Google Scholar

Bleves, S. et al. Protein secretion systems in Pseudomonas aeruginosa: a wealth of pathogenic weapons. Int. J. Med. Microbiol. 300, 534–543 (2010).

Article CAS PubMed Google Scholar

Zoued, A. et al. Architecture and assembly of the Type VI secretion system. Biochim Biophys. Acta. 1843, 1664–1673 (2014).

Hood, R. D. et al. A type VI secretion system of Pseudomonas aeruginosa targets a toxin to bacteria. Cell Host Microbe. 7, 25–37 (2010).

Wang, J., Brodmann, M. & Basler, M. Assembly and subcellular localization of bacterial type VI secretion systems. Annu. Rev. Microbiol. 73, 621–638 (2019).

Article CAS PubMed Google Scholar

Basler, M. Type VI secretion system: secretion by a contractile nanomachine. Philos. Trans. R. Soc. Lond. B Biol. Sci. 370, 1–11 (2015).

Berni, B. et al. A type VI secretion system trans-kingdom effector is required for the delivery of a novel antibacterial toxin in Pseudomonas aeruginosa. Front. Microbiol. 10, 1218 (2019).

Article PubMed PubMed Central Google Scholar

Huang, H. et al. An integrated genomic regulatory network of virulence-related transcriptional factors in Pseudomonas aeruginosa. Nat. Commun. 10, 2931 (2019).

Article PubMed PubMed Central CAS Google Scholar

Li, C. et al. T6SS secretes an LPS-binding effector to recruit OMVs for exploitative competition and horizontal gene transfer. ISME J. 16, 500–510 (2022).

Article CAS PubMed Google Scholar

Lesic, B. et al. Quorum sensing differentially regulates Pseudomonas aeruginosa type VI secretion locus I and homologous loci II and III, which are required for pathogenesis. Microbiology. 155, 2845–2855 (2009).

Sana, T. G. et al. The second type VI secretion system of Pseudomonas aeruginosa strain PAO1 is regulated by quorum sensing and Fur and modulates internalization in epithelial cells. J. Biol. Chem. 287, 27095–27105 (2012).

Article CAS PubMed PubMed Central Google Scholar

Maura, D. et al. Evidence for direct control of virulence and defense gene circuits by the Pseudomonas aeruginosa quorum sensing regulator, MvfR. Sci. Rep. 6, 34083 (2016).

Article CAS PubMed PubMed Central Google Scholar

Han, Y. et al. A Pseudomonas aeruginosa type VI secretion system regulated by CueR facilitates copper acquisition. PLoS Pathog. 15, e1008198 (2019).

Article PubMed PubMed Central CAS Google Scholar

Gao, P. et al. oprC impairs host defense by increasing the quorum-sensing-mediated virulence of Pseudomonas aeruginosa. Front. Immunol. 11, 1696 (2020).

Article CAS PubMed PubMed Central Google Scholar

Wang, T. et al. A type VI secretion system delivers a cell wall amidase to target bacterial competitors. Mol. Microbiol. 114, 308–321 (2020).

Article CAS PubMed PubMed Central Google Scholar

Shao, X. et al. Novel therapeutic strategies for treating Pseudomonas aeruginosa infection. Expert Opin. Drug Disco. 15, 1403–1423 (2020).

Article CAS Google Scholar

Horna, G. & Ruiz, J. Type 3 secretion system of Pseudomonas aeruginosa. Microbiol. Res. 246, 126719 (2021).

Article CAS PubMed Google Scholar

Sana, T. G., Berni, B. & Bleves, S. The T6SSs of Pseudomonas aeruginosa strain PAO1 and their effectors: beyond bacterial-cell targeting. Front. Cell Infect. Microbiol. 6, 61 (2016).

Article PubMed PubMed Central CAS Google Scholar

Records, A. R. & Gross, D. C. Sensor kinases RetS and LadS regulate Pseudomonas syringae type VI secretion and virulence factors. J. Bacteriol. 192, 3584–3596 (2010).

Article CAS PubMed PubMed Central Google Scholar

Dadashi, M. et al. Putative RNA ligase RtcB affects the Switch between T6SS and T3SS in Pseudomonas aeruginosa. Int. J. Mol. Sci. 22, 1–21 (2021).

Xia, Y. et al. YbeY controls the type III and type VI secretion systems and biofilm formation through RetS in Pseudomonas aeruginosa. Appl. Environ. Microbiol. 87, e02171–20 (2020).

Diaz, M. R., King, J. M. & Yahr, T. L. Intrinsic and extrinsic regulation of type III secretion gene expression in Pseudomonas Aeruginosa. Front. Microbiol. 2, 89 (2011).

CAS PubMed PubMed Central Google Scholar

Kaminski, A. et al. Pseudomonas aeruginosa ExoS induces intrinsic apoptosis in target host cells in a manner that is dependent on its GAP domain activity. Sci. Rep. 8, 14047 (2018).

Article PubMed PubMed Central CAS Google Scholar

Kroken, A. R. et al. Exotoxin S secreted by internalized Pseudomonas aeruginosa delays lytic host cell death. PLoS Pathog. 18, e1010306 (2022).

Article CAS PubMed PubMed Central Google Scholar

Jia, J., Wang, Y., Zhou, L. & Jin, S. Expression of Pseudomonas aeruginosa toxin ExoS effectively induces apoptosis in host cells. Infect. Immun. 74, 6557–6570 (2006).

Article CAS PubMed PubMed Central Google Scholar

Finck-Barbancon, V. & Frank, D. W. Multiple domains are required for the toxic activity of Pseudomonas aeruginosa ExoU. J. Bacteriol. 183, 4330–4344 (2001).

Article CAS PubMed PubMed Central Google Scholar

Hardy, K. S. et al. ExoU induces lung endothelial cell damage and activates pro-inflammatory caspase-1 during Pseudomonas aeruginosa Infection. Toxins. 14, 152 (2022).

Sutterwala, F. S. et al. Immune recognition of Pseudomonas aeruginosa mediated by the IPAF/NLRC4 inflammasome. J. Exp. Med. 204, 3235–3245 (2007).

Article CAS PubMed PubMed Central Google Scholar

Lindsey, A. S. et al. Analysis of pulmonary vascular injury and repair during Pseudomonas aeruginosa infection-induced pneumonia and acute respiratory distress syndrome. Pulm. Circ. 9, 2045894019826941 (2019).

Article PubMed PubMed Central CAS Google Scholar

Foulkes, D. M. et al. Pseudomonas aeruginosa toxin ExoU as a therapeutic target in the treatment of bacterial infections. Microorganisms. 7, 707 (2019).

Wood, S., Goldufsky, J. & Shafikhani, S. H. Pseudomonas aeruginosa ExoT induces atypical anoikis apoptosis in target host cells by transforming Crk adaptor protein into a cytotoxin. PLoS Pathog. 11, e1004934 (2015).

Article PubMed PubMed Central CAS Google Scholar

Mohamed, M. F. et al. Pseudomonas aeruginosa ExoT induces G1 cell cycle arrest in melanoma cells. Cell Microbiol. 23, e13339 (2021).

Article CAS PubMed Google Scholar

Winsor, G. L. et al. Enhanced annotations and features for comparing thousands of Pseudomonas genomes in the Pseudomonas genome database. Nucleic Acids Res. 44, D646–D653 (2016).

Article CAS PubMed Google Scholar

He, C. et al. Bacterial nucleotidyl cyclase inhibits the host innate immune response by suppressing TAK1 activation. Infect. Immun. 85, e00239–17 (2017).

Jeon, J., Kim, Y. J., Shin, H. & Ha, U. H. T3SS effector ExoY reduces inflammasome-related responses by suppressing bacterial motility and delaying activation of NF-kappaB and caspase-1. FEBS J. 284, 3392–3403 (2017).

Article CAS PubMed Google Scholar

Yang, X., Long, M. & Shen, X. Effector(-)immunity pairs provide the T6SS nanomachine its offensive and defensive capabilities. Molecules. 23, 1009 (2018).

Russell, A. B. et al. Type VI secretion delivers bacteriolytic effectors to target cells. Nature. 475, 343–347 (2011).

Pérez-Lorente, A. I., Molina-Santiago, C., de Vicente, A. & Romero, D. Sporulation activated via σW protects Bacillus from a Tse1 peptidoglycan hydrolase T6SS effector. Preprint at bioRxiv https://doi.org/10.1101/2022.02.23.481616 (2022).

Coulthurst, S. The Type VI secretion system: a versatile bacterial weapon. Microbiology. 165, 503–515 (2019).

Sana, T. G. et al. Internalization of Pseudomonas aeruginosa strain PAO1 into epithelial cells is promoted by interaction of a T6SS effector with the microtubule network. mBio. 6, e00712 (2015).

Kollman, J. M., Merdes, A., Mourey, L. & Agard, D. A. Microtubule nucleation by gamma-tubulin complexes. Nat. Rev. Mol. Cell Biol. 12, 709–721 (2011).

Article CAS PubMed PubMed Central Google Scholar

Kierbel, A., Gassama-Diagne, A., Mostov, K. & Engel, J. N. The phosphoinositol-3-kinase-protein kinase B/Akt pathway is critical for Pseudomonas aeruginosa strain PAK internalization. Mol. Biol. Cell. 16, 2577–2585 (2005).

Naskar, S., Hohl, M., Tassinari, M. & Low, H. H. The structure and mechanism of the bacterial type II secretion system. Mol. Microbiol. 115, 412–424 (2021).

Article CAS PubMed Google Scholar

Lee, P. A., Tullman-Ercek, D. & Georgiou, G. The bacterial twin-arginine translocation pathway. Annu. Rev. Microbiol. 60, 373–395 (2006).

Article PubMed PubMed Central CAS Google Scholar

Nivaskumar, M. & Francetic, O. Type II secretion system: a magic beanstalk or a protein escalator. Biochim. Biophys. Acta. 1843, 1568–1577 (2014).

Green, E. R. & Mecsas, J. Bacterial secretion systems: an overview. Microbiol. Spectr. 4, VMBF-0012-2015 (2016).

Veenendaal, A. K., van der Does, C. & Driessen, A. J. The protein-conducting channel SecYEG. Biochim. Biophys. Acta. 1694, 81–95 (2004).

Goosens, V. J., Monteferrante, C. G. & van Dijl, J. M. The Tat system of Gram-positive bacteria. Biochim. Biophys. Acta. 1843, 1698–1706 (2014).

Filloux, A., Michel, G. & Bally, M. GSP-dependent protein secretion in gram-negative bacteria: the Xcp system of Pseudomonas aeruginosa. FEMS Microbiol. Rev. 22, 177–198 (1998).

Article CAS PubMed Google Scholar

Swietnicki, W. et al. Identification of a potent inhibitor of type II secretion system from Pseudomonas aeruginosa. Biochem. Biophys. Res. Commun. 513, 688–693 (2019).

Article CAS PubMed Google Scholar

Costa, T. R. et al. Secretion systems in Gram-negative bacteria: structural and mechanistic insights. Nat. Rev. Microbiol. 13, 343–359 (2015).

Article CAS PubMed Google Scholar

Meuskens, I., Saragliadis, A., Leo, J. C. & Linke, D. Type V secretion systems: an overview of passenger domain functions. Front. Microbiol. 10, 1163 (2019).

Article PubMed PubMed Central Google Scholar

Stubenrauch, C. J. & Lithgow, T. The TAM: a translocation and assembly module of the beta-barrel assembly machinery in bacterial outer membranes. EcoSal Plus. 8, ESP-0036-2018 (2019).

