Next Article in Journal
Untargeted Metabolomics Insights into Newborns with Congenital Zika Infection
Next Article in Special Issue
Sylvatic Canine Morbillivirus in Captive Panthera Highlights Viral Promiscuity and the Need for Better Prevention Strategies
Previous Article in Journal
Virulence of Clinical Candida Isolates
Previous Article in Special Issue
Rabies in Our Neighbourhood: Preparedness for an Emerging Infectious Disease
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Avian Pathogenic Escherichia coli (APEC): An Overview of Virulence and Pathogenesis Factors, Zoonotic Potential, and Control Strategies

Center for Food Animal Health, Department of Animal Sciences, The Ohio State University, Wooster, OH 44691, USA
*
Author to whom correspondence should be addressed.
Pathogens 2021, 10(4), 467; https://doi.org/10.3390/pathogens10040467
Submission received: 22 March 2021 / Revised: 5 April 2021 / Accepted: 9 April 2021 / Published: 12 April 2021
(This article belongs to the Special Issue Epidemiology, Surveillance and Control of Infectious Diseases)

Abstract

:
Avian pathogenic Escherichia coli (APEC) causes colibacillosis in avian species, and recent reports have suggested APEC as a potential foodborne zoonotic pathogen. Herein, we discuss the virulence and pathogenesis factors of APEC, review the zoonotic potential, provide the current status of antibiotic resistance and progress in vaccine development, and summarize the alternative control measures being investigated. In addition to the known virulence factors, several other factors including quorum sensing system, secretion systems, two-component systems, transcriptional regulators, and genes associated with metabolism also contribute to APEC pathogenesis. The clear understanding of these factors will help in developing new effective treatments. The APEC isolates (particularly belonging to ST95 and ST131 or O1, O2, and O18) have genetic similarities and commonalities in virulence genes with human uropathogenic E. coli (UPEC) and neonatal meningitis E. coli (NMEC) and abilities to cause urinary tract infections and meningitis in humans. Therefore, the zoonotic potential of APEC cannot be undervalued. APEC resistance to almost all classes of antibiotics, including carbapenems, has been already reported. There is a need for an effective APEC vaccine that can provide protection against diverse APEC serotypes. Alternative therapies, especially the virulence inhibitors, can provide a novel solution with less likelihood of developing resistance.

1. Introduction

Avian pathogenic Escherichia coli (APEC), an extra-intestinal pathogenic E. coli (ExPEC), causes diverse local and systemic infections in poultry, including chickens, turkeys, ducks, and many other avian species [1]. The most common infections caused by APEC in chickens are perihepatitis, airsacculitis, pericarditis, egg peritonitis, salphingitis, coligranuloma, omphalitis, cellulitis, and osteomyelitis/arthritis; these are commonly referred as avian colibacillosis [2]. APEC also causes swollen head syndrome in chickens and osteomyelitis complex in turkeys [2]. Colibacillosis is one of the leading causes of mortality (up to 20%) and morbidity in poultry and also results in decreased meat (2% decline in live weight, 2.7% deterioration in feed conversion ratio) and egg production (up to 20%), decreased hatching rates, and increased condemnation of carcasses (up to 43%) at slaughter [1,3,4]. Furthermore, APEC is responsible for high mortality (up to 53.5%) in young chickens [4]. Taken together, along with the treatment expenses, APEC costs the poultry industry hundreds of millions of dollars in economic losses worldwide [5]. In the United States (US), it has been estimated that economic losses to the broiler industry can be as high as $40 million annually only due to carcass condemnation [6].
APEC can affect all species of poultry in all types of production systems [3]. APEC is also prevalent (9.52% to 36.73%) in all age groups of chickens [7]. Broiler chickens between the ages of 4 and 6 weeks are more susceptible [1], whereas layer chickens can be affected by APEC throughout the grow and lay periods, particularly around the peak egg production and late lay period [1]. In the US, it is estimated that at least 30% of commercial flocks are affected by APEC at any point of time [8]. Multiple APEC serotypes have been associated with colibacillosis cases in the field outbreaks; however, three serotypes (O78, O2, and O1) account for the majority (more than 80%) of the cases [1,5]. APEC leads to systemic infections in chickens either as a primary pathogen or secondary to viral (infectious bronchitis (IBV), Newcastle disease (NDV), avian influenza (AIV)) and Mycoplasma (Mycoplasma gallisepticum (MG)) infections, immunosuppressive disease (infectious bursal disease (IBD)), or environmental stresses (overcrowding, high level of dust and ammonia) by entering through oral and respiratory routes [1,5]. Interestingly, studies have shown that APEC can colonize the gastrointestinal and respiratory tracts of chickens without causing disease and only translocate to extra-intestinal sites in the presence of stressors (production-related stress, immunosuppression, and concurrent infections) as an opportunistic pathogen [2,9]. APEC invades the gastrointestinal and respiratory tracts through abraded tracheal and intestinal epithelium in the presence of stressors and reaches bloodstream and internal organs [1,2,3]. Chickens get infected through contaminated feed and water and can spread to other birds through the feco-oral or aerosol route [1,2,3]. Furthermore, APEC can be vertically transmitted from infected breeders via contaminated eggs [1,2,3]. An overview of APEC infection in chickens is shown in Scheme 1.
APEC utilizes different virulence and pathogenesis factors to cause disease in chickens, primarily adhesins, invasins, protectins, iron acquisition systems, and toxins [2]. These factors facilitate adhesion, invasion, evasion from the host immune responses, colonization, proliferation, and systemic dissemination of APEC, thereby allowing the establishment of infection in chickens [2]. In addition to these factors, several other bacterial factors including but not limited to secretion systems (type III and VI), quorum sensing (QS) systems, transcriptional regulators, two-component systems, and metabolism-associated genes also contribute to APEC pathogenesis in chickens [10,11,12,13,14,15,16,17,18]. An in-depth understanding of these factors and their roles in APEC pathogenesis will help to develop new effective preventative and therapeutic treatments.
Recent studies suggest APEC (particularly isolates belonging to sequence types ST95 and ST131 or O1, O2, and O18 serogroups) as a potential foodborne zoonotic pathogen as well as a source or reservoir of extra-intestinal infections in humans [4,19,20,21]. Particularly, APEC shares genetic similarity with human ExPECs, uropathogenic E. coli (UPEC), and neonatal meningitis E. coli (NMEC) and possesses UPEC-and NMEC-defining virulence genes with the ability to cause urinary tract infections (UTI) and meningitis in mice and rat models [4,22]. Furthermore, the detection of APEC-specific ColV (colicin V) plasmids in human ExPEC isolates suggests a possible zoonotic transmission of APEC from poultry to humans [21]. Therefore, APEC is a pathogen of importance to the poultry industry and public health.
Antibiotics (tetracyclines, sulfonamides, and aminoglycosides) are frequently used to control colibacillosis in chickens [23]. However, increasing resistance of APEC to different classes of antibiotics, including medically important antibiotics (β-lactams, colistin, and carbapenems), suggests challenges ahead in using antibiotics to control APEC infections in chickens [24]. Furthermore, there is no effective vaccine available to protect chickens against APEC infections, which is mainly due to the diversity of APEC serotypes associated with colibacillosis cases in field outbreaks [5]. Currently, only two vaccines (live-attenuated APEC O78 ΔaroA Poulvac® E. coli vaccine and inactivated Nobilis® E. coli vaccine containing F11 fimbrial and FT flagellar antigens) are commercially available for use in chickens [5,25]. These scenarios necessitate the development of new and alternative therapies to control APEC infections in chickens. Probiotics, bacteriophages, and different new therapies (innate immune stimulants, growth and QS inhibitors, and antimicrobial peptides) have shown promising efficacy in reducing APEC infections in chickens [26,27,28,29,30]; however, none of these has advanced into field applications to date.
Here, we discuss the virulence and pathogenesis factors of APEC and their roles in systemic infections in chickens, review the zoonotic potential, provide the current status of antibiotic resistance in APEC and progress in vaccines development, and summarize the alternative control measures under investigation. The PubMed database was used for advanced search (PubMed Advanced Search Builder) of the articles using key words: “avian pathogenic Escherichia coli”, “virulence”, “pathogenesis”, “human infections”, “control”, “antibiotics”, “antimicrobials”, “resistance”, “vaccine”, “probiotics” and “phages”. Single key word “avian pathogenic Escherichia coli” and two key words consisting of “avian pathogenic Escherichia coli” and either of “virulence”, “pathogenesis”, “human infections”, “control”, “antibiotics”, “antimicrobials”, “resistance”, “vaccine”, “probiotics” or “phages” was used for each advanced search.

2. Virulence and Pathogenesis Factors

APEC possesses or utilizes different virulence and pathogenesis factors or mechanisms to cause colibacillosis in poultry [2,10,11,12,13,14,15,16,17,18,31]. These factors include but are not limited to adhesins, invasins, protectins, iron acquisition systems, toxins, two-component systems, a quorum-sensing (QS) system, transcriptional regulators, secretion systems, and genes associated with metabolism [2,10,11,12,13,14,15,16,17,18,31]. These factors play various roles in APEC infections, including attachment to host cells, invasion of the host cells, survival inside the phagocytic (macrophages) cells, colonization of tissues, persistence in the bloodstream, proliferation/replication inside the cells, cell lysis and damage, sequestering metals from body fluids for growth, resistance to serum bactericidal activity and oxidative and environmental stresses, motility, and biofilm formation [2,10,11,12,13,14,15,16,17,18,31]. Table 1 provides the list of virulence and pathogenesis factors defined or characterized in APEC to date along with their roles in APEC pathogenesis/infection.

2.1. Adhesins

Adhesins are appendages or cell-surface components of bacteria that facilitate adhesion or adherence to other cells or to surfaces, usually in the host they are living in or infecting [2,31]. Adherence is required for colonizing a new host and is an essential step in bacterial pathogenesis or infection [2,31]. Adherence in APEC is facilitated primarily by type 1 fimbriae, P fimbriae, and S fimbriae [2,31]. Several genes encoding these fimbriae and additional adhesins, fimH, fimC (type 1 fimbriae), papA, papC, papEF, papG I, papG II, papGIII, felA (P fimbriae), sfa/sfaS (S fimbriae), focGE (F1C fimbriae), afaIBC (afimbriae), lpfA, lpf0141, lpf0154 (long polar fimbriae), mat/ecpA (fimbrillin), flgE (flagellar hook), crl, csg (curli), tsh (temperature-sensitive haemagglutinin), bmaE (M hemagglutinin), hra/hrlA/hek (heat-resistant agglutinin), iha (IrgA homologue adhesin), yqiG (putative outer membrane usher protein), and kii (K capsule encoding genes) have been reported in APEC [32,33,34,35,36,37,38,39,40,41]. These adhesins also mediate motility, biofilm formation, and APEC survival in macrophages [31]. Furthermore, the fimbriae-encoding gene, yfcO, facilitates adhesion, colonization, and resistance to environmental stresses [42], whereas yadC, promotes adhesion, intracellular survival, and motility [43]. Similarly, autotransporter adhesin genes (aatA, aatB, upaB) contribute to adhesion, colonization, and biofilm formation [44,45]. By screening a random transposon mutant library, multiple other genes (fdtA, rluD, yjhB, ecpR, and fdeC) were found to be responsible for adhesion to chicken and human cell lines [46].

2.2. Invasins

Invasins are a class of proteins associated with the entry of pathogens into host cells [2,31]. Invasins play a role in promoting entry during the initial stage of the infection [2,31]. Multiple genes encoding invasins, ibeA (also called ibe10), ibeB (invasion protein), tia (toxigenic invasion locus), and gimB (genetic island associated with neonatal meningitis) have been reported in APEC isolates [35,47,48]. In addition, invasins also contribute to APEC resistance to oxidative stress induced by macrophages, biofilm formation, colonization, and proliferation in the host [47,48]. IbeR, a regulator of ibeRAT operon, contributes to invasion, resistance to serum and environmental stresses, and the expression of virulence genes [49]. Similarly, ychO, a putative invasin gene, plays a role in motility, adhesion, invasion, biofilm formation, and the expression of membrane proteins and metabolism genes [50].

2.3. Iron Acquisition Systems

Iron is an essential micronutrient required for bacterial growth and proliferation inside the host, once bacteria successfully colonize and/or invade the host [2,31]. APEC possesses different iron acquisition systems consisting of multiple siderophores (aerobactin, salmochelin, yersiniabactin) and transporters to sequester iron from the body fluids [2,31]. Several genes encoding the iron uptake and transport systems, iucCD, iutA, aerJ (aerobactin), iroBCDEN (salmochelin), fyuA (yersiniabactin), sitABCD, mntH (iron and manganese transporter), irp2 (iron repressible protein), feoB (ferrous ion transporter), fepC (ferric enterobactin transporter), ireA (iron-regulated virulence gene), eitABCD (putative iron transporter), chuA (outer membrane hemin receptor), and bfr (bacterioferritin) have been reported in APEC [36,37,40,51,52,53,54,55,56,57]. In addition, these siderophores and transporters also mediate APEC adhesion, invasion, resistance to environmental stresses, the expression of other virulence genes, colonization, and persistence in the host [52,56,58,59]. Furthermore, enterobactin synthesis and transport genes (entE and entS) in coordination with gene encoding outer membrane efflux protein (tolC) also facilitate invasion, colonization, and persistence [60].

2.4. Protectins

Protectins protect bacteria from the host immune system as well as various unfavorable conditions [2,31]. In particular, protectins include bacterial capsules, outer membrane proteins, and lipopolysaccharide (LPS) components, and they provide protection against phagocytic engulfment by macrophages and complement-mediated bactericidal effect in the host serum [2,31]. Several genes encoding multiple protectins, iss (increased serum survival), traT (complement resistance protein), ompT (outer membrane protease), kpsMT(K1), kpsMT(II), kpsMT(III), neuC, neuS, neuD (capsule), kfiC-K5 (glycosyl transferase), and betA (choline dehydrogenase) have been reported in APEC [36,41,51,53,61]. These protectins also mediate APEC adhesion, invasion, intracellular survival, colonization, and proliferation in the host, in addition to protection from host defense [61]. The outer membrane proteins, YbjX and PagP, also play a role in resistance to serum and environmental stresses, invasion, and intracellular survival [62,63]. Similarly, OmpA, another outer membrane protein, also promotes APEC survival in macrophages [64]. The genes involved in LPS biosynthesis, wzy (O-antigen polymerase) and waaL (O-antigen ligase), facilitate intracellular survival and resistance to phagocytosis and environmental stresses along with adhesion, invasion, colonization, motility, and biofilm formation [65,66]. Similarly, lpxM (myristoyl transferase), a gene involved in lipid A biosynthesis, plays a role in invasion, intracellular survival, colonization, and regulation of cytokine genes expression and nitric oxide production [67]. Whereas, sodA (superoxide dismutase) protects APEC from reactive oxygen species (ROS)-mediated host defense and promotes biofilm formation [68].

2.5. Toxins

Toxins are biological poisons that assist in the bacterial ability to invade and cause damage to the tissues [2,31]. Several genes encoding multiple types of toxins, hlyF, hlyA, hlyE (putative avian hemolysin), vat (vacuolating autotransporter toxin), sat (secreted autotransporter toxin), cdtB, cdtS (cytolethal distending factor), astA, EAST-1 (heat-stable enterotoxin), stx2f (shiga toxin variant), pic (serine protease autotransporter), espC (serine protease), and ace4/35 (acetylcholine esterase) have been reported in APEC [35,38,39,40,41,51,53,69,70,71,72,73]. These toxins also facilitate the colonization, motility, biofilm formation, agglutination, induction of vacuolization, and formation of outer membrane vesicles [69].

2.6. Other Virulence and Pathogenesis Factors

Other virulence and pathogenesis factors of APEC include the QS system, transcriptional regulators, two-component systems, secretion systems, and genes associated with bacterial metabolism [10,11,12,13,14,15,16,17,18]. These factors assist in different processes of APEC pathogenesis/infection, including adhesion, invasion, colonization, persistence, interbacterial competitions, resistance to host defenses, and modulation of host immune responses [10,11,12,13,14,15,16,17,18], thereby facilitating the APEC proliferation and establishment of disease in the host.

2.6.1. Quorum-Sensing (QS) System

Quorum sensing is an autoinducer (small hormone-like organic molecules)-based cell-to-cell communication system in bacteria that regulates the expression of various genes associated with motility, biofilm formation, virulence, and others [74]. QS in APEC is mediated by a LuxS synthesized autoinducer-2 (AI-2) molecule and regulated by LsrABCDFGK operon [10,74,75]. Lsr operon, LuxS, and AI-2 along with ptsI (phosphotransferase system), and Pfs (activated methyl cycle pathway) play various roles in APEC pathogenesis, including motility, biofilm formation, adherence, invasion, colonization, intracellular survival, persistence, cell damage, and the expression of virulence genes [10,74,75,76,77].

2.6.2. Secretion Systems

The secretion systems are cell-associated systems that are present on cell membranes of bacteria and function to secrete proteins into host cells, thereby causing damage to the host cells [11,12]. The secreted proteins promote the bacterial virulence either by directly intoxicating the host cells or by enhancing attachment to host cells, establishing replicative niche by scavenging resources and by competing with other microorganisms. Among the different bacterial secretion systems, two secretion systems (type III and VI) contribute to APEC pathogenesis [11,12]. The regulators (EtrA and YqeI) [12,78] and ATPase (EivC) [79] of the type III secretion system 2 (ETT2) play a role in motility, adhesion, intracellular survival, proliferation, colonization, resistance to phagocytosis and serum bactericidal activity, expression of fimbriae genes, and the downregulation of pro-inflammatory cytokine responses. Similarly, different components of type VI secretion system, DotU (organelle trafficking protein), IcmF (intracellular multiplication factor), Hcp (hemolysis co-regulation protein), CpxR, CpxA (envelope stress response system), ClpV (ATPase), and VrgG (secreted protein) mediate interbacterial competition, adhesion, invasion, intracellular survival, colonization, motility, biofilm formation, production of type 1 fimbriae, resistance to serum bactericidal activity, and modulation of intracellular host responses (IL-8, IL-1β) [11,40,80,81,82,83,84,85].

2.6.3. Two-Component Systems

Two-component systems (TCS) are major signaling proteins in bacteria that enable bacteria to respond to changing environments by altering the expression of genes [86]. Different TCSs have been reported with a role in APEC pathogenesis [86,87,88,89]. A membrane-associated TCS, PhoPQ, plays a role in biofilm formation, motility, adhesion, invasion, intracellular survival, systemic dissemination, and the expression of virulence genes and genes associated with flagellar assembly, ABC transporters, quorum sensing, and bacterial chemotaxis [13,86,90]. Similarly, another membrane-associated TCS, BasSR, is involved in biofilm formation and APEC virulence and colonization in vivo [91]. KdpDE, a TCS regulating potassium transport, mediates the expression of flagella-related genes, flagellum formation, motility, and resistance to serum bactericidal activity [87]. Likewise, a TCS regulating nitrogen metabolism, RstAB, contributes to iron acquisition, acid resistance, intracellular survival, and colonization [88,92]. Another TCS, BarA-UvrY, plays a role in the adhesion, invasion, persistence, intracellular survival, resistance to serum bactericidal activity and oxidative stress, and regulation of exopolysaccharide production and expression of type 1 and P fimbriae [89].

2.6.4. Transcriptional Regulators

Multiple transcriptional regulators have shown a role in APEC pathogenesis [14,15,37,93,94,95]. The AutA and AutR, two global transcriptional regulators, mediate the expression of K1 capsule and acid resistance systems, and change in adaptive lifestyle to facilitate infection [14]. FNR (fumarate and nitrate reduction), another global transcriptional regulator, facilitates the adhesion, invasion, expression of type 1 fimbriae and type VI secretion system, and resistance to oxidative stress [15]. McbR (MqsR-controlled colonic acid and biofilm regulator) plays a role in biofilm formation and stress response [94], whereas tyrR (a transcriptional regulator involved in the biosynthesis and transport of aromatic amino acids) promotes invasion, motility, and intracellular survival [37]. YjjQ (transcriptional regulator, LuxR family) contributes to flagellar motility [93], and RfaH, a transcriptional anti-terminator, contributes to invasion, intracellular survival, and resistance to serum bactericidal activity [95].