Leo, J. C., Grin, I. & Linke, D. Type V secretion: mechanism(s) of autotransport through the bacterial outer membrane. Philos. Trans. R. Soc. Lond. B Biol. Sci. 367, 1088–1101 (2012).

Article CAS PubMed PubMed Central Google Scholar

Guerin, J. et al. Two-partner secretion: combining efficiency and simplicity in the secretion of large proteins for bacteria-host and bacteria-bacteria interactions. Front Cell Infect. Microbiol. 7, 148 (2017).

Article PubMed PubMed Central CAS Google Scholar

Lyskowski, A., Leo, J. C. & Goldman, A. Structure and biology of trimeric autotransporter adhesins. Adv. Exp. Med. Biol. 715, 143–158 (2011).

Article CAS PubMed Google Scholar

Leo, J. C., Oberhettinger, P., Schutz, M. & Linke, D. The inverse autotransporter family: intimin, invasin and related proteins. Int. J. Med. Microbiol. 305, 276–282 (2015).

Article CAS PubMed Google Scholar

Juhas, M., Crook, D. W. & Hood, D. W. Type IV secretion systems: tools of bacterial horizontal gene transfer and virulence. Cell Microbiol. 10, 2377–2386 (2008).

Article CAS PubMed PubMed Central Google Scholar

Tammam, S. et al. PilMNOPQ from the Pseudomonas aeruginosa type IV pilus system form a transenvelope protein interaction network that interacts with PilA. J. Bacteriol. 195, 2126–2135 (2013).

Article CAS PubMed PubMed Central Google Scholar

Liu, Y. et al. Transposon insertion sequencing reveals T4SS as the major genetic trait for conjugation transfer of multi-drug resistance pEIB202 from Edwardsiella. BMC Microbiol. 17, 112 (2017).

Article PubMed PubMed Central CAS Google Scholar

Juhas, M. et al. Genomic islands: tools of bacterial horizontal gene transfer and evolution. FEMS Microbiol. Rev. 33, 376–393 (2009).

Article CAS PubMed Google Scholar

Arlehamn, C. S. & Evans, T. J. Pseudomonas aeruginosa pilin activates the inflammasome. Cell Microbiol. 13, 388–401 (2011).

Article CAS PubMed Google Scholar

Yang, D. et al. Paeonol attenuates quorum-sensing regulated virulence and biofilm formation in Pseudomonas aeruginosa. Front. Microbiol. 12, 692474 (2021).

Article PubMed PubMed Central Google Scholar

Elshaer, S. L. & Shaaban, M. I. Inhibition of quorum sensing and virulence factors of Pseudomonas aeruginosa by biologically synthesized gold and selenium nanoparticles. Antibiotics. 10, 1461 (2021).

Fuqua, C. & Greenberg, E. P. Listening in on bacteria: acyl-homoserine lactone signalling. Nat. Rev. Mol. Cell Biol. 3, 685–695 (2002).

Article CAS PubMed Google Scholar

Rutherford, S. T. & Bassler, B. L. Bacterial quorum sensing: its role in virulence and possibilities for its control. Cold Spring Harb. Perspect. Med. 2, a012427 (2012).

Hense, B. A. & Schuster, M. Core principles of bacterial autoinducer systems. Microbiol. Mol. Biol. Rev. 79, 153–169 (2015).

Article CAS PubMed PubMed Central Google Scholar

Gerdt, J. P. & Blackwell, H. E. Competition studies confirm two major barriers that can preclude the spread of resistance to quorum-sensing inhibitors in bacteria. ACS Chem. Biol. 9, 2291–2299 (2014).

Article CAS PubMed PubMed Central Google Scholar

Allen, R. C., Popat, R., Diggle, S. P. & Brown, S. P. Targeting virulence: can we make evolution-proof drugs? Nat. Rev. Microbiol. 12, 300–308 (2014).

Article CAS PubMed Google Scholar

Oh, J., Li, X. H., Kim, S. K. & Lee, J. H. Post-secretional activation of Protease IV by quorum sensing in Pseudomonas aeruginosa. Sci. Rep. 7, 4416 (2017).

Article PubMed PubMed Central CAS Google Scholar

Kaufmann, G. F. et al. Revisiting quorum sensing: discovery of additional chemical and biological functions for 3-oxo-N-acylhomoserine lactones. Proc. Natl Acad. Sci. USA. 102, 309–314 (2005).

Dietrich, L. E. et al. The phenazine pyocyanin is a terminal signalling factor in the quorum sensing network of Pseudomonas aeruginosa. Mol. Microbiol. 61, 1308–1321 (2006).

Article CAS PubMed Google Scholar

Sadikot, R. T., Blackwell, T. S., Christman, J. W. & Prince, A. S. Pathogen-host interactions in Pseudomonas aeruginosa pneumonia. Am. J. Respir. Crit. Care Med. 171, 1209–1223 (2005).

Article PubMed PubMed Central Google Scholar

Blus-Kadosh, I., Zilka, A., Yerushalmi, G. & Banin, E. The effect of pstS and phoB on quorum sensing and swarming motility in Pseudomonas aeruginosa. PLoS One. 8, e74444 (2013).

Cooley, M., Chhabra, S. R. & Williams, P. N-Acylhomoserine lactone-mediated quorum sensing: a twist in the tail and a blow for host immunity. Chem. Biol. 15, 1141–1147 (2008).

Article CAS PubMed Google Scholar

Schuster, M., Sexton, D. J., Diggle, S. P. & Greenberg, E. P. Acyl-homoserine lactone quorum sensing: from evolution to application. Annu. Rev. Microbiol. 67, 43–63 (2013).

Article CAS PubMed Google Scholar

Venturi, V. Regulation of quorum sensing in Pseudomonas. FEMS Microbiol. Rev. 30, 274–291 (2006).

Article CAS PubMed Google Scholar

Welsh, M. A. & Blackwell, H. E. Chemical probes of quorum sensing: from compound development to biological discovery. FEMS Microbiol. Rev. 40, 774–794 (2016).

Article CAS PubMed PubMed Central Google Scholar

Lee, J. & Zhang, L. The hierarchy quorum sensing network in Pseudomonas aeruginosa. Protein Cell. 6, 26–41 (2015).

LaSarre, B. & Federle, M. J. Exploiting quorum sensing to confuse bacterial pathogens. Microbiol. Mol. Biol. Rev. 77, 73–111 (2013).

Article CAS PubMed PubMed Central Google Scholar

Schuster, M. & Greenberg, E. P. Early activation of quorum sensing in Pseudomonas aeruginosa reveals the architecture of a complex regulon. BMC Genomics. 8, 287 (2007).

Fuqua, C. The QscR quorum-sensing regulon of Pseudomonas aeruginosa: an orphan claims its identity. J. Bacteriol. 188, 3169–3171 (2006).

Article CAS PubMed PubMed Central Google Scholar

Liang, H. et al. Molecular mechanisms of master regulator VqsM mediating quorum-sensing and antibiotic resistance in Pseudomonas aeruginosa. Nucleic Acids Res. 42, 10307–10320 (2014).

Article CAS PubMed PubMed Central Google Scholar

Diggle, S. P. et al. Advancing the quorum in Pseudomonas aeruginosa: MvaT and the regulation of N-acylhomoserine lactone production and virulence gene expression. J. Bacteriol. 184, 2576–2586 (2002).

Article CAS PubMed PubMed Central Google Scholar

Castang, S., McManus, H. R., Turner, K. H. & Dove, S. L. H-NS family members function coordinately in an opportunistic pathogen. Proc. Natl Acad. Sci. USA. 105, 18947–18952 (2008).

Rampioni, G. et al. Contribution of the RsaL global regulator to Pseudomonas aeruginosa virulence and biofilm formation. FEMS Microbiol. Lett. 301, 210–217 (2009).

Article CAS PubMed Google Scholar

Rampioni, G. et al. RsaL provides quorum sensing homeostasis and functions as a global regulator of gene expression in Pseudomonas aeruginosa. Mol. Microbiol. 66, 1557–1565 (2007).

Article CAS PubMed Google Scholar

Kang, H. et al. Crystal structure of Pseudomonas aeruginosa RsaL bound to promoter DNA reaffirms its role as a global regulator involved in quorum-sensing. Nucleic Acids Res. 45, 699–710 (2017).

Article CAS PubMed Google Scholar

Balasubramanian, D. et al. The regulatory repertoire of Pseudomonas aeruginosa AmpC ss-lactamase regulator AmpR includes virulence genes. PLoS One. 7, e34067 (2012).

Zhao, J. et al. Structural and molecular mechanism of CdpR involved in quorum-sensing and bacterial virulence in Pseudomonas aeruginosa. PLoS Biol. 14, e1002449 (2016).

Article PubMed PubMed Central CAS Google Scholar

Cao, Q. et al. A novel signal transduction pathway that modulates rhl quorum sensing and bacterial virulence in Pseudomonas aeruginosa. PLoS Pathog. 10, e1004340 (2014).

Article PubMed PubMed Central CAS Google Scholar

Yang, N. et al. The Crc protein participates in down-regulation of the Lon gene to promote rhamnolipid production and rhl quorum sensing in Pseudomonas aeruginosa. Mol. Microbiol. 96, 526–547 (2015).

Article CAS PubMed Google Scholar

Zhang, Y. et al. Pseudomonas aeruginosa regulatory protein AnvM controls pathogenicity in anaerobic environments and impacts host defense. MBio. 10, e01362–01319 (2019).

Siehnel, R. et al. A unique regulator controls the activation threshold of quorum-regulated genes in Pseudomonas aeruginosa. Proc. Natl Acad. Sci. USA. 107, 7916–7921 (2010).

Seet, Q. & Zhang, L. H. Anti-activator QslA defines the quorum sensing threshold and response in Pseudomonas aeruginosa. Mol. Microbiol. 80, 951–965 (2011).

Article CAS PubMed Google Scholar

Ghosh, S. et al. Phytocompound mediated blockage of quorum sensing cascade in ESKAPE pathogens. Antibiotics. 11, 61 (2022).

Piewngam, P., Chiou, J., Chatterjee, P. & Otto, M. Alternative approaches to treat bacterial infections: targeting quorum-sensing. Expert Rev. Anti. Infect. Ther. 18, 499–510 (2020).

Article CAS PubMed Google Scholar

Beier, D. & Gross, R. Regulation of bacterial virulence by two-component systems. Curr. Opin. Microbiol. 9, 143–152 (2006).

Article CAS PubMed Google Scholar

Balasubramanian, D., Schneper, L., Kumari, H. & Mathee, K. A dynamic and intricate regulatory network determines Pseudomonas aeruginosa virulence. Nucleic Acids Res. 41, 1–20 (2013).

Article CAS PubMed Google Scholar

Rodrigue, A. et al. Two-component systems in Pseudomonas aeruginosa: why so many? Trends Microbiol. 8, 498–504 (2000).

Article CAS PubMed Google Scholar

Francis, V. I., Stevenson, E. C. & Porter, S. L. Two-component systems required for virulence in Pseudomonas aeruginosa. FEMS Microbiol. Lett. 364, 1–22 (2017).

Bem, A. E. et al. Bacterial histidine kinases as novel antibacterial drug targets. ACS Chem. Biol. 10, 213–224 (2015).

Article CAS PubMed Google Scholar

Ma, S., Wozniak, D. J. & Ohman, D. E. Identification of the histidine protein kinase KinB in Pseudomonas aeruginosa and its phosphorylation of the alginate regulator algB. J. Biol. Chem. 272, 17952–17960 (1997).

Article CAS PubMed Google Scholar

Goswami, M., Espinasse, A. & Carlson, E. E. Disarming the virulence arsenal of Pseudomonas aeruginosa by blocking two-component system signaling. Chem. Sci. 9, 7332–7337 (2018).

Article CAS PubMed PubMed Central Google Scholar

Yu, L. et al. A novel copper-sensing two-component system for inducing Dsb gene expression in bacteria. Science Bulletin. 67, 198–212 (2022).

Wang, Y. et al. Histamine activates HinK to promote the virulence of Pseudomonas aeruginosa. Science Bulletin. 66, 1101–1118 (2021).

Broder, U. N., Jaeger, T. & Jenal, U. LadS is a calcium-responsive kinase that induces acute-to-chronic virulence switch in Pseudomonas aeruginosa. Nat. Microbiol. 2, 16184 (2016).

Article CAS PubMed Google Scholar

Sall, K. M. et al. A gacS deletion in Pseudomonas aeruginosa cystic fibrosis isolate CHA shapes its virulence. PLoS One. 9, e95936 (2014).

Morris, E. R. et al. Structural rearrangement in an RsmA/CsrA ortholog of Pseudomonas aeruginosa creates a dimeric RNA-binding protein, RsmN. Structure. 21, 1659–1671 (2013).

Mikkelsen, H., McMullan, R. & Filloux, A. The Pseudomonas aeruginosa reference strain PA14 displays increased virulence due to a mutation in ladS. PLoS One. 6, e29113 (2011).

Kazmierczak, B. I., Schniederberend, M. & Jain, R. Cross-regulation of Pseudomonas motility systems: the intimate relationship between flagella, pili and virulence. Curr. Opin. Microbiol. 28, 78–82 (2015).

Article CAS PubMed PubMed Central Google Scholar

Rao, F., Yang, Y., Qi, Y. & Liang, Z. X. Catalytic mechanism of cyclic di-GMP-specific phosphodiesterase: a study of the EAL domain-containing RocR from Pseudomonas aeruginosa. J. Bacteriol. 190, 3622–3631 (2008).

Article CAS PubMed PubMed Central Google Scholar

Mikkelsen, H., Hui, K., Barraud, N. & Filloux, A. The pathogenicity island encoded PvrSR/RcsCB regulatory network controls biofilm formation and dispersal in Pseudomonas aeruginosa PA14. Mol. Microbiol. 89, 450–463 (2013).

Article CAS PubMed PubMed Central Google Scholar

Gellatly, S. L. et al. The Pseudomonas aeruginosa PhoP-PhoQ two-component regulatory system is induced upon interaction with epithelial cells and controls cytotoxicity and inflammation. Infect. Immun. 80, 3122–3131 (2012).

Article CAS PubMed PubMed Central Google Scholar

Mulcahy, H., Charron-Mazenod, L. & Lewenza, S. Extracellular DNA chelates cations and induces antibiotic resistance in Pseudomonas aeruginosa biofilms. PLoS Pathog. 4, e1000213 (2008).

Article PubMed PubMed Central CAS Google Scholar

Lewenza, S. Extracellular DNA-induced antimicrobial peptide resistance mechanisms in Pseudomonas aeruginosa. Front. Microbiol. 4, 21 (2013).

Article PubMed PubMed Central Google Scholar

Turner, K. H. et al. Requirements for Pseudomonas aeruginosa acute burn and chronic surgical wound infection. PLoS Genet. 10, e1004518 (2014).

Article PubMed PubMed Central CAS Google Scholar

Hickman, J. W. & Harwood, C. S. Identification of FleQ from Pseudomonas aeruginosa as a c-di-GMP-responsive transcription factor. Mol. Microbiol. 69, 376–389 (2008).

Article CAS PubMed PubMed Central Google Scholar

O’Toole, G. A. & Wong, G. C. Sensational biofilms: surface sensing in bacteria. Curr. Opin. Microbiol. 30, 139–146 (2016).

Article PubMed PubMed Central CAS Google Scholar

Luo, Y. et al. A hierarchical cascade of second messengers regulates Pseudomonas aeruginosa surface behaviors. MBio. 6, e02456–14 (2015).

Cao, Q. et al. Mutation-induced remodeling of the BfmRS two-component system in Pseudomonas aeruginosa clinical isolates. Sci. Signal. 13, eaaz1529 (2020).

Nitzan, M., Rehani, R. & Margalit, H. Integration of Bacterial Small RNAs in Regulatory Networks. Annu Rev. Biophys. 46, 131–148 (2017).

Article CAS PubMed Google Scholar

Mikulik, K. Structure and functional properties of prokaryotic small noncoding RNAs. Folia Microbiol. 48, 443–468 (2003).

Article CAS Google Scholar

Papenfort, K. et al. SigmaE-dependent small RNAs of Salmonella respond to membrane stress by accelerating global omp mRNA decay. Mol. Microbiol. 62, 1674–1688 (2006).

Article CAS PubMed PubMed Central Google Scholar

Klein, G. & Raina, S. Small regulatory bacterial RNAs regulating the envelope stress response. Biochem. Soc. Trans. 45, 417–425 (2017).

Article CAS PubMed PubMed Central Google Scholar

Michaux, C., Verneuil, N., Hartke, A. & Giard, J. C. Physiological roles of small RNA molecules. Microbiology. 160, 1007–1019 (2014).

Pita, T., Feliciano, J. R. & Leitao, J. H. Small noncoding regulatory RNAs from Pseudomonas aeruginosa and Burkholderia cepacia complex. Int. J. Mol. Sci. 19, 3759 (2018).

Sonnleitner, E. et al. Reduced virulence of a hfq mutant of Pseudomonas aeruginosa O1. Micro. Pathog. 35, 217–228 (2003).

Article CAS Google Scholar

Zhao, K., Li, Y., Yue, B. & Wu, M. Genes as early responders regulate quorum-sensing and control bacterial cooperation in Pseudomonas aeruginosa. PLoS One. 9, e101887 (2014).

Gomez-Lozano, M. et al. Diversity of small RNAs expressed in Pseudomonas species. Environ. Microbiol. Rep. 7, 227–236 (2015).

Article CAS PubMed Google Scholar

Lin, P. et al. High-throughput screen reveals sRNAs regulating crRNA biogenesis by targeting CRISPR leader to repress Rho termination. Nat. Commun. 10, 3728 (2019).

Article PubMed PubMed Central CAS Google Scholar

Pang, Z. et al. Antibiotic resistance in Pseudomonas aeruginosa: mechanisms and alternative therapeutic strategies. Biotechnol. Adv. 37, 177–192 (2019).

Article CAS PubMed Google Scholar

Hancock, R. E. & Speert, D. P. Antibiotic resistance in Pseudomonas aeruginosa: mechanisms and impact on treatment. Drug Resist. Updat. 3, 247–255 (2000).

Breidenstein, E. B., de la Fuente-Nunez, C. & Hancock, R. E. Pseudomonas aeruginosa: all roads lead to resistance. Trends Microbiol. 19, 419–426 (2011).

Article CAS PubMed Google Scholar

Lambert, P. A. Mechanisms of antibiotic resistance in Pseudomonas aeruginosa. J. R. Soc. Med. 95, 22–26 (2002).

CAS PubMed PubMed Central Google Scholar

Welte, W., Nestel, U., Wacker, T. & Diederichs, K. Structure and function of the porin channel. Kidney Int. 48, 930–940 (1995).

Article CAS PubMed Google Scholar

Hancock, R. E. & Brinkman, F. S. Function of pseudomonas porins in uptake and efflux. Annu. Rev. Microbiol. 56, 17–38 (2002).

Article CAS PubMed Google Scholar

Angeletti, S. et al. Multi-drug resistant Pseudomonas aeruginosa nosocomial strains: molecular epidemiology and evolution. Micro. Pathog. 123, 233–241 (2018).

Article CAS Google Scholar

Samanta, S. et al. Getting drugs through small pores: exploiting the porins pathway in Pseudomonas aeruginosa. ACS Infect. Dis. 4, 1519–1528 (2018).

Article CAS PubMed Google Scholar

Song, F., Wang, H., Sauer, K. & Ren, D. Cyclic-di-GMP and oprF are involved in the response of Pseudomonas aeruginosa to substrate material stiffness during attachment on polydimethylsiloxane (PDMS). Front. Microbiol. 9, 110 (2018).

Article PubMed PubMed Central Google Scholar

Simm, R. et al. GGDEF and EAL domains inversely regulate cyclic di-GMP levels and transition from sessility to motility. Mol. Microbiol. 53, 1123–1134 (2004).

Article CAS PubMed Google Scholar

Valentini, M. & Filloux, A. Biofilms and Cyclic di-GMP (c-di-GMP) signaling: lessons from Pseudomonas aeruginosa and other bacteria. J. Biol. Chem. 291, 12547–12555 (2016).

Article CAS PubMed PubMed Central Google Scholar

Li, H. et al. Structure and function of OprD protein in Pseudomonas aeruginosa: from antibiotic resistance to novel therapies. Int. J. Med. Microbiol. 302, 63–68 (2012).

Article CAS PubMed Google Scholar

Lee, J. et al. Refinement of OprH-LPS interactions by molecular simulations. Biophys. J. 112, 346–355 (2017).

Article PubMed PubMed Central CAS Google Scholar

Young, M. L., Bains, M., Bell, A. & Hancock, R. E. Role of Pseudomonas aeruginosa outer membrane protein OprH in polymyxin and gentamicin resistance: isolation of an OprH-deficient mutant by gene replacement techniques. Antimicrob. Agents Chemother. 36, 2566–2568 (1992).

Article CAS PubMed PubMed Central Google Scholar

Li, X. Z., Nikaido, H. & Poole, K. Role of mexA-mexB-oprM in antibiotic efflux in Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 39, 1948–1953 (1995).

Article CAS PubMed PubMed Central Google Scholar

Farjah, A. et al. Immunological evaluation of an alginate-based conjugate as a vaccine candidate against Pseudomonas aeruginosa. APMIS. 123, 175–183 (2015).

Li, X. Z., Plesiat, P. & Nikaido, H. The challenge of efflux-mediated antibiotic resistance in Gram-negative bacteria. Clin. Microbiol. Rev. 28, 337–418 (2015).

Article PubMed PubMed Central Google Scholar

Dreier, J. & Ruggerone, P. Interaction of antibacterial compounds with RND e ffl ux pumps in Pseudomonas aeruginosa. Front. Microbiol. 6, 660 (2015).

Article PubMed PubMed Central Google Scholar

Jeannot, K. et al. Resistance and virulence of Pseudomonas aeruginosa clinical strains overproducing the MexCD-OprJ efflux pump. Antimicrob. Agents Chemother. 52, 2455–2462 (2008).

Article CAS PubMed PubMed Central Google Scholar

Llanes, C. et al. Role of the MexEF-OprN efflux system in low-level resistance of Pseudomonas aeruginosa to ciprofloxacin. Antimicrob. Agents Chemother. 55, 5676–5684 (2011).

Article CAS PubMed PubMed Central Google Scholar

Guenard, S. et al. Multiple mutations lead to MexXY-OprM-dependent aminoglycoside resistance in clinical strains of Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 58, 221–228 (2014).