2.6.5. Metabolism-Associated Genes

Different genes associated with bacterial metabolism contribute to APEC pathogenesis [16,17,18,96]. The operon, acs-yjcH-actP, encoding acetate assimilation system facilitates intracellular survival, proliferation, colonization, and the production of pro-inflammatory cytokines and nitric oxide [96]. Similarly, PotE (putrescine transporter) and NirC (nitrite transporter), involved in polyamine biosynthesis and putrescine transport, and nitrogen metabolism and cytoplasmic detoxification, respectively, mediate adhesion and colonization [16,17]. ArcA (aerobic respiratory control), involved in citrate transport and metabolism, plays a role in motility and chemotaxis [18].

2.6.6. Miscellaneous

Various other bacterial components, such as genes encoding prophage, porins, enzymes, hypothetical protein, and transport systems also play a role in APEC pathogenesis. Prophage phiv142-3, particularly orf20 gene, and phiv205-1 contributes to resistance to serum and environmental stresses, adhesion, invasion, intracellular survival, colonization, biofilm formation, and formation of flagella and type 1 fimbriae [97,98,99]. Outer membrane porins, OmpF and OmpC, facilitate adhesion, invasion, colonization, and proliferation [100]. The hypothetical protein, YicS, plays a role in motility, biofilm formation, and invasion [101]. While, cpdB (2′, 3′-cyclic phosphodiesterase) mediates colonization [102], the phosphate transport system (pstSCAB), particularly pstB, plays a role in colonization as well as resistance to serum bactericidal activity and oxidative stress [103]. The transfer-mRNA-small protein B, tmRNA-SmpB, mediates colonization, persistence, replication, and intracellular survival [104], and mliC, a lysozyme inhibitor, plays a role in resistance to serum bactericidal activity [105]. Additionally, different virulence genes of unknown/not clearly known functions in APEC (but defined in other bacteria), malX (enzyme II of phosphotransferase system recognizing glucose and maltose) [53], etsB (putative ABC transport system) [54], uidA (beta-glucuronidase), usp (uropathogenic specific protein), and hemF (oxygen-dependent coproporphyrinogen-III oxidase) have been also reported. malX plays a role in sugar transport, whereas etsB is type VI secretion system-associated gene. uidA, usp, and hemF encode beta-glucuronidase that break down carbohydrates, non-specific nuclease that cleaves nucleic acids, and coproporphyrinogen-III oxidase involved in heme biosynthesis, respectivley. Furthermore, other genes such as HJ fliC (flagellin), clbB, clbN (colibactin) [39,40], frz (carbohydrate metabolic operon), sopB (plasmid partitioning protein) [106], cvaABC, cvi (colicin V operon) [34], cib/cibI (colicin Ib), cbi (colicin M immunity protein) [57], cba (colicin B activity protein), and cma (colicin M activity protein) have been also reported. fliC encodes flagellin essential for flagellar motility, whereas clbB and clbN are colibactin genes involved in fatty acid biosynthesis. Colicin genes (cvaABC, cvi, cib/cibI, cbi, cba, and cma) are associated with bacteriocin production to exert cytotoxic effects against other bacteria in the niche. frz plays a role in sugar (fructose) transport, whereas sopB is associated with plasmid replication. Moreover, genes such as yfcV (fimbrial protein), gad (glutamate decarboxylase), mchBCF (microcin H47), mcmA(microcin M), bor (bacteriophage lambda bor protein), air (enteroaggregative immunoglobulin repeat protein), eilA (homolog of transcriptional regulator hilA) [107], celB (permease IIC component), and pabB (aminodeoxychorismate synthase) [40] have been also reported. yfcV encodes uncharacterized fimbrial-like protein, whereas gad is involved in acid tolerance. Microcin genes (mchBCF and mcmA) are associated with antibacterial activity against closely related species. bor is involved in serum resistance, whereas air encodes protease facilitating mucin resistance. eilA is a component of a type III secretion system, whereas celB and pabB are involved in sugar transport and tetrahydrofolate biosynthesis, respectively. In addition, genes such as capU (cap locus protein), cif (cell cycle inhibiting factor), tir (translocated intimin receptor), tccp (tir cytoskeleton coupling protein), nleB (non-LEE encoded effector B) [57], iaL (invasion-associated locus), cjrC (putative siderophore receptor) [108], and mig-14p (antimicrobial resistance protein) [41] have been also reported in APEC. capU is putative hexosyltransferase with uncharacterized function, whereas cif belongs to bacterial toxins that arrest host cell division. tir and tccp are involved in bacterial adherence to host cells, whereas nleB is a component of a type III secretion system. iaL is involved in cell penetration, whereas cjrC is a putative siderophore receptor. mig-14p plays a role in resistance to antimicrobial peptides in host cells.

2.7. Genes Essential for Systemic Infections and Adaptation in Chickens

Identifying the genes essential for systemic infections and adaptation is crucial to develop rational treatments against the infections. Nelwike et al. (2012) [109] investigated the APEC genes induced during systemic infections in chickens using RIVET (recombination-based in vivo expression technology). Genes involved in metabolism, cell envelope and integrity, transport systems, and virulence (metH, lysA, pntA, purL, serS, ybjE, ycdK (rutC), wcaJ, gspL, sdsR, irp2, eitD, ylbE, yjiY, tkt1, and phage-related genes) were upregulated in APEC isolated from infected chickens. Similarly, Dozois et al. (2003) [110] studied the APEC genes expressed in infected tissues in chickens using SCOTS (selective capture of transcribed sequences) technology. Genes involved in adherence (pilN, pilQ, tsh, hpb, TcfD, Z5222), LPS synthesis (waaO, waaY), iron acquisition (iutA, iucA, iucD, iroC), plasmid function (ColE2, traK, traG, traT, SopA, psiA), phage-related (hkaG, hkbV, hkbQ, Z3370, Int), and of unknown functions (CC0532, TM0427, YPO3000, rhsH, RSp0733) were highly expressed in APEC infected tissues. On the other hand, Zhang et al. (2019) [111] identified essential genes for APEC adaptation in chickens using the TraDIS (transposon-directed insertion site sequencing) strategy. Genes involved in metabolism, transport, regulation and stress response, RNA processing and translation, cell division and DNA replication, cell envelope biogenesis, and unknown functions were essential to cause disease in chickens. Particularly, genes involved in biotin synthesis (bioABFCD), Rnf electron transport complex (rnfA, rfnE, and gene nth encoding endonuclease III), and cre two-component system (creABCD) were important for the adaptation of APEC in chickens. In another study, genes identified as upregulated through microarray analysis (yehD, potF, flgE, tyrR, and bfr) in APEC isolated from chicken showing swollen head syndrome were essential for adhesion, invasion, survival inside macrophages, and motility [37]. These studies provide insights on APEC pathophysiological processes during systemic infections in chickens.
Overall, multiple virulence and pathogenesis factors of APEC are involved in causing colibacillosis in poultry. As a result of the involvement of multiple virulence and pathogenesis factors, there is a hindrance in developing therapeutics broadly effective against APEC infections. In-depth understanding of these factors as well as unraveling the new factors will help develop the effective therapeutics against colibacillosis in poultry. Furthermore, several of these factors have coordinated and overlapping functions, which necessitates a holistic strategy to formulate an ideal anti-APEC therapeutics. For instance, developing therapeutics targeting iron acquisition systems [112], QS system [113], bacterial metabolism [114], and secretion systems [115] can provide solutions to mitigate APEC infections in poultry in the future.

3. Zoonotic Potential

APEC belongs to the ExPEC subgroup of E. coli, similar to UPEC and NMEC [4]. Multiple studies have reported APEC as a potential foodborne zoonotic pathogen as well as a source or reservoir of extra-intestinal infections in humans [21,22,41,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141,142,143]. This is in particular due to its shared genetic similarity with human ExPECs, the presence of common or human ExPEC-defining virulence genes, and the ability to cause UTI and meningitis in rodent models, similar to UPEC and NMEC.

3.1. Genetic Similarity and Commonality in Virulence Genes

Multiple studies have shown that APEC shares genetic similarities with human ExPECs (UPEC and NMEC) and avian-associated ColV plasmids (for instance, pAPEC-078-ColV, pAPEC-O2-ColV, pAPEC-O1, p1ColV5155) essential for poultry adaptation are present in human ExPEC isolates. The phylogenetic (single nucleotide polymorphisms; SNP) comparison of whole-genomes of 323 APEC and human ExPEC isolates belonging to sequence type ST95 revealed genetic overlap (no distinct clustering) between APEC and certain human ExPECs, indicating that certain ExPEC clones may have the potential to cause infection in both poultry and humans [116]. Especially, ColV plasmids specific to APEC were present in those human ExPEC isolates, and 10 virulence genes (iucC, iucD, iutA, cvaA, etsA, hlyF, ompT, cvaB, cvaC, and cvi) were common between them. Another study compared the whole genomes of 48 APEC and UPEC isolates belonging to ST131-H22 [21]. The high genetic similarity (no distinct lineage) was observed in SNP-based phylogenetic analysis between UPEC and APEC isolates together with the presence of ColV plasmids in UPEC isolates. The ColV/ColBM plasmids were also present in 34.5% of APEC and 18.9% of human ExPEC isolates when whole genomes of 551 mcr-1 positive APEC and human ExPEC isolates were analyzed [118]. Therefore, the presence of poultry-specific ColV plasmids in human ExPEC isolates might suggest a zoonotic transmission of APEC from poultry to humans.
Other studies have also shown the commonality in virulence genes between APEC and human ExPEC isolates. Four virulence genes (iss, iutA, ompT, and papGII) were common when the presence of virulence genes was investigated in 200 UPEC and APEC isolates [117]. In another study, several virulence genes (iroN, traT, iucD, cvi/cva, ibeA, gimB, tia, neuC, kpsMTII, tsh, iss, sitD, chuA, fyuA, irp2, vat, malX, and pic) were present in APEC, UPEC, and NMEC isolates [123]. The screening of virulence genes in 27 APEC isolates belonging to ST73 found different virulence genes (pap, sfa, usp, cnf1, kpsMTII, hlyA, and ibeA) known to be specific to UPEC and NMEC [119]. The human ExPEC-defining virulence markers (papAH, papC, papEF, papG, papG allele II, fyuA, kpsMTII, K1, usp, ibeA, and iutA) were also highly prevalent in 25 APEC isolates belonging to ST95, ST127, ST131, ST141, ST420, and ST69 [120]. Similarly, different human ExPEC-defining virulence markers (papA, papC, papEF, papG, hra, astA, hlyF, tsh, fyuA, ireA, iroN, iutA, cvaC, iss, kpsMT K1, kii, traT, ibeA, ompT, malX, usp, and fimH) were prevalent in 129 ExPEC isolates from meat and shell eggs [121]. Four human ExPEC-defining virulence markers (papA, papC, kpsMII, and iutA) were also prevalent in APEC isolates from broiler chickens and meat [122]. Virulence determinants/markers associated with UPEC (growth in human urine, iha, foc, cnf, papC, papGII, iroN, iutA, kpsMTII, cvaC, hlyF, vat, malX, usp) and NMEC (kpsMTK1 and ibeA) were present in ExPEC isolates from chicken samples, including feces and cecal contents [143]. Therefore, the presence of common virulence traits between APEC and human ExPEC isolates can provide substantial evidence of zoonotic threats posed by APEC to humans.
Several other phylogenetic studies have also shown the similarity of APEC with human ExPEC isolates. The phylogenomic tree constructed based on whole genomes of 47 E. coli strains revealed that APEC isolates belonging to serogroups O1:K1 and O2:K1 share significant genetic similarities/overlap (same cluster) with human ExPEC O18:K1 strains [124]. This finding is further supported by several other studies [22,125,126,127,128,129,136]. The ExPEC (APEC, UPEC, and NMEC) isolates belonging to ST95 and serogroups O1, O2, and O18 were clustered together when the phylogeny of ExPEC isolates was constructed based on the possession of different genes/traits or multilocus sequence typing (MLST), suggesting the potential transmission of certain ExPEC strains between poultry and humans [125,127]. Similarly, APEC isolates belonging to serogroup O18 were similar to NMEC strains when compared using MLST and pulsed-field gel electrophoresis (PFGE) [22,129]. In other studies, APEC isolates belonging to ST95 and serogroup O1 were also similar to UPEC and NMEC isolates when compared using MLST, PFGE, and whole genome analysis, suggesting the zoonotic potential of these isolates to humans with no host-specificity [126,128]. In several other studies performed using MLST and PFGE, APEC isolates belonging to sequence type other than ST95, such as ST359 [130], ST23 [131], ST10, ST117, ST746 [132], ST617, ST23 [135], or other serogroups such as O45 [133,134] were also similar to human ExPEC isolates. Therefore, the APEC isolates belonging to certain STs or serogroups could pose a significant zoonotic risk to humans.

3.2. Ability to Cause Disease with Similar Clinical Manifestations

Multiple studies have shown that APEC can cause UTI and meningitis similar to UPEC and NMEC, respectively [22,124,137,138,139,140]. The E. coli isolates from chicken meat and shell eggs were lethal similar to UPEC in a mouse model of UTI, caused sepsis in a mouse sepsis model, and infected the cerebrospinal fluid (CSF) similar to NMEC in a neonatal rat meningitis model [137]. Furthermore, these isolates also possessed swimming (motility) and biofilm-forming ability in urine, and they adhere, invade, and survive intracellularly in human kidney and bladder cells, similar to UPEC. These isolates also possessed K1 capsule and ibeA, which are essential virulence factors for NMEC pathogenesis. Similar findings were also observed in other studies [22,41,124,138,139,140]. The E. coli isolates from chicken feces or from colibacillosis lesions were also lethal, caused sepsis, and infected CSF in rodent model studies [41,138]. In another study, E. coli isolated from colibacillosis cases and belonging to phylogroup F were able to cause disease (sepsis, meningitis, and UTI) in animal models of human infections [41]. The APEC and ExPEC (UPEC and NMEC) isolates, belonging to ST95 when compared, both were equally able to adhere and invade kidney cells, form strong biofilm, and resist the bactericidal activity of serum [139], reinforcing the understanding that APEC isolates belonging to ST95 pose a potential zoonotic risk to humans. Other similar studies have found that APEC isolates, particularly belonging to serogroup O18, survive in human serum, and they bind and enter human macrophages and human cerebral microvascular endothelial cells, similar to NMEC [22,140]. These isolates also induce neuronal apoptosis in mice, suggesting that APEC O18 isolates utilize similar pathogenic mechanisms as NMEC to cause meningitis in mice. Furthermore, APEC isolates, belonging to serogroup O1:K1 and O2:K1, also cause sepsis and meningitis in rodent models, suggesting that APEC O1:K1 and O2:K1 can have zoonotic potential [124].
Conversely, UPEC isolates also induced colibacillosis in chickens [141,142]. The experimental infection of laying hens with UPEC isolate caused salphingitis similar to APEC [141]. UPEC also induced similar symptoms and lesions comparable to those caused by APEC in chickens [144]. In another study, UPEC isolates belonging to serogroups O4, O74, O1, and O75 were 100% lethal to chickens [142]. The NMEC isolates were also lethal to chick embryos and caused colisepticemia in chickens, similar to APEC [22].
Overall, the potential of APEC ST95 and ST131 strains to cause UTI and meningitis in humans through the consumption of contaminated poultry products signifies the zoonotic nature of APEC. Furthermore, the zoonotic potential of APEC isolates, especially belonging to serogroups O1, O2, and O18, cannot be underestimated; therefore, interventions appropriate to mitigate the transmission of APEC to humans should be undertaken to combat the food safety threat posed by APEC to humans. Future investigations should consider determining the link between APEC and potential zoonotic transmission to humans.

4. Control Strategies

The control of APEC infections in poultry relies on antibiotic medication and vaccination, other than managing the environmental stressors, applying the biosecurity measures, and vaccinating the chickens against the viral and immunosuppressive diseases [1,2,145]. Probiotics, bacteriophages, and different new alternatives (innate immune stimulants, virulence and growth inhibitors, and antimicrobial peptides) have been also evaluated [26,27,28,29,30] with a goal to develop effective preventative and therapeutic treatments to control colibacillosis in chickens. Potential checkpoints for controlling APEC infection in chickens are shown in Scheme 1.

4.1. Management and Biosecurity Measures

The management of environmental stressors such as ammonia and dust in poultry houses by maintaining good litter and air quality are some of the key factors in preventing APEC infections in poultry houses [1,2]. Proper ventilation as well as maintaining optimum temperature, humidity, and bird density help mitigate environmental stress in chickens [1,2]. Furthermore, the elimination of pre-disposing factors by vaccinating chickens against MG, IBV, NDV, and IBD reduces the incidence of APEC infections [1,2]. Good nutrition and birds with enhanced immune systems are also likely contributors to reducing the incidence of colibacillosis [1,2]. Moreover, the vertical transmission of APEC can be prevented at the breeding level or at the top of the production pyramid by different intervention measures such as developing breeds with increased resistance to APEC infections, cleaning and disinfection of hatching eggs, and minimizing the use of floor eggs [145]. The horizontal transmission of APEC can be limited by using all-in-all-out production systems, systematic culling of weak chicks at first week, and implementing effective sanitation programs [145]. The proper and efficient biosecurity measures together with feed and water (chlorination) decontamination and disinfection of poultry houses, feed mills, farm equipment, and premises are necessary to prevent APEC entry into farms [1,2]. The biosafety measures such as preventing access of vectors such as houseflies, wild birds, and rodents are also necessary to keep APEC out of poultry facilities [145].

4.2. Antibiotics

Antibiotics are commonly used to control APEC infections in poultry [23]. Many antibiotics belonging to different classes, such as tetracyclines (tetracycline, oxytetracycline, chlortetracycline), sulfonamides (sulfadimethoxine, trimethoprim, sulfadiazine, sulfamethazine, sulfaquinoxaline, ormethoprim), aminoglycosides (apramycin, gentamicin, neomycin, spectinomycin), penicillins (amoxicillin, ampicillin), cephalosporins (ceftiofur), quinolones (danofloxacin, sarafloxacin, enrofloxacin), polymyxins (colistin), chloramphenicols (florfenicol), macrolides (erythromycin), and lincosamides (lincomycin) have been used in poultry industry worldwide for the control of APEC infections [23]. However, APEC resistance to multiple antibiotics has been reported [32,34,35,51,53,54,70,71,72,73,107,108,108,146,147,148,149,150,151,152,153,154,155,156,157,158,159,160,161,162,163,164,165,166,167,168,169,170,171,172], which limits the use of these antibiotics and suggests a challenge ahead in using these antibiotics. Table 2 provides a summary of antibiotic resistance and resistance genes (mechanisms) reported worldwide in APEC isolates from chickens in the last five years (2015 to 2020). These data indicate APEC resistance to almost all classes of antibiotics, except carbapenems. Resistance to imipenem has been also recently reported [108,108,171]. The resistance is most commonly seen with ampicillin, tetracycline, trimethoprim, sulfamethoxazole, and streptomycin antibiotics. Importantly, a high level of APEC resistance to medically important antibiotics, such as β-lactams and colistin, have also been reported, which might pose a high risk to humans because of the transmission of antibiotic-resistant bacteria and genes through the food chain [173]. The strategies employed by the US and European Union (EU) to restrict the non-therapeutic use (for growth promotion) of antibiotics in food–animal production and to limit the therapeutic use (for treatment) of medically important antibiotics could aid in mitigating this risk [174]; however, benefits of such measures in curbing antibiotic resistance issues may take time to realize. The development of antibacterials solely for animal uses without cross-resistance potential or the identification of antibiotic alternatives as a replacement for antibiotics could aid in combating antibiotic resistance issues in the food–animal production.