Article PubMed PubMed Central CAS Google Scholar

Biswas, R., Panja, A. S. & Bandopadhyay, R. Molecular mechanism of antibiotic resistance: the untouched area of future hope. Indian J. Microbiol. 59, 254–259 (2019).

Article PubMed PubMed Central CAS Google Scholar

Rafiee, R., Eftekhar, F., Tabatabaei, S. A. & Minaee Tehrani, D. Prevalence of extended-spectrum and metallo beta-lactamase production in AmpC beta-lactamase producing Pseudomonas aeruginosa isolates from burns. Jundishapur J. Microbiol. 7, e16436 (2014).

Article PubMed PubMed Central Google Scholar

Paterson, D. L. & Bonomo, R. A. Extended-spectrum beta-lactamases: a clinical update. Clin. Microbiol. Rev. 18, 657–686 (2005).

Article CAS PubMed PubMed Central Google Scholar

Rawat, D. & Nair, D. Extended-spectrum beta-lactamases in Gram negative bacteria. J. Glob. Infect. Dis. 2, 263–274 (2010).

Article PubMed PubMed Central Google Scholar

Poirel, L. et al. Acquisition of extended-spectrum beta-lactamase GES-6 leading to resistance to ceftolozane-tazobactam combination in Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 63, e01809-18 (2019).

Ramirez, M. S. & Tolmasky, M. E. Aminoglycoside modifying enzymes. Drug Resist. Updat. 13, 151–171 (2010).

Article CAS PubMed PubMed Central Google Scholar

Ratjen, F., Brockhaus, F. & Angyalosi, G. Aminoglycoside therapy against Pseudomonas aeruginosa in cystic fibrosis: a review. J. Cyst. Fibros. 8, 361–369 (2009).

Article CAS PubMed Google Scholar

Hachler, H., Santanam, P. & Kayser, F. H. Sequence and characterization of a novel chromosomal aminoglycoside phosphotransferase gene, aph (3′)-IIb, in Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 40, 1254–1256 (1996).

Article CAS PubMed PubMed Central Google Scholar

Hainrichson, M. et al. Overexpression and initial characterization of the chromosomal aminoglycoside 3′-O-phosphotransferase APH(3’)-IIb from Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 51, 774–776 (2007).

Article CAS PubMed PubMed Central Google Scholar

Poole, K. Aminoglycoside resistance in Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 49, 479–487 (2005).

Article CAS PubMed PubMed Central Google Scholar

Subedi, D., Vijay, A. K. & Willcox, M. Overview of mechanisms of antibiotic resistance in Pseudomonas aeruginosa: an ocular perspective. Clin. Exp. Optom. 101, 162–171 (2018).

Article PubMed Google Scholar

Woodford, N. & Ellington, M. J. The emergence of antibiotic resistance by mutation. Clin. Microbiol. Infect. 13, 5–18 (2007).

Article CAS PubMed Google Scholar

Lerminiaux, N. A. & Cameron, A. D. S. Horizontal transfer of antibiotic resistance genes in clinical environments. Can. J. Microbiol. 65, 34–44 (2019).

Article CAS PubMed Google Scholar

Aldred, K. J., Kerns, R. J. & Osheroff, N. Mechanism of quinolone action and resistance. Biochemistry. 53, 1565–1574 (2014).

Berrazeg, M. et al. Mutations in beta-lactamase AmpC increase resistance of Pseudomonas aeruginosa isolates to antipseudomonal cephalosporins. Antimicrob. Agents Chemother. 59, 6248–6255 (2015).

Article CAS PubMed PubMed Central Google Scholar

Husnik, F. & McCutcheon, J. P. Functional horizontal gene transfer from bacteria to eukaryotes. Nat. Rev. Microbiol. 16, 67–79 (2018).

Article CAS PubMed Google Scholar

Bonomo, R. A. & Szabo, D. Mechanisms of multidrug resistance in Acinetobacter species and Pseudomonas aeruginosa. Clin. Infect. Dis. 43, S49–S56 (2006).

Article CAS PubMed Google Scholar

Motta, S. S., Cluzel, P. & Aldana, M. Adaptive resistance in bacteria requires epigenetic inheritance, genetic noise, and cost of efflux pumps. PLoS One. 10, e0118464 (2015).

Maurice, N. M., Bedi, B. & Sadikot, R. T. Pseudomonas aeruginosa biofilms: host response and clinical implications in lung infections. Am. J. Respir. Cell Mol. Biol. 58, 428–439 (2018).

Article CAS PubMed PubMed Central Google Scholar

Cohen, N. R., Lobritz, M. A. & Collins, J. J. Microbial persistence and the road to drug resistance. Cell Host Microbe. 13, 632–642 (2013).

Fisher, R. A., Gollan, B. & Helaine, S. Persistent bacterial infections and persister cells. Nat. Rev. Microbiol. 15, 453–464 (2017).

Article CAS PubMed Google Scholar

Li, R. et al. Annexin A2 regulates autophagy in Pseudomonas aeruginosa infection through the Akt1-mTOR-ULK1/2 signaling pathway. J. Immunol. 195, 3901–3911 (2015).

Article CAS PubMed Google Scholar

Zhou, X. et al. MicroRNA-302b augments host defense to bacteria by regulating inflammatory responses via feedback to TLR/IRAK4 circuits. Nat. Commun. 5, 3619 (2014).

Article PubMed CAS Google Scholar

Ye, Y. et al. Atg7 deficiency impairs host defense against Klebsiella pneumoniae by impacting bacterial clearance, survival and inflammatory responses in mice. Am. J. Physiol. Lung Cell Mol. Physiol. 307, L355–L363 (2014).

Article CAS PubMed PubMed Central Google Scholar

Imbert, P. R. et al. A Pseudomonas aeruginosa TIR effector mediates immune evasion by targeting UBAP1 and TLR adaptors. EMBO J. 36, 1869–1887 (2017).

Article CAS PubMed PubMed Central Google Scholar

McIsaac, S. M., Stadnyk, A. W. & Lin, T. J. Toll-like receptors in the host defense against Pseudomonas aeruginosa respiratory infection and cystic fibrosis. J. Leukoc. Biol. 92, 977–985 (2012).

Article CAS PubMed Google Scholar

Kawai, T. & Akira, S. Signaling to NF-kappaB by Toll-like receptors. Trends Mol. Med. 13, 460–469 (2007).

Article CAS PubMed Google Scholar

Ernst, R. K. et al. Specific lipopolysaccharide found in cystic fibrosis airway Pseudomonas aeruginosa. Science. 286, 1561–1565 (1999).

Hajjar, A. M. et al. Human Toll-like receptor 4 recognizes host-specific LPS modifications. Nat. Immunol. 3, 354–359 (2002).

Article CAS PubMed Google Scholar

da Silva Correia, J. et al. Lipopolysaccharide is in close proximity to each of the proteins in its membrane receptor complex. transfer from CD14 to TLR4 and MD-2. J. Biol. Chem. 276, 21129–21135 (2001).

Article PubMed Google Scholar

Power, M. R. et al. The development of early host response to Pseudomonas aeruginosa lung infection is critically dependent on myeloid differentiation factor 88 in mice. J. Biol. Chem. 279, 49315–49322 (2004).

Article CAS PubMed Google Scholar

Zhou, C. M. et al. Annexin A2 regulates unfolded protein response via IRE1-XBP1 axis in macrophages during P. aeruginosa infection. J. Leukoc. Biol. 110, 375–384 (2020).

Pu, Q. et al. Atg7 deficiency intensifies inflammasome activation and pyroptosis in Pseudomonas Sepsis. J. Immunol. 198, 3205–3213 (2017).

Article CAS PubMed Google Scholar

Pu, Q. et al. Interaction among inflammasome, autophagy and non-coding RNAs: new horizons for drug. Precis Clin. Med. 2, 166–182 (2019).

Article PubMed PubMed Central Google Scholar

Wu, Q. et al. Bacterial Type I CRISPR-Cas systems influence inflammasome activation in mammalian host by promoting autophagy. Immunology. 158, 240–251 (2019).

Zhou, C. M. et al. Identification of cGAS as an innate immune sensor of extracellular bacterium Pseudomonas aeruginosa. iScience. 24, 101928 (2021).

Qin, S. et al. Small-molecule inhibitor of 8-oxoguanine DNA glycosylase 1 regulates inflammatory responses during Pseudomonas aeruginosa infection. J. Immunol. 205, 2231–2242 (2020).

Article CAS PubMed Google Scholar

Zhou, X. et al. Transient receptor potential channel 1 deficiency impairs host defense and proinflammatory responses to bacterial infection by regulating protein kinase calpha signaling. Mol. Cell Biol. 35, 2729–2739 (2015).

Article CAS PubMed PubMed Central Google Scholar

Wu, M. et al. Host DNA repair proteins in response to Pseudomonas aeruginosa in lung epithelial cells and in mice. Infect. Immun. 79, 75–87 (2011).

Article CAS PubMed Google Scholar

Huang, T. et al. MicroRNA-302/367 cluster impacts host antimicrobial defense via regulation of mitophagic response against Pseudomonas aeruginosa infection. Front. Immunol. 11, 569173 (2020).

Article CAS PubMed PubMed Central Google Scholar

Hayashi, F. et al. The innate immune response to bacterial flagellin is mediated by Toll-like receptor 5. Nature. 410, 1099–1103 (2001).

Barton, G. M., Kagan, J. C. & Medzhitov, R. Intracellular localization of Toll-like receptor 9 prevents recognition of self DNA but facilitates access to viral DNA. Nat. Immunol. 7, 49–56 (2006).

Article CAS PubMed Google Scholar

Martinon, F., Burns, K. & Tschopp, J. The inflammasome: a molecular platform triggering activation of inflammatory caspases and processing of proIL-beta. Mol. Cell. 10, 417–426 (2002).

Sharma, D. & Kanneganti, T. D. The cell biology of inflammasomes: mechanisms of inflammasome activation and regulation. J. Cell Biol. 213, 617–629 (2016).

Article CAS PubMed PubMed Central Google Scholar

Shi, J. et al. Cleavage of GSDMD by inflammatory caspases determines pyroptotic cell death. Nature. 526, 660–665 (2015).

Liu, X. et al. Inflammasome-activated gasdermin D causes pyroptosis by forming membrane pores. Nature. 535, 153–158 (2016).

Man, S. M. & Kanneganti, T. D. Regulation of inflammasome activation. Immunol. Rev. 265, 6–21 (2015).

Article CAS PubMed PubMed Central Google Scholar

Lin, C. K. & Kazmierczak, B. I. Inflammation: a double-edged sword in the response to Pseudomonas aeruginosa infection. J. Innate Immun. 9, 250–261 (2017).

Article CAS PubMed PubMed Central Google Scholar

Kofoed, E. M. & Vance, R. E. Innate immune recognition of bacterial ligands by NAIPs determines inflammasome specificity. Nature. 477, 592–595 (2011).

Zhao, Y. et al. The NLRC4 inflammasome receptors for bacterial flagellin and type III secretion apparatus. Nature. 477, 596–600 (2011).

Miao, E. A. et al. Innate immune detection of the type III secretion apparatus through the NLRC4 inflammasome. Proc. Natl Acad. Sci. USA. 107, 3076–3080 (2010).

Grandjean, T. et al. The human NAIP-NLRC4-inflammasome senses the Pseudomonas aeruginosa T3SS inner-rod protein. Int. Immunol. 29, 377–384 (2017).

Article CAS PubMed Google Scholar

Jabir, M. S. et al. Mitochondrial damage contributes to Pseudomonas aeruginosa activation of the inflammasome and is downregulated by autophagy. Autophagy. 11, 166–182 (2015).

Pang, Z., Sun, G., Junkins, R. D. & Lin, T. J. AIM2 inflammasome is dispensable for the host defense against Pseudomonas aeruginosa infection. Cell Mol. Biol. 61, 63–70 (2015).

CAS PubMed Google Scholar

Yang, J. et al. Bacterial secretant from Pseudomonas aeruginosa dampens inflammasome activation in a quorum sensing-dependent manner. Front. Immunol. 8, 333 (2017).

PubMed PubMed Central Google Scholar

Virreira Winter, S. & Zychlinsky, A. The bacterial pigment pyocyanin inhibits the NLRP3 inflammasome through intracellular reactive oxygen and nitrogen species. J. Biol. Chem. 293, 4893–4900 (2018).

Article PubMed PubMed Central Google Scholar

Abais, J. M. et al. Redox regulation of NLRP3 inflammasomes: ROS as trigger or effector? Antioxid. Redox Signal. 22, 1111–1129 (2015).

Arlehamn, C. S. et al. The role of potassium in inflammasome activation by bacteria. J. Biol. Chem. 285, 10508–10518 (2010).

Article CAS PubMed PubMed Central Google Scholar

Miao, E. A. et al. Pseudomonas aeruginosa activates caspase 1 through Ipaf. Proc. Natl Acad. Sci. USA. 105, 2562–2567 (2008).

Kupz, A. et al. NLRC4 inflammasomes in dendritic cells regulate noncognate effector function by memory CD8(+) T cells. Nat. Immunol. 13, 162–169 (2012).

Article CAS PubMed Google Scholar

Faure, E. et al. Pseudomonas aeruginosa type-3 secretion system dampens host defense by exploiting the NLRC4-coupled inflammasome. Am. J. Respir. Crit. Care Med. 189, 799–811 (2014).

Article CAS PubMed Google Scholar

Cohen, T. S. & Prince, A. S. Activation of inflammasome signaling mediates pathology of acute P. aeruginosa pneumonia. J. Clin. Invest. 123, 1630–1637 (2013).

Article CAS PubMed PubMed Central Google Scholar

Sun, L. et al. Cyclic GMP-AMP synthase is a cytosolic DNA sensor that activates the type I interferon pathway. Science. 339, 786–791 (2013).

Collins, A. C. et al. Cyclic GMP-AMP synthase is an innate immune DNA sensor for Mycobacterium tuberculosis. Cell Host Microbe. 17, 820–828 (2015).

Chen, K. et al. Stimulator of interferon genes promotes host resistance against Pseudomonas aeruginosa keratitis. Front. Immunol. 9, 1225 (2018).

Article PubMed PubMed Central CAS Google Scholar

Zhou, C. et al. Identification of cGAS as an innate immune sensor of extracellular bacterium Pseudomonas aeruginosa. iScience. 24, 101928 (2020).

Zhang, F., Wen, Y. & Guo, X. CRISPR/Cas9 for genome editing: progress, implications and challenges. Hum. Mol. Genet. 23, R40–R46 (2014).

Article CAS PubMed Google Scholar

Makarova, K. S. et al. An updated evolutionary classification of CRISPR-Cas systems. Nat. Rev. Microbiol. 13, 722–736 (2015).

Article CAS PubMed PubMed Central Google Scholar

Koonin, E. V., Makarova, K. S. & Zhang, F. Diversity, classification and evolution of CRISPR-Cas systems. Curr. Opin. Microbiol. 37, 67–78 (2017).

Article CAS PubMed PubMed Central Google Scholar

Stanley, S. Y. & Maxwell, K. L. Phage-encoded anti-CRISPR defenses. Annu. Rev. Genet. 52, 445-464 (2018).

Sternberg, S. H., Richter, H., Charpentier, E. & Qimron, U. Adaptation in CRISPR-Cas systems. Mol. Cell. 61, 797–808 (2016).

Marino, N. D. et al. Discovery of widespread type I and type V CRISPR-Cas inhibitors. Science. 362, 240–242 (2018).

van Belkum, A. et al. Phylogenetic distribution of CRISPR-Cas systems in antibiotic-resistant Pseudomonas aeruginosa. mBio. 6, e01796–01715 (2015).

Vorontsova, D. et al. Foreign DNA acquisition by the I-F CRISPR-Cas system requires all components of the interference machinery. Nucleic Acids Res. 43, 10848–10860 (2015).

Article CAS PubMed PubMed Central Google Scholar

Marraffini, L. A. CRISPR-Cas immunity in prokaryotes. Nature. 526, 55–61 (2015).

Cady, K. C. et al. The CRISPR/Cas adaptive immune system of Pseudomonas aeruginosa mediates resistance to naturally occurring and engineered phages. J. Bacteriol. 194, 5728–5738 (2012).

Article CAS PubMed PubMed Central Google Scholar

Louwen, R. et al. The role of CRISPR-Cas systems in virulence of pathogenic bacteria. Microbiol. Mol. Biol. Rev. 78, 74–88 (2014).

Article PubMed PubMed Central Google Scholar

Li, R. et al. Type I CRISPR-Cas targets endogenous genes and regulates virulence to evade mammalian host immunity. Cell Res. 26, 1273–1287 (2016).

Article CAS PubMed PubMed Central Google Scholar

Hoyland-Kroghsbo, N. M. et al. Quorum sensing controls the Pseudomonas aeruginosa CRISPR-Cas adaptive immune system. Proc. Natl Acad. Sci. USA. 114, 131–135 (2017).

Alseth, E. O. et al. Bacterial biodiversity drives the evolution of CRISPR-based phage resistance. Nature. 574, 549–552 (2019).

Vasquez-Rifo, A. et al. The Pseudomonas aeruginosa accessory genome elements influence virulence towards Caenorhabditis elegans. Genome Biol. 20, 270 (2019).

Article CAS PubMed PubMed Central Google Scholar

Heussler, G. E. et al. Clustered regularly interspaced short palindromic repeat-dependent, biofilm-specific death of Pseudomonas aeruginosa mediated by increased expression of phage-related genes. mBio. 6, e00129–00115 (2015).

Sampson, T. R. et al. A CRISPR-Cas system enhances envelope integrity mediating antibiotic resistance and inflammasome evasion. Proc. Natl Acad. Sci. USA. 111, 11163–11168 (2014).

Pawluk, A. et al. Inactivation of CRISPR-Cas systems by anti-CRISPR proteins in diverse bacterial species. Nat. Microbiol. 1, 16085 (2016).

Article CAS PubMed Google Scholar

Bondy-Denomy, J., Pawluk, A., Maxwell, K. L. & Davidson, A. R. Bacteriophage genes that inactivate the CRISPR/Cas bacterial immune system. Nature. 493, 429–432 (2013).

Pawluk, A. et al. Naturally occurring off-switches for CRISPR-Cas9. Cell. 167, 1829–1838. e1829 (2016).

Harrington, L. B. et al. A broad-spectrum inhibitor of CRISPR-Cas9. Cell. 170, 1224–1233. e1215 (2017).

Rauch, B. J. et al. Inhibition of CRISPR-Cas9 with bacteriophage proteins. Cell. 168, 150–158. e110 (2017).

Watters, K. E. et al. Systematic discovery of natural CRISPR-Cas12a inhibitors. Science. 362, 236–239 (2018).

Lin, P. et al. CRISPR-Cas13 inhibitors block RNA editing in bacteria and mammalian cells. Mol. Cell. 78, 850–861. e855 (2020).

Lin, P. et al. Type III CRISPR-based RNA editing for programmable control of SARS-CoV-2 and human coronaviruses. Nucleic. Acids Res. 50, e47 (2022).

Chin, W. et al. A macromolecular approach to eradicate multidrug resistant bacterial infections while mitigating drug resistance onset. Nat. Commun. 9, 917 (2018).

Article PubMed PubMed Central CAS Google Scholar

Biswaro, L. S. et al. Antimicrobial peptides and nanotechnology, recent advances and challenges. Front. Microbiol. 9, 855 (2018).

Article PubMed PubMed Central Google Scholar

Eljaaly, K. et al. Plazomicin: a novel aminoglycoside for the treatment of resistant Gram-negative bacterial infections. Drugs. 79, 243–269 (2019).

Sutcliffe, J. A., O’Brien, W., Fyfe, C. & Grossman, T. H. Antibacterial activity of eravacycline (TP-434), a novel fluorocycline, against hospital and community pathogens. Antimicrob. Agents Chemother. 57, 5548–5558 (2013).

Article CAS PubMed PubMed Central Google Scholar

Sader, H. S. et al. WCK 5222 (cefepime/zidebactam) antimicrobial activity tested against Gram-negative organisms producing clinically relevant beta-lactamases. J. Antimicrob. Chemother. 72, 1696–1703 (2017).

Article CAS PubMed Google Scholar

Del Tordello, E., Danilchanka, O., McCluskey, A. J. & Mekalanos, J. J. Type VI secretion system sheaths as nanoparticles for antigen display. Proc. Natl Acad. Sci. USA. 113, 3042–3047 (2016).

Zhang, C. et al. Synthesis of magnetite hybrid nanocomplexes to eliminate bacteria and enhance biofilm disruption. Biomater. Sci. 7, 2833–2840 (2019).

Article CAS PubMed Google Scholar

Salman, M. et al. Synergistic effect of silver nanoparticles and polymyxin B against biofilm produced by Pseudomonas aeruginosa isolates of pus samples in vitro. Artif. Cells Nanomed. Biotechnol. 47, 2465–2472 (2019).

Article CAS PubMed Google Scholar

Safari Zanjani, L. et al. Exotoxin A-PLGA nanoconjugate vaccine against Pseudomonas aeruginosa infection: protectivity in murine model. World J. Microbiol. Biotechnol. 35, 94 (2019).

Article PubMed CAS Google Scholar

Xiao, Y. et al. Engineering nanoparticles for targeted delivery of nucleic acid therapeutics in tumor. Mol. Ther. Methods Clin. Dev. 12, 1–18 (2019).

Article PubMed CAS Google Scholar

Anselmo, A. C. & Mitragotri, S. Nanoparticles in the clinic. Bioeng. Transl. Med. 1, 10–29 (2016).

Article PubMed PubMed Central Google Scholar

Adler-Moore, J. et al. Preclinical safety, tolerability, pharmacokinetics, pharmacodynamics, and antifungal activity of liposomal amphotericin B. Clin. Infect. Dis. 68, S244–S259 (2019).

Article CAS PubMed PubMed Central Google Scholar

Oliver, S. E. et al. The advisory committee on immunization practices’ interim recommendation for use of Pfizer-BioNTech COVID-19 vaccine—United States, December 2020. MMWR Morb. Mortal. Wkly Rep. 69, 1922–1924 (2020).

Article CAS PubMed PubMed Central Google Scholar

Hoggarth, A. et al. Mechanistic research holds promise for bacterial vaccines and phage therapies for Pseudomonas aeruginosa. Drug Des. Devel Ther. 13, 909–924 (2019).

Article CAS PubMed PubMed Central Google Scholar

Forti, F. et al. Design of a broad-range bacteriophage cocktail that reduces Pseudomonas aeruginosa biofilms and treats acute infections in two animal models. Antimicrob. Agents Chemother. 62, e02573–17 (2018).