4.3. Vaccines

Various vaccine candidates, mostly live-attenuated and recombinant vaccines, have been investigated to protect chickens against APEC infections [5,175,176,177,178,179,180,181,182,183,184,185,186,187,188,189,190,191,192,193,194,195,196,197,198,199,200,201,202,203,204,205,206]. Table 3 provides the summary of vaccines tested to date along with their main findings. In the past, inactivated vaccines were tested; however, recent studies have focused mostly evaluating live-attenuated and recombinant vaccines in chickens. The varying degrees of protection, ranging from none to partial to complete, have been achieved using these vaccines. Among the tested vaccines, multiple vaccines such as outer membrane vesicles (OMVs), bacterial ghost vaccines, recombinant iss, recombinant antigen (rAg) vaccine containing ExPEC antigens, Salmonella-delivered vaccines containing APEC antigens, ΔaroA, and ΔtonB/Δfur were able to reduce the mortality, lesions, and bacterial burden as well as stimulate the antibody (immunoglobulins; IgG and IgA) responses in chickens.
Despite multiple vaccine candidates showing proven efficacy in chickens in experimental studies, only two vaccines (live-attenuated APEC O78 ΔaroA Poulvac® E. coli vaccine and inactivated Nobilis® E. coli vaccine containing F11 fimbrial and FT flagellar antigens) are commercially available currently for use in chickens [5]. However, the major drawback of these vaccines is the lack of protection against heterogenous APEC infections [5]. The ideal APEC vaccine should be able to confer cross-protection against multiple APEC serotypes and be deliverable by mass-immunization methods, such as oral (feed or water) or spray routes [5]. The identification of common/conserved virulence and pathogenesis mechanisms employed by diverse APEC serotypes to cause disease in chickens would facilitate the development of new broad-spectrum vaccines. For instance, vaccine developed using outer membrane iron receptors required for iron acquisition (FyuA, Hma, IreA, IutA) protects against UPEC infections in a murine model [112]. In another study, a tetravalent conjugate vaccine developed using O antigens of predominant UPEC serotypes provides broad protection against UPEC infections [207]. Furthermore, the knowledge gained on new virulence and pathogenesis factors should be exploited to design potent vaccine candidates. For example, the type VI secretion system (vgrG) and quorum sensing system (luxS) can be the new vaccine targets because of their substantial involvement in APEC virulence and pathogenesis.

4.4. Probiotics

Different probiotics have been tested for their efficacy to prevent APEC infections in chickens [26,175,208,209]. The efficacy of Lactobacillus plantarum B1 was evaluated against E. coli (K88) infection by supplementing in the broilers feed (2 × 109 CFU/kg) [26]. Broilers fed with L. plantarum B1 showed significantly decreased cecal E. coli counts and increased growth performance, villus height to crypt depth ratio, ileal mucosal sIgA concentration, and cecal lactic acid bacteria counts. Similarly, the efficacy of L. plantarum 15-1 and fructooligosaccharides (FOS) combination was evaluated against APEC (O78) infection by supplementing in the broilers feed (1 × 108 CFU/kg) [208]. The decrease in mortality and serum diamine oxidase and increase in IgA and IgG concentrations was observed in broilers fed with probiotic and FOS mix. The effects of Enterococcus faecalis-1 was assessed in broiler chickens challenged with APEC (O78) by inoculating orally in drinking water for 3 days (1 × 108 CFU) from days 1 to 3 of growth [210]. E. faecalis-1 supplementation significantly improved the growth performance and immune response, reduced the mortality, and decreased the visceral organs invasion by APEC O78. Likewise, the efficacy of multi-strain commercial probiotic mix (Bacillus subtilis, Clostridium butyricum, and L. plantarum) was tested against APEC 078 infection by supplementing in the broilers feed [209]. There was significant decrease in mortality (13.6% to 0%) and APEC counts in liver and spleen and increase in growth performance and lactobacilli population in broilers fed with the probiotic mix. Another commercial probiotic mix (B. subtilis, L. acidophilus, Pediococcus acidilactici, Pediococcus pentosaceus, and Saccharomyces pastorianus) was also tested in combination with recombinant attenuated Salmonella vaccine (RASV) for protection against APEC (O78:K80) and Salmonella infection by supplementing in feed in layer chickens [175]. Chickens showed reduced signs of airsacculitis, perihepatitis, and pericarditis and lower APEC load in the blood. These studies suggest that different probiotics belonging to genus Lactobacillus, Bacillus, Clostridium, and Pediococcus show efficacy in preventing APEC infections as well as improve the growth performance, maintain the healthy intestinal microbiota, and enhance the intestinal mucosal immunity. Furthermore, there are multiple probiotic products (for example, Sav-A-Chick® Probiotic Poultry Supplement, Probios®, HealthyGutTM PROBIOTICS, and SuperDFM- Poultry) commercially available for use in maintaining intestinal health and boosting immune status in poultry. These probiotics contain different beneficial microorganisms, such as B. subtilis, B. licheniformis, L. plantarum, L. casei, L. acidophilus, L. brevis, L. reuteri, Enterococcus faecium, E. thermophilus, P. acidilactici, P. pentosaceus, Bifidobacterium bifidum, B. animalis, Propionibacterium shermanii, and P. freudenerichii. Even though these probiotics are not indicated specifically for APEC, they can reduce the incidence of APEC infections in poultry farms due to their broad-spectrum effect against enteric pathogens. Furthermore, the identification of new probiotics with superior potential to protect against APEC infections is necessary, which can provide alternatives to antibiotics, thereby also mitigating the development of antibiotic resistance. For instance, the next-generation probiotics specific to APEC can be developed by investigating the microbiome of healthy and APEC infected chickens followed by the identification of beneficial bacteria crucial to resist APEC infection in chickens [211].

4.5. Bacteriophages

To date, multiple studies have been conducted to evaluate the preventative and therapeutic efficacy of phages against APEC infections in chickens [27,212,213,214]. The efficacy of phage mixture (SPR02 and DAF6) was evaluated in APEC (O2) challenged chickens by spray and intramuscular administrations [214]. The phage treatment three days prior to APEC challenge significantly reduced (40% to 3%) the mortality of chickens. Similarly, phage treatment at 24 h and 48 h post-challenge also reduced the mortality rate (55% to below 20%). The efficacy of phage cocktail (phi F78E Myoviridiae, phi F258E Siphoviridae, and phi F61E Myoviridae) was tested in experimentally (O78) and naturally infected flocks refractive to antibiotic treatment by oral or spray administration [213]. The treatment with phage cocktail reduced mortality by 25% in experimentally infected chickens and decreased the flocks’ mortality level to below 0.5% in flocks infected naturally with APEC. Similarly, the efficacy of another phage cocktail (TM1, TM2, TM3, and TM4) was evaluated in APEC challenged (O78:K80 and O2:K1) chickens by administering through intramuscular injection [212]. The phage cocktail treatment reduced the mortality (46.6% to 13.6%), APEC load in lung, and APEC lesions in lung, liver, and heart, and increased the body weight of chickens. The efficacy of phage-loaded chitosan nanoparticles (C-ΦKAZ14 NPs; Myxoviridae, T4-like coliphage) was also evaluated in APEC-challenged (O1:K1:H7) chickens by oral administration [27]. The C-ΦKAZ14 NP treatment decreased the mortality (58.33% to 16.7%), intestinal colonization of APEC (2.30 × 109 ± 0.02 to 0.79 × 103 ± 0.10 CFU/mL), and fecal shedding (2.35 × 109 ± 0.05 to 1.58 × 103 ± 0.06 CFU/mL). C-ΦKAZ14 NP treatment also increased the body weight of chickens as well as ameliorated the clinical signs and symptoms. These studies suggest that the phage therapy can be a valuable approach to control APEC infections in chickens. However, no treatment involving phages has yet advanced into field applications. This is partly due to challenges in large-scale production and controversies associated with approval for use in poultry production [215]. If proven safe and effective in field settings, phages can serve as a valuable alternative to antibiotics in poultry production.

4.6. New Alternative Approaches

Apart from antibiotics, vaccines, probiotics, and bacteriophages, different novel approaches, including but not limited to innate immune stimulants, virulence and growth inhibitors, and antimicrobial peptides have been studied for their protective effects against APEC infections in chickens [28,29,216,217,218,219,220,221,222].

4.6.1. Innate Immune Stimulants

Innate immune stimulants can activate the innate immune responses by acting as pathogen-associated molecular patterns (PAMPs) and binding to pattern recognition receptors (PRRs); thereby, preventing the host from infection [217]. The synthetic oligodeoxynucleotides containing unmethylated cytosine-phosphodiester-guanine motifs (CpG-ODN) was tested to protect neonatal chickens against APEC (O2) infection by administering through the intrapulmonary route using compressor nebulizer in an acrylic chamber [217]. Higher survival, better clinical conditions, and lower bacterial load were observed in chickens that received CpG-ODN. Furthermore, CpG-ODN also induced systemic antibacterial immune responses, including upregulation of expression of pro-inflammatory (IL-1β, LPS-induced tumor necrosis factor, IL-18) and anti-inflammatory (IL-10, IL-4) cytokine genes, and enrichment and maturation (higher CD40 and MHCII expression) of monocytes/macrophages and CD4+ and CD8+ T-cell subsets. Similarly, the prophylactic potential of three in ovo-administered innate immune stimulants and immune adjuvants (CpG-ODN, polyinosinic:polycytidic acid, and polyphosphazene) was evaluated to protect young chickens from yolk sac infection caused by APEC (O2) [28]. CpG-ODN increased the survival (>80%) of young chickens. Overall, these studies demonstrate that CpG-ODN can induce protective immunity against APEC-induced yolk sac infection in neonatal chickens. The evaluation in adult chickens to prevent APEC-associated infections is warranted to develop innate immune stimulants as new non-antibiotic measures for colibacillosis prevention.

4.6.2. APEC Virulence and Growth Inhibitors

Virulence inhibitors disarm/attenuate pathogens by inhibiting virulence mechanisms, such as the QS system, unlike the antibiotics that inhibit the bacterial growth [223]. Therefore, virulence inhibitors can overcome the limitations of current antibiotics, viz. resistance and killing of commensal bacteria, and making pathogens susceptible to natural host defenses; thus, they are superior to conventional antibiotics [223]. On the other hand, growth inhibitors possessing novel scaffolds with newer antibacterial targets with less likelihood for resistance development such as those targeting bacterial membrane can be promising new antibacterial agents [224].
The protective effect of baicalin (medicinal ingredient isolated from dry roots of Scutellaria baicalensis), an APEC QS inhibitor, was evaluated against APEC-induced acute lung injury (O78) in chickens [218]. The pre-treatment of baicalin (200 mg/kg) significantly reduced the mortality, lesion in lung, lung wet/dry ratio, myeloperoxidase activity, and levels of pro-inflammatory cytokines (IL-1β, TNF-α, and IL-6) in the lung. In another study, the effect of rutin (flavonoid extracted from plants), an APEC QS inhibitor, was investigated on AI-2 secretion (APEC O78), biofilm formation, and the expression of virulence genes [29]. Rutin significantly reduced the AI-2 secretion, biofilm formation, and expression of virulence genes together with a decrease in adhesion and damage to chicken type II pneumocytes. Multiple APEC small molecule QS inhibitors, mostly piperazines, effective in reducing AI-2 secretion, biofilm formation, motility, the expression of virulence genes, and intracellular survival in macrophage and epithelial cells were identified through screening of the small molecule (SM) library [222]. The treatment with these QS inhibitors also increased the survival of wax moth (Galleria mellonella) larvae and decreased the intra-larval APEC load. Among the identified QS inhibitors, two QS inhibitors [QSI-5 (C-5) and QSI-10 (C-10)] [222] significantly reduced the APEC-induced mortality, lesions, and APEC load in internal organs of chickens (data not published). The effect of andrographolide (Andrographis paniculata) in reducing cell damage caused by APEC O78 was investigated in chicken type II pneumocytes [220]. Andrographolide significantly reduced AI-2 secretion, expression of virulence genes, the release of lactate dehydrogenase (LDH), F-actin cytoskeleton polymerization, and adhesion to chicken type II pneumocytes.
Multiple APEC small molecule growth inhibitors possessing pyrrolidinyl, imidazole, piperidine, quinoline, and nitrophenyl scaffolds were identified in our study through screening of the SM library [221,225]. These inhibitors were bactericidal to APEC, at very low concentrations, affected APEC membrane integrity, and decreased the intracellular survival of APEC in macrophage and epithelial cells. The treatment with these inhibitors also increased the survival of wax moth larvae and decreased the intra-larval APEC load. Among the identified growth inhibitors, two growth inhibitors [GI-7 (SM7) and GI-10 (SM10)] [221] significantly reduced the APEC-induced mortality, lesions, and APEC load in internal organs of chickens (data not published).
Altogether, various virulence and growth inhibitors have shown the potential to be developed as novel anti-APEC therapeutics. However, further efforts are needed to advance these inhibitors into field applications.

4.6.3. Antimicrobial Peptides

Antimicrobial peptides (AMPs), regarded as new category of therapeutic agents, are short and generally positive charged peptides [226]. AMPs have fast and selective antimicrobial action, even against antibiotic-resistant bacteria, with low propensity for resistance development, which makes them ideal candidates for antibacterial development [226]. The protective effect of prophylactic in ovo treatment of D analog of chicken cathelicidin-2 (D-CATH-2; host defense peptide) was evaluated against APEC (O78:K80) infection in chickens [219]. The treatment of D-CATH-2 reduced the mortality (by 63%) and bacterial load (>90%) along with the increment of IgM level and peripheral blood lymphocytes and heterophils. In another study, the efficacy of orally administered surfactin–amoxicillin combination was evaluated against APEC (O78) infection in chickens [216]. Surfactin (lipopeptide) combination (0.01 mg/g) enhanced the efficacy (significant decrease in mortality, APEC load, and APEC lesions compared to amoxicillin treatment alone) of otherwise ineffective amoxicillin. The treatment also resulted in upregulation of expression of pro-inflammatory and anti-inflammatory cytokine genes (IL-1β, TNF-α, IL-10 and IL-13). In other studies, peptides (A3, P5, cecropin A-D-Asn, cLF36) have shown efficacy in decreasing the E. coli load in the chicken gut [30,227]. These studies demonstrate that antimicrobial peptides can be developed either as an alternative to antibiotics or as an adjuvant to antibiotics to enhance the efficacy of antibiotics.

5. Conclusions and Future Perspectives

APEC is a most common bacterial pathogen of poultry that causes significant economic losses to the poultry industry worldwide. The effective control of APEC is beneficial to both animal and human health. Multiple virulence and pathogenesis factors of APEC are involved in a coordinated way to cause systemic infections in poultry; therefore, a holistic approach encompassing all factors such as iron acquisition systems, QS system, bacterial metabolism, and secretion systems is necessary to formulate effective anti-APEC therapeutics in the future. Further investigation is necessary to provide concrete evidence for zoonotic transmission of APEC to humans. APEC isolates, particularly belonging to ST95 and ST131 or O1, O2, and O18 serogroups might serve as a source of human extra-intestinal infections. As a result of the significant antibiotic resistance issues and high risk of transmission of antibiotic-resistant bacteria and genes to humans, the development of antibacterials solely for animal uses without cross-resistance to current antibiotics might provide a solution for the future. Furthermore, there is a need for an ideal APEC vaccine that can provide cross-protection against multiple APEC serotypes. Knowledge gained on virulence and pathogenesis mechanisms of APEC should be exploited to identify the new vaccine candidates. Lastly, alternative therapies should also be considered for further development, especially the small molecule virulence inhibitors or growth inhibitors and AMPs with novel targets that could be effective at controlling colibacillosis in poultry.

Author Contributions

Conceptualization, G.R. and D.K.; Writing—Original Draft Preparation, D.K., G.R., D.L., S.R.; Writing—Review and Editing, D.K., G.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by U.S. Department of Agriculture, Grants 2015-68004-23131 and 2020-6701-31401. The APC was funded by Center for Food Animal Health, Department of Animal Sciences, The Ohio State University.

Data Availability Statement

Not applicable.