Hall, A. R. et al. Effects of sequential and simultaneous applications of bacteriophages on populations of Pseudomonas aeruginosa in vitro and in wax moth larvae. Appl. Environ. Microbiol. 78, 5646–5652 (2012).

Article CAS PubMed PubMed Central Google Scholar

Alves, D. R. et al. A novel bacteriophage cocktail reduces and disperses Pseudomonas aeruginosa biofilms under static and flow conditions. Micro. Biotechnol. 9, 61–74 (2016).

Article CAS Google Scholar

Wright, A., Hawkins, C. H., Anggard, E. E. & Harper, D. R. A controlled clinical trial of a therapeutic bacteriophage preparation in chronic otitis due to antibiotic-resistant Pseudomonas aeruginosa; a preliminary report of efficacy. Clin. Otolaryngol. 34, 349–357 (2009).

Article CAS PubMed Google Scholar

Merabishvili, M. et al. Quality-controlled small-scale production of a well-defined bacteriophage cocktail for use in human clinical trials. PLoS One. 4, e4944 (2009).

Niu, Y. et al. A type I-F anti-CRISPR protein inhibits the CRISPR-Cas surveillance complex by ADP-ribosylation. Mol. Cell. 80, 512–524. e515 (2020).

Bikard, D. et al. Exploiting CRISPR-Cas nucleases to produce sequence-specific antimicrobials. Nat. Biotechnol. 32, 1146–1150 (2014).

Article CAS PubMed PubMed Central Google Scholar

Nick, J. A. et al. Host and pathogen response to bacteriophage engineered against Mycobacterium abscessus lung infection. Cell. 185, 1860–1874 (2022).

Dedrick, R. M. et al. Engineered bacteriophages for treatment of a patient with a disseminated drug-resistant Mycobacterium abscessus. Nat. Med. 25, 730–733 (2019).

Article CAS PubMed PubMed Central Google Scholar

Perche, F. et al. Cardiolipin-based lipopolyplex platform for the delivery of diverse nucleic acids into Gram-negative bacteria. Pharmaceuticals. 12, 81 (2019).

Wang, T., Wei, J. J., Sabatini, D. M. & Lander, E. S. Genetic screens in human cells using the CRISPR-Cas9 system. Science. 343, 80–84 (2014).

Tan, X. X., Actor, J. K. & Chen, Y. Peptide nucleic acid antisense oligomer as a therapeutic strategy against bacterial infection: proof of principle using mouse intraperitoneal infection. Antimicrob. Agents Chemother. 49, 3203–3207 (2005).

Article CAS PubMed PubMed Central Google Scholar

Shim, G. et al. Therapeutic gene editing: delivery and regulatory perspectives. Acta Pharm. Sin. 38, 738–753 (2017).

Article CAS Google Scholar

Cockrell, A. S. & Kafri, T. Gene delivery by lentivirus vectors. Mol. Biotechnol. 36, 184–204 (2007).

Article CAS PubMed Google Scholar

Fakhiri, J., Nickl, M. & Grimm, D. Rapid and simple screening of CRISPR guide RNAs (gRNAs) in cultured cells using adeno-associated viral (AAV) vectors. Methods Mol. Biol. 1961, 111–126 (2019).

Article CAS PubMed Google Scholar

Lau, C. H. & Suh, Y. In vivo genome editing in animals using AAV-CRISPR system: applications to translational research of human disease. F1000Res. 6, 2153 (2017).

Article PubMed PubMed Central CAS Google Scholar

Delucia, A. M. et al. Lipopolysaccharide (LPS) inner-core phosphates are required for complete LPS synthesis and transport to the outer membrane in Pseudomonas aeruginosa PAO1. MBio. 2, e00142–11 (2011).

Burrows, L. L. Pseudomonas aeruginosa twitching motility: type IV pili in action. Annu. Rev. Microbiol. 66, 493–520 (2012).

Article CAS PubMed Google Scholar

Campodonico, V. L. et al. Evaluation of flagella and flagellin of Pseudomonas aeruginosa as vaccines. Infect. Immun. 78, 746–755 (2010).

Article CAS PubMed Google Scholar

Hegerle, N. et al. Development of a broad spectrum glycoconjugate vaccine to prevent wound and disseminated infections with Klebsiella pneumoniae and Pseudomonas aeruginosa. PLoS One. 13, e0203143 (2018).

Sharma, A., Krause, A. & Worgall, S. Recent developments for Pseudomonas vaccines. Hum. Vaccin. 7, 999–1011 (2011).

Alionte, L. G. et al. Pseudomonas aeruginosa LasA protease and corneal infections. Curr. Eye Res. 22, 266–271 (2001).

Article CAS PubMed Google Scholar

Allured, V. S., Collier, R. J., Carroll, S. F. & McKay, D. B. Structure of exotoxin A of Pseudomonas aeruginosa at 3.0-Angstrom resolution. Proc. Natl Acad. Sci. USA. 83, 1320–1324 (1986).

Arhin, A. & Boucher, C. The outer membrane protein OprQ and adherence of Pseudomonas aeruginosa to human fibronectin. Microbiol. (Read.). 156, 1415–1423 (2010).

Baumann, U., Wu, S., Flaherty, K. M. & McKay, D. B. Three-dimensional structure of the alkaline protease of Pseudomonas aeruginosa: a two-domain protein with a calcium binding parallel beta roll motif. EMBO J. 12, 3357–3364 (1993).

Article CAS PubMed PubMed Central Google Scholar

Berka, R. M., Gray, G. L. & Vasil, M. L. Studies of phospholipase C (heat-labile hemolysin) in Pseudomonas aeruginosa. Infect. Immun. 34, 1071–1074 (1981).

Article CAS PubMed PubMed Central Google Scholar

Bezzerri, V. et al. Phospholipase C-beta3 is a key modulator of IL-8 expression in cystic fibrosis bronchial epithelial cells. J. Immunol. 186, 4946–4958 (2011).

Article CAS PubMed Google Scholar

Blier, A. S. et al. C-type natriuretic peptide modulates quorum sensing molecule and toxin production in Pseudomonas aeruginosa. Microbiol. 157, 1929–1944 (2011).

Boyd, C. D. et al. Structural features of the Pseudomonas fluorescens biofilm adhesin LapA required for LapG-dependent cleavage, biofilm formation, and cell surface localization. J. Bacteriol. 196, 2775–2788 (2014).

Article PubMed PubMed Central CAS Google Scholar

Broomfield, R. J., Morgan, S. D., Khan, A. & Stickler, D. J. Crystalline bacterial biofilm formation on urinary catheters by urease-producing urinary tract pathogens: a simple method of control. J. Med. Microbiol. 58, 1367–1375 (2009).

Article CAS PubMed Google Scholar

Bulman, Z., Le, P., Hudson, A. O. & Savka, M. A. A novel property of propolis (bee glue): anti-pathogenic activity by inhibition of N-acyl-homoserine lactone mediated signaling in bacteria. J. Ethnopharmacol. 138, 788–797 (2011).

Article CAS PubMed Google Scholar

Burrows, L. L. The therapeutic pipeline for Pseudomonas aeruginosa infections. ACS Infect. Dis. 4, 1041–1047 (2018).

Article CAS PubMed Google Scholar

Cai, X. et al. Structural and functional characterization of Pseudomonas aeruginosa CupB chaperones. PLoS One. 6, e16583 (2011).

Carterson, A. J. et al. The transcriptional regulator AlgR controls cyanide production in Pseudomonas aeruginosa. J. Bacteriol. 186, 6837–6844 (2004).

Article CAS PubMed PubMed Central Google Scholar

Castric, P. A. Glycine metabolism by Pseudomonas aeruginosa: hydrogen cyanide biosynthesis. J. Bacteriol. 130, 826–831 (1977).

Article CAS PubMed PubMed Central Google Scholar

Cathcart, G. R. et al. Novel inhibitors of the Pseudomonas aeruginosa virulence factor LasB: a potential therapeutic approach for the attenuation of virulence mechanisms in pseudomonal infection. Antimicrob. Agents Chemother. 55, 2670–2678 (2011).

Article CAS PubMed PubMed Central Google Scholar

Cota-Gomez, A. et al. PlcR1 and PlcR2 are putative calcium-binding proteins required for secretion of the hemolytic phospholipase C of Pseudomonas aeruginosa. Infect. Immun. 65, 2904–2913 (1997).

Article CAS PubMed PubMed Central Google Scholar

Cowell, B. A., Evans, D. J. & Fleiszig, S. M. Actin cytoskeleton disruption by ExoY and its effects on Pseudomonas aeruginosa invasion. FEMS Microbiol. Lett. 250, 71–76 (2005).

Article CAS PubMed Google Scholar

Cox, C. D. & Adams, P. Siderophore activity of pyoverdin for Pseudomonas aeruginosa. Infect. Immun. 48, 130–138 (1985).

Article CAS PubMed PubMed Central Google Scholar

de Lima, C. D. et al. ExoU activates NF-kappaB and increases IL-8/KC secretion during Pseudomonas aeruginosa infection. PLoS One. 7, e41772 (2012).

de Mattos, K. A., Sarno, E. N., Pessolani, M. C. & Bozza, P. T. Deciphering the contribution of lipid droplets in leprosy: multifunctional organelles with roles in Mycobacterium leprae pathogenesis. Mem. Inst. Oswaldo Cruz. 107, 156–166 (2012).

de Regt, A. K. et al. Overexpression of CupB5 activates alginate overproduction in Pseudomonas aeruginosa by a novel AlgW-dependent mechanism. Mol. Microbiol. 93, 415–425 (2014).

Article PubMed PubMed Central CAS Google Scholar

Ding, J. et al. Structural insights into the Pseudomonas aeruginosa type VI virulence effector Tse1 bacteriolysis and self-protection mechanisms. J. Biol. Chem. 287, 26911–26920 (2012).

Article CAS PubMed PubMed Central Google Scholar

Everett, M. J. & Davies, D. T. Pseudomonas aeruginosa elastase (LasB) as a therapeutic target. Drug Discov. Today. 26, 2108–2123 (2021).

Feldman, M. et al. Role of flagella in pathogenesis of Pseudomonas aeruginosa pulmonary infection. Infect. Immun. 66, 43–51 (1998).

Article CAS PubMed PubMed Central Google Scholar

Feltzer, R. E., Gray, R. D., Dean, W. L. & Pierce, W. M. Jr. Alkaline proteinase inhibitor of Pseudomonas aeruginosa. Interaction of native and N-terminally truncated inhibitor proteins with Pseudomonas metalloproteinases. J. Biol. Chem. 275, 21002–21009 (2000).

Article CAS PubMed Google Scholar

Galdino, A. C. M. et al. Disarming Pseudomonas aeruginosa virulence by the inhibitory action of 1,10-phenanthroline-5,6-dione-based compounds: elastase B (LasB) as a chemotherapeutic target. Front. Microbiol. 10, 1701 (2019).

Article PubMed PubMed Central Google Scholar

Garcia-Contreras, R. et al. Rhamnolipids stabilize quorum sensing mediated cooperation in Pseudomonas aeruginosa. FEMS Microbiol. Lett. 367, fnaa080 (2020).

Garner, A. L. et al. 3-Hydroxy-1-alkyl-2-methylpyridine-4(1H)-thiones: inhibition of the Pseudomonas aeruginosa virulence factor LasB. ACS Med. Chem. Lett. 3, 668–672 (2012).

Article CAS PubMed PubMed Central Google Scholar

Garrity-Ryan, L. et al. The arginine finger domain of ExoT contributes to actin cytoskeleton disruption and inhibition of internalization of Pseudomonas aeruginosa by epithelial cells and macrophages. Infect. Immun. 68, 7100–7113 (2000).

Article CAS PubMed PubMed Central Google Scholar

Geiser, T. K. et al. Pseudomonas aeruginosa ExoT inhibits in vitro lung epithelial wound repair. Cell Microbiol. 3, 223–236 (2001).

Article CAS PubMed Google Scholar

Grande, K. K., Gustin, J. K., Kessler, E. & Ohman, D. E. Identification of critical residues in the propeptide of LasA protease of Pseudomonas aeruginosa involved in the formation of a stable mature protease. J. Bacteriol. 189, 3960–3968 (2007).

Article CAS PubMed PubMed Central Google Scholar

Gupta, R. K., Chhibber, S. & Harjai, K. Acyl homoserine lactones from culture supernatants of Pseudomonas aeruginosa accelerate host immunomodulation. PLoS One. 6, e20860 (2011).

Gutierrez-Gomez, U., Soto-Aceves, M. P., Servin-Gonzalez, L. & Soberon-Chavez, G. Overproduction of rhamnolipids in Pseudomonas aeruginosa PA14 by redirection of the carbon flux from polyhydroxyalkanoate synthesis and overexpression of the rhlAB-R operon. Biotechnol. Lett. 40, 1561–1566 (2018).

Article CAS PubMed Google Scholar

Howe, T. R. & Iglewski, B. H. Isolation and characterization of alkaline protease-deficient mutants of Pseudomonas aeruginosa in vitro and in a mouse eye model. Infect. Immun. 43, 1058–1063 (1984).

Article CAS PubMed PubMed Central Google Scholar

Islamieh, D. I., Afshar, D. & Esmaeili, D. Effect of Satureja khuzistanica essential oil (SKEO) extract on expression of lasA and lasB genes in Pseudomonas aeruginosa. Iran. J. Microbiol. 11, 55–59 (2019).

PubMed PubMed Central Google Scholar

Ivanov, I. E. et al. Atomic force and super-resolution microscopy support a role for LapA as a cell-surface biofilm adhesin of Pseudomonas fluorescens. Res. Microbiol. 163, 685–691 (2012).

Article CAS PubMed PubMed Central Google Scholar

Jepkorir, G. et al. Structural, NMR spectroscopic, and computational investigation of hemin loading in the hemophore HasAp from Pseudomonas aeruginosa. J. Am. Chem. Soc. 132, 9857–9872 (2010).

Article CAS PubMed PubMed Central Google Scholar

Kazmierczak, B. I. & Engel, J. N. Pseudomonas aeruginosa ExoT acts in vivo as a GTPase-activating protein for RhoA, Rac1, and Cdc42. Infect. Immun. 70, 2198–2205 (2002).

Article CAS PubMed PubMed Central Google Scholar

Kida, Y. Roles of Pseudomonas aeruginosa-derived proteases as a virulence factor. Nihon Saikingaku Zasshi. 68, 313–323 (2013).

Kim, D. et al. Identification of arylsulfonamides as ExoU inhibitors. Bioorg. Med. Chem. Lett. 24, 3823–3825 (2014).

Article CAS PubMed Google Scholar

Kim, S. et al. Pseudomonas aeruginosa bacteriophage PA1O requires type IV pili for infection and shows broad bactericidal and biofilm removal activities. Appl. Environ. Microbiol. 78, 6380–6385 (2012).

Article CAS PubMed PubMed Central Google Scholar

Konig, B. et al. Role of Pseudomonas aeruginosa lipase in inflammatory mediator release from human inflammatory effector cells (platelets, granulocytes, and monocytes. Infect Immun. 64, 3252–3258 (1996).

Kumar, R. et al. Replacing the axial ligand tyrosine 75 or its hydrogen bond partner histidine 83 minimally affects hemin acquisition by the hemophore HasAp from Pseudomonas aeruginosa. Biochemistry. 53, 2112–2125 (2014).

Lau, G. W., Hassett, D. J., Ran, H. & Kong, F. The role of pyocyanin in Pseudomonas aeruginosa infection. Trends Mol. Med. 10, 599–606 (2004).

Article CAS PubMed Google Scholar

Li, W. et al. Norepinephrine represses the expression of toxA and the siderophore genes in Pseudomonas aeruginosa. FEMS Microbiol. Lett. 299, 100–109 (2009).

Article CAS PubMed Google Scholar

Liu, P. V. Extracellular toxins of Pseudomonas aeruginosa. J. Infect. Dis. 130, S94–S99 (1974).

Article PubMed Google Scholar

Lu, D. et al. Structural insights into the T6SS effector protein Tse3 and the Tse3-Tsi3 complex from Pseudomonas aeruginosa reveal a calcium-dependent membrane-binding mechanism. Mol. Microbiol. 92, 1092–1112 (2014).

Article CAS PubMed Google Scholar

Miller, L. C. et al. Development of potent inhibitors of pyocyanin production in Pseudomonas aeruginosa. J. Med. Chem. 58, 1298–1306 (2015).

Article CAS PubMed PubMed Central Google Scholar

Mohammadzamani, Z. et al. Inhibitory effects of Cinnamaldehyde, Carvacrol, and honey on the expression of exoS and ampC genes in multidrug-resistant Pseudomonas aeruginosa isolated from burn wound infections. Micro. Pathog. 140, 103946 (2020).

Article CAS Google Scholar

Monds, R. D. et al. Di-adenosine tetraphosphate (Ap4A) metabolism impacts biofilm formation by Pseudomonas fluorescens via modulation of c-di-GMP-dependent pathways. J. Bacteriol. 192, 3011–3023 (2010).

Article CAS PubMed PubMed Central Google Scholar

Moroz, O. V. et al. The structure of a calcium-dependent phosphoinositide-specific phospholipase C from Pseudomonas sp. 62186, the first from a Gram-negative bacterium. Acta Crystallogr D. Struct. Biol. 73, 32–44 (2017).

Article CAS PubMed Google Scholar

Ostroff, R. M., Wretlind, B. & Vasil, M. L. Mutations in the hemolytic-phospholipase C operon result in decreased virulence of Pseudomonas aeruginosa PAO1 grown under phosphate-limiting conditions. Infect. Immun. 57, 1369–1373 (1989).

Article CAS PubMed PubMed Central Google Scholar

Paranchych, W. et al. Fimbriae (pili): molecular basis of Pseudomonas aeruginosa adherence. Clin. Invest. Med. 9, 113–118 (1986).

CAS PubMed Google Scholar

Park, S. & Galloway, D. R. Pseudomonas aeruginosa LasD processes the inactive LasA precursor to the active protease form. Arch. Biochem. Biophys. 357, 8–12 (1998).

Article CAS PubMed Google Scholar

Piepenbrink, K. H. & Sundberg, E. J. Motility and adhesion through type IV pili in Gram-positive bacteria. Biochem Soc. Trans. 44, 1659–1666 (2016).

Article CAS PubMed PubMed Central Google Scholar

Pier, G. B. Pseudomonas aeruginosa lipopolysaccharide: a major virulence factor, initiator of inflammation and target for effective immunity. Int. J. Med. Microbiol. 297, 277–295 (2007).

Article CAS PubMed PubMed Central Google Scholar

Rabin, S. D. & Hauser, A. R. Functional regions of the Pseudomonas aeruginosa cytotoxin ExoU. Infect. Immun. 73, 573–582 (2005).

Article CAS PubMed PubMed Central Google Scholar

Raissy, H. H. et al. A proof of concept study to detect urease producing bacteria in lungs using aerosolized (13)C-urea. Pediatr. Allergy Immunol. Pulmonol. 29, 68–73 (2016).

Article PubMed PubMed Central Google Scholar

Raju, H., Sundararajan, R. & Sharma, R. The structure of BrlR reveals a potential pyocyanin binding site. FEBS Lett. 592, 256–262 (2018).

Article CAS PubMed Google Scholar

Rao, L. et al. Pseudomonas aeruginosa survives in epithelia by ExoS-mediated inhibition of autophagy and mTOR. EMBO Rep. 22, e50613 (2021).

Article CAS PubMed Google Scholar

Rietsch, A., Vallet-Gely, I., Dove, S. L. & Mekalanos, J. J. ExsE, a secreted regulator of type III secretion genes in Pseudomonas aeruginosa. Proc. Natl Acad. Sci. USA. 102, 8006–8011 (2005).

Robb, C. S., Robb, M., Nano, F. E. & Boraston, A. B. The structure of the toxin and type six secretion system substrate Tse2 in Complex with its immunity protein. Structure. 24, 277–284 (2016).

Rosenau, F. et al. Lipase LipC affects motility, biofilm formation and rhamnolipid production in Pseudomonas aeruginosa. FEMS Microbiol. Lett. 309, 25–34 (2010).

CAS PubMed Google Scholar

Saalim, M., Villegas-Moreno, J. & Clark, B. R. Bacterial Alkyl-4-quinolones: discovery, structural diversity and biological properties. Molecules. 25, 5689 (2020).

Sarkisova, S. et al. Calcium-induced virulence factors associated with the extracellular matrix of mucoid Pseudomonas aeruginosa biofilms. J. Bacteriol. 187, 4327–4337 (2005).

Article CAS PubMed PubMed Central Google Scholar

Schultz, M. J. et al. Impairment of host defence by exotoxin A in Pseudomonas aeruginosa pneumonia in mice. J. Med. Microbiol. 50, 822–827 (2001).

Article CAS PubMed Google Scholar

Shiau, J. W. et al. Mice immunized with DNA encoding a modified Pseudomonas aeruginosa exotoxin A develop protective immunity against exotoxin intoxication. Vaccine. 19, 1106–1112 (2000).

Sun, J. & Barbieri, J. T. Pseudomonas aeruginosa ExoT ADP-ribosylates CT10 regulator of kinase (Crk) proteins. J. Biol. Chem. 278, 32794–32800 (2003).

Article CAS PubMed Google Scholar

Sun, J. et al. The Pseudomonas aeruginosa protease LasB directly activates IL-1beta. EBioMedicine. 60, 102984 (2020).

Suzuki, T. et al. Role of pvdE pyoverdine synthesis in Pseudomonas aeruginosa keratitis. Cornea. 37, S99–S105 (2018).

Terada, L. S. et al. Pseudomonas aeruginosa hemolytic phospholipase C suppresses neutrophil respiratory burst activity. Infect. Immun. 67, 2371–2376 (1999).

Article CAS PubMed PubMed Central Google Scholar

Uehara, H. et al. Structures of the heme acquisition protein HasA with iron(III)-5,15-diphenylporphyrin and derivatives thereof as an artificial prosthetic group. Angew. Chem. Int. Ed. Engl. 56, 15279–15283 (2017).

Article CAS PubMed Google Scholar

Urban, A., Leipelt, M., Eggert, T. & Jaeger, K. E. DsbA and DsbC affect extracellular enzyme formation in Pseudomonas aeruginosa. J. Bacteriol. 183, 587–596 (2001).

Article CAS PubMed PubMed Central Google Scholar

van den Berg, B. Crystal structure of a full-length autotransporter. J. Mol. Biol. 396, 627–633 (2010).

Article PubMed CAS Google Scholar

Vareechon, C. et al. Pseudomonas aeruginosa effector ExoS inhibits ROS production in human neutrophils. Cell Host Microbe. 21, 611–618. e615 (2017).