Acknowledgments

The research in Rajashekara laboratory is supported by National Institute for Food and Agriculture (NIFA), U.S. Department of Agriculture.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dho-Moulin, M.; Fairbrother, J.M. Avian pathogenic Escherichia coli (APEC). Vet. Res. 1999, 30, 299–316. [Google Scholar]
  2. Dziva, F.; Stevens, M.P. Colibacillosis in poultry: Unravelling the molecular basis of virulence of avian pathogenic Escherichia coli in their natural hosts. Avian Pathol. 2008, 37, 355–366. [Google Scholar] [CrossRef] [Green Version]
  3. Guabiraba, R.; Schouler, C. Avian colibacillosis: Still many black holes. FEMS Microbiol. Lett. 2015, 362, fnv118. [Google Scholar] [CrossRef]
  4. Mellata, M. Human and avian extraintestinal pathogenic Escherichia coli: Infections, zoonotic risks, and antibiotic resistance trends. Foodborne Pathog. Dis. 2013, 10, 916–932. [Google Scholar] [CrossRef] [Green Version]
  5. Ghunaim, H.; Abu-Madi, M.A.; Kariyawasam, S. Advances in vaccination against avian pathogenic Escherichia coli respiratory disease: Potentials and limitations. Vet. Microbiol. 2014, 172, 13–22. [Google Scholar] [CrossRef]
  6. de Brito, B.G.; Gaziri, L.C.J.; Vidotto, M.C. Virulence Factors and Clonal Relationships among Escherichia coli Strains Isolated from Broiler Chickens with Cellulitis. Infect. Immun. 2003, 71, 4175. [Google Scholar] [CrossRef] [Green Version]
  7. Lutful Kabir, S.M. Avian colibacillosis and salmonellosis: A closer look at epidemiology, pathogenesis, diagnosis, control and public health concerns. Int. J. Environ. Res. Public Health 2010, 7, 89–114. [Google Scholar] [CrossRef] [Green Version]
  8. Johnson, T.J.; Wannemuehler, Y.; Doetkott, C.; Johnson, S.J.; Rosenberger, S.C.; Nolan, L.K. Identification of Minimal Predictors of Avian Pathogenic Escherichia coli Virulence for Use as a Rapid Diagnostic Tool. J. Clin. Microbiol. 2008, 46, 3987–3996. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Collingwood, C.; Kemmett, K.; Williams, N.; Wigley, P. Is the Concept of Avian Pathogenic Escherichia coli as a Single Pathotype Fundamentally Flawed? Front. Vet. Sci. 2014, 1, 5. [Google Scholar] [CrossRef] [Green Version]
  10. Palaniyandi, S.; Mitra, A.; Herren, C.D.; Zhu, X.; Mukhopadhyay, S. LuxS contributes to virulence in avian pathogenic Escherichia coli O78:K80:H9. Vet. Microbiol. 2013, 166, 567–575. [Google Scholar] [CrossRef]
  11. Ma, J.; Bao, Y.; Sun, M.; Dong, W.; Pan, Z.; Zhang, W.; Lu, C.; Yao, H. Two functional type VI secretion systems in avian pathogenic Escherichia coli are involved in different pathogenic pathways. Infect. Immun. 2014, 82, 3867–3879. [Google Scholar] [CrossRef] [Green Version]
  12. Wang, S.; Xu, X.; Liu, X.; Wang, D.; Liang, H.; Wu, X.; Tian, M.; Ding, C.; Wang, G.; Yu, S. Escherichia coli type III secretion system 2 regulator EtrA promotes virulence of avian pathogenic Escherichia coli. Microbiology 2017, 163, 1515–1524. [Google Scholar] [CrossRef] [PubMed]
  13. Li, Q.; Yin, L.; Xue, M.; Wang, Z.; Song, X.; Shao, Y.; Liu, H.; Tu, J.; Qi, K. The transcriptional regulator PhoP mediates the tolC molecular mechanism on APEC biofilm formation and pathogenicity. Avian Pathol. 2020, 49, 211–220. [Google Scholar] [CrossRef]
  14. Zhuge, X.; Tang, F.; Zhu, H.; Mao, X.; Wang, S.; Wu, Z.; Lu, C.; Dai, J.; Fan, H. AutA and AutR, Two Novel Global Transcriptional Regulators, Facilitate Avian Pathogenic Escherichia coli Infection. Sci. Rep. 2016, 6, 25085. [Google Scholar] [CrossRef] [Green Version]
  15. Barbieri, N.L.; Vande Vorde, J.A.; Baker, A.R.; Horn, F.; Li, G.; Logue, C.M.; Nolan, L.K. FNR Regulates the Expression of Important Virulence Factors Contributing to the Pathogenicity of Avian Pathogenic Escherichia coli. Front. Cell Infect. Microbiol. 2017, 7, 265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Guerra, P.R.; Herrero-Fresno, A.; Pors, S.E.; Ahmed, S.; Wang, D.; Thofner, I.; Antenucci, F.; Olsen, J.E. The membrane transporter PotE is required for virulence in avian pathogenic Escherichia coli (APEC). Vet. Microbiol. 2018, 216, 38–44. [Google Scholar] [CrossRef]
  17. De Paiva, J.B.; Leite, J.L.; da Silva, L.P.; Rojas, T.C.; de Pace, F.; Conceicao, R.A.; Sperandio, V.; da Silveira, W.D. Influence of the major nitrite transporter NirC on the virulence of a Swollen Head Syndrome avian pathogenic E. coli (APEC) strain. Vet. Microbiol. 2015, 175, 123–131. [Google Scholar] [CrossRef]
  18. Jiang, F.; An, C.; Bao, Y.; Zhao, X.; Jernigan, R.L.; Lithio, A.; Nettleton, D.; Li, L.; Wurtele, E.S.; Nolan, L.K.; et al. ArcA Controls Metabolism, Chemotaxis, and Motility Contributing to the Pathogenicity of Avian Pathogenic Escherichia coli. Infect. Immun. 2015, 83, 3545–3554. [Google Scholar] [CrossRef] [Green Version]
  19. Markland, S.M.; LeStrange, K.J.; Sharma, M.; Kniel, K.E. Old Friends in New Places: Exploring the Role of Extraintestinal E. coli in Intestinal Disease and Foodborne Illness. Zoonoses Public Health 2015, 62, 491–496. [Google Scholar] [CrossRef]
  20. Bélanger, L.; Garenaux, A.; Harel, J.; Boulianne, M.; Nadeau, E.; Dozois, C.M. Escherichia coli from animal reservoirs as a potential source of human extraintestinal pathogenic E. coli. FEMS Immunol. Med Microbiol. 2011, 62, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Liu, C.M.; Stegger, M.; Aziz, M.; Johnson, T.J.; Waits, K.; Nordstrom, L.; Gauld, L.; Weaver, B.; Rolland, D.; Statham, S.; et al. Escherichia coli ST131-H22 as a Foodborne Uropathogen. mBio 2018, 9. [Google Scholar] [CrossRef] [Green Version]
  22. Tivendale, K.A.; Logue, C.M.; Kariyawasam, S.; Jordan, D.; Hussein, A.; Li, G.; Wannemuehler, Y.; Nolan, L.K. Avian-pathogenic Escherichia coli strains are similar to neonatal meningitis E. coli strains and are able to cause meningitis in the rat model of human disease. Infect. Immun. 2010, 78, 3412–3419. [Google Scholar] [CrossRef] [Green Version]
  23. Agunos, A.; Léger, D.; Carson, C. Review of antimicrobial therapy of selected bacterial diseases in broiler chickens in Canada. Can. Vet. J. 2012, 53, 1289–1300. [Google Scholar]
  24. Nhung, N.T.; Chansiripornchai, N.; Carrique-Mas, J.J. Antimicrobial Resistance in Bacterial Poultry Pathogens: A Review. Front. Vet. Sci. 2017, 4, 126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Gregersen, R.H.; Christensen, H.; Ewers, C.; Bisgaard, M. Impact of Escherichia coli vaccine on parent stock mortality, first week mortality of broilers and population diversity of E. coli in vaccinated flocks. Avian Pathol. 2010, 39, 287–295. [Google Scholar] [CrossRef]
  26. Wang, S.; Peng, Q.; Jia, H.M.; Zeng, X.F.; Zhu, J.L.; Hou, C.L.; Liu, X.T.; Yang, F.J.; Qiao, S.Y. Prevention of Escherichia coli infection in broiler chickens with Lactobacillus plantarum B1. Poult. Sci. 2017, 96, 2576–2586. [Google Scholar] [CrossRef] [PubMed]
  27. Kaikabo, A.A.; AbdulKarim, S.M.; Abas, F. Evaluation of the efficacy of chitosan nanoparticles loaded PhiKAZ14 bacteriophage in the biological control of colibacillosis in chickens. Poult. Sci. 2017, 96, 295–302. [Google Scholar] [CrossRef] [PubMed]
  28. Allan, B.; Wheler, C.; Koster, W.; Sarfraz, M.; Potter, A.; Gerdts, V.; Dar, A. In Ovo Administration of Innate Immune Stimulants and Protection from Early Chick Mortalities due to Yolk Sac Infection. Avian Dis. 2018, 62, 316–321. [Google Scholar] [CrossRef]
  29. Peng, L.Y.; Yuan, M.; Cui, Z.Q.; Wu, Z.M.; Yu, Z.J.; Song, K.; Tang, B.; Fu, B.D. Rutin inhibits quorum sensing, biofilm formation and virulence genes in avian pathogenic Escherichia coli. Microb. Pathog. 2018, 119, 54–59. [Google Scholar] [CrossRef]
  30. Wang, S.; Zeng, X.; Yang, Q.; Qiao, S. Antimicrobial Peptides as Potential Alternatives to Antibiotics in Food Animal Industry. Int. J. Mol. Sci. 2016, 17, 603. [Google Scholar] [CrossRef]
  31. Sarowska, J.; Futoma-Koloch, B.; Jama-Kmiecik, A.; Frej-Madrzak, M.; Ksiazczyk, M.; Bugla-Ploskonska, G.; Choroszy-Krol, I. Virulence factors, prevalence and potential transmission of extraintestinal pathogenic Escherichia coli isolated from different sources: Recent reports. Gut Pathog. 2019, 11, 10. [Google Scholar] [CrossRef] [Green Version]
  32. Awad, A.M.; El-Shall, N.A.; Khalil, D.S.; Abd El-Hack, M.E.; Swelum, A.A.; Mahmoud, A.H.; Ebaid, H.; Komany, A.; Sammour, R.H.; Sedeik, M.E. Incidence, Pathotyping, and Antibiotic Susceptibility of Avian Pathogenic Escherichia coli among Diseased Broiler Chicks. Pathogens 2020, 9, 114. [Google Scholar] [CrossRef] [Green Version]
  33. Mohamed, L.; Ge, Z.; Yuehua, L.; Yubin, G.; Rachid, K.; Mustapha, O.; Junwei, W.; Karine, O. Virulence traits of avian pathogenic (APEC) and fecal (AFEC) E. coli isolated from broiler chickens in Algeria. Trop. Anim. Health Prod. 2018, 50, 547–553. [Google Scholar] [CrossRef] [PubMed]
  34. Azam, M.; Mohsin, M.; Sajjad Ur, R.; Saleemi, M.K. Virulence-associated genes and antimicrobial resistance among avian pathogenic Escherichia coli from colibacillosis affected broilers in Pakistan. Trop. Anim. Health Prod. 2019, 51, 1259–1265. [Google Scholar] [CrossRef] [PubMed]
  35. Maciel, J.F.; Matter, L.B.; Trindade, M.M.; Camillo, G.; Lovato, M.; de Avila Botton, S.; Castagna de Vargas, A. Virulence factors and antimicrobial susceptibility profile of extraintestinal Escherichia coli isolated from an avian colisepticemia outbreak. Microb. Pathog. 2017, 103, 119–122. [Google Scholar] [CrossRef] [PubMed]
  36. Silveira, F.; Maluta, R.P.; Tiba, M.R.; de Paiva, J.B.; Guastalli, E.A.; da Silveira, W.D. Comparison between avian pathogenic (APEC) and avian faecal (AFEC) Escherichia coli isolated from different regions in Brazil. Vet. J. 2016, 217, 65–67. [Google Scholar] [CrossRef] [PubMed]
  37. De Paiva, J.B.; da Silva, L.P.; Casas, M.R.; Conceicao, R.A.; Nakazato, G.; de Pace, F.; Sperandio, V.; da Silveira, W.D. In vivo influence of in vitro up-regulated genes in the virulence of an APEC strain associated with swollen head syndrome. Avian Pathol. 2016, 45, 94–105. [Google Scholar] [CrossRef] [Green Version]
  38. Wang, X.; Cao, C.; Huan, H.; Zhang, L.; Mu, X.; Gao, Q.; Dong, X.; Gao, S.; Liu, X. Isolation, identification, and pathogenicity of O142 avian pathogenic Escherichia coli causing black proventriculus and septicemia in broiler breeders. Infect. Genet. Evol. 2015, 32, 23–29. [Google Scholar] [CrossRef] [PubMed]
  39. Aslam, M.; Toufeer, M.; Narvaez Bravo, C.; Lai, V.; Rempel, H.; Manges, A.; Diarra, M.S. Characterization of Extraintestinal Pathogenic Escherichia coli isolated from retail poultry meats from Alberta, Canada. Int. J. Food Microbiol. 2014, 177, 49–56. [Google Scholar] [CrossRef]
  40. Delannoy, S.; Schouler, C.; Souillard, R.; Yousfi, L.; Le Devendec, L.; Lucas, C.; Bougeard, S.; Keita, A.; Fach, P.; Galliot, P.; et al. Diversity of Escherichia coli strains isolated from day-old broiler chicks, their environment and colibacillosis lesions in 80 flocks in France. Vet. Microbiol. 2021, 252, 108923. [Google Scholar] [CrossRef]
  41. Zhuge, X.; Zhou, Z.; Jiang, M.; Wang, Z.; Sun, Y.; Tang, F.; Xue, F.; Ren, J.; Dai, J. Chicken-source Escherichia coli within phylogroup F shares virulence genotypes and is closely related to extraintestinal pathogenic E. coli causing human infections. Transbound Emerg. Dis. 2020. [Google Scholar] [CrossRef]
  42. Li, Y.; Wang, H.; Ren, J.; Chen, L.; Zhuge, X.; Hu, L.; Li, D.; Tang, F.; Dai, J. The YfcO fimbriae gene enhances adherence and colonization abilities of avian pathogenic Escherichia coli in vivo and in vitro. Microb. Pathog. 2016, 100, 56–61. [Google Scholar] [CrossRef]
  43. Verma, R.; Rojas, T.C.; Maluta, R.P.; Leite, J.L.; da Silva, L.P.; Nakazato, G.; Dias da Silveira, W. Fimbria-Encoding Gene yadC Has a Pleiotropic Effect on Several Biological Characteristics and Plays a Role in Avian Pathogenic Escherichia coli Pathogenicity. Infect. Immun. 2016, 84, 187–193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Zhuge, X.; Wang, S.; Fan, H.; Pan, Z.; Ren, J.; Yi, L.; Meng, Q.; Yang, X.; Lu, C.; Dai, J. Characterization and functional analysis of AatB, a novel autotransporter adhesin and virulence factor of avian pathogenic Escherichia coli. Infect. Immun. 2013, 81, 2437–2447. [Google Scholar] [CrossRef] [Green Version]
  45. Zhu-Ge, X.K.; Pan, Z.H.; Tang, F.; Mao, X.; Hu, L.; Wang, S.H.; Xu, B.; Lu, C.P.; Fan, H.J.; Dai, J.J. The effects of upaB deletion and the double/triple deletion of upaB, aatA, and aatB genes on pathogenicity of avian pathogenic Escherichia coli. Appl. Microbiol. Biotechnol. 2015, 99, 10639–10654. [Google Scholar] [CrossRef]
  46. Ali, A.; Kolenda, R.; Khan, M.M.; Weinreich, J.; Li, G.; Wieler, L.H.; Tedin, K.; Roggenbuck, D.; Schierack, P. Novel Avian Pathogenic Escherichia coli Genes Responsible for Adhesion to Chicken and Human Cell Lines. Appl. Environ. Microbiol. 2020, 86. [Google Scholar] [CrossRef]
  47. Wang, S.; Niu, C.; Shi, Z.; Xia, Y.; Yaqoob, M.; Dai, J.; Lu, C. Effects of ibeA deletion on virulence and biofilm formation of avian pathogenic Escherichia coli. Infect. Immun. 2011, 79, 279–287. [Google Scholar] [CrossRef] [Green Version]
  48. Wang, S.; Shi, Z.; Xia, Y.; Li, H.; Kou, Y.; Bao, Y.; Dai, J.; Lu, C. IbeB is involved in the invasion and pathogenicity of avian pathogenic Escherichia coli. Vet. Microbiol. 2012, 159, 411–419. [Google Scholar] [CrossRef]
  49. Wang, S.; Bao, Y.; Meng, Q.; Xia, Y.; Zhao, Y.; Wang, Y.; Tang, F.; ZhuGe, X.; Yu, S.; Han, X.; et al. IbeR facilitates stress-resistance, invasion and pathogenicity of avian pathogenic Escherichia coli. PLoS ONE 2015, 10, e0119698. [Google Scholar] [CrossRef]
  50. Pilatti, L.; Boldrin de Paiva, J.; Rojas, T.C.; Leite, J.L.; Conceicao, R.A.; Nakazato, G.; Dias da Silveira, W. The virulence factor ychO has a pleiotropic action in an Avian Pathogenic Escherichia coli (APEC) strain. BMC Microbiol. 2016, 16, 35. [Google Scholar] [CrossRef] [Green Version]
  51. Thomrongsuwannakij, T.; Blackall, P.J.; Djordjevic, S.P.; Cummins, M.L.; Chansiripornchai, N. A comparison of virulence genes, antimicrobial resistance profiles and genetic diversity of avian pathogenic Escherichia coli (APEC) isolates from broilers and broiler breeders in Thailand and Australia. Avian Pathol. 2020, 49, 457–466. [Google Scholar] [CrossRef]
  52. Tu, J.; Xue, T.; Qi, K.; Shao, Y.; Huang, B.; Wang, X.; Zhou, X. The irp2 and fyuA genes in High Pathogenicity Islands are involved in the pathogenesis of infections caused by avian pathogenic Escherichia coli (APEC). Pol. J. Vet. Sci. 2016, 19, 21–29. [Google Scholar] [CrossRef] [Green Version]
  53. Xu, X.; Sun, Q.; Zhao, L. Virulence Factors and Antibiotic Resistance of Avian Pathogenic Escherichia coli in Eastern China. J. Vet. Res. 2019, 63, 317–320. [Google Scholar] [CrossRef] [Green Version]
  54. Varga, C.; Brash, M.L.; Slavic, D.; Boerlin, P.; Ouckama, R.; Weis, A.; Petrik, M.; Philippe, C.; Barham, M.; Guerin, M.T. Evaluating Virulence-Associated Genes and Antimicrobial Resistance of Avian Pathogenic Escherichia coli Isolates from Broiler and Broiler Breeder Chickens in Ontario, Canada. Avian Dis. 2018, 62, 291–299. [Google Scholar] [CrossRef]
  55. Paixao, A.C.; Ferreira, A.C.; Fontes, M.; Themudo, P.; Albuquerque, T.; Soares, M.C.; Fevereiro, M.; Martins, L.; Correa de Sa, M.I. Detection of virulence-associated genes in pathogenic and commensal avian Escherichia coli isolates. Poult. Sci. 2016, 95, 1646–1652. [Google Scholar] [CrossRef]
  56. Sabri, M.; Caza, M.; Proulx, J.; Lymberopoulos, M.H.; Bree, A.; Moulin-Schouleur, M.; Curtiss, R., 3rd; Dozois, C.M. Contribution of the SitABCD, MntH, and FeoB metal transporters to the virulence of avian pathogenic Escherichia coli O78 strain chi7122. Infect. Immun. 2008, 76, 601–611. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Azam, M.; Mohsin, M.; Johnson, T.J.; Smith, E.A.; Johnson, A.; Umair, M.; Saleemi, M.K.; Sajjad Ur, R. Genomic landscape of multi-drug resistant avian pathogenic Escherichia coli recovered from broilers. Vet. Microbiol. 2020, 247, 108766. [Google Scholar] [CrossRef] [PubMed]
  58. Zhang, Z.; Jiang, S.; Liu, Y.; Sun, Y.; Yu, P.; Gong, Q.; Zeng, H.; Li, Y.; Xue, F.; Zhuge, X.; et al. Identification of ireA, 0007, 0008, and 2235 as TonB-dependent receptors in the avian pathogenic Escherichia coli strain DE205B. Vet. Res. 2020, 51, 5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Li, Y.; Dai, J.; Zhuge, X.; Wang, H.; Hu, L.; Ren, J.; Chen, L.; Li, D.; Tang, F. Iron-regulated gene ireA in avian pathogenic Escherichia coli participates in adhesion and stress-resistance. BMC Vet. Res. 2016, 12, 167. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Mu, X.; Gao, R.; Xiao, W.; Gao, Q.; Cao, C.; Xu, H.; Gao, S.; Liu, X. EntE, EntS and TolC synergistically contributed to the pathogenesis of APEC strain E058. Microb. Pathog. 2020, 141, 103990. [Google Scholar] [CrossRef]
  61. Hejair, H.M.A.; Ma, J.; Zhu, Y.; Sun, M.; Dong, W.; Zhang, Y.; Pan, Z.; Zhang, W.; Yao, H. Role of outer membrane protein T in pathogenicity of avian pathogenic Escherichia coli. Res. Vet. Sci. 2017, 115, 109–116. [Google Scholar] [CrossRef]