Wang, T. et al. Complex structure of type VI peptidoglycan muramidase effector and a cognate immunity protein. Acta Crystallogr D. Biol. Crystallogr. 69, 1889–1900 (2013).

Wibowo, J. P. et al. A novel mechanism of inhibition by phenylthiourea on PvdP, a tyrosinase synthesizing pyoverdine of Pseudomonas aeruginosa. Int. J. Biol. Macromol. 146, 212–221 (2020).

Article CAS PubMed Google Scholar

Wilderman, P. J. et al. Characterization of an endoprotease (PrpL) encoded by a PvdS-regulated gene in Pseudomonas aeruginosa. Infect. Immun. 69, 5385–5394 (2001).

Article CAS PubMed PubMed Central Google Scholar

Wilhelm, S. et al. The autotransporter esterase EstA of Pseudomonas aeruginosa is required for rhamnolipid production, cell motility, and biofilm formation. J. Bacteriol. 189, 6695–6703 (2007).

Article CAS PubMed PubMed Central Google Scholar

Wurtele, M. et al. Structure of the ExoS GTPase activating domain. FEBS Lett. 491, 26–29 (2001).

Article CAS PubMed Google Scholar

Yarlagadda, V. & Wright, G. D. Membrane-active rhamnolipids overcome aminoglycoside resistance. Cell Chem. Biol. 26, 1333–1334 (2019).

Article CAS PubMed Google Scholar

Yu, H. et al. Elastase LasB of Pseudomonas aeruginosa promotes biofilm formation partly through rhamnolipid-mediated regulation. Can. J. Microbiol. 60, 227–235 (2014).

Article CAS PubMed Google Scholar

Zhang, H., Gao, Z. Q., Su, X. D. & Dong, Y. H. Crystal structure of type VI effector Tse1 from Pseudomonas aeruginosa. FEBS Lett. 586, 3193–3199 (2012).

Article CAS PubMed Google Scholar

Zhao, F. et al. Production of rhamnolipids by Pseudomonas aeruginosa is inhibited by H2S but resumes in a co-culture with P. stutzeri: applications for microbial enhanced oil recovery. Biotechnol. Lett. 37, 1803–1808 (2015).

Article CAS PubMed Google Scholar

Zhao, F. et al. Oxygen effects on rhamnolipids production by Pseudomonas aeruginosa. Micro. Cell Fact. 17, 39 (2018).

Article CAS Google Scholar

Zheng, Z., Ma, D., Yahr, T. L. & Chen, L. The transiently ordered regions in intrinsically disordered ExsE are correlated with structural elements involved in chaperone binding. Biochem. Biophys. Res. Commun. 417, 129–134 (2012).

Article CAS PubMed Google Scholar

Wang, Y. et al. Optimization of antitumor immunotherapy mediated by type III secretion system-based live attenuated bacterial vectors. J. Immunother. 35, 223–234 (2012).

Article CAS PubMed Google Scholar

Lin, Y. M. et al. Outer membrane protein I of Pseudomonas aeruginosa is a target of cationic antimicrobial peptide/protein. J. Biol. Chem. 285, 8985–8994 (2010).

Article CAS PubMed PubMed Central Google Scholar

Cowell, B. A. et al. Mutation of lasA and lasB reduces Pseudomonas aeruginosa invasion of epithelial cells. Microbiology. 149, 2291–2299 (2003).

Barequet, I. S. et al. Pseudomonas aeruginosa LasA protease in treatment of experimental staphylococcal keratitis. Antimicrob. Agents Chemother. 48, 1681–1687 (2004).

Article CAS PubMed PubMed Central Google Scholar

Kilmury, S. L. N. & Burrows, L. L. The Pseudomonas aeruginosa PilSR Two-Component System Regulates Both Twitching and Swimming Motilities. mBio. 9, e01310-18 (2018).

Mikkelsen, H., Ball, G., Giraud, C. & Filloux, A. Expression of Pseudomonas aeruginosa CupD fimbrial genes is antagonistically controlled by RcsB and the EAL-containing PvrR response regulators. PLoS One. 4, e6018 (2009).

Yu, H. et al. Identification of the algZ gene upstream of the response regulator algR and its participation in control of alginate production in Pseudomonas aeruginosa. J. Bacteriol. 179, 187–193 (1997).

Article CAS PubMed PubMed Central Google Scholar

Miller, C. L. et al. RsmW, Pseudomonas aeruginosa small non-coding RsmA-binding RNA upregulated in biofilm versus planktonic growth conditions. BMC Microbiol. 16, 155 (2016).

Article PubMed PubMed Central CAS Google Scholar

Chambonnier, G. et al. The hybrid histidine kinase LadS forms a multicomponent signal transduction system with the GacS/GacA two-component system in Pseudomonas aeruginosa. PLoS Genet. 12, e1006032 (2016).

Article PubMed PubMed Central CAS Google Scholar

Petrova, O. E., Schurr, J. R., Schurr, M. J. & Sauer, K. The novel Pseudomonas aeruginosa two-component regulator BfmR controls bacteriophage-mediated lysis and DNA release during biofilm development through PhdA. Mol. Microbiol. 81, 767–783 (2011).

Article CAS PubMed PubMed Central Google Scholar

Giraud, C. et al. The PprA-PprB two-component system activates CupE, the first non-archetypal Pseudomonas aeruginosa chaperone-usher pathway system assembling fimbriae. Environ. Microbiol. 13, 666–683 (2011).

Article CAS PubMed Google Scholar

Reis, R. S., Pereira, A. G., Neves, B. C. & Freire, D. M. Gene regulation of rhamnolipid production in Pseudomonas aeruginosa—a review. Bioresour. Technol. 102, 6377–6384 (2011).

Article CAS PubMed Google Scholar

Dasgupta, N. et al. A four-tiered transcriptional regulatory circuit controls flagellar biogenesis in Pseudomonas aeruginosa. Mol. Microbiol. 50, 809–824 (2003).

Article CAS PubMed Google Scholar

Marko, V. A., Kilmury, S. L. N., MacNeil, L. T. & Burrows, L. L. Pseudomonas aeruginosa type IV minor pilins and PilY1 regulate virulence by modulating FimS-AlgR activity. PLoS Pathog. 14, e1007074 (2018).

Article PubMed PubMed Central CAS Google Scholar

Cadoret, F., Ball, G., Douzi, B. & Voulhoux, R. Txc, a new type II secretion system of Pseudomonas aeruginosa strain PA7, is regulated by the TtsS/TtsR two-component system and directs specific secretion of the CbpE chitin-binding protein. J. Bacteriol. 196, 2376–2386 (2014).

Article PubMed PubMed Central CAS Google Scholar

Daddaoua, A. et al. GtrS and GltR form a two-component system: the central role of 2-ketogluconate in the expression of exotoxin A and glucose catabolic enzymes in Pseudomonas aeruginosa. Nucleic Acids Res. 42, 7654–7663 (2014).

Article CAS PubMed PubMed Central Google Scholar

Marden, J. N. et al. An unusual CsrA family member operates in series with RsmA to amplify posttranscriptional responses in Pseudomonas aeruginosa. Proc. Natl Acad. Sci. USA. 110, 15055–15060 (2013).

Ryan Kaler, K. M., Nix, J. C. & Schubot, F. D. RetS inhibits Pseudomonas aeruginosa biofilm formation by disrupting the canonical histidine kinase dimerization interface of GacS. J. Biol. Chem. 297, 101193 (2021).

Article CAS PubMed PubMed Central Google Scholar

Heurlier, K. et al. Positive control of swarming, rhamnolipid synthesis, and lipase production by the posttranscriptional RsmA/RsmZ system in Pseudomonas aeruginosa PAO1. J. Bacteriol. 186, 2936–2945 (2004).

Article CAS PubMed PubMed Central Google Scholar

Chen, G. et al. Oligoribonuclease is required for the type III secretion system and pathogenesis of Pseudomonas aeruginosa. Microbiol. Res. 188-189, 90–96 (2016).

Article CAS PubMed Google Scholar

Sultan, M., Arya, R. & Kim, K. K. Roles of two-component systems in Pseudomonas aeruginosa virulence. Int. J. Mol. Sci. 22, 12152 (2021).

Yeung, A. T., Bains, M. & Hancock, R. E. The sensor kinase CbrA is a global regulator that modulates metabolism, virulence, and antibiotic resistance in Pseudomonas aeruginosa. J. Bacteriol. 193, 918–931 (2011).

Article CAS PubMed Google Scholar

Lehman, H. K. & Segal, B. H. The role of neutrophils in host defense and disease. J. Allergy Clin. Immunol. 145, 1535–1544 (2020).

Article CAS PubMed PubMed Central Google Scholar

Shin, H. S. et al. Pseudomonas aeruginosa-dependent upregulation of TLR2 influences host responses to a secondary Staphylococcus aureus infection. Pathog. Dis. 69, 149–156 (2013).

Article CAS PubMed Google Scholar

Download references

This work was supported by National Institutes of Health Grants R01 AI109317-06 and AI138203-3 to M.W. Some icons or graphic element in the Figures (Figs. 1–7) are adapted from BioRender.com (2022). Retrieved from https://app.biorender.com/. Final schematic illustrations were created and integrated by our original design.

These authors contributed equally: Shugang Qin, Wen Xiao, Chuanmin Zhou.

Department of Critical Care Medicine, Frontiers Science Center for Disease-related Molecular Network, State Key Laboratory of Biotherapy and Cancer Center, West China Hospital, Sichuan University, Chengdu, China

Shugang Qin, Wen Xiao, Xiangrong Song & Min Wu

State Key Laboratory of Virology, School of Public Health, Wuhan University, Wuhan, 430071, P.R. China

Chuanmin Zhou

Department of Biomedical Sciences, School of Medicine and Health Sciences, University of North Dakota, Grand Forks, ND, 58203, USA

Chuanmin Zhou, Qinqin Pu & Min Wu

Department of Biomedical Sciences, City University of Hong Kong, Hong Kong, People's Republic of China

Xin Deng

State Key Laboratory of Drug Research, Shanghai Institute of Materia Medica, Chinese Academy of Sciences, Shanghai, 201203, China

Lefu Lan

College of Life Sciences, Northwest University, Xi’an, ShaanXi, 710069, China

Haihua Liang

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

M.W. conceived, supervised, and revised the paper; S.Q. organized figures and formatted the paper; S.Q., X.S., W.X., C.Z., Q.P., X.D., L.L., H.L., and M.W. participated in different parts of writing.

Correspondence to Xiangrong Song or Min Wu.

The authors have no financial conflict of interest. Xiangrong Song and Min Wu are editorial board members of Signal Transduction and Targeted Therapy, but they have not been involved in the process of the manuscript handling.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and Permissions

Qin, S., Xiao, W., Zhou, C. et al. Pseudomonas aeruginosa: pathogenesis, virulence factors, antibiotic resistance, interaction with host, technology advances and emerging therapeutics. Sig Transduct Target Ther 7, 199 (2022). https://doi.org/10.1038/s41392-022-01056-1

Download citation

Received: 01 September 2021

Revised: 04 June 2022

Accepted: 08 June 2022

Published: 25 June 2022

DOI: https://doi.org/10.1038/s41392-022-01056-1

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

Scientific Reports (2023)

International Journal of Peptide Research and Therapeutics (2023)

International Microbiology (2023)

World Journal of Microbiology and Biotechnology (2023)

Aquaculture International (2023)

Envoyer une demande
Envoyer