  62. Song, X.; Qiu, M.; Jiang, H.; Xue, M.; Hu, J.; Liu, H.; Zhou, X.; Tu, J.; Qi, K. ybjX mutation regulated avian pathogenic Escherichia coli pathogenicity though stress-resistance pathway. Avian Pathol. 2020, 49, 144–152. [Google Scholar] [CrossRef]
  63. Song, X.; Hou, M.; Tu, J.; Xue, M.; Shao, Y.; Jiang, H.; Liu, H.; Xue, T.; Wang, G.; Qi, K. Outer membrane proteins YbjX and PagP co-regulate motility in Escherichia coli via the bacterial chemotaxis pathway. Res. Vet. Sci. 2019, 125, 279–284. [Google Scholar] [CrossRef]
  64. Nielsen, D.W.; Ricker, N.; Barbieri, N.L.; Allen, H.K.; Nolan, L.K.; Logue, C.M. Outer membrane protein A (OmpA) of extraintestinal pathogenic Escherichia coli. BMC Res. Notes 2020, 13, 51. [Google Scholar] [CrossRef] [PubMed]
  65. Zuo, J.; Tu, C.; Wang, Y.; Qi, K.; Hu, J.; Wang, Z.; Mi, R.; Yan, H.; Chen, Z.; Han, X. The role of the wzy gene in lipopolysaccharide biosynthesis and pathogenesis of avian pathogenic Escherichia coli. Microb. Pathog. 2019, 127, 296–303. [Google Scholar] [CrossRef] [PubMed]
  66. Han, Y.; Han, X.; Wang, S.; Meng, Q.; Zhang, Y.; Ding, C.; Yu, S. The waaL gene is involved in lipopolysaccharide synthesis and plays a role on the bacterial pathogenesis of avian pathogenic Escherichia coli. Vet. Microbiol. 2014, 172, 486–491. [Google Scholar] [CrossRef] [PubMed]
  67. Xu, H.; Ling, J.; Gao, Q.; He, H.; Mu, X.; Yan, Z.; Gao, S.; Liu, X. Role of the lpxM lipid A biosynthesis pathway gene in pathogenicity of avian pathogenic Escherichia coli strain E058 in a chicken infection model. Vet. Microbiol. 2013, 166, 516–526. [Google Scholar] [CrossRef]
  68. Gao, Q.; Xia, L.; Wang, X.; Ye, Z.; Liu, J.; Gao, S. SodA Contributes to the Virulence of Avian Pathogenic Escherichia coli O2 Strain E058 in Experimentally Infected Chickens. J. Bacteriol. 2019, 201. [Google Scholar] [CrossRef] [Green Version]
  69. Murase, K.; Martin, P.; Porcheron, G.; Houle, S.; Helloin, E.; Penary, M.; Nougayrede, J.P.; Dozois, C.M.; Hayashi, T.; Oswald, E. HlyF Produced by Extraintestinal Pathogenic Escherichia coli Is a Virulence Factor That Regulates Outer Membrane Vesicle Biogenesis. J. Infect. Dis. 2016, 213, 856–865. [Google Scholar] [CrossRef] [Green Version]
  70. Zhao, S.; Wang, C.L.; Chang, S.K.; Tsai, Y.L.; Chou, C.H. Characterization of Escherichia coli Isolated from Day-old Chicken Fluff in Taiwanese Hatcheries. Avian Dis. 2019, 63, 9–16. [Google Scholar] [CrossRef]
  71. Ibrahim, R.A.; Cryer, T.L.; Lafi, S.Q.; Basha, E.A.; Good, L.; Tarazi, Y.H. Identification of Escherichia coli from broiler chickens in Jordan, their antimicrobial resistance, gene characterization and the associated risk factors. BMC Vet. Res. 2019, 15, 159. [Google Scholar] [CrossRef] [PubMed]
  72. Sgariglia, E.; Aconiti Mandolini, N.; Napoleoni, M.; Medici, L.; Fraticelli, R.; Conquista, M.; Gianfelici, P.; Staffolani, M.; Fisichella, S.; Capuccella, M.; et al. Antibiotic resistance pattern and virulence genesin avian pathogenic Escherichia coli (APEC) from different breeding systems. Vet. Ital. 2019, 55, 26–33. [Google Scholar] [CrossRef]
  73. Dou, X.; Gong, J.; Han, X.; Xu, M.; Shen, H.; Zhang, D.; Zhuang, L.; Liu, J.; Zou, J. Characterization of avian pathogenic Escherichia coli isolated in eastern China. Gene 2016, 576, 244–248. [Google Scholar] [CrossRef] [PubMed]
  74. Zuo, J.; Yin, H.; Hu, J.; Miao, J.; Chen, Z.; Qi, K.; Wang, Z.; Gong, J.; Phouthapane, V.; Jiang, W.; et al. Lsr operon is associated with AI-2 transfer and pathogenicity in avian pathogenic Escherichia coli. Vet. Res. 2019, 50, 109. [Google Scholar] [CrossRef] [Green Version]
  75. Cui, Z.Q.; Wu, Z.M.; Fu, Y.X.; Xu, D.X.; Guo, X.; Shen, H.Q.; Wei, X.B.; Yi, P.F.; Fu, B.D. Autoinducer-2 of quorum sensing is involved in cell damage caused by avian pathogenic Escherichia coli. Microb. Pathog. 2016, 99, 247–252. [Google Scholar] [CrossRef] [PubMed]
  76. Wu, X.; Lv, X.; Lu, J.; Yu, S.; Jin, Y.; Hu, J.; Zuo, J.; Mi, R.; Huang, Y.; Qi, K.; et al. The role of the ptsI gene on AI-2 internalization and pathogenesis of avian pathogenic Escherichia coli. Microb. Pathog. 2017, 113, 321–329. [Google Scholar] [CrossRef]
  77. Xu, D.; Zuo, J.; Chen, Z.; Lv, X.; Hu, J.; Wu, X.; Qi, K.; Mi, R.; Huang, Y.; Miao, J.; et al. Different activated methyl cycle pathways affect the pathogenicity of avian pathogenic Escherichia coli. Vet. Microbiol. 2017, 211, 160–168. [Google Scholar] [CrossRef]
  78. Xue, M.; Xiao, Y.; Fu, D.; Raheem, M.A.; Shao, Y.; Song, X.; Tu, J.; Xue, T.; Qi, K. Transcriptional Regulator YqeI, Locating at ETT2 Locus, Affects the Pathogenicity of Avian Pathogenic Escherichia coli. Animals 2020, 10, 1658. [Google Scholar] [CrossRef]
  79. Wang, S.; Liu, X.; Xu, X.; Yang, D.; Wang, D.; Han, X.; Shi, Y.; Tian, M.; Ding, C.; Peng, D.; et al. Escherichia coli Type III Secretion System 2 ATPase EivC Is Involved in the Motility and Virulence of Avian Pathogenic Escherichia coli. Front. Microbiol. 2016, 7, 1387. [Google Scholar] [CrossRef] [Green Version]
  80. Yi, Z.; Wang, D.; Xin, S.; Zhou, D.; Li, T.; Tian, M.; Qi, J.; Ding, C.; Wang, S.; Yu, S. The CpxR regulates type VI secretion system 2 expression and facilitates the interbacterial competition activity and virulence of avian pathogenic Escherichia coli. Vet. Res. 2019, 50, 40. [Google Scholar] [CrossRef] [Green Version]
  81. Matter, L.B.; Ares, M.A.; Abundes-Gallegos, J.; Cedillo, M.L.; Yanez, J.A.; Martinez-Laguna, Y.; De la Cruz, M.A.; Giron, J.A. The CpxRA stress response system regulates virulence features of avian pathogenic Escherichia coli. Environ. Microbiol. 2018, 20, 3363–3377. [Google Scholar] [CrossRef]
  82. Wang, S.; Dai, J.; Meng, Q.; Han, X.; Han, Y.; Zhao, Y.; Yang, D.; Ding, C.; Yu, S. DotU expression is highly induced during in vivo infection and responsible for virulence and Hcp1 secretion in avian pathogenic Escherichia coli. Front. Microbiol. 2014, 5, 588. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. de Pace, F.; Boldrin de Paiva, J.; Nakazato, G.; Lancellotti, M.; Sircili, M.P.; Guedes Stehling, E.; Dias da Silveira, W.; Sperandio, V. Characterization of IcmF of the type VI secretion system in an avian pathogenic Escherichia coli (APEC) strain. Microbiology 2011, 157, 2954–2962. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Ding, X.; Zhang, Q.; Wang, H.; Quan, G.; Zhang, D.; Ren, W.; Liao, Y.; Xia, P.; Zhu, G. The different roles of hcp1 and hcp2 of the type VI secretion system in Escherichia coli strain CE129. J. Basic Microbiol. 2018, 58, 938–946. [Google Scholar] [CrossRef] [PubMed]
  85. Song, X.; Hou, M.; Jiang, H.; Shen, X.; Xue, M.; Shao, Y.; Wang, L.; He, Q.; Zheng, L.; Tu, J.; et al. Hcp2a of type VI secretion system contributes to IL8 and IL1β expression of chicken tracheal epithelium by affecting APEC colonization. Res. Vet. Sci. 2020, 132, 279–284. [Google Scholar] [CrossRef]
  86. Tu, J.; Huang, B.; Zhang, Y.; Xue, T.; Li, S.; Qi, K. Modulation of virulence genes by the two-component system PhoP-PhoQ in avian pathogenic Escherichia coli. Pol. J. Vet. Sci. 2016, 19, 31–40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Xue, M.; Raheem, M.A.; Gu, Y.; Lu, H.; Song, X.; Tu, J.; Xue, T.; Qi, K. The KdpD/KdpE two-component system contributes to the motility and virulence of avian pathogenic Escherichia coli. Res. Vet. Sci. 2020, 131, 24–30. [Google Scholar] [CrossRef] [PubMed]
  88. Gao, Q.; Su, S.; Li, X.; Wang, H.; Liu, J.; Gao, S. Transcriptional analysis of RstA/RstB in avian pathogenic Escherichia coli identifies its role in the regulation of hdeD-mediated virulence and survival in chicken macrophages. Vet. Microbiol. 2020, 241, 108555. [Google Scholar] [CrossRef]
  89. Herren, C.D.; Mitra, A.; Palaniyandi, S.K.; Coleman, A.; Elankumaran, S.; Mukhopadhyay, S. The BarA-UvrY two-component system regulates virulence in avian pathogenic Escherichia coli O78:K80:H9. Infect. Immun. 2006, 74, 4900–4909. [Google Scholar] [CrossRef] [Green Version]
  90. Yin, L.; Li, Q.; Xue, M.; Wang, Z.; Tu, J.; Song, X.; Shao, Y.; Han, X.; Xue, T.; Liu, H.; et al. The role of the phoP transcriptional regulator on biofilm formation of avian pathogenic Escherichia coli. Avian Pathol. 2019, 48, 362–370. [Google Scholar] [CrossRef]
  91. Yu, L.; Wang, H.; Han, X.; Li, W.; Xue, M.; Qi, K.; Chen, X.; Ni, J.; Deng, R.; Shang, F.; et al. The two-component system, BasSR, is involved in the regulation of biofilm and virulence in avian pathogenic Escherichia coli. Avian Pathol. 2020, 49, 532–546. [Google Scholar] [CrossRef] [PubMed]
  92. Gao, Q.; Ye, Z.; Wang, X.; Mu, X.; Gao, S.; Liu, X. RstA is required for the virulence of an avian pathogenic Escherichia coli O2 strain E058. Infect. Genet. Evol. 2015, 29, 180–188. [Google Scholar] [CrossRef] [PubMed]
  93. Wiebe, H.; Gurlebeck, D.; Gross, J.; Dreck, K.; Pannen, D.; Ewers, C.; Wieler, L.H.; Schnetz, K. YjjQ Represses Transcription of flhDC and Additional Loci in Escherichia coli. J. Bacteriol. 2015, 197, 2713–2720. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Yu, L.; Li, W.; Qi, K.; Wang, S.; Chen, X.; Ni, J.; Deng, R.; Shang, F.; Xue, T. McbR is involved in biofilm formation and H2O2 stress response in avian pathogenic Escherichia coli X40. Poult. Sci. 2019, 98, 4094–4103. [Google Scholar] [CrossRef]
  95. Gao, Q.; Xu, H.; Wang, X.; Zhang, D.; Ye, Z.; Gao, S.; Liu, X. RfaH promotes the ability of the avian pathogenic Escherichia coli O2 strain E058 to cause avian colibacillosis. J. Bacteriol. 2013, 195, 2474–2480. [Google Scholar] [CrossRef] [Green Version]
  96. Zhuge, X.; Sun, Y.; Jiang, M.; Wang, J.; Tang, F.; Xue, F.; Ren, J.; Zhu, W.; Dai, J. Acetate metabolic requirement of avian pathogenic Escherichia coli promotes its intracellular proliferation within macrophage. Vet. Res. 2019, 50, 31. [Google Scholar] [CrossRef] [Green Version]
  97. Li, D.; Chen, Y.; Qian, X.; Liu, Y.; Ren, J.; Xue, F.; Sun, J.; Tang, F.; Dai, J. orf20 in prophage phiv142-3 contributes to the adhesion and colonization ability of avian pathogenic Escherichia coli strain DE142 by affecting the formation of flagella and I fimbriae. Vet. Microbiol. 2019, 235, 301–309. [Google Scholar] [CrossRef] [PubMed]
  98. Li, D.; Tang, F.; Xue, F.; Ren, J.; Liu, Y.; Yang, D.; Dai, J. Prophage phiv142-3 enhances the colonization and resistance to environmental stresses of avian pathogenic Escherichia coli. Vet. Microbiol. 2018, 218, 70–77. [Google Scholar] [CrossRef]
  99. Liu, Y.; Gong, Q.; Qian, X.; Li, D.; Zeng, H.; Li, Y.; Xue, F.; Ren, J.; Zhu Ge, X.; Tang, F.; et al. Prophage phiv205-1 facilitates biofilm formation and pathogenicity of avian pathogenic Escherichia coli strain DE205B. Vet. Microbiol. 2020, 247, 108752. [Google Scholar] [CrossRef]
  100. Hejair, H.M.A.; Zhu, Y.; Ma, J.; Zhang, Y.; Pan, Z.; Zhang, W.; Yao, H. Functional role of ompF and ompC porins in pathogenesis of avian pathogenic Escherichia coli. Microb. Pathog. 2017, 107, 29–37. [Google Scholar] [CrossRef]
  101. Verma, R.; Rojas, T.C.G.; Maluta, R.P.; Leite, J.L.; Nakazato, G.; de Silveira, W.D. Role of hypothetical protein YicS in the pathogenicity of Avian Pathogenic Escherichia coli in vivo and in vitro. Microbiol. Res. 2018, 214, 28–36. [Google Scholar] [CrossRef]
  102. Liu, H.; Chen, L.; Si, W.; Wang, C.; Zhu, F.; Li, G.; Liu, S. Physiology and pathogenicity of cpdB deleted mutant of avian pathogenic Escherichia coli. Res. Vet. Sci. 2017, 111, 21–25. [Google Scholar] [CrossRef] [PubMed]
  103. Lamarche, M.G.; Dozois, C.M.; Daigle, F.; Caza, M.; Curtiss, R., 3rd; Dubreuil, J.D.; Harel, J. Inactivation of the pst system reduces the virulence of an avian pathogenic Escherichia coli O78 strain. Infect. Immun. 2005, 73, 4138–4145. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Mu, X.; Huan, H.; Xu, H.; Gao, Q.; Xiong, L.; Gao, R.; Gao, S.; Liu, X. The transfer-messenger RNA-small protein B system plays a role in avian pathogenic Escherichia coli pathogenicity. J. Bacteriol. 2013, 195, 5064–5071. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Vanderkelen, L.; Ons, E.; Van Herreweghe, J.M.; Callewaert, L.; Goddeeris, B.M.; Michiels, C.W. Role of lysozyme inhibitors in the virulence of avian pathogenic Escherichia coli. PLoS ONE 2012, 7, e45954. [Google Scholar] [CrossRef] [Green Version]
  106. Mbanga, J.; Nyararai, Y.O. Virulence gene profiles of avian pathogenic Escherichia coli isolated from chickens with colibacillosis in Bulawayo, Zimbabwe. Onderstepoort J. Vet. Res. 2015, 82, e1–e8. [Google Scholar] [CrossRef] [Green Version]
  107. Hayashi, W.; Tanaka, H.; Taniguchi, Y.; Iimura, M.; Soga, E.; Kubo, R.; Matsuo, N.; Kawamura, K.; Arakawa, Y.; Nagano, Y.; et al. Acquisition of mcr-1 and Cocarriage of Virulence Genes in Avian Pathogenic Escherichia coli Isolates from Municipal Wastewater Influents in Japan. Appl. Environ. Microbiol. 2019, 85. [Google Scholar] [CrossRef]
  108. Saha, O.; Hoque, M.N.; Islam, O.K.; Rahaman, M.M.; Sultana, M.; Hossain, M.A. Multidrug-Resistant Avian Pathogenic Escherichia coli Strains and Association of Their Virulence Genes in Bangladesh. Microorganisms 2020, 8, 1135. [Google Scholar] [CrossRef] [PubMed]
  109. Tuntufye, H.N.; Lebeer, S.; Gwakisa, P.S.; Goddeeris, B.M. Identification of Avian pathogenic Escherichia coli genes that are induced in vivo during infection in chickens. Appl. Environ. Microbiol. 2012, 78, 3343–3351. [Google Scholar] [CrossRef] [Green Version]
  110. Dozois, C.M.; Daigle, F.; Curtiss, R., 3rd. Identification of pathogen-specific and conserved genes expressed in vivo by an avian pathogenic Escherichia coli strain. Proc. Natl. Acad. Sci. USA 2003, 100, 247–252. [Google Scholar] [CrossRef] [Green Version]
  111. Zhang, H.; Chen, X.; Nolan, L.K.; Zhang, W.; Li, G. Identification of Host Adaptation Genes in Extraintestinal Pathogenic Escherichia coli during Infection in Different Hosts. Infect. Immun. 2019, 87. [Google Scholar] [CrossRef]
  112. Forsyth, V.S.; Himpsl, S.D.; Smith, S.N.; Sarkissian, C.A.; Mike, L.A.; Stocki, J.A.; Sintsova, A.; Alteri, C.J.; Mobley, H.L.T. Optimization of an Experimental Vaccine To Prevent Escherichia coli Urinary Tract Infection. mBio 2020, 11, e00555-20. [Google Scholar] [CrossRef]
  113. Linciano, P.; Cavalloro, V.; Martino, E.; Kirchmair, J.; Listro, R.; Rossi, D.; Collina, S. Tackling Antimicrobial Resistance with Small Molecules Targeting LsrK: Challenges and Opportunities. J. Med. Chem. 2020, 63, 15243–15257. [Google Scholar] [CrossRef] [PubMed]
  114. Murima, P.; McKinney, J.D.; Pethe, K. Targeting Bacterial Central Metabolism for Drug Development. Chem. Biol. 2014, 21, 1423–1432. [Google Scholar] [CrossRef] [PubMed]
  115. Marshall, N.C.; Finlay, B.B. Targeting the type III secretion system to treat bacterial infections. Expert Opin. Ther. Targets 2014, 18, 137–152. [Google Scholar] [CrossRef] [PubMed]
  116. Jorgensen, S.L.; Stegger, M.; Kudirkiene, E.; Lilje, B.; Poulsen, L.L.; Ronco, T.; Pires Dos Santos, T.; Kiil, K.; Bisgaard, M.; Pedersen, K.; et al. Diversity and Population Overlap between Avian and Human Escherichia coli Belonging to Sequence Type 95. mSphere 2019, 4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Najafi, S.; Rahimi, M.; Nikousefat, Z. Extra-intestinal pathogenic Escherichia coli from human and avian origin: Detection of the most common virulence-encoding genes. Vet. Res. Forum 2019, 10, 43–49. [Google Scholar] [CrossRef] [PubMed]
  118. Zhuge, X.; Jiang, M.; Tang, F.; Sun, Y.; Ji, Y.; Xue, F.; Ren, J.; Zhu, W.; Dai, J. Avian-source mcr-1-positive Escherichia coli is phylogenetically diverse and shares virulence characteristics with E. coli causing human extra-intestinal infections. Vet. Microbiol. 2019, 239, 108483. [Google Scholar] [CrossRef]
  119. Cunha, M.P.V.; Saidenberg, A.B.; Moreno, A.M.; Ferreira, A.J.P.; Vieira, M.A.M.; Gomes, T.A.T.; Knobl, T. Pandemic extra-intestinal pathogenic Escherichia coli (ExPEC) clonal group O6-B2-ST73 as a cause of avian colibacillosis in Brazil. PLoS ONE 2017, 12, e0178970. [Google Scholar] [CrossRef] [Green Version]
  120. Johnson, J.R.; Porter, S.B.; Johnston, B.; Thuras, P.; Clock, S.; Crupain, M.; Rangan, U. Extraintestinal Pathogenic and Antimicrobial-Resistant Escherichia coli, Including Sequence Type 131 (ST131), from Retail Chicken Breasts in the United States in 2013. Appl. Environ. Microbiol. 2017, 83. [Google Scholar] [CrossRef] [Green Version]
  121. Mitchell, N.M.; Johnson, J.R.; Johnston, B.; Curtiss, R., 3rd; Mellata, M. Zoonotic potential of Escherichia coli isolates from retail chicken meat products and eggs. Appl. Environ. Microbiol. 2015, 81, 1177–1187. [Google Scholar] [CrossRef] [Green Version]
  122. Jakobsen, L.; Spangholm, D.J.; Pedersen, K.; Jensen, L.B.; Emborg, H.D.; Agerso, Y.; Aarestrup, F.M.; Hammerum, A.M.; Frimodt-Moller, N. Broiler chickens, broiler chicken meat, pigs and pork as sources of ExPEC related virulence genes and resistance in Escherichia coli isolates from community-dwelling humans and UTI patients. Int. J. Food Microbiol. 2010, 142, 264–272. [Google Scholar] [CrossRef] [PubMed]
  123. Ewers, C.; Li, G.; Wilking, H.; Kiessling, S.; Alt, K.; Antao, E.M.; Laturnus, C.; Diehl, I.; Glodde, S.; Homeier, T.; et al. Avian pathogenic, uropathogenic, and newborn meningitis-causing Escherichia coli: How closely related are they? Int. J. Med. Microbiol. 2007, 297, 163–176. [Google Scholar] [CrossRef] [PubMed]
  124. Zhu Ge, X.; Jiang, J.; Pan, Z.; Hu, L.; Wang, S.; Wang, H.; Leung, F.C.; Dai, J.; Fan, H. Comparative genomic analysis shows that avian pathogenic Escherichia coli isolate IMT5155 (O2:K1:H5; ST complex 95, ST140) shares close relationship with ST95 APEC O1:K1 and human ExPEC O18:K1 strains. PLoS ONE 2014, 9, e112048. [Google Scholar] [CrossRef] [Green Version]
  125. Johnson, T.J.; Wannemuehler, Y.; Johnson, S.J.; Stell, A.L.; Doetkott, C.; Johnson, J.R.; Kim, K.S.; Spanjaard, L.; Nolan, L.K. Comparison of extraintestinal pathogenic Escherichia coli strains from human and avian sources reveals a mixed subset representing potential zoonotic pathogens. Appl. Environ. Microbiol. 2008, 74, 7043–7050. [Google Scholar] [CrossRef] [Green Version]
  126. Mora, A.; Lopez, C.; Dabhi, G.; Blanco, M.; Blanco, J.E.; Alonso, M.P.; Herrera, A.; Mamani, R.; Bonacorsi, S.; Moulin-Schouleur, M.; et al. Extraintestinal pathogenic Escherichia coli O1:K1:H7/NM from human and avian origin: Detection of clonal groups B2 ST95 and D ST59 with different host distribution. BMC Microbiol. 2009, 9, 132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Moulin-Schouleur, M.; Reperant, M.; Laurent, S.; Bree, A.; Mignon-Grasteau, S.; Germon, P.; Rasschaert, D.; Schouler, C. Extraintestinal pathogenic Escherichia coli strains of avian and human origin: Link between phylogenetic relationships and common virulence patterns. J. Clin. Microbiol. 2007, 45, 3366–3376. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Johnson, T.J.; Kariyawasam, S.; Wannemuehler, Y.; Mangiamele, P.; Johnson, S.J.; Doetkott, C.; Skyberg, J.A.; Lynne, A.M.; Johnson, J.R.; Nolan, L.K. The genome sequence of avian pathogenic Escherichia coli strain O1:K1:H7 shares strong similarities with human extraintestinal pathogenic E. coli genomes. J. Bacteriol. 2007, 189, 3228–3236. [Google Scholar] [CrossRef] [Green Version]
  129. Moulin-Schouleur, M.; Schouler, C.; Tailliez, P.; Kao, M.R.; Bree, A.; Germon, P.; Oswald, E.; Mainil, J.; Blanco, M.; Blanco, J. Common virulence factors and genetic relationships between O18:K1:H7 Escherichia coli isolates of human and avian origin. J. Clin. Microbiol. 2006, 44, 3484–3492. [Google Scholar] [CrossRef] [Green Version]
  130. Maluta, R.P.; Logue, C.M.; Casas, M.R.; Meng, T.; Guastalli, E.A.; Rojas, T.C.; Montelli, A.C.; Sadatsune, T.; de Carvalho Ramos, M.; Nolan, L.K.; et al. Overlapped sequence types (STs) and serogroups of avian pathogenic (APEC) and human extra-intestinal pathogenic (ExPEC) Escherichia coli isolated in Brazil. PLoS ONE 2014, 9, e105016. [Google Scholar] [CrossRef]
  131. Danzeisen, J.L.; Wannemuehler, Y.; Nolan, L.K.; Johnson, T.J. Comparison of multilocus sequence analysis and virulence genotyping of Escherichia coli from live birds, retail poultry meat, and human extraintestinal infection. Avian Dis. 2013, 57, 104–108. [Google Scholar] [CrossRef]
  132. Bergeron, C.R.; Prussing, C.; Boerlin, P.; Daignault, D.; Dutil, L.; Reid-Smith, R.J.; Zhanel, G.G.; Manges, A.R. Chicken as reservoir for extraintestinal pathogenic Escherichia coli in humans, Canada. Emerg. Infect. Dis. 2012, 18, 415–421. [Google Scholar] [CrossRef] [PubMed]
  133. Mora, A.; Viso, S.; Lopez, C.; Alonso, M.P.; Garcia-Garrote, F.; Dabhi, G.; Mamani, R.; Herrera, A.; Marzoa, J.; Blanco, M.; et al. Poultry as reservoir for extraintestinal pathogenic Escherichia coli O45:K1:H7-B2-ST95 in humans. Vet. Microbiol. 2013, 167, 506–512. [Google Scholar] [CrossRef] [PubMed]
  134. Peigne, C.; Bidet, P.; Mahjoub-Messai, F.; Plainvert, C.; Barbe, V.; Medigue, C.; Frapy, E.; Nassif, X.; Denamur, E.; Bingen, E.; et al. The plasmid of Escherichia coli strain S88 (O45:K1:H7) that causes neonatal meningitis is closely related to avian pathogenic E. coli plasmids and is associated with high-level bacteremia in a neonatal rat meningitis model. Infect. Immun. 2009, 77, 2272–2284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Giufre, M.; Graziani, C.; Accogli, M.; Luzzi, I.; Busani, L.; Cerquetti, M. Escherichia coli of human and avian origin: Detection of clonal groups associated with fluoroquinolone and multidrug resistance in Italy. J. Antimicrob. Chemother. 2012, 67, 860–867. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Rodriguez-Siek, K.E.; Giddings, C.W.; Doetkott, C.; Johnson, T.J.; Fakhr, M.K.; Nolan, L.K. Comparison of Escherichia coli isolates implicated in human urinary tract infection and avian colibacillosis. Microbiology 2005, 151, 2097–2110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Mellata, M.; Johnson, J.R.; Curtiss, R., 3rd. Escherichia coli isolates from commercial chicken meat and eggs cause sepsis, meningitis and urinary tract infection in rodent models of human infections. Zoonoses Public Health 2018, 65, 103–113. [Google Scholar] [CrossRef]
  138. Stromberg, Z.R.; Johnson, J.R.; Fairbrother, J.M.; Kilbourne, J.; Van Goor, A.; Curtiss, R.R.; Mellata, M. Evaluation of Escherichia coli isolates from healthy chickens to determine their potential risk to poultry and human health. PLoS ONE 2017, 12, e0180599. [Google Scholar] [CrossRef] [Green Version]
  139. Nandanwar, N.; Janssen, T.; Kuhl, M.; Ahmed, N.; Ewers, C.; Wieler, L.H. Extraintestinal pathogenic Escherichia coli (ExPEC) of human and avian origin belonging to sequence type complex 95 (STC95) portray indistinguishable virulence features. Int. J. Med. Microbiol. 2014, 304, 835–842. [Google Scholar] [CrossRef]
  140. Krishnan, S.; Chang, A.C.; Hodges, J.; Couraud, P.O.; Romero, I.A.; Weksler, B.; Nicholson, B.A.; Nolan, L.K.; Prasadarao, N.V. Serotype O18 avian pathogenic and neonatal meningitis Escherichia coli strains employ similar pathogenic strategies for the onset of meningitis. Virulence 2015, 6, 777–786. [Google Scholar] [CrossRef] [Green Version]
  141. Mortensen, S.; Johansen, A.E.; Thofner, I.; Christensen, J.P.; Pors, S.E.; Fresno, A.H.; Moller-Jensen, J.; Olsen, J.E. Infectious potential of human derived uropathogenic Escherichia coli UTI89 in the reproductive tract of laying hens. Vet. Microbiol. 2019, 239, 108445. [Google Scholar] [CrossRef]
  142. Gao, Q.; Zhang, D.; Ye, Z.; Zhu, X.; Yang, W.; Dong, L.; Gao, S.; Liu, X. Virulence traits and pathogenicity of uropathogenic Escherichia coli isolates with common and uncommon O serotypes. Microb. Pathog. 2017, 104, 217–224. [Google Scholar] [CrossRef]
  143. Meena, P.R.; Yadav, P.; Hemlata, H.; Tejavath, K.K.; Singh, A.P. Poultry-origin extraintestinal Escherichia coli strains carrying the traits associated with urinary tract infection, sepsis, meningitis and avian colibacillosis in India. J. Appl. Microbiol. 2020. [Google Scholar] [CrossRef]
  144. Zhao, L.; Gao, S.; Huan, H.; Xu, X.; Zhu, X.; Yang, W.; Gao, Q.; Liu, X. Comparison of virulence factors and expression of specific genes between uropathogenic Escherichia coli and avian pathogenic E. coli in a murine urinary tract infection model and a chicken challenge model. Microbiology 2009, 155, 1634–1644. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Christensen, H.; Bachmeier, J.; Bisgaard, M. New strategies to prevent and control avian pathogenic Escherichia coli (APEC). Avian Pathol. 2020, 1–30. [Google Scholar] [CrossRef]
  146. Dhaouadi, S.; Soufi, L.; Hamza, A.; Fedida, D.; Zied, C.; Awadhi, E.; Mtibaa, M.; Hassen, B.; Cherif, A.; Torres, C.; et al. Co-occurrence of mcr-1 mediated colistin resistance and beta-lactamases encoding genes in Multidrug-resistant Escherichia coli from broiler chickens with colibacillosis in Tunisia. J. Glob. Antimicrob. Resist. 2020. [Google Scholar] [CrossRef]
  147. Kim, Y.B.; Yoon, M.Y.; Ha, J.S.; Seo, K.W.; Noh, E.B.; Son, S.H.; Lee, Y.J. Molecular characterization of avian pathogenic Escherichia coli from broiler chickens with colibacillosis. Poult. Sci. 2020, 99, 1088–1095. [Google Scholar] [CrossRef]
  148. Meguenni, N.; Chanteloup, N.; Tourtereau, A.; Ahmed, C.A.; Bounar-Kechih, S.; Schouler, C. Virulence and antibiotic resistance profile of avian Escherichia coli strains isolated from colibacillosis lesions in central of Algeria. Vet. World 2019, 12, 1840–1848. [Google Scholar] [CrossRef] [PubMed]
  149. Sarba, E.J.; Kelbesa, K.A.; Bayu, M.D.; Gebremedhin, E.Z.; Borena, B.M.; Teshale, A. Identification and antimicrobial susceptibility profile of Escherichia coli isolated from backyard chicken in and around ambo, Central Ethiopia. BMC Vet. Res. 2019, 15, 85. [Google Scholar] [CrossRef] [PubMed]
  150. Kurnia, R.S.; Indrawati, A.; Mayasari, N.; Priadi, A. Molecular detection of genes encoding resistance to tetracycline and determination of plasmid-mediated resistance to quinolones in avian pathogenic Escherichia coli in Sukabumi, Indonesia. Vet. World 2018, 11, 1581–1586. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Amer, M.M.; Mekky, H.M.; Amer, A.M.; Fedawy, H.S. Antimicrobial resistance genes in pathogenic Escherichia coli isolated from diseased broiler chickens in Egypt and their relationship with the phenotypic resistance characteristics. Vet. World 2018, 11, 1082–1088. [Google Scholar] [CrossRef]
  152. Gao, J.; Duan, X.; Li, X.; Cao, H.; Wang, Y.; Zheng, S.J. Emerging of a highly pathogenic and multi-drug resistant strain of Escherichia coli causing an outbreak of colibacillosis in chickens. Infect. Genet. Evol. 2018, 65, 392–398. [Google Scholar] [CrossRef] [PubMed]
  153. Vounba, P.; Kane, Y.; Ndiaye, C.; Arsenault, J.; Fairbrother, J.M.; Bada Alambedji, R. Molecular Characterization of Escherichia coli Isolated from Chickens with Colibacillosis in Senegal. Foodborne Pathog. Dis. 2018, 15, 517–525. [Google Scholar] [CrossRef] [PubMed]
  154. Hoepers, P.G.; Silva, P.L.; Rossi, D.A.; Valadares Junior, E.C.; Ferreira, B.C.; Zuffo, J.P.; Koerich, P.K.; Fonseca, B.B. The association between extended spectrum beta-lactamase (ESBL) and ampicillin C (AmpC) beta-lactamase genes with multidrug resistance in Escherichia coli isolates recovered from turkeys in Brazil. Br. Poult. Sci. 2018, 59, 396–401. [Google Scholar] [CrossRef] [PubMed]
  155. Subedi, M.; Luitel, H.; Devkota, B.; Bhattarai, R.K.; Phuyal, S.; Panthi, P.; Shrestha, A.; Chaudhary, D.K. Antibiotic resistance pattern and virulence genes content in avian pathogenic Escherichia coli (APEC) from broiler chickens in Chitwan, Nepal. BMC Vet. Res. 2018, 14, 113. [Google Scholar] [CrossRef] [PubMed]
  156. Azam, M.; Ehsan, I.; Sajjad Ur, R.; Saleemi, M.K.; Javed, M.R.; Mohsin, M. Detection of the colistin resistance gene mcr-1 in avian pathogenic Escherichia coli in Pakistan. J. Glob. Antimicrob. Resist. 2017, 11, 152–153. [Google Scholar] [CrossRef] [PubMed]
  157. Younis, G.; Awad, A.; Mohamed, N. Phenotypic and genotypic characterization of antimicrobial susceptibility of avian pathogenic Escherichia coli isolated from broiler chickens. Vet. World 2017, 10, 1167–1172. [Google Scholar] [CrossRef] [Green Version]
  158. Yassin, A.K.; Gong, J.; Kelly, P.; Lu, G.; Guardabassi, L.; Wei, L.; Han, X.; Qiu, H.; Price, S.; Cheng, D.; et al. Antimicrobial resistance in clinical Escherichia coli isolates from poultry and livestock, China. PLoS ONE 2017, 12, e0185326. [Google Scholar] [CrossRef] [Green Version]
  159. Ozaki, H.; Matsuoka, Y.; Nakagawa, E.; Murase, T. Characteristics of Escherichia coli isolated from broiler chickens with colibacillosis in commercial farms from a common hatchery. Poult. Sci. 2017, 96, 3717–3724. [Google Scholar] [CrossRef]
  160. Lima Barbieri, N.; Nielsen, D.W.; Wannemuehler, Y.; Cavender, T.; Hussein, A.; Yan, S.G.; Nolan, L.K.; Logue, C.M. mcr-1 identified in Avian Pathogenic Escherichia coli (APEC). PLoS ONE 2017, 12, e0172997. [Google Scholar] [CrossRef] [Green Version]
  161. Perreten, V.; Strauss, C.; Collaud, A.; Gerber, D. Colistin Resistance Gene mcr-1 in Avian-Pathogenic Escherichia coli in South Africa. Antimicrob Agents Chemother 2016, 60, 4414–4415. [Google Scholar] [CrossRef] [Green Version]
  162. Sola-Gines, M.; Cameron-Veas, K.; Badiola, I.; Dolz, R.; Majo, N.; Dahbi, G.; Viso, S.; Mora, A.; Blanco, J.; Piedra-Carrasco, N.; et al. Diversity of Multi-Drug Resistant Avian Pathogenic Escherichia coli (APEC) Causing Outbreaks of Colibacillosis in Broilers during 2012 in Spain. PLoS ONE 2015, 10, e0143191. [Google Scholar] [CrossRef] [Green Version]
  163. Cavicchio, L.; Dotto, G.; Giacomelli, M.; Giovanardi, D.; Grilli, G.; Franciosini, M.P.; Trocino, A.; Piccirillo, A. Class 1 and class 2 integrons in avian pathogenic Escherichia coli from poultry in Italy. Poult. Sci. 2015, 94, 1202–1208. [Google Scholar] [CrossRef] [PubMed]
  164. Hornsey, M.; Betts, J.W.; Mehat, J.W.; Wareham, D.W.; van Vliet, A.H.M.; Woodward, M.J.; La Ragione, R.M. Characterization of a colistin-resistant Avian Pathogenic Escherichia coli ST69 isolate recovered from a broiler chicken in Germany. J. Med. Microbiol. 2019, 68, 111–114. [Google Scholar] [CrossRef] [PubMed]
  165. Bourely, C.; Chauvin, C.; Jouy, E.; Cazeau, G.; Jarrige, N.; Leblond, A.; Gay, E. Comparative epidemiology of E. coli resistance to third-generation cephalosporins in diseased food-producing animals. Vet. Microbiol. 2018, 223, 72–78. [Google Scholar] [CrossRef] [PubMed]
  166. Halfaoui, Z.; Menoueri, N.M.; Bendali, L.M. Serogrouping and antibiotic resistance of Escherichia coli isolated from broiler chicken with colibacillosis in center of Algeria. Vet. World 2017, 10, 830–835. [Google Scholar] [CrossRef] [Green Version]
  167. Ievy, S.; Islam, M.S.; Sobur, M.A.; Talukder, M.; Rahman, M.B.; Khan, M.F.R.; Rahman, M.T. Molecular Detection of Avian Pathogenic Escherichia coli (APEC) for the First Time in Layer Farms in Bangladesh and Their Antibiotic Resistance Patterns. Microorganisms 2020, 8, 1021. [Google Scholar] [CrossRef] [PubMed]
  168. Saidenberg, A.B.S.; Stegger, M.; Price, L.B.; Johannesen, T.B.; Aziz, M.; Cunha, M.P.V.; Moreno, A.M.; Knöbl, T. mcr-Positive Escherichia coli ST131-H22 from Poultry in Brazil. Emerg. Infect. Dis. 2020, 26, 1951–1954. [Google Scholar] [CrossRef]
  169. Li, Y.; Chen, L.; Wu, X.; Huo, S. Molecular characterization of multidrug-resistant avian pathogenic Escherichia coli isolated from septicemic broilers. Poult. Sci. 2015, 94, 601–611. [Google Scholar] [CrossRef] [PubMed]
  170. Koutsianos, D.; Athanasiou, L.V.; Dimitriou, T.; Nikolaidis, M.; Tsadila, C.; Amoutzias, G.; Mossialos, D.; Koutoulis, K.C. Antibiotic Resistance Patterns and mcr-1 Detection in Avian Pathogenic Escherichia coli Isolates from Commercial Layer and Layer Breeder Flocks Demonstrating Colibacillosis in Greece. Microb. Drug Resist. 2020. [Google Scholar] [CrossRef]
  171. Bista, S.; Thapa Shrestha, U.; Dhungel, B.; Koirala, P.; Gompo, T.R.; Shrestha, N.; Adhikari, N.; Joshi, D.R.; Banjara, M.R.; Adhikari, B.; et al. Detection of Plasmid-Mediated Colistin Resistant mcr-1 Gene in Escherichia coli Isolated from Infected Chicken Livers in Nepal. Animals 2020, 10, 2060. [Google Scholar] [CrossRef]
  172. Temmerman, R.; Garmyn, A.; Antonissen, G.; Vanantwerpen, G.; Vanrobaeys, M.; Haesebrouck, F.; Devreese, M. Evaluation of Fluoroquinolone Resistance in Clinical Avian Pathogenic Escherichia coli Isolates from Flanders (Belgium). Antibiotics 2020, 9, 800. [Google Scholar] [CrossRef]
  173. Osman, K.M.; Kappell, A.D.; Elhadidy, M.; ElMougy, F.; El-Ghany, W.A.A.; Orabi, A.; Mubarak, A.S.; Dawoud, T.M.; Hemeg, H.A.; Moussa, I.M.I.; et al. Poultry hatcheries as potential reservoirs for antimicrobial-resistant Escherichia coli: A risk to public health and food safety. Sci. Rep. 2018, 8, 5859. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Tang, K.L.; Caffrey, N.P.; Nóbrega, D.B.; Cork, S.C.; Ronksley, P.E.; Barkema, H.W.; Polachek, A.J.; Ganshorn, H.; Sharma, N.; Kellner, J.D.; et al. Restricting the use of antibiotics in food-producing animals and its associations with antibiotic resistance in food-producing animals and human beings: A systematic review and meta-analysis. Lancet Planet. Health 2017, 1, e316–e327. [Google Scholar] [CrossRef]
  175. Redweik, G.A.J.; Stromberg, Z.R.; Van Goor, A.; Mellata, M. Protection against avian pathogenic Escherichia coli and Salmonella Kentucky exhibited in chickens given both probiotics and live Salmonella vaccine. Poult. Sci. 2020, 99, 752–762. [Google Scholar] [CrossRef] [PubMed]
  176. Wang, H.; Liang, K.; Kong, Q.; Liu, Q. Immunization with outer membrane vesicles of avian pathogenic Escherichia coli O78 induces protective immunity in chickens. Vet. Microbiol. 2019, 236, 108367. [Google Scholar] [CrossRef]
  177. Hoseini Shahidi, R.; Hashemi Tabar, G.; Bassami, M.R.; Jamshidi, A.; Dehghani, H. The design and application of a bacterial ghost vaccine to evaluate immune response and defense against avian pathogenic Escherichia coli O2:K1 serotype. Res. Vet. Sci. 2019, 125, 153–161. [Google Scholar] [CrossRef]
  178. Hu, J.; Zuo, J.; Chen, Z.; Fu, L.; Lv, X.; Hu, S.; Shi, X.; Jing, Y.; Wang, Y.; Wang, Z.; et al. Use of a modified bacterial ghost lysis system for the construction of an inactivated avian pathogenic Escherichia coli vaccine candidate. Vet. Microbiol. 2019, 229, 48–58. [Google Scholar] [CrossRef] [PubMed]
  179. Ebrahimi-Nik, H.; Bassami, M.R.; Mohri, M.; Rad, M.; Khan, M.I. Bacterial ghost of avian pathogenic E. coli (APEC) serotype O78:K80 as a homologous vaccine against avian colibacillosis. PLoS ONE 2018, 13, e0194888. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Han, Y.; Liu, Q.; Willias, S.; Liang, K.; Li, P.; Cheng, A.; Kong, Q. A bivalent vaccine derived from attenuated Salmonella expressing O-antigen polysaccharide provides protection against avian pathogenic Escherichia coli O1 and O2 infection. Vaccine 2018, 36, 1038–1046. [Google Scholar] [CrossRef]
  181. Van Goor, A.; Stromberg, Z.R.; Mellata, M. A recombinant multi-antigen vaccine with broad protection potential against avian pathogenic Escherichia coli. PLoS ONE 2017, 12, e0183929. [Google Scholar] [CrossRef] [Green Version]
  182. Lee, J.H.; Chaudhari, A.A.; Oh, I.G.; Eo, S.K.; Park, S.Y.; Jawale, C.V. Immune responses to oral vaccination with Salmonella-delivered avian pathogenic Escherichia coli antigens and protective efficacy against colibacillosis. Can. J. Vet. Res. 2015, 79, 229–234. [Google Scholar]
  183. Sadeyen, J.R.; Wu, Z.; Davies, H.; van Diemen, P.M.; Milicic, A.; La Ragione, R.M.; Kaiser, P.; Stevens, M.P.; Dziva, F. Immune responses associated with homologous protection conferred by commercial vaccines for control of avian pathogenic Escherichia coli in turkeys. Vet. Res. 2015, 46, 5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Holden, K.M.; Browning, G.F.; Noormohammadi, A.H.; Markham, P.; Marenda, M.S. Avian pathogenic Escherichia coli DeltatonB mutants are safe and protective live-attenuated vaccine candidates. Vet. Microbiol. 2014, 173, 289–298. [Google Scholar] [CrossRef]
  185. La Ragione, R.M.; Woodward, M.J.; Kumar, M.; Rodenberg, J.; Fan, H.; Wales, A.D.; Karaca, K. Efficacy of a live attenuated Escherichia coli O78:K80 vaccine in chickens and turkeys. Avian Dis. 2013, 57, 273–279. [Google Scholar] [CrossRef]
  186. Nagano, T.; Kitahara, R.; Nagai, S. An attenuated mutant of avian pathogenic Escherichia coli serovar O78: A possible live vaccine strain for prevention of avian colibacillosis. Microbiol. Immunol. 2012, 56, 605–612. [Google Scholar] [CrossRef]
  187. Yaguchi, K.; Ohgitani, T.; Noro, T.; Kaneshige, T.; Shimizu, Y. Vaccination of chickens with liposomal inactivated avian pathogenic Escherichia coli (APEC) vaccine by eye drop or coarse spray administration. Avian Dis. 2009, 53, 245–249. [Google Scholar] [CrossRef] [PubMed]
  188. Lynne, A.M.; Kariyawasam, S.; Wannemuehler, Y.; Johnson, T.J.; Johnson, S.J.; Sinha, A.S.; Lynne, D.K.; Moon, H.W.; Jordan, D.M.; Logue, C.M.; et al. Recombinant Iss as a potential vaccine for avian colibacillosis. Avian Dis. 2012, 56, 192–199. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Salehi, T.Z.; Tabatabaei, S.; Karimi, V.; Fasaei, B.N.; Derakhshandeh, A.; Jahromi, A.O.N. Assessment of immunity against avian colibacillosis induced by an aroA mutant containing increased serum survival gene in broilers. Braz. J. Microbiol. 2012, 43, 363–370. [Google Scholar] [CrossRef] [Green Version]
  190. Fernandes Filho, T.; Favaro, C., Jr.; Ingberman, M.; Beirao, B.C.; Inoue, A.; Gomes, L.; Caron, L.F. Effect of spray Escherichia coli vaccine on the immunity of poultry. Avian Dis. 2013, 57, 671–676. [Google Scholar] [CrossRef] [PubMed]
  191. Han, Y.; Liu, Q.; Yi, J.; Liang, K.; Wei, Y.; Kong, Q. A biologically conjugated polysaccharide vaccine delivered by attenuated Salmonella Typhimurium provides protection against challenge of avian pathogenic Escherichia coli O1 infection. Pathog. Dis. 2017, 75. [Google Scholar] [CrossRef]
  192. Vandemaele, F.; Bleyen, N.; Abuaboud, O.; vanderMeer, E.; Jacobs, A.; Goddeeris, B.M. Immunization with the biologically active lectin domain of PapGII induces strong adhesion-inhibiting antibody responses but not protection against avian pathogenic Escherichia coli. Avian Pathol. 2006, 35, 238–249. [Google Scholar] [CrossRef]
  193. Vandemaele, F.; Ververken, C.; Bleyen, N.; Geys, J.; D’Hulst, C.; Addwebi, T.; van Empel, P.; Goddeeris, B.M. Immunization with the binding domain of FimH, the adhesin of type 1 fimbriae, does not protect chickens against avian pathogenic Escherichia coli. Avian Pathol. 2005, 34, 264–272. [Google Scholar] [CrossRef]
  194. Roland, K.; Karaca, K.; Sizemore, D. Expression of Escherichia coli antigens in Salmonella typhimurium as a vaccine to prevent airsacculitis in chickens. Avian Dis. 2004, 48, 595–605. [Google Scholar] [CrossRef]
  195. Lynne, A.M.; Foley, S.L.; Nolan, L.K. Immune response to recombinant Escherichia coli Iss protein in poultry. Avian Dis. 2006, 50, 273–276. [Google Scholar] [CrossRef] [Green Version]
  196. Kariyawasam, S.; Wilkie, B.N.; Gyles, C.L. Construction, characterization, and evaluation of the vaccine potential of three genetically defined mutants of avian pathogenic Escherichia coli. Avian Dis. 2004, 48, 287–299. [Google Scholar] [CrossRef] [PubMed]
  197. Kariyawasam, S.; Wilkie, B.N.; Hunter, D.B.; Gyles, C.L. Systemic and mucosal antibody responses to selected cell surface antigens of avian pathogenic Escherichia coli in experimentally infected chickens. Avian Dis. 2002, 46, 668–678. [Google Scholar] [CrossRef]
  198. Tuntufye, H.N.; Ons, E.; Pham, A.D.; Luyten, T.; Van Gerven, N.; Bleyen, N.; Goddeeris, B.M. Escherichia coli ghosts or live E. coli expressing the ferri-siderophore receptors FepA, FhuE, IroN and IutA do not protect broiler chickens against avian pathogenic E. coli (APEC). Vet. Microbiol. 2012, 159, 470–478. [Google Scholar] [CrossRef]
  199. Ma, S.T.; Ding, G.J.; Huang, X.W.; Wang, Z.W.; Wang, L.; Yu, M.L.; Shi, W.; Jiang, Y.P.; Tang, L.J.; Xu, Y.G.; et al. Immunogenicity in chickens with orally administered recombinant chicken-borne Lactobacillus saerimneri expressing FimA and OmpC antigen of O78 avian pathogenic Escherichia coli. J. Med. Microbiol. 2018, 67, 441–451. [Google Scholar] [CrossRef] [PubMed]
  200. Amoako, K.K.; Prysliak, T.; Potter, A.A.; Collinson, S.K.; Kay, W.W.; Allan, B.J. Attenuation of an avian pathogenic Escherichia coli strain due to a mutation in the rpsL gene. Avian Dis. 2004, 48, 19–25. [Google Scholar] [CrossRef]
  201. Hu, R.; Li, J.; Zhao, Y.; Lin, H.; Liang, L.; Wang, M.; Liu, H.; Min, Y.; Gao, Y.; Yang, M. Exploiting bacterial outer membrane vesicles as a cross-protective vaccine candidate against avian pathogenic Escherichia coli (APEC). Microb. Cell Fact. 2020, 19, 119. [Google Scholar] [CrossRef]
  202. Koutsianos, D.; Gantelet, H.; Franzo, G.; Lecoupeur, M.; Thibault, E.; Cecchinato, M.; Koutoulis, K.C. An Assessment of the Level of Protection Against Colibacillosis Conferred by Several Autogenous and/or Commercial Vaccination Programs in Conventional Pullets upon Experimental Challenge. Vet. Sci. 2020, 7, 80. [Google Scholar] [CrossRef]
  203. Hu, R.; Liu, H.; Wang, M.; Li, J.; Lin, H.; Liang, M.; Gao, Y.; Yang, M. An OMV-Based Nanovaccine Confers Safety and Protection against Pathogenic Escherichia coli via Both Humoral and Predominantly Th1 Immune Responses in Poultry. Nanomaterials 2020, 10, 2293. [Google Scholar] [CrossRef]
  204. Soleymani, S.; Tavassoli, A.; Hashemi Tabar, G.; Kalidari, G.A.; Dehghani, H. Design, development, and evaluation of the efficacy of a nucleic acid-free version of a bacterial ghost candidate vaccine against avian pathogenic E. coli (APEC) O78:K80 serotype. Vet. Res. 2020, 51, 144. [Google Scholar] [CrossRef] [PubMed]
  205. Mohammed, G.M.; ElZorkany, H.E.; Farroh, K.Y.; Abd El-Aziz, W.R.; Elshoky, H.A. Potential improvement of the immune response of chickens against E. coli vaccine by using two forms of chitosan nanoparticles. Int. J. Biol Macromol. 2020, 167, 395–404. [Google Scholar] [CrossRef]
  206. Cox, G.J.; Griffith, B.; Reed, M.; Sandstrom, J.D.; Peterson, M.P.; Emery, D.; Straub, D.E. A vaccine to prevent egg layer peritonitis in chickens. Avian Dis. 2020. [Google Scholar] [CrossRef]
  207. Huttner, A.; Gambillara, V. The development and early clinical testing of the ExPEC4V conjugate vaccine against uropathogenic Escherichia coli. Clin. Microbiol. Infect. 2018, 24, 1046–1050. [Google Scholar] [CrossRef] [PubMed]
  208. Ding, S.; Wang, Y.; Yan, W.; Li, A.; Jiang, H.; Fang, J. Effects of Lactobacillus plantarum 15-1 and fructooligosaccharides on the response of broilers to pathogenic Escherichia coli O78 challenge. PLoS ONE 2019, 14, e0212079. [Google Scholar] [CrossRef] [Green Version]
  209. Tarabees, R.; Gafar, K.M.; El-Sayed, M.S.; Shehata, A.A.; Ahmed, M. Effects of Dietary Supplementation of Probiotic Mix and Prebiotic on Growth Performance, Cecal Microbiota Composition, and Protection Against Escherichia coli O78 in Broiler Chickens. Probiot. Antimicrob. Proteins 2019, 11, 981–989. [Google Scholar] [CrossRef]
  210. Tarabees, R.; El-Sayed, M.S.; Shehata, A.A.; Diab, M.S. Effects of the Probiotic Candidate E. faecalis-1, the Poulvac E. coli Vaccine, and their Combination on Growth Performance, Caecal Microbial Composition, Immune Response, and Protection against E. coli O78 Challenge in Broiler Chickens. Probiot. Antimicrob. Proteins 2020, 12, 860–872. [Google Scholar] [CrossRef] [PubMed]
  211. Chang, C.-J.; Lin, T.-L.; Tsai, Y.-L.; Wu, T.-R.; Lai, W.-F.; Lu, C.-C.; Lai, H.-C. Next generation probiotics in disease amelioration. J. Food Drug Anal. 2019, 27, 615–622. [Google Scholar] [CrossRef]
  212. Naghizadeh, M.; Karimi Torshizi, M.A.; Rahimi, S.; Dalgaard, T.S. Synergistic effect of phage therapy using a cocktail rather than a single phage in the control of severe colibacillosis in quails. Poult. Sci. 2019, 98, 653–663. [Google Scholar] [CrossRef] [PubMed]
  213. Oliveira, A.; Sereno, R.; Azeredo, J. In vivo efficiency evaluation of a phage cocktail in controlling severe colibacillosis in confined conditions and experimental poultry houses. Vet. Microbiol. 2010, 146, 303–308. [Google Scholar] [CrossRef] [Green Version]
  214. Huff, W.E.; Huff, G.R.; Rath, N.C.; Balog, J.M.; Donoghue, A.M. Alternatives to antibiotics: Utilization of bacteriophage to treat colibacillosis and prevent foodborne pathogens. Poult. Sci. 2005, 84, 655–659. [Google Scholar] [CrossRef] [PubMed]
  215. Żbikowska, K.; Michalczuk, M.; Dolka, B. The Use of Bacteriophages in the Poultry Industry. Animals 2020, 10, 872. [Google Scholar] [CrossRef] [PubMed]
  216. Liu, J.; Wang, X.; Shi, W.; Qian, Z.; Wang, Y. Sensitization of avian pathogenic Escherichia coli to amoxicillin in vitro and in vivo in the presence of surfactin. PLoS ONE 2019, 14, e0222413. [Google Scholar] [CrossRef] [PubMed]
  217. Goonewardene, K.; Ahmed, K.A.; Gunawardana, T.; Popowich, S.; Kurukulasuriya, S.; Karunarathna, R.; Gupta, A.; Ayalew, L.E.; Lockerbie, B.; Foldvari, M.; et al. Mucosal delivery of CpG-ODN mimicking bacterial DNA via the intrapulmonary route induces systemic antimicrobial immune responses in neonatal chicks. Sci. Rep. 2020, 10, 5343. [Google Scholar] [CrossRef]
  218. Peng, L.Y.; Yuan, M.; Song, K.; Yu, J.L.; Li, J.H.; Huang, J.N.; Yi, P.F.; Fu, B.D.; Shen, H.Q. Baicalin alleviated APEC-induced acute lung injury in chicken by inhibiting NF-kappaB pathway activation. Int. Immunopharmacol. 2019, 72, 467–472. [Google Scholar] [CrossRef]
  219. Cuperus, T.; van Dijk, A.; Matthijs, M.G.; Veldhuizen, E.J.; Haagsman, H.P. Protective effect of in ovo treatment with the chicken cathelicidin analog D-CATH-2 against avian pathogenic E. coli. Sci. Rep. 2016, 6, 26622. [Google Scholar] [CrossRef]
  220. Guo, X.; Zhang, L.Y.; Wu, S.C.; Xia, F.; Fu, Y.X.; Wu, Y.L.; Leng, C.Q.; Yi, P.F.; Shen, H.Q.; Wei, X.B.; et al. Andrographolide interferes quorum sensing to reduce cell damage caused by avian pathogenic Escherichia coli. Vet. Microbiol. 2014, 174, 496–503. [Google Scholar] [CrossRef]
  221. Kathayat, D.; Helmy, Y.A.; Deblais, L.; Rajashekara, G. Novel small molecules affecting cell membrane as potential therapeutics for avian pathogenic Escherichia coli. Sci. Rep. 2018, 8, 15329. [Google Scholar] [CrossRef] [PubMed]
  222. Helmy, Y.A.; Deblais, L.; Kassem, I.I.; Kathayat, D.; Rajashekara, G. Novel small molecule modulators of quorum sensing in avian pathogenic Escherichia coli (APEC). Virulence 2018, 9, 1640–1657. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Totsika, M. Disarming pathogens: Benefits and challenges of antimicrobials that target bacterial virulence instead of growth and viability. Future Med. Chem. 2017, 9, 267–269. [Google Scholar] [CrossRef] [Green Version]
  224. Belete, T.M. Novel targets to develop new antibacterial agents and novel alternatives to antibacterial agents. Hum. Microbiome J. 2019, 11, 100052. [Google Scholar] [CrossRef]
  225. Kathayat, D.; Antony, L.; Deblais, L.; Helmy, Y.A.; Scaria, J.; Rajashekara, G. Small Molecule Adjuvants Potentiate Colistin Activity and Attenuate Resistance Development in Escherichia coli by Affecting pmrAB System. Infect. Drug Resist. 2020, 13, 2205–2222. [Google Scholar] [CrossRef] [PubMed]
  226. Mahlapuu, M.; Håkansson, J.; Ringstad, L.; Björn, C. Antimicrobial Peptides: An Emerging Category of Therapeutic Agents. Front. Cell. Infect. Microbiol. 2016, 6, 194. [Google Scholar] [CrossRef] [Green Version]
  227. Daneshmand, A.; Kermanshahi, H.; Sekhavati, M.H.; Javadmanesh, A.; Ahmadian, M. Antimicrobial peptide, cLF36, affects performance and intestinal morphology, microflora, junctional proteins, and immune cells in broilers challenged with E. coli. Sci. Rep. 2019, 9, 14176. [Google Scholar] [CrossRef]
Scheme 1. Schematic diagram showing overview of Avian pathogenic Escherichia coli (APEC) infection in chickens along with infection control checkpoints. After entry through oral, nasal, or cloacal routes, APEC colonizes the mucosal sites of gastrointestinal, respiratory, and reproductive tracts without causing disease in chickens. In the presence of concurrent viral or mycoplasma infections or under immunosuppressive or stressed conditions, APEC invades the mucosal layers and reach extra-intestinal organs (heart, liver, lung, spleen, kidney, reproductive organs, etc.) resulting in multi-systemic infections, which are commonly referred to as colibacillosis. Colibacillosis leads to high morbidity and mortality, production losses, and condemnation of carcasses as well as foodborne transmission risk to humans. Colibacillosis can be prevented by the management of stressors, biosecurity measures, and vaccination against APEC or associated viral infections. Chicken breeds with high intrinsic resistance to APEC can be developed through genetic technologies. Vertical transmission of APEC from breeders through contaminated eggs should be monitored to prevent APEC entry into chicken flocks. Antibiotics are commonly used to treat chicken flocks affected with colibacillosis.
Scheme 1. Schematic diagram showing overview of Avian pathogenic Escherichia coli (APEC) infection in chickens along with infection control checkpoints. After entry through oral, nasal, or cloacal routes, APEC colonizes the mucosal sites of gastrointestinal, respiratory, and reproductive tracts without causing disease in chickens. In the presence of concurrent viral or mycoplasma infections or under immunosuppressive or stressed conditions, APEC invades the mucosal layers and reach extra-intestinal organs (heart, liver, lung, spleen, kidney, reproductive organs, etc.) resulting in multi-systemic infections, which are commonly referred to as colibacillosis. Colibacillosis leads to high morbidity and mortality, production losses, and condemnation of carcasses as well as foodborne transmission risk to humans. Colibacillosis can be prevented by the management of stressors, biosecurity measures, and vaccination against APEC or associated viral infections. Chicken breeds with high intrinsic resistance to APEC can be developed through genetic technologies. Vertical transmission of APEC from breeders through contaminated eggs should be monitored to prevent APEC entry into chicken flocks. Antibiotics are commonly used to treat chicken flocks affected with colibacillosis.
Pathogens 10 00467 sch001
Table 1. APEC virulence and pathogenesis factors and their role in systemic infections.
Table 1. APEC virulence and pathogenesis factors and their role in systemic infections.
Virulence FactorsGenes/Proteins InvolvedRole in Pathogenesis/InfectionReference
AdhesinsfimH, fimC, papA, papC, papEF, papG I, papG II, papGIII, felA, sfa/sfaS, afaIBC, focGE, lpfA, lpf0141, lpf0154, flgE, crl, csg, bmaE, tsh, mat/ecpA, hra/hrlA/hek, iha, yqiG, kiiAdhesion, colonization, biofilm formation, motility, intracellular survival[32,33,34,35,36,37,38,39,41]
yfc O Adhesion, colonization, resistance to environmental stresses [42]
yad C Adhesion, intracellular survival, motility[43]
aat A, aatB, upaB Adhesion, colonization, biofilm formation[44,45]
fdtA, rluD, yjhB, ecpR, fdeCAdhesion[46]
InvasinsibeA, ibeB, tia, gimB Invasion, resistance to oxidative stress, colonization, proliferation, biofilm formation [35,47,48]
IbeR Invasion, resistance to serum and environmental stresses, expression of virulence genes[49]
ych O Motility, adhesion, invasion, biofilm formation, expression of membrane proteins and metabolism genes[50]
Iron acquisition systemsiutA, iucC, iucD, aerJ, iucA, iucB, iroBCDEN, fyuA, sitABCD, mntH, feoB, irp2, ireA, eitABCD, fepC, chuA, bfrIron and manganese uptake from the host, adhesion, invasion, colonization, persistence, expression of virulence genes, resistance to environmental stresses [36,37,51,52,53,54,55,56,57]
entE, entS, tolC Invasion, colonization, persistence [60]
Protectinsiss, traT, ompT, kpsMT(K1), kpsMT(II), kpsMT(III), neuC, neuS, neuD, kfiC-K5, betAProtect from serum bactericidal activity and phagocytosis, adhesion, invasion, intracellular survival, colonization, proliferation[36,41,51,53,61]
YbjX, PagPResistance to serum and environmental stresses, invasion, intracellular survival[62,63]
OmpAIntracellular survival[64]
wzy Adhesion, invasion, intracellular survival, colonization [65]
waa L Motility, resistance to phagocytosis and environmental stresses, adhesion, invasion, biofilm formation[65]
sod A Protect against ROS-mediated host defenses, biofilm formation [68]
lpx M Invasion, intracellular survival, colonization, regulation of expression of cytokine genes and nitric oxide production [67]
ToxinshlyF, hlyA, hlyE, cdtB, cdtS, vat, sat, stx2f, astA, pic, EAST-1, espC, ace4/35Cell lysis and damage, induce host cell vacuolization, colonization, motility, biofilm formation, agglutination, formation of outer membrane vesicles[35,38,39,41,51,53,69,70,71,72,73]
Other virulence and pathogenesis factors
Quorum-sensing system (AI-2)LuxS, LsrABCDFGK, ptsI, Pfs Motility, biofilm formation, adherence, invasion, colonization, intracellular survival, persistence, expression of virulence genes, cell damage [10,74,75,76,77]
Secretion systemsDotU, CpxRA, IcmF, Hcp, ClpV, VrgG (Type VI)Interbacterial competition, adhesion, invasion, intracellular survival, colonization, motility, biofilm formation, production of type 1 fimbriae, resistance to serum bactericidal activity, modulation of intracellular host responses (IL-18, IL-1β)[11,80,81,82,83,84,85]
EtrA, YqeI, EivC (Type III) Motility, intracellular survival, resistance to phagocytosis and serum bactericidal activity, proliferation, expression of fimbriae genes, downregulation of pro-inflammatory cytokines [12,78,79]
Two-component systemsPhoPQ, tolCBiofilm formation, motility, adhesion, invasion, intracellular survival, systemic infection, expression of virulence genes and genes associated with flagellar assembly, ABC transporters, quorum sensing, and bacterial chemotaxis [13,86,90]
BasSRBiofilm formation, APEC virulence and colonization in vivo[91]
KdpDEExpression of flagella-related genes, flagellum formation, motility and resistance to serum bactericidal activity[87]
RstAB, hdeDIron acquisition, acid resistance, intracellular survival, colonization[88,92]
BarA-UvrY Adhesion, invasion, persistence, intracellular survival, resistance to serum bactericidal activity and oxidative stress, regulation of exopolysaccharide production and type 1 and P fimbriae [89]
Transcriptional regulators AutA/AutR Expression of K1 capsule and acid resistance systems, adaptive lifestyle change[14]
FNR Adhesion, invasion, expression of type 1 fimbriae and type VII secretion system, resistance to oxidative stress[15]
YjjQ Flagellar motility[93]
McbR Biofilm formation, response to H2O2[94]
tyrRInvasion, motility, intracellular survival[37]
RfaH Invasion, intracellular survival, resistance to serum bactericidal activity[95]
Metabolism-associated genes acs -yjcH-actP Intracellular survival, proliferation, colonization, production of pro-inflammatory cytokines and nitric oxide[96]
PotE, PotF Colonization, adhesion[16]
NirC Adhesion, colonization [17]
ArcA Chemotaxis, motility [18]
Miscellaneous OmpF, OmpC Adhesion, invasion, colonization, proliferation[100]
Prophage phiv142-3 (orf20) and phiv205-1Resistance to serum and environmental stresses, adhesion, invasion, intracellular survival, colonization, biofilm formation, formation of flagella and I fimbriae [97,98,99]
YicS Motility, biofilm formation, invasion[101]
cpd B Colonization [102]
pst B Resistance to serum bactericidal activity and oxidative stress, colonization[103]
tmRNA-SmpB Colonization, persistence, replication, intracellular survival[104]
mli C Resistance to serum bactericidal activity[105]
malX, frz, cvaABC, cvi, cba, cib/cibI, cbi, cma, eaeA, sopB, yfcV, gad, mchBCF, mcmA, bor, air, eilA, celB, pabB, capU, cif, tir, tccp, nleB, iaL, cjrC, mig-14pUnknown/not clearly known functions[34,39,40,41,53,54,57,106,107,108]
Genes essential for systemic infections and adaptation metH, lysA, pntA, purL, serS, ybjE, ycdK (rutC), wcaJ, gspL, sdsR, irp2, eitD, ylbE, yjiY, tkt1, pilN, pilQ, tsh, hpb, TcfD, Z5222, waaO, waaY, iutA, iucA, iucD, iroC, ColE2, traK, traG, traT, SopA, psiA, hkaG, hkbV, hkbQ, Z3370, Int, CC0532, TM0427, YPO3000, rhsH, RSp0733, bioABFCD, rnfA, rfnE, gene encoding endonuclease III, creABCD, yehD, potF, flgE, tyrR, bfrSystemic APEC infections and adaptation [37,109,110,111]
Table 2. Antibiotic resistance reported in APEC isolates worldwide from 2015 to 2020.
Table 2. Antibiotic resistance reported in APEC isolates worldwide from 2015 to 2020.
Study LocationStudy SamplesResistance ReportedResistance Genes ReportedReference
BelgiumBroilersEFXgyrA, parC, parE, PMQR[172]
NepalChickensCLT, TTC, CFX, CXT, IPMmcr-1[171]
GreeceLayersCLTmcr-1[170]
BangladeshChickensAPC, DXC, TTC, NFN, CFX, NDA, CXT, IPM, GNC, CMC, SFN, AZT, PMB [108]
BrazilChickensCLTmcr-5, mcr-9[168]
BangladeshLayersAPC, TTC, CMC, EHC, EFX, CFX, SPM, CLT, GNC [167]
Thailand and AustraliaBroilers and broiler breedersAXC, CFU, CMC, EFX, FFC, GNC, NDA, TMP-SFM, TTCgyrA, parC[51]
TunisiaChicken’s fresh carcassesNDA, FMQ, EFX, DXC, TTC, FFC, TMP-SFM, SPM, AXC, AXC-CVA; CFT, CFZ, ATN, CFP, CLTmcr-1, blaCTX-M-1, blaTEM,blaSHV, tetA, dfraI, floR, cmlA, aadA, strA, strB, sul1[146]
JapanMunicipal wastewater
Influents
CLTmcr-1[107]
EgyptBroilersAPC, AXC-CVA, TTC, CLT, DXC, SMC, FFC, CFT, CFX [32]
KoreaBroilers APC, NDA, TTC, CPT, CPXblaCTX-M-1, blaCTX-M-15,blaTEM-1,aac(3)-II, qnrA, qnrS[147]
Germany BroilerCLTmcr-1[164]
AlgeriaBroilersNDA, AXC, APC, TCC, PPA, TMP-SFM [148]
ChinaChickens APC, CFT, CMC, GNC, KMC, SPM, TMP-SFM, NDA, TTC [53]
TaiwanDay-old
hatchery chicks
APC, AXC, CPX, FFC, TMP-SFM [70]
JordanBroilersTMP-SFM, FFC, AXC, DXC, SCCint1, sul1, sul2, tetA, blaTEM[71]
ItalyBroilers/turkey
/goose/guinea fowl/duck/
pigeon/layers
/capon
APC, NDA, AXC-CVA, CFT, CFX, CFN, CFZ, CMC, EFX, GNC, KMC, NDA, SPM, TTC, FMQ, TMP-SFM [72]
FranceBroilers/ducks
/turkeys/layers
CFU [165]
EthiopiaBackyard
chickens
APC, CFT, CXC, CRX, TTC [149]
PakistanBroilersAPC, TTC, CFX [34]
IndonesiaChickensTTC, OTC, CFX, CFX, NFX tetA, tetB, qnrA, qnrB, qnrS[150]
CanadaBroilers
/Breeders
TTC, APC, SCC, GNC, CFU, KMC, APC, TMP-SFM [54]
EgyptBroilersOTC, KMC, APC, CDC, SPM, EFX, CMC, CFT, GNC, EHC, OXC, TMP-SFMCITM, ere, aac(3)-(IV), tetA, tetB, dfr(A1), aad (A1)[151]
ChinaLayerCRX, TMC, MMC, CFN, SPM, APCstrA, strB, blaCMY-2, blaCTX-M-19, blaTEM-1B, fosA, mphA, floR, sul2, tetA, tetB[152]
SenegalChickens TTC, SSX, APC, TMP-SFM, SPM, CFU, CTX, CFXgyrB, parC, blaCTX-M,blaOXA-1, blaTEM,blaSHV, tetA, dhrfVII[153]
AlgeriaBroilersTTC, TMP-SFM, EFX, NDA, APC, DXC, CMC, AXC-CVA, FMQ [166]
BrazilTurkeysEHC, APC, TTC, OTC, LCC, SCC, TMP-SFMblaCTX-M-2,blaCTX-M-8, blaCTX-M-8/25, blaTEM,blaSHV, blaCMY-2[154]
NepalBroilersAPC, CMX, DXC, CLT [155]
EgyptBroilersPNC, CFP, NFX, CMCsul1, blaTEM,blaSHV,tetA, aadA[157]
PakistanBroilersCLTmcr-1[156]
ChinaChickensTTC, NDA, TMP-SFM, APC, AKC, ATN, CFZ, CFT, CMC, CFX [158]
JapanBroilersAPC, CFT, OTC, SPM, NDA, EFX, TMP, KMC, GNC, CMCblaCTX-M-2,blaCTX-M-14, blaCTX-M-65, blaTEM-1,blaSHV-12, blaCMY-2, aacA4, aadA2, aadA5, catB3, dfrA1, dfrA17, parC, aac(6′)-Ib-cr, oqxAB[159]
China/
Egypt
Chickens CLTmcr-1[160]
BrazilChickens APC, TTC, GNC, NMC, EFX, NFX, TMP-SFM [35]
South AfricaBroilersCLTmcr-1[161]
SpainBroilersCFT, CFZ, XCT, NDA, CFX, TTC, KMC, SPM, TMP-SFM, GNC, APC, FFCblaCTX-M-14, blaSHV-2, blaCMY-2, qnrA, aac(6′)-Ib-cr[162]
ChinaChickensTTC, NDA, APC, SFZ, SPM, TMP, CMX [73]
ItalyBroilers/layers
/turkeys
SPM, TMPaadA
dfrA
[163]
ChinaBroilersAPC, KMC, CFX, LFX, SPM, GNC, OFX, NFX, CTXblaCTX-M-15, blaOXA-30, blaTEM, blaSHV-2, blaCMY-2, qnrA, qnrB, qnrS, strA, strB, aph(3′)-II1, aac(3)-IIa, aac(6′)-Ib, ant(3″)-Ia[169]
NDA: nalidixic acid; FMQ: flumequine; EFX: enrofloxacin; CFX: ciprofloxacin; NFX: norfloxacin; LFX: levofloxacin; OFX: ofloxacin; DXC: doxycycline, TTC: tetracycline; OTC: oxytetracycline; FFC: florfenicol; TMP-SFM: trimethoprim-sulfamethoxazole; SSX: sulfisoxazole; SFZ: sulphafurazole; SFN: sulfonamide, CMX: co-trimoxazole; SPM: streptomycin; SMC: spiramycin; GNC: gentamicin; AKC: amikacin; KMC: kanamycin; SCC: spectinomycin; LCC: lincomycin; APC: apramycin; CDC: clindamycin; EHC: erythromycin; TMC: tobramycin; MMC: medemycin; NMC: neomycin; AZT: azithromycin; PNC: penicillin; APC: ampicillin; AXC: amoxicillin; TTC: ticarcillin; CXC: cloxacillin; OXC: oxacillin; AXC-CVA: amoxicillin-clavunilic acid; CFT: cefotaxime; CFZ: ceftazidime; CPT: cephatholin; CPX: cephalexin; CXT: cefoxitin; CFN: cefazolin; CRX: cefuroxime; CFU: ceftiofur; CTX: ceftriaxone; ATN: aztreonam; CFP: cefepime; CLT: colistin; PPA: pipedimic acid; CMC: chloramphenicol; NFN: nitrofurantoin; PMB: polymyxin B; IPM: imipenem.
Table 3. Various vaccine candidates tested against APEC infections in chickens.
Table 3. Various vaccine candidates tested against APEC infections in chickens.
Vaccine DescriptionChallengeRoute Main Findings Reference
Live-attenuated APEC O78 ΔaroA (Poulvac® E. coli) and autogenous (O18, O78, and O111) vaccine (alone and combination)Layers (O78, O18, and O111)Spray, I/M-Combination of live- attenuated and autogenous vaccine significantly reduce the mortality, lesion scores, and APEC load
-Live-attenuated vaccine alone does not provide significant protection
[202]
Live-attenuated Salmonella-delivered vaccine secreting APEC antigens (PapA, PapG, IutA, ClpG)Layers (JOL718)Oral-High IgA, IFN- γ, IL-2, and IL-6 levels
-Complete protection against JOL718 challenge
-Less lesions in liver
[182]
Live-attenuated APEC O78 ΔaroA vaccine (Poulvac® E. coli) and formalin-inactivated APEC O78 bacterinTurkeys (EC34195nal®, O78:K80)Spray and S/C, respectively-2 log10 CFU/g reduction in APEC load in spleen, liver, and kidney
-No lesions
-Induce transcription of IL-13 and TGF-β4
-Repress transcription of IFN- γ
-Modulate CXCLi2 expression
-Elevate IgA and IgY levels
-Higher ability of splenocytes to proliferate
[183]
Live-attenuated APEC O78 ΔaroA vaccine (Poulvac® E. coli)Broilers Spray-High CD4+TCRVβ1+ cells
-High naïve and memory CD8 cells
-Stimulate high state of immunocompetence
-Stimulate production of IgA
[190]
Live-attenuated ΔtonB and ΔtonB/Δfur vaccine (E956)Layers (E956)Spray -Less lesions
In airsacs, heart, and liver
[184]
Siderophore receptor and porin (SRP®) APEC vaccine for egg layer peritonitisLayers (O1, O2, and O78)I/V, I/VAG, I/T, I/P-Complete protection against mortality
- Substantially reduce APEC colonization and lesions
[206]
Live-attenuated O78:K80 ΔaroA vaccine Chickens and turkeys (O78 and X)Spray or spray followed by oral booster -Significant reduction in mortality
-homologous and heterologous protection
[185]
Live-attenuated O78:K80 ΔaroA vaccineBroilers (χ1378 and 02:K12)Aerosol-High weight gain
-No antibody responses
-No homologous and heterologous protection
[189]
Live-attenuated APEC O78:J29—Δcrp vaccine (AESN1331)Layers (O78:J46)Spray, eye drop, and in ovo -Significant reduction in mortality
-Less lesion scores in heart and liver
[186]
Live-attenuated S. typhimurium vaccine expressing O78 LPS and E. coli type 1 fimbriae Layers (χ7122, χ7252, and χ7096)Oral and spray-Lower lesion scores in airsacs
-No heterologous protection
[194]
ΔgalE, ΔpurA, and ΔaroA (APECO78:EC99) vaccineBroilers (EC99 and EC317)Spray-High IgY level
-Less lesion scores and APEC load
-No heterologous protection
[196]
Live E. coli vaccine with rspL mutation (EC844)Broilers (EC317)Aerosol and oral-Lower APEC lesions with three doses[200]
Recombinant Salmonella vaccine strain S740 (pSS28) containing APEC O1 and APEC O2 O-antigensLayers (O1 C24-2, O2 CE37)Oral followed by I/M booster-High IgG and IgA levels
-73.33% and 66.67% protection against O2 CE37 and O1 C24-2 challenge, respectively
-High opsonophagocytosis and serum bactericidal activity
-Less lesions in liver and spleen
[180]
Recombinant antigen (rAg) vaccine containing ExPEC proteins (OmpA, OmpT, TraT, EtsC)Layers (O2)S/C-Higher IgY, IL-1β, IL-6, IL-18, IFN-γ, IL-4, IFN-β, and IL-8 levels
-Increased serum bactericidal activity against multiple APEC strains
-Less lesions in airsacs, liver, and heart
-Less APEC load in heart and liver
[181]
Recombinant attenuated S. Typhimurium
vaccine (RASV) χ9373
Layers χ7122 (O78:K80)Oral-High IgY level
-Lower signs of airsacculitis
-Lower lesion scores in heart and liver
-Lower APEC load in blood
-Increase blood and serum bactericidal activity against multiple APEC strains
[175]
Recombinant Salmonella vaccine strain S740 containing APEC O1 O-antigen (pSS27)Layers (O1:C24-2)Oral followed by and I/M booster-High IgG and IgA levels
->50% protection against O1:C24-2 challenge
[191]
Recombinant iss vaccineBroilers (O1, O2, and χ7122) I/M -High IgA and IgG levels
-Less lesions in heart, liver, and airsacs
-Heterologous protection
[188]
Recombinant iss vaccineLayers (O2 and O78)S/C-High humoral response
-Lower APEC lesions
-Heterologous protection
[195]
E. coli BL21:D3 expressing FepA, FhuE, IroN, and IutA antigensBroilers (CH2)I/N-High IgG level
-No protection against CH2 challenge
[198]
Lactobacillus saerimineri expressing O78 FimC and OmpC antigensLayers (O78)Oral-High IgG and IgA levels
-Lower mortality
[199]
APEC O78, O1, and O2 outer membrane vesicles (OMVs) Broilers (O78, O1, and O2)I/M-High IgG level
-High IFN-γ, IL-17, and IL-10 levels
->90% protective efficacy
-Reduce bacterial burden in liver and lung
-No effect on growth performance
[201]
APEC O78 OMVs Layers (O78)I/M-High IgG level
-High opsonophagocytosis and serum bactericidal activity
-Complete protection against O78 challenge
[176]
APEC O2 OMVsBroilers (O2)I/M-Protective against homologous infection
-Enhance specific (IgY) antibody response
-Elicit IFN-γ mediated Th1 responses
[203]
Bacterial ghost vaccine (O2:K1) constructed using PhiX 174 lysis gene EBroilers (O2:K1)S/C, spray-Less lesion scores in airsacs, liver, and heart
-High IgG and sIgA levels
-High IFN-γ
[177]
Bacterial ghost vaccine (O2: DE17ΔluxaroA) constructed using PhiX 174 lysis gene ELayers (O2: DE17)S/C-Complete protection against O2:DE17 challenge
-Less protection against O2: CE35 challenge
-No lesions in liver, heart, spleen, and lung
-Higher IgG, IFN-γ, and TNF-α levels
[178]
Bacterial ghost vaccine (O78:K80) constructed using PhiX 174 lysis gene EBroilers (O78:K80)S/C, spray-Reduce lesions in airsacs, liver, and heart
-High IgY, IgA, and IFN-γ levels
[179]
Bacterial ghost vaccine (O78:K80) containing E-lysis and S nuclease genesBroilers (O78:K80)S/C, spray-Lower lesion scores
-High antibody (IgY and IgA) and IFN-γ levels
-Increase proinflammatory (IL-6, IL-1 β, and TNFSF15) cytokines production
[204]
Chitosan nanoparticles containing APEC O1 and O78 outer membrane proteins (OMPs) and flagellar antigensBroilers (O1 and O78)S/C-High antibody titer
->80% protection against O1 and O78 challenges
[205]
PapGII196 vaccineBroilers (CH2)I/M-High IgG level
-No protection against CH2 challenge
[192]
FimH (FimH156) vaccine Broilers (CH2)I/M and I/N-High IgG and IgA level
-No protection against CH2 challenge
[193]
APEC O78:EC99 cell surface antigens (FimA, PapG, IutA, and LPS)Broilers (EC99)I/N-High IgA, IgG, and IgM levels
-Less mortality
[197]
Liposomal inactivated APEC O78:KAI-2 vaccineLayers (APEC O78:PDI386)Eye drop or spray-High IgG and IgA levels
-Less mortality, APEC load, and lesions
[187]
S/C: subcutaneous; I/M: intramuscular; I/P: intraperitoneal; I/N: intranasal; I/V: intravenous; I/VAG: intravaginal; I/T: intratracheal.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kathayat, D.; Lokesh, D.; Ranjit, S.; Rajashekara, G. Avian Pathogenic Escherichia coli (APEC): An Overview of Virulence and Pathogenesis Factors, Zoonotic Potential, and Control Strategies. Pathogens 2021, 10, 467. https://doi.org/10.3390/pathogens10040467

AMA Style

Kathayat D, Lokesh D, Ranjit S, Rajashekara G. Avian Pathogenic Escherichia coli (APEC): An Overview of Virulence and Pathogenesis Factors, Zoonotic Potential, and Control Strategies. Pathogens. 2021; 10(4):467. https://doi.org/10.3390/pathogens10040467

Chicago/Turabian Style

Kathayat, Dipak, Dhanashree Lokesh, Sochina Ranjit, and Gireesh Rajashekara. 2021. "Avian Pathogenic Escherichia coli (APEC): An Overview of Virulence and Pathogenesis Factors, Zoonotic Potential, and Control Strategies" Pathogens 10, no. 4: 467. https://doi.org/10.3390/pathogens10040467

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop