Abstract
Gram-negative bacteria assemble a variety of surface structures, including the hair-like organelles known as pili or fimbriae. Pili typically function in adhesion and mediate interactions with various surfaces, with other bacteria, and with other types of cells such as host cells. The chaperone/usher (CU) pathway assembles a widespread class of adhesive and virulence-associated pili. Pilus biogenesis by the CU pathway requires a dedicated periplasmic chaperone and integral outer membrane protein termed the usher, which forms a multifunctional assembly and secretion platform. This review addresses the molecular and biochemical aspects of the CU pathway in detail, focusing on the type 1 and P pili expressed by uropathogenic Escherichia coli as model systems. We provide an overview of representative CU pili expressed by E. coli and Salmonella, and conclude with a discussion of potential approaches to develop antivirulence therapeutics that interfere with pilus assembly or function.
OVERVIEW AND CLASSIFICATION OF PILI
Bacteria assemble a wide variety of proteinaceous structures on their surfaces, with functions including motility, adhesion, and the secretion of proteins and other molecules (1–3). Bacterial surface structures distinct from the motility organelles, flagella (Latin for whip), were first described in the late 1940s and early 1950s, with the advent of the electron microscope (4). The term fimbriae (Latin for thread or fiber) was coined by Duguid and coworkers to describe the surface structures necessary to bind and agglutinate erythrocytes (5). Soon thereafter, Brinton used the term pili (Latin for hair) to describe the nonflagellar surface structures expressed by Escherichia coli (6). In 1975, Ottow suggested that “pili” be reserved for those structures involved in bacterial mating and “fimbriae” for structures involved in adhesion (4). However, his recommendation did not stick and the terms remain used interchangeably. Here, we will generally use the term pili.
Pili are hair-like organelles that decorate the bacterial surface. Pili are typically involved in adhesion and function in a range of interactions between bacteria, bacteria and other cells, and bacteria and their surrounding environment. These functions include the formation of microcolonies and biofilms, colonization of surfaces, and receptor-mediated adhesion to host cells (1). Some types of pili also function in motility and the uptake of DNA or phage (7). By acting outside a bacterium’s capsule or other protective surface structure, pili may increase the functional reach of bacteria and confer adhesive functions while preserving the barrier properties of the cellular envelope. The ability of pili to act distantly from the cell surface also may facilitate bacterial evasion of immune surveillance and detection or uptake by host cells.
Pilus Classification Schemes
Pilus classification schemes have changed over the years. In 1965, Brinton distinguished six types of pili in E. coli (8). The following year, Duguid and colleagues proposed a classification scheme based on pilus morphology and hemagglutination potential (9). This scheme comprised seven pilus types (types 1 through 6 and F). In subsequent schemes, pili were classified based on their abilities to agglutinate human red blood cells of different blood groups in the presence or absence of mannosides. For example, P pili of uropathogenic E. coli (UPEC) bind the disaccharide Galα (1–4)Gal linkage on erythrocytes of the P blood group system and are mannose-resistant (10, 11), whereas Dr pili (also mannose-resistant) bind CD55 on DR blood group erythrocytes (12, 13). This classification scheme led to the term type 1 pili, which remains in current use, to refer to mannose-sensitive bacterial surface fibers. However, genetic analyses revealed that hemagglutination-based classification schema are arbitrary, because they may assign pili encoded by homologous genes into different groups, and pili encoded by distinct systems into the same group (14–17). Additional classification systems based on serology have emerged (18). Such schemes have been particularly useful to classify the variety of pilus antigens expressed by E. coli. However, because of the high specificity of serological classification, this method is not useful in assigning individual pili to larger groups of related surface structures. In his 1975 landmark review, Ottow proposed an updated classification scheme that distinguished six groups of pili (4). For example, group 4 included polarly inserted pili involved in bacterial locomotion. Only parts of the systems prescribed by Duguid and Ottow have endured over the years. For instance, type 1 pili, the classification designated by Duguid, remains the common descriptor for pili that mediate mannose-sensitive hemagglutination. Ottow’s “group 4” designation has evolved to “type IV” pili, the designation given to pili that mediate twitching motility and are capable of extending and retracting (7). Additional classification has been based on the pilus architecture (see “Architecture of Chaperone/Usher Pili” below) (19–21). For example, pili may be classified as either homopolymeric or heteropolymeric. Heteropolymeric pili (e.g., type 1 and P pili) are composed of a number of different pilus subunits of different stoichiometry. The architecture of heteropolymeric pili generally consists of a relatively rigid pilus rod, with a flexible distal tip. The adhesin subunit in heteropolymeric pili is present in a single copy at the most distal location of the tip. Thus, heteropolymeric pili are “monoadhesive.” Homopolymeric pili (e.g., Dr and AFA-III pili) are composed of a multiply incorporated single subunit and are generally less structured than heteropolymeric pili. Homopolymeric pili are thin and resemble the distal tip of the heteropolymeric pilus. The homopolymeric pilus subunit also functions as an adhesin, and thus these pili are “polyadhesive.”
Modern classification schemes based on the biochemistry and genetics of pilus assembly systems have proven more useful and durable for assigning pili into distinct classes. These schemes are based on the pilus assembly mechanism and group pili according to the sequence homology of the genes encoding the machinery required for pilus biogenesis. For Gram-negative bacteria, such as E. coli and Salmonella, this results in the classification of pili into at least four different groups: chaperone/usher (CU) pili, curli, type IV pili, and conjugative F pili (22–26). Pili assembled by the CU pathway are the focus of this review.
Pili Assembled by a Chaperone- and Usher-Dependent Mechanism
The CU pathway serves to assemble and secrete a superfamily of adhesive and virulence-associated surface structures in Gram-negative bacteria (27, 28). Pili are polymeric fibers assembled from multiple subunit proteins. The assembly of pili by the CU pathway involves the binding of nascent pilus subunits by a dedicated chaperone in the bacterial periplasm, and the subsequent polymerization of subunits into the pilus fiber at the outer membrane (OM) by an integral OM channel protein termed the usher. Genetic loci coding for CU pili are located both chromosomally and on plasmids, and a given bacterial genome may contain multiple different CU loci. A systematic effort by Nuccio and Bäumler (29) categorized all CU pathways into phylogenetic clades on the basis of usher gene sequence, yielding six clades: α, β, γ (subdivided into γ1, γ2, γ3, and γ4), κ, π, and σ. For the α, κ, π, and σ clades, the clade designations were assigned to reflect a particular quality of the clade or a prominent member as follows: α-pili, alternate CU family; κ-pili, K88 (F4) pili; π-pili, pyelonephritis-associated pili (P pili); and σ-pili, spore coat protein U from Myxococcus xanthus.
In addition to the six phylogenic clades based on usher gene sequence (29), the CU superfamily has been subdivided by using other criteria. The α-phylogenetic clade corresponds to a subfamily of CU pili previously termed the “alternate” CU family. The “alternate” designation was initially made based on the fact that the gene sequences for these pili possess little homology to constituents of “classical” CU pili such as the type 1 and P pili of UPEC. However, subsequent studies revealed that the alternate and classical CU pili are in fact related and share similar assembly mechanisms (29–31). The alternate CU family comprises pili expressed by enterotoxigenic E. coli (ETEC) and Salmonella enterica serotype Typhi, including colonization factor antigen I (CFA/I) and Typhi colonization factor (Tcf) pili (31–36). The alternate CU family is sometimes referred to as “class 5” pili, from a classification scheme based on sequence analysis of the pilus subunit proteins (37). An additional division of the CU superfamily into two subfamilies has been made based on conserved sequence differences in a region of the pilus chaperones. These differences relate to the length of the loop connecting the chaperone’s F1 and G1 β-strands (see “Mechanism of Pilus Assembly by the Chaperone/Usher Pathway” below). Those systems whose chaperones contain a short loop belong to the F1-G1-short (FGS) subfamily; those whose chaperones contain large loops are categorized in the F1-G1-long (FGL) subfamily (38–40). The FGS subfamily assembles a range of pilus structures, including rigid, helical rods with distinct tip fibers such as the type 1 and P pili. The FGL subfamily assembles thin, fibrillar, or nonfibrillar surface structures (38, 39). The FGL CU subfamily forms the γ3-phylogenetic clade in the classification described by Nuccio and Bäumler, and has also been termed “class 2” pili based on sequence analysis of the subunit proteins (29, 37).
CU pili are prevalent among the Enterobacteriaceae, and most of the known types of pili expressed by E. coli and Salmonella belong to the CU family. The CU pathway assembles a variety of surface structures, ranging from rigid, rod-like pili to amorphous (afimbrial) and capsular-like structures. This architectural diversity is reflected by a broad functional diversity among CU pili. The pili assembled by E. coli and Salmonella are representative of the structural and functional diversity encoded by CU pathways (Fig. 1, Table 1). Two pili in particular have served as model systems for the study of pilus assembly by the CU pathway: the adhesive type 1 and P pili expressed by UPEC (41, 42). UPEC is the primary causative agent of urinary tract infections (UTI), one of the most common infections worldwide, affecting more than 50% of women during their lifetime (43–45). Type 1 and P pili confer UPEC with the ability to adhere to the bladder and kidney, respectively, withstand turbulent urine flow, and colonize the urinary tract. The type 1 and P pili expressed by UPEC are two of the best-characterized pilus assembly systems, and this review will focus on these pili as models.
Figure 1.

Model for pilus biogenesis by the CU pathway. Pilus subunits translocate from the cytoplasm to the periplasm as unfolded polypeptides via the Sec system. Subunit folding occurs upon interaction with the pilus chaperone (yellow) in the periplasm. Chaperone-subunit complexes then interact with the OM usher for exchange of chaperone-subunit for subunit-subunit interactions, ordered assembly of the pilus fiber, and secretion through the usher channel to the cell surface. The usher is depicted with its β-barrel channel domain in the OM and its plug, N, C1, and C2 domains labeled. The N domain forms the initial binding site for chaperone-subunit complexes, and the C domains provide a second binding site for the assembling pilus fiber. Chaperone-adhesin complexes have the highest affinity for the usher and initiate pilus assembly by binding to the usher N domain, with subsequent handoff to the usher C domains. Repeated rounds of chaperone-subunit targeting to the usher N domain and subunit-subunit interaction then lead to assembly and secretion of the pilus fiber in a top-down manner. Models of fully assembled type 1 (Fim), P (Pap), and Afa pilus fibers are shown.
Table 1.
CU pili present in E. coli and Salmonella
| Pilus | Adhesin | Chaperone/usher | Gene cluster | Associated disease or function | Receptor | Refs |
|---|---|---|---|---|---|---|
| Escherichia coli | ||||||
| P | PapG | PapD/PapC | papAHCDJKEFG | Pyelonephritis or cystitis | Galα1,4 Galβ, GbO3, GbO4, GbO5, kidney epithelium, erythrocytes | 41 |
| Type 1 | FimH | FimC/FimD | fimAICDFGH | Cystitis, biofilm formation | Uroplakin UP1a, β1α3 integrins, laminin, CD48, collagen (type I and IV), bladder and kidney epithelium, buccal cells, erythrocytes, mast cells, neutrophils, macrophages | 42 |
| Afa/Dr | AfaE | Afa(Dra)B/Afa(Dra)C | afa(dra)BCDPE | Pyelonephritis | Dr blood group, α5β1 integrin, CD55/DAF, CEACAMS, uroepithelium, erythrocytes | 90, 147, 343 |
| AAF-I | AggA | AggD/AggC | aggDCBA | Diarrhea | Hep-2 cells | 344 |
| AAF-II | AafA | AafD/AafC | aafBCDA | Diarrhea | Intestinal mucosa | 223 |
| AAF-III | Agg-3A | Agg-3D/Agg-3C | agg-3DCBA | Diarrhea | Intestinal mucosa | 215 |
| AFA-III | Afa3-E | Afa-3B/Afa-3C | afa-3BCDE | Diarrhea, cystitis | Dr blood group, α5β1integrin, CD55/DAF, CEACAMs, uroepithelium, erythrocytes | 345 |
| Afa-8 | Afa-8E | Afa-8B/Afa-8C | afa-8BCDE | Diarrhea/septicemia | DAF | 346 |
| AF/R1 | AfrE | AfrC/AfrB | afrABCDE | Diarrhea in rabbits | Rabbit sucrose-isomaltase protein complex | 347 |
| Auf | AufG | AufB,F/AufC | aufABCDEFG | N.D.,a auf identified in a UPEC strain | N.D. | 348 |
| Csh | N.D. | CshC/CshB | cshABCDEFG | N.D. | N.D. | 349 |
| CS1, CFA/I | CooD | CooB/CooC | cooBACD | Diarrhea | Intestinal epithelium, bovine erythrocytes | 283 |
| CS2, CFA/II | CotD | CotB/CotC | cotBACD | Diarrhea | Intestinal epithelium, bovine erythrocytes | 129 |
| CS3 | CS3-3 | CS3-1/CS3-2 | CS3-1,-2,-3 | Diarrhea | Adhesion to intestinal mucosa | 350 |
| CS4 | CsaE | CsaA/CsaC | csaABCE | Diarrhea | Intestinal epithelium | 32 |
| CS5 | N.D. | CsfB,F/CsfC | csfABCEFD | Diarrhea | Intestinal epithelium | 351 |
| CS6 | (probably CssB) | CssC/CssD | cssABCD | Diarrhea | sulfatide (SO3-3Galβ1Cer) | 352 |
| CS12 | CswG | CswB,C,E/CswD | cswABCDEFG | Diarrhea | Human intestinal epithelium | 353 |
| CS14 | CsuD | CsuB/CsuC | csuBA1A2CD | Diarrhea | Intestinal epithelium | 354 |
| CS17 | CsbD | CsbB/CsbC | csbBACD | Diarrhea | GlcNAc β1,2 Man, Epithelial intestinal cells | 354 |
| CS18 | FotG | FotB/FotD | fotABCDEFG | Diarrhea | Intestinal epithelium | 355 |
| CS19 | CsdD | CsdB/CsdC | csdBACD | Diarrhea | Intestinal epithelium | 356 |
| CS20 | N.D. | N.D. | N.D. | N.D.; associated with an Indian ETEC strain | Human intestinal epithelium | 357 |
| CS31 | ClpG | ClpE/ClpD | clpBCDEFGHI | Diarrhea | Intestinal epithelium | 358 |
| DAF | DafaE | DafaB/DafaC | dafaABCDE | Diarrhea | N.D. | 359 |
| ECP (Yag) | N.D. | YagV/YagX | yagZYXWV | N.D. | HEp-2 and HeLa epithelial cells | 360 |
| Fst | FstE | N.D. | N.D. | N.D. | Fibronectin, rat basolateral membranes of kidney tubules | 361, 362 |
| F1C | FocH | FocC/FocD | focAICDFGH | Cystitis, pyelonephritis | GalNAc β 1,4 Galβ (bladder endothelium; distal tubules, collecting ducts, glomeruli, endothelium of kidney) | 363 |
| F7 (1) | FsoE | N.D. | N.D. | N.D. | Fibronectin, rat basolateral membranes of kidney tubules | 361, 362 |
| F9 (Yde) | YdeQ | ND/YdeT | ydeTSRQ | N.D., biofilm formation | Bovine epithelial cells, mouse model of chronic cystitis | 364, 365 |
| F17 | F17G | F17D/F17C | F17aADCG | Diarrhea | N-Acetylglucosamine, intestinal villi | 366 |
| F107 (F18ab) | FedF | FedC/FedB | fedABCEF | Edema and diarrhea in pigs | Blood group H type 1 determinant; pig intestinal epithelium | 367 |
| F1651 | N.D. | N.D. | N.D. | N.D. | N.D. (found in a porcine/bovine pathogenic E. coli strain 4787 | 368 |
| F1845 | Probably DaaE | DaaB/DaaC | daaFABCDPE | Diarrhea | Decay-accelerating factor (DAF) | 231 |
| Hda | HdaA | HdaD/HdaC | hdaDCBA | N.D. | Aggregative adhesion to Hep-2 cells | 216 |
| K88 (F4) | FanH | FanE/FanD | fanCDEFGH | Diarrhea in neonatal piglets | NeuGc α2,3 Gal β1,4 GlcNAc | 369 |
| K99 (F5) | FaeH | FaeE/FaeD | faeCDEFGHIJ | Diarrhea in neonatal calves, lambs, piglets | NeuGc α2,3 Gal β1,4 GlcNAc | 369 |
| Lda | LdaH | LdaE/LdaD | ldaCDEFGHI | Diarrhea | Diffuse adhesion on Hep-2 cells | 370 |
| Lpf | LpfD | LpfB/LpfC | lpfABCD | Associated with diarrheagenic strains | Intestinal epithelium | 174 |
| Nfa | NfaA | NfaE/NfaC | nfaEDCBA | Urinary tract infections, newborn meningitis | Neutrophils | 90, 371 |
| PCF071 | CosD | CosB/CosC | cosBACD | Diarrhea | Intestinal epithelium | 372 |
| Pix | PixG | PixD,J/PixC | pixAHCDJFG | N.D., associated with pyelonephritis isolate | N.D. | 47 |
| PRF | PrfG | PapD,J/PrfC | prfAHCDJKEFG | N.D. | N.D. | 373 |
| REPEC | RalG | RalE/RalD | ralCDEFGHI | Diarrhea in rabbits | Rabbit intestinal epithelium | 374 |
| S | SfaH | SfaE/SfaF | sfaADEFGSH | Urinary tract infections, newborn meningitis | NeuAc α 2,3 Galβ sialic acid residues, plasminogen, bladder and kidney epithelial cells, erythrocytes, endothelial cells | 263 |
| Sfm | SfmH | SfmC/SfmD | sfmACDHF | N.D. | Bladder epithelium | 48 |
| Sfp | SfpG | SfpD,J/ SfpC | sfpAHCDJFG | Diarrhea | N.D. | 167 |
| Stg | StgD | StgB/StgC | stgABCD | Respiratory tract infection in birds | Avian respiratory tissues | 375 |
| Yad | YadC | EcpD/HtrE | yadNecpDhtrEyadMLKC | UTI in mouse model | Abiotic surfaces, bladder epithelium | 376 |
| Ybg | N.D. | YbgP/YbgQ | ybgDQPO | N.D. | N.D. | 48 |
| Ycb | YcbT | YcbR,F/YcbS | ycbQRSTUVF | N.D. | Abiotic surfaces | (48) |
| Yeh | YehA | YehC/YehB | yehDCBA | UTI in mouse model | Abiotic surfaces, embryonic kidney cells | 48, 376 |
| Yfc | N.D. | YfcS/YfcT,U | yfcVUTSRQPO | N.D. | N.D. | 48 |
| Ygi | YgiL | YgiH/YgiG | ygiLyqiGHI | UTI | Human kidney epithelium, abiotic surfaces | 48 |
| Yhc | N.D. | YhcA/YhcD | gltFyhcADEF | N.D. | N.D. | 47 |
| Yqi | N.D. | YqiH/YqiG | yqiLGHI | N.D. | N.D. | 29 |
| Yra | YraK | YraI/YraJ | yraHIJK | N.D. | Bladder epithelium | 48 |
| 987P | FasG | FasB,C,E/FasD | fasABCDEFG | Diarrhea in neonatal pigs | Brush border vesicle-derived Histone H1 proteins of neonatal piglet intestinal epithelium | 377, 378 |
| Salmonella spp. | ||||||
| Bcf | BcfD | BcfB/BcfC | bcfABCDEFGH | Intestinal persistence in mice | N.D. | 49, 292 |
| Fae | FaeG | FaeE/FaeD | faeBCDEFGHIJKA | N.D. | N.D. | 379 |
| Lpf | LpfD | LpfB/LpfC | lpfABCDE | Intestinal persistence in mice, typhoid in mouse model, gastroenteritis, biofilm formation | Peyer’s patches in murine model | 290, 292 |
| Mrk | MrkD | MrkB/MrkC | mrkABCDFJI | N.D. | N.D. | 379 |
| Pef | PefD | PefC/PefA | pefBACDEFI | Gastroenteritis, biofilm formation | Lewis X blood group antigen | 299 |
| Peg | PegD | PegB/PegC | pegABCDE | Intestinal and oviduct colonization in chickens; egg contamination; systemic infection in chickens and mice | N.D. | 380, 381 |
| Peh | PehD | PehB/PehC | pehABCDE | N.D. | N.D. | 379 |
| Saf | SafD | SafB/SafC | safABCD | Gastroenteritis, enteric typhoid fever | N.D. | 382 |
| Sba | SbaH | SbaC/SbaD | sbaABCDEFGH | N.D. | N.D. | 379, 383 |
| Sbc | SbcF | SbcD/SbcC | sbcABCDEFGH | N.D. | N.D. | 379 |
| Sbb | SbbD | SbbC/SbbB | sbbABCD | N.D. | N.D. | 379, 384 |
| Sdk | SdkD | SdkC/SdkB | sdkABCD | N.D. | N.D. | 379 |
| Stj | StjA | StjC/StjB | stdABCD | Intestinal persistence in mice | N.D. | 385 |
| Sdc | SdcE | SdcC/SdcD | sdcABCDE | N.D. | N.D. | 379 |
| Sdd | SddD | SddB/SddC | sddABCDEF | N.D. | N.D. | 379 |
| Sde | SdeD | SdeB/SdeC | sdeABCDE | N.D. | N.D. | 379 |
| Sdf | SdfD | SdfB/SdfC | sdfABCDEF | N.D. | N.D. | 379 |
| Sdg | SdgG | SdgB/SdgC | sdgABCDEFG | N.D. | N.D. | 379 |
| Sdh | SdhG | SdhB/SdhC | sdhABCDEFG | N.D. | N.D. | 379 |
| Sdi | SdiD | SdiB/SdiC | stiABCD | N.D. | N.D. | 379, 384 |
| Sdj | SdjD | SdjB/SdjC | sdjABCD_yhoN | N.D. | N.D. | 379 |
| Sef14/18 | SefD | SefB/SefC | sefABCDR | Enteric typhoid fever | N.D. | 386 |
| Sta | StaG | StaB/StaC | staABCDEFG | N.D. | N.D. | 49 |
| Stb | StbD | StbB/StbC | stbABCDEF | Intestinal persistence in mice | N.D. | 49, 292 |
| Stc | Stc | StcB/StcC | stcABCD_yhoN | Intestinal persistence in mice | N.D. | 49, 292, 300 |
| Std | Std | StdA/StdB | stdABCD | Intestinal persistence in mice | α(1,2)-fucose | 49, 292 |
| Ste | SteG | SteB/SteC | steABCDEFGHIJ | N.D. | N.D. | 49 |
| Stf | StfH | StfC/StfB | stfABCDEFGH | Systemic and fatal infection in inbred mice | N.D. | 289, 387 |
| Stg | StgD | StgB/StgC | stgABCD | N.D. | Association, invasion, and permeabilization of HEp-2 epithelial cells and macrophage-like cells | 49, 388 |
| Sth | SthE | SthB/SthC | sthABCDE | Intestinal persistence in mice | N.D. | 49, 292 |
| Sti | StiH | StiB/StiC | stiABCH | N.D. | N.D. | 17, 379 |
| Stk | StkG | StkB/StkC | stkABCDEFG | N.D. | N.D. | 51 |
| Tcf | TcfD | TcfA/TcfC | tcfABCD | N.D. | N.D. | (49) |
| Type 1 | FimH | FimD/FimC | fimAICDHFZYW | Intestinal persistence in mice and pigs, biofilm formation | Mannose, adherence to laminin via mannose, horse and chicken erythrocytes, enterocytes, glycoprotein-2 | 289, 389 |
N.D., not determined.
CHAPERONE/USHER GENETIC LOCI AND REGULATION
Genes encoding CU pili are found on both the bacterial chromosome and plasmids. The pilus genes are typically clustered together with a similar organization: an upstream regulatory region followed by a single downstream operon that encodes the pilus structural proteins, adhesive component(s), and assembly machinery (Fig. 2). CU pilus gene clusters are often located in proximity to other genes encoding virulence factors, in regions termed pathogenicity islands that have signatures suggesting acquisition by horizontal gene transfer (46). A genomic analysis found that E. coli strains encode as many as 17 CU gene clusters and that UPEC strains encode 9 to 12 intact CU pilus operons (47). Many of these CU loci are cryptic and do not express pili under laboratory growth conditions, and thus the functions of their gene products remain elusive (48). However, evidence suggests that many, if not most, of these cryptic operons encode products that are functional for adhesion and may contribute to ability of E. coli to colonize a wide variety of environmental niches (48). Salmonella strains similarly encode multiple pilus gene clusters, many of which are also not expressed under laboratory growth conditions (17, 49, 50).
Figure 2.

CU pilus gene clusters and regulatory regions. (A) Gene clusters, including upstream regulatory regions, are shown for P (pap), type 1 (fim), and Dr/Afa pili, with the functions of the genes indicated. (B) Regulatory region of the pap gene cluster, shown in phase-OFF and phase-ON states. Lrp binding to the GATCproximal site turns off expression from the papBA promoter. Binding of Lrp, together with PapI, to the GATCdistal site allows Dam methylation of GATCproximal, resulting in phase-ON expression. Production of PapB during phase-ON expression initiates a positive feedback loop through upregulation of PapI. (C) Regulatory region of the fim gene cluster, showing the phase-OFF and phase-ON orientations of the fimS switch region. The left and right inverted repeat sequences (IRL and IRR) that flank the fimS switch are indicated. Binding of H-NS maintains fimS in the phase-OFF position, whereas binding of IHF and Lrp favors phase-ON expression.
Pilus gene sequences are among the most polymorphic regions in the genomes of E. coli and Salmonella. Comparisons of pilus repertoires of S. enterica serotypes Typhimurium (LT2), Typhi (CT18 and Ty2), and Paratyphi A (SARB64) illustrate that pilus gene sequences were lost or acquired frequently during the divergence of their lineages. The strictly human adapted serotypes Typhi and Paratyphi A carry stop codons or frameshift mutations in a number of their pilus gene clusters, which may reflect their host restriction or the decreased importance of intestinal persistence during systemic infections, such as typhoid fever (49, 51, 52). Alternatively, the loss of these pili may reflect a negative impact of expression of these structures during infection of the host. In E. coli, the tip adhesin subunits of pili are under strong evolutionary pressure, selecting for functionally adaptive amino acid replacements (53–55). This is consistent with the importance of adhesin-receptor interactions in determining bacterial tropism and the ability to colonize different niches.
The expression of CU pili is highly regulated, subject to ON/OFF phase variation and responsive to environmental cues (56–59). Regulatory cross talk occurs among the different CU gene clusters present within a given bacterium (60–63). This cross talk among pilus gene clusters likely serves as a way to limit unnecessary expression of surface structures and ensure expression of only one adhesin at time. Such regulation may be particularly important in the context of infection, to prevent recognition of bacterial surface structures by the host immune system and to control bacterial tropism to specific colonization sites. Furthermore, expression of adhesive pili is inversely correlated with the expression of flagella for motility (64). This allows bacteria to stick (colonize a specific site) or swim (turn off adhesion and move to new site). Here, we provide descriptions of the regulation of type 1 and P pilus expression by UPEC, as examples of CU pilus regulatory mechanisms.
Type 1 Pilus Regulation
Type 1 pili are encoded by the fim chromosomal gene cluster (Fig. 2A). Phase variable expression of type 1 pili is dictated by the orientation of fimS, an invertible DNA segment containing the promoter for the fim operon (Fig. 2A). The FimB and FimE recombinases, part of the λ-integrase (tyrosine-dependent site-specific recombinase) family, directly influence fimS orientation. FimB and FimE act on two 9-base pair inverted repeat sequences (IRR and IRL) that flank the invertible 314-base pair fimS DNA central switch, influencing the orientation of the switch and thus phase variation of the fim operon (65, 66). The orientation in which fim transcription occurs is referred to as phase-ON, and the nontranscriptional orientation is phase-OFF. The FimB and FimE recombinases have opposing influences on the fimS switch. FimE primarily flips the switch from the phase-ON to phase-OFF orientation, whereas FimB promotes fimS switching to either orientation (67, 68). This process of DNA inversion presumably involves the formation of a FimB or FimE complex with the fimS sequence, generation of a Holliday junction, and the breaking, rejoining, and resolution of the DNA strands. Phase variation is reversible, and the rate of switching has been calculated at 10−3 to 10−4 per cell per generation (69, 70). As a result of phase variation, cultures of UPEC exhibit a mixture of type 1 piliated and nonpiliated cells. This may ensure that a subpopulation of bacteria is always present to adapt to the local environmental niche.
The fimB and fimE genes, located upstream of the fim operon, are controlled by independent promoters (Fig. 2A). Because of their direct influence on fim phase variation, the regulation of both recombinases is key for type 1 pilus expression. For example, biofilm formation is influenced by the iron-sulfur cluster regulator (IscR) through modulation of type 1 pilus expression (71). Under iron-limiting conditions, IscR binds to the fimE promoter region, resulting in increased fimE expression and switching of fimS to the phase-OFF position. Conversely, when iron is available, fimE expression is downregulated, thereby positively influencing fim gene expression and pilus production. This model is consistent with the observation of decreased E. coli biofilm and pilus production under iron-limiting conditions (71). In contrast to IscR working through fimE, the global regulators ppGpp and DksA influence type 1 pilus production through transcriptional control of fimB (67, 72). As examples of the cross talk among CU pilus gene clusters, PapB, which regulates P pilus expression, also affects orientation of the fimS switch (60, 63, 73), and SfaB and SfaX, which are S pili regulatory proteins, negatively influence fim phase-ON switching (73, 74).
The fimS switch orientation is also influenced by other factors, including the bacterial nucleoproteins leucine-responsive regulatory protein (Lrp), integration host factor (IHF), and histone-like nucleoid-structuring (H-NS) protein (75). IHF and Lrp facilitate phase-ON switching, whereas H-NS maintains fimS in the phase-OFF position (Fig. 2C). The degree of DNA supercoiling, controlled by DNA gyrase activity, is postulated to be related to Lrp, IHF, and H-NS binding (75, 76). Changes in DNA gyrase activity and Lrp and IHF levels during stationary growth phase likely facilitate phase-ON switching of the fimS region, which is consistent with the early observation of robust pilus production by cultures maintained under static growth conditions (77). Type 1 pilus expression is also correlated with the osmolality and pH of the growth conditions. Increased osmolality and an acidic pH repress fim transcription (78). Such observations lead to a model consistent with infection of the human bladder and kidney; the bladder has a pH relatively higher and an osmolality relatively lower than the human kidney. Thus, type 1 pili, which facilitate colonization of the bladder, are likely upregulated in response to cues within the bladder niche and downregulated in bacteria that have ascended to the kidneys. In addition, a link between oxygen tension and type 1 pilus expression has been demonstrated in the context of bacterial biofilms grown in culture, with the fim promoter switched to phase-ON at the air-exposed surface of the biofilm, but switched OFF in deeper layers of the biofilm (79, 80).
P Pilus Regulation
P pili are encoded by the pap chromosomal gene cluster (Fig. 2A). P pili are regulated by a distinct mechanism compared to type 1 pili. Expression of the pap operon from the papBA promoter is regulated by the PapI and PapB proteins (81). Rather than an invertible switch element, P pilus expression is dictated by the methylation pattern of two GATC sites located upstream of the papBA promoter. Methylation at these sites, in conjunction with the action of PapI, influences binding of the global regulator Lrp (82). At high Lrp and low PapI concentrations, Lrp binds efficiently to the GATC site proximal to the papBA promoter (GATCproximal) (Fig. 2B). Lrp binding to GATCproximal results in inhibition of pap transcription (83, 84). This binding can only occur in the absence of GATCproximal methylation, and therefore Lrp competes with the DNA adenine methylase (Dam) for binding (Fig. 2B). PapI binds to Lrp and is required for phase-ON expression (85). P pilus phase variation is dictated by Lrp concentration, which in turn is dependent on cell growth stage and conditions, and Lrp accessibility to the promoter. When cells reach early stationary phase, Lrp concentration increases, thereby promoting the transition from phase-ON to phase-OFF. At low Lrp concentrations, for example, soon after cell replication and DNA division, and in the presence of PapI, Lrp binds to the hemimethylated distal GATC site (GATCdistal), thereby preventing Dam from accessing this site for additional methylation, and thus encouraging Dam methylation of GATCproximal, which in turn contributes to phase-ON variation (Fig. 2B). In addition, phase-ON expression results in PapB expression, which initiates a positive feedback loop through upregulation of PapI (86).
P pili, as with other CU pili, are subject to multiple layers of regulation. Consistent with the pathophysiology of the human urinary tract, the phase-OFF to phase-ON transition in P pili occurs in response to growth in urine with a high-amino-acid content (87). This regulation may allow pap genes to remain phase-OFF within the fecal flora, but switch to phase-ON upon exposure to the urinary environment. The CpxAR signal transduction system also modulates P pilus phase variation. CpxR competes for binding to both the proximal and distal papBA promoter sites by binding at positions adjacent to the GATC regions, inhibiting expression of the pap operon (88). This additional mode of control has been observed in response to an alkali environment and is not altered by methylation (88). UPEC also use small regulatory RNAs (sRNAs), which have short half-lives and thus allow rapid responses, to control P pilus expression. The sRNA repertoire of UTI89, a pathogenic UPEC strain, differs significantly for bacteria grown in laboratory culture versus the mouse bladder infection model (89). A novel sRNA, termed papR (P-fimbriae regulator), was shown to influence P pilus phase variation (89). The papR sRNA mediates posttranscriptional repression of PapI, and a papR deletion strain has increased bladder and kidney adhesive properties.
ARCHITECTURE OF CHAPERONE/USHER PILI
Structures assembled by the CU pathway range in shape from rigid, helical rods, which are typically thought of as classical pili, to thin, flexible fibers that, in some cases, form amorphous, capsular-like, “afimbrial” structures. The type 1 and P pili of UPEC are prototypical examples of classical pilus structures. Afa/Dr pili, expressed by various pathogenic E. coli strains (90, 91), are examples of thin, flexible pilus structures. The overall architectures of mature type 1 and P pili are shown in Fig. 1. Both pilus fibers consist of a rigid, helical rod with a diameter of ∼7 to 8 nm, forming the bulk of the 1- to 2-μm pilus length. The pilus rod is anchored in the bacterial OM and is composed of thousands of copies of the major subunit protein, FimA (type 1 pili) or PapA (P pili). The major subunit protein polymerizes into a linear fiber, which then further folds into a one-start, right-handed helix. The helical rod is terminated by the PapH subunit in P pili (Fig. 1), and the terminating subunit also has a role in anchoring the pilus fiber in the OM (92). The FimI protein might perform the equivalent function as a rod terminator for type 1 pili, but this remains to be confirmed (93, 94). The end of the pilus rod distal from the bacterial surface contains the adhesive tip portion of the pilus (Fig. 1). The pilus tip is an open linear fiber, with the most distal of the subunits being the adhesive subunit or adhesin. Type 1 pilus tip fibers contain the FimH adhesin, followed by the FimG and F adaptor subunits. P pilus tip fibers contain the PapG adhesin, followed by the PapF, E, and K subunits. The P pilus tip fiber is longer and more flexible than the type 1 pilus tip, because of the incorporation of 5 to 10 PapE subunits per P pilus tip. In contrast to the type 1 and P pili, Afa/Dr are thin (∼2 nm in diameter) flexible, linear fibers composed of a single-subunit protein, AfaE (DraE), which also functions as the adhesive subunit (95–99) (Fig. 1). A second pilus-associated protein, AfaD (DraD), is present in a single copy at the distal end of the AfaE pilus fiber and is thought to function as an invasin for uptake into host cells (99, 100). The DraD protein may exist in a secreted form separate from the pilus but located on the bacterial surface (100–102). Pili such as the type 1 and P pili that contain a single distal adhesin have been termed monoadhesive pili (21). In contrast to monoadhesive pili, polyadhesive pili such as Afa/Dr possess a major subunit protein that provides adhesive properties and receptor-binding sites along the exposed surface of the pilus. Thus, for polyadhesive pili, the entire pilus fiber may act in adhesion (20, 99, 103, 104).
Structures of Pilus Subunits and Adhesins
All pilus subunits belonging to the CU superfamily possess a common structural domain termed the pilin domain, which consists of an incomplete immunoglobulin (Ig)-like fold (Fig. 3). Canonical Ig folds contain seven β-strands arranged into two sheets to form a β-sandwich (105). However, the pilin domain lacks the seventh, C-terminal β-strand (the G strand) (106–109). Tip-located adhesin subunits contain a second, N-terminal adhesin or lectin domain in addition to the pilin domain (Fig. 3). The adhesin domain, which has the receptor-binding site, typically has a complete Ig-like fold termed a β-barrel jelly-roll fold. The adhesin domain likely evolved through duplication of the pilin domain, allowing specialization of the adhesin domain for receptor binding. Although sharing similar overall structures, adhesin domains from different CU pili vary significantly in sequence and display marked differences in their receptor-binding sites and receptor-binding mechanisms, reflective of their diverse functions and adaptation to specific niches (110).
Figure 3.

Structures of pilus chaperone and subunit proteins. (A) Structure of a FimC-H chaperone-adhesin complex (PDB ID: 1QUN) from the type 1 pilus system. The FimC chaperone is in yellow and the FimH adhesin in green, with the FimH lectin and pilin domains indicated. The chaperone is engaged in donor strand complementation (DSC) with the subunit. The chaperone donates its G1 β-strand (in blue) to complete the Ig fold of the FimH pilin domain. (B) Structure of a FimG-FimH subunit-subunit complex (PDB ID: 4J3O) from the type 1 pilus system. FimH is depicted as in (A) and FimG is in orange. The N-terminal extension (Nte) of FimG is engaged in donor strand exchange (DSE) with FimH. The Nte of FimG completes the Ig fold of the FimH pilin domain. (C) Topology diagrams depicting the Ig folds of the FimG (orange) and FimF (red) pilin domains. FimF is depicted in DSC with the donated G1-strand of the FimC chaperone, which is inserted parallel to the FimF F-strand. FimG is depicted in DSE with the Nte of FimF, which is inserted antiparallel to the FimG F-strand. (D) and (E) Structures of the lectin domains of the FimH (type 1 pili; PDB ID: 1KLF) and PapG (P pili; PDB ID: 1J8R) adhesins, with bound mannose and globoside molecules, respectively. The sugars are depicted in dark gray stick representation.
The type 1 and P pilus adhesins are illustrative of the variations in adhesin structure. Type 1 pili bind to mannosylated receptors via the tip-located FimH adhesin. The FimH adhesin domain has a typical β-barrel jelly-roll fold (Fig. 3D). The mannose-binding site is located at the tip of the adhesin domain, formed by a deep, negatively charged pocket surrounded by a hydrophobic ridge (106, 111). Because the type 1 pilus tip fiber is relatively short and inflexible, this positions the FimH-binding site at the most distal part of the mature pilus, in a location where it can readily bind host cells. P pili bind to Gal(α1-4)Gal receptors via the tip-located PapG adhesin. The PapG adhesin domain also has a typical β-barrel jelly-roll fold. However, in contrast to FimH, the PapG receptor-binding site lies in a shallow pocket on one side of the adhesin domain, formed by three β-strands and a connecting loop region (Fig. 3E) (112, 113). The P pilus tip fiber is longer and more flexible than the FimH tip fiber (Fig. 1). The side-on orientation of the PapG-binding site, in combination with the length and flexibility of the P pilus tip, facilitates interaction of PapG with the globoside moieties that are oriented parallel to the membrane surface (112). For polyadhesive pili such as Afa/Dr, the major subunit protein also functions as the adhesin, and consists of a single pilin domain. The high stability of the DraE subunit, as well as other pilus subunits, has been attributed to a disulfide bond in a noncanonical position in the Ig fold (114). The AfaE and DraE subunits possess distinct receptor-binding sites located on opposite sides of the pilin domain, which recognize CD55/decay accelerating factor (DAF) and members of the carcinoembryonic antigen family (CEACAM) (21, 99, 103). These binding sites are repetitively presented along the length of the assembled pilus fiber.
Mechanics of Pilus-Mediated Adhesion
Pili are exquisitely adapted to facilitate binding to specific receptors, under variable environments and conditions, with both the adhesin subunit and overall pilus architecture contributing to pilus function. The mechanics of the type 1 pilus adhesin FimH have been well studied. To mediate colonization in the urinary tract, type 1 pili must be able to withstand the shear forces generated by the flow of urine. Application of shear stress shifts FimH-mediated binding from a “loose” to a “firm” mode, with increased receptor affinity in the latter state (115). This FimH binding behavior is due to a catch-bond mechanism (116). Catch bonds are similar in concept to a Chinese finger trap toy, which grips tighter as it is stretched. The determining factor of “tight” versus “loose” binding is the connective region between the pilin and adhesin domains of FimH. FimH in its relaxed state (with an interdomain angle of 142°) has an open mannose binding pocket, which is poised for receptor binding (117, 118). Application of shear force causes FimH to become stretched (180° interdomain angle). This leads to a twisting of the β-sandwich fold of the lectin domain that imparts tighter FimH binding by causing the mannose pocket to clamp shut.
The catch-bond behavior allows type 1-piliated bacteria to “stick and roll” on the mannose-coated bladder epithelium, providing functions of both migration (rolling) and attachment (sticking) during times of both urinary stasis and turbulence, respectively (116, 119). Recent evidence has further elucidated roles for the low- and high-affinity states of FimH. In the context of high flow (for example, during micturition), the low-affinity state of FimH facilitates initial binding because of its capacity to form bonds rapidly. The shear stress then switches FimH into the high-affinity state, which is slower to release from its ligand. This nonequilibrium cycle that makes use of the rapid binding and slow unbinding of two kinetic conformational variants suggests a model wherein FimH possesses a higher effective affinity than it would if the active and inactive variants existed at equilibrium (120). The reduced affinity of FimH under low shear may also provide a mechanism for bacteria to sample receptors in the urinary tract, because shear stress will only be generated on FimH if the bacteria bind to an immobilized receptor such as on the bladder wall, allowing the bacteria to avoid latching onto soluble mannose moieties in the urine such as Tamm-Horsfall protein (121).
The kidney is a less turbulent environment compared with the bladder lumen, and P pili are adapted to function in this niche. Binding of the P pilus adhesin PapG to the globoseries of glycolipids on human kidney epithelial cells, via the Gal(α1-4)Gal linkage, facilitates attachment of UPEC during an ascending UTI and is one of the initiating steps of pyelonephritis. Rather than the catch-bond mechanism used by FimH, PapG uses what is termed a slip bond. The slip bond between PapG and the galabiose moiety is weaker than the FimH-mannose interaction. The PapG interaction has a shorter bond lifetime that decreases as force increases (122, 123). This binding mechanism is adapted for bacterial rolling and rapid spreading in the lower shear environment of the kidney.
In addition to the binding behavior of the pilus adhesin, the helical pilus rod is important for resistance to shear forces such as encountered during urine flow, through the rod’s structural properties of compliance and flexibility (124). Studies of pilus rod stretching and flexibility have been performed for both the type 1 and P pilus systems (125–127). The entire stretching process is reversible (126, 127). Presumably, this reversible nature allows bacteria to regain proximity to host cells after exposure to shear force, as in micturition. The P pilus rod exhibits remarkable reversible stretching properties (127). The PapG-receptor bond is weak enough for rolling but strong enough to initiate unstacking of the PapA helical rod, and thus the redistribution of force throughout the pilus (122). Recently, the molecular basis for this property of the PapA rod was determined via a high-resolution cryo-EM structure (128). The structure revealed that each PapA subunit in the helical rod makes contact with ten additional PapA subunits, five below and five above. These extensive subunit-subunit interfaces are mostly composed of polar interactions, which are weaker than the hydrophobic interactions that mediate subunit-subunit polymerization. Thus, the quaternary structure of the pilus rod may be linearized as these polar interactions are disrupted, but the strong subunit-subunit interactions remain intact, and the pili do not break. Upon reduction of shear stress, the pilus rod can then recoil to its helical conformation. Comparing the unwinding forces of different pili illustrates the adaptation of bacteria to binding in various niches. Type 1 and P pili of UPEC exhibit unwinding forces of 21 to 30 pN, and Klebsiella pneumoniae type 3 pili exhibit an unwinding force of >65 pN (129). In comparison, the CS, CFA/I, and CFA/II pili, which facilitate gut colonization by E. coli, exhibit lower unwinding forces of <15 pN (129). These values reflect the relative turbidities of the urinary tract, respiratory tract, and gastrointestinal tract.
MECHANISM OF PILUS ASSEMBLY BY THE CHAPERONE/USHER PATHWAY
All pili assembled by the CU pathway share a common mechanism of biogenesis, which depends on a dedicated periplasmic chaperone and an OM-localized assembly and secretion channel termed the usher. Pilus biogenesis at the bacterial OM occurs in the absence of input from an external energy source such as ATP, which is not available in the periplasm (130, 131). Instead, the CU pathway harnesses protein-protein interactions to drive fiber assembly and secretion in a defined order. The CU pathway is perhaps the best understood bacterial secretion system, and the UPEC type 1 and P pilus systems have served as central model systems for these molecular and structural studies.
Donor Strand Complementation
Individual pilus subunits are synthesized with an N-terminal signal sequence that directs them to the Sec general secretory pathway for translocation from the cytoplasm to the periplasm (Fig. 1) (132). The signal sequence is cleaved in the periplasm, and the subunits undergo disulfide bond formation in a process catalyzed by the periplasmic oxidoreductase DsbA (130, 133). Following disulfide bond formation, subunits are bound by a dedicated periplasmic chaperone (FimC for type 1 pili, PapD for P pili). Pilus chaperones are ∼25 kDa, boomerang-shaped proteins composed of two Ig-like domains (Fig. 3A). The chaperone specifically recognizes and binds unfolded, but disulfide-bridged (oxidized), pilus subunits, ensuring that the subunits have productively interacted with DsbA prior to chaperone engagement (134). Engagement with the pilus chaperone is required for subunit folding. In the absence of the chaperone, subunits misfold, aggregate, and are degraded by the DegP periplasmic protease (135–139). The chaperone forms stable, binary complexes with the subunits in the periplasm, preventing premature subunit-subunit interactions and maintaining the subunits in an assembly-competent state.
Atomic-resolution structures have been solved for many different chaperone-subunit complexes, permitting a detailed understanding of this key step in the pilus assembly pathway. For type 1 and P pili, structures have been solved for all chaperone-subunit complexes except FimC-G; the chaperone-subunit interface is conserved for each of these complexes (106–108, 134, 140–144). As discussed in “Architecture of Chaperone/Usher Pili” above, each pilus subunit possesses an incomplete Ig-like fold termed the pilin domain, which is characterized by a missing C-terminal β-strand (the seventh or G β-strand). The absence of this strand leads to an exposed hydrophobic groove on the surface of the subunit. The chaperone completes this hydrophobic groove of the pilin domain with its own G1 β-strand, simulating a complete Ig-like fold in the subunit (Fig. 3A and C). This in trans β-strand donation process is termed donor-strand complementation (DSC). Through the chaperone’s contribution of the missing structural information to the subunit, the chaperone provides stabilization and thus catalysis of subunit folding in the periplasm, increasing the folding rate by at least 4 orders of magnitude relative to chaperone-independent folding (134).
The subunit groove caused by the missing β-strand contains five binding pockets, termed P1 to P5. The donor G1 β-strand of the chaperone contains an alternating sequence of four bulky, hydrophobic amino acids, termed residues P1 to P4. In DSC, the chaperone inserts its P1 to P4 residues into the P1 to P4 pockets of the subunit’s hydrophobic groove (the P5 pocket of the subunit is left unoccupied). The donated G1 β-strand from the chaperone is inserted in a parallel orientation to the F strand of the pilin domain, thus completing the subunit’s Ig fold in a noncanonical manner (Fig. 3C). This parallel orientation, coupled with the bulky size of the P1 to P4 residues inserted by the chaperone, maintains the subunit in a stable, high-energy intermediate state, which keeps the subunit structurally competent for assembly into the pilus fiber (108, 109, 145).
The type 1 and P pilus chaperones (FimC and PapD) belong to the FGS subfamily, with a short loop connecting their F1 and G1 β-strands. In comparison with these chaperones, the longer F1-G1 loop of FGL chaperones allows the G1 donor strand to fill the P5 binding pocket of the bound subunit, but this interaction is weaker than at the other pockets (145, 146). Compared with the FGS chaperones, FGL chaperones assemble simpler, homopolymeric pilus fibers, comprising only one or two different subunits (such as the Afa/Dr pili). FGL chaperones characteristically possess two conserved cysteines (one in the F1 β-strand and one in the G1 β-strand) which form a disulfide bond that is indispensable for chaperone-subunit complex formation (139, 147). Correlative studies have compared the size of the F1-G1 loop with the number of subunits the chaperone is capable of binding; the longer the F1-G1 sequence, the more specifically the chaperone binds (20, 109). The FGS chaperones may have evolved to generate less stringent specificity, because pili such as the type 1 or P pili are composed of up to 7 different subunits, each of which must be bound by the chaperone in the periplasm.
Donor Strand Exchange
Following formation in the periplasm, chaperone-subunit complexes next target the OM-spanning usher (FimD for type 1 pili, PapC for P pili), where chaperone-subunit interactions are exchanged for subunit-subunit interactions in a mechanism termed donor strand exchange (DSE) (108, 109). The usher catalyzes the formation of subunit-subunit interactions, and the mature pilus grows through a channel provided by the usher, starting with the adhesin and followed by the rest of the tip fiber and finally the pilus rod (Fig. 1).
Each pilus subunit (except the adhesin) contains an N-terminal extension (Nte; Fig. 3), which, much like the chaperone G1 β-strand, contains a series of alternating hydrophobic residues. The subunit Nte ranges from 10 to 20 amino acids in length and is disordered when the subunit is in complex with its chaperone. In the DSE reaction, the Nte of an incoming subunit displaces the chaperone donor strand and occupies the P1- or P2-to-P5 pockets of the preceding subunit’s hydrophobic groove (Fig. 3). The exchange of chaperone-subunit for subunit-subunit interactions is initiated by the Nte’s P5 residue binding to the P5 pocket (the only pocket left unoccupied by the chaperone) of the preceding subunit and proceeding toward P1 (146). This mechanism of exchange has thus been termed “zip-in, zip-out.” The P4 residue of the Nte is a strictly conserved glycine, the only amino acid small enough to fit in the P4 pocket (a bulky aromatic residue juts into this groove). This strict glycine conservation ensures that the donated Nte is properly oriented and registered when binding its partner. In P pili, the PapH subunit acts to terminate assembly of the pilus rod and facilitates anchoring of the pilus to the OM (92, 142). The mechanism by which PapH acts as an efficient terminator is well understood. The hydrophobic groove of the PapH pilin domain lacks a P5 pocket. The P1-P4 pockets are intact, and thus PapH can undergo DSC with the chaperone, but not DSE with a subsequent subunit. Therefore, PapH remains bound to the chaperone on the periplasmic face of the usher, anchoring the pilus fiber (Fig. 1).
In contrast to DSC, in which the chaperone’s G1 β-strand is inserted parallel to the F strand of the subunit’s pilin domain, the DSE reaction results in the donated Nte being inserted in a canonical, antiparallel orientation (Fig. 3C). In addition, the Nte donates smaller residues than the bulky residues inserted by the chaperone’s G1 β-strand. Together, these factors allow DSE to produce a complex residing at a lower energy state compared with DSC, resulting in a highly stable subunit-subunit interaction (136). The subunit-subunit interaction has a binding affinity among the strongest noncovalent interactions in nature, with a Kd of 1.5 × 10−20 M and a dissociation rate of 3 × 10−9 years (136, 148). Thus, the assembled pilus fiber, which is an array of Ig folds linked to each other noncovalently through DSE, is an extremely stable organelle that is well suited for bacterial adhesion in varied and stressful environments.
The Pilus Usher
Assembly of the pilus fiber takes place on the periplasmic face of the OM usher protein, with the usher also providing the channel for secretion of the fiber to the cell surface (Fig. 1). The usher catalyzes pilus assembly by a rate of over 1000-fold compared with spontaneous pilus assembly in the absence of the usher (149). The catalytic activity of the usher is thought to be due in large part to its ability to orient two periplasmic chaperone-subunit complexes relative to each other in such a way that they undergo DSE efficiently. The usher is a multidomain, membrane-spanning protein that is present in the OM in oligomeric form. Several structural snapshots of pilus assembly at the usher are now available: the FimD and PapC usher channel domains in their apo states; the isolated FimD N domain bound to FimC-H or FimC-F chaperone-subunit complexes; and the active FimD usher bound to FimC-H (FimD-C-H) or to an assembling pilus tip fiber (FimD-C-F-G-H) (Figs. 4 and 5) (117, 140, 150–153). From these snapshots, as well as accumulated biochemical, genetic, and computational evidence, the stages and mechanism of pilus assembly at the usher are becoming well defined.
Figure 4.

Structures of apo and activated FimD usher from the type 1 pilus system. (A) and (C) Structure of apo FimD (PDB ID: 3OHN), shown from side (A) and top (C) views. The transmembrane β-barrel channel domain is pictured in light blue and the plug domain in pink. The plug domain is positioned laterally within the β-barrel domain, closing the usher channel. The N, C1, and C2 domains are not present in this structure. (B and D) Structure of activated FimD (PDB ID: 3RFZ), shown from side (B) and top (D) views. The channel and plug domains are depicted as in (A), the N domain is in dark blue, the C1 domain is in cyan, and the C2 domain is in purple. In the activated usher, the plug is expelled from the channel and resides adjacent to the N domain in the periplasm.
Figure 5.

Type 1 pilus assembly cycle at the FimD usher and corresponding crystal structures. (A) Cartoon depictions of domain organization and color coding for the Fim proteins shown in panels (B through F). (B) Structure of apo FimD (PDB ID: 3OHN) with the transmembrane β-barrel channel closed by the plug domain. (C) Structure of the FimD N domain bound to a FimC-FimH pilin domain complex (PDB ID: 1ZE3). (D) Structure of the activated FimD usher with bound FimC-H chaperone-adhesin complex (PDB ID: 3RFZ). The FimH lectin domain is inserted inside the usher channel and the FimH pilin domain and bound FimC chaperone are located at the usher C domains. The FimD plug domain resides adjacent to the N domain in the periplasm. Structure of the FimD-C-F-G-H type 1 pilus assembly intermediate (PDB ID: 4J3O). The FimF-G-H pilus tip fiber is traversing the usher channel, with FimH exposed to the cell surface, and FimF bound by FimC located at the usher C domains. (F) Model for type 1 pilus assembly at the FimD usher. In its resting (apo) state, the FimD plug domain resides laterally within the usher channel (structure shown in B). The plug closes the usher channel and also functions to mask the C domains. Pilus assembly initiates with the binding of a FimC-H chaperone-adhesin complexes to the FimD N domain (step 1; structure shown in C). FimC-H binding to the N domain activates the usher by triggering opening of the plug domain and unmasking of the C domains. FimC-H then undergoes handoff from the N to the C domains, concomitant with insertion of the FimH lectin domain into the usher channel (step 2; structure shown in D). The usher N domain functions to recruit the next chaperone-subunit complex, FimC-G, from the periplasm (step 3). The FimC-G complex bound at usher N domain is perfectly positioned to undergo DSE with FimC-H bound at the C domains, forming the first link in the pilus fiber and displacing FimC from FimH. FimC-G is then handed off from the N to the C domains, concomitant with movement of the nascent pilus fiber through the usher channel toward the cell surface (step 4). Repeated cycles of chaperone-subunit recruitment and DSE (step 5) then result in assembly and secretion of the pilus tip (structure shown in E) and finally the pilus rod.
Each usher protomer comprises a 24-stranded β-barrel channel domain, a plug domain that serves as a channel gate, an N-terminal (N) periplasmic domain, and two C-terminal (C1 and C2) periplasmic domains (Fig. 4). In the resting (apo) usher, the plug domain closes the channel pore, preventing nonspecific influx or efflux of proteins (Fig. 4A and C). The usher’s N and C periplasmic domains serve as chaperone-subunit binding sites during pilus assembly. Chaperone-subunit complexes first bind to the N domain of the usher. A quality control mechanism ensures that “empty” chaperones do not bind the usher. Research on the Yersinia F1 capsule CU system revealed a conserved two-proline “lock,” in which the lock occludes the chaperone’s usher-binding interface (154). Only upon chaperone binding to a subunit is the lock rotated away from the usher-binding interface, allowing docking at the usher N domain.
The usher must undergo a global conformational change to initiate pilus assembly. In its resting state, the usher channel is occluded by the plug domain. In the type 1 pilus system, the usher can only be activated by binding of a FimC-H chaperone-adhesin complex to the usher N domain. Activation results in expulsion of the plug from the channel lumen to a periplasmic location, adjacent to the N domain (Fig. 4B and D) (151). During activation, the geometry of the usher channel changes from reniform to round, and the channel area increases by about 100 Å (Fig. 4) (151). Together, these changes allow the usher to accommodate folded pilus subunits within its channel. In the closed, apo state of the usher, the plug domain also functions to mask the usher C domains (Fig. 5F) (155). Therefore, plug expulsion not only exposes the channel lumen, but also frees the C domains and allows them to participate in chaperone-subunit binding.
Following activation, the chaperone-adhesin complex bound to the usher N domain must be “handed off” to the usher C domains (Fig. 5F). The usher N domain has lower affinity for the chaperone-adhesin complex compared with the C domains, and this handoff is likely driven by differential affinity (155). Additionally, the usher C2 domain may facilitate handoff by destabilizing chaperone-subunit binding to the N domain (156). Following handoff, the N domain is cleared of the chaperone-adhesin complex and available to recruit the next chaperone-subunit complex (FimC-G for type 1 pili) from the periplasm (Fig. 5D and F). Together with the plug domain (in the activated usher, the plug domain resides adjacent to the N domain), the N domain serves as a platform for the recruitment of additional chaperone-subunit complexes (151, 156). The newly recruited complex bound at the N domain is thought to be oriented perfectly for its Nte to undergo DSE with the previously recruited chaperone-subunit complex (FimC-H) bound at the C domains (Fig. 5F) (151). DSE displaces the chaperone from the subunit bound at the usher C domains, resulting in formation of the first link of the pilus fiber (FimG-H). The process continues with handoff of the chaperone-subunit complex bound at the N domain to the C domains, concomitant with translation of the nascent pilus fiber through the usher channel toward the cell surface (Fig. 5F). The usher N and C domains bind to the same surface of the chaperone; therefore, handoff requires rotation of the chaperone-subunit complex and pilus fiber (140, 150, 151). Computation analysis indicated that this rotation is facilitated by a helical low-energy pathway through the usher channel (117). Iterative rounds of chaperone-subunit recruitment to the usher N domain, DSE between subunits, and handoff from the usher N to the C domains drive pilus elongation and secretion, resulting in assembly of the tip fiber (Fig. 5E) and finally the rod. The usher channel is only large enough to accommodate a linear fiber of folded pilus subunits. Therefore, the pilus rod adopts its final helical quaternary conformation on the cell surface, upon exiting the usher channel.
The model for the subunit incorporation cycle at the usher as described above and in Fig. 5 shows that a monomeric usher is sufficient for pilus biogenesis, having sites for both recruitment and assembly of the pilus fiber (the N and C domains). However, multiple lines of evidence indicate that the usher is present in the OM as dimeric or higher-order oligomeric complexes (151, 152, 155, 157–159). How the usher oligomer contributes to pilus biogenesis remains to be determined. However, recent evidence indicates that individual usher protomers are capable of functioning in trans to assemble adhesive pili in vivo (155). The usher oligomer might function in an asymmetric manner during pilus assembly, with the individual protomers serving to recruit chaperone-subunit complexes to the OM, but only one usher providing the active translocation channel. Such a mechanism could facilitate pilus biogenesis by increasing the local concentration of chaperone-subunit complexes at the OM. This could be particularly important during assembly of the pilus rod, where thousands of copies of the major pilus subunit (FimA or PapA) must be incorporated.
Ordered Assembly of the Pilus Fiber
The order of assembly of pilus subunits is of great importance for pilus function to ensure proper presentation of the pilus adhesin. Regulation of the sequence of pilus assembly is due to (i) selective activation of the usher; (ii) differential affinities of chaperone-subunit complexes for the usher; (iii) the rate of DSE for each subunit-subunit pairing; and (iv) periplasmic concentrations of chaperone-subunit complexes. In its apo state, the usher C domains are masked and the usher is thought to sample periplasmic chaperone-subunit complexes via its N domain (155). Pilus assembly begins with the adhesin, and chaperone-adhesin complexes have highest affinity for the usher compared with the other chaperone-subunit complexes (156, 160–162). In the type 1 pilus system, activation of the usher (release of the plug domain and unmasking of the C domains) only occurs upon binding of a FimC-H chaperone-adhesin complex to the usher N domain. This provides a mechanism to ensure the assembly of functional pili, with the adhesin at the tip (155). Following activation, the recruitment platform formed by the usher N and plug domains likely facilitates ordered pilus assembly, because the N domain has differing affinities for each chaperone-subunit complex (156, 160–162). However, rates of DSE between subunit-subunit pairs is likely the determining factor in ordering of the pilus fiber. Not surprisingly, DSE is faster (by 2- to 50-fold) for cognate subunits (those that are adjacent in the mature pilus) than for noncognates. Additionally, the relative number of correctly versus incorrectly paired subunits is similar in catalyzed (with usher) and uncatalyzed (no usher) reactions (149, 163). This suggests that the basis for DSE rate differences lies in the specificity of a subunit’s Nte for the pilin domain of the neighboring subunit. Finally, a mathematical model incorporating both experimentally determined DSE rates and usher-binding affinities indicated that order accuracy was also highly dependent on the periplasmic concentration of subunits (163). For example, using the experimentally measured usher-binding affinities and DSE rates, a tenfold excess of FimC-G over FimC-F was needed to obtain a correctly sequenced type 1 pilus tip fiber. Notably, the native periplasmic concentrations of chaperone-subunit complexes have not been thoroughly investigated, and therefore, the extent to which this affects pilus biogenesis remains to be determined.
CHAPERONE/USHER PILUS REPERTOIRES IN E. COLI AND SALMONELLA
In this section, we present a survey of CU pili expressed by E. coli, followed by CU pili of Salmonella spp. Our overview highlights those pili that are assembled by the CU pathway, are fairly well characterized, and have been shown to be involved in human pathology. Detailed descriptions are provided for the type 1 and P pili expressed by UPEC, because these CU pili have been extensively studied. Table 1 provides a list of CU pili present in E. coli and Salmonella.
E. coli CU Pili
CU pili and their component proteins facilitate E. coli pathogenesis through a variety of mechanisms. The prototypical type 1 and P pili are found in UPEC and mediate adhesion to the urinary tract. Type 1 pili are widely present among other E. coli strains as well, and contribute to biofilm formation and colonization of different sites within the host. E. coli strains such as UPEC and ETEC express a variety of additional CU pili to promote adherence, biofilm formation, and other functions. Unlike ETEC and UPEC, which express multiple pili when cultured in vitro, isolates of enterohemorrhagic E. coli (EHEC) express only a few surface structures detectable by electron microscopy upon laboratory culture (164–168). However, whole-genome sequencing of two E. coli O157:H7 isolates revealed the presence of at least 17 pilus gene clusters (47, 169, 170). The construction of transcriptional fusions with 13 of these gene clusters showed that most of these genes are not expressed in vitro (171). Sequence analysis also indicates that some O157:H7 isolates are unable to produce functional type 1 pili (111, 172). Nevertheless, analyses of cloned operons and phenotypes resulting from mutational inactivation demonstrate that many of the EHEC CU loci encode functional pili (173–177). These studies reinforce the concept that E. coli strains express multiple CU pili, contributing to the diverse capabilities of these bacteria to recognize different receptors and cause different pathologies.
Type 1 pili
The chromosomal fimBEAICDFGH gene cluster coding for type 1 pili (Fig. 2) is widely distributed among E. coli strains, including nonpathogenic and laboratory strains. As discussed in “Architecture of Chaperone/Usher Pili” above, FimA is the major pilin subunit that forms the helical pilus rod proximal to the cell surface, FimF and FimG are adapter subunits present in the distal tip fibrillum, and FimH is the tip-associated adhesin (Fig. 1). FimC is the pilus chaperone and FimD the usher protein for type 1 pili. FimI’s function is unknown, although it may be the rod-terminating subunit, analogous to PapH of P pili (93, 94). FimB and FimE are regulatory proteins involved in phase variation of type 1 pilus expression, as described in “Chaperone/Usher Genetic Loci and Regulation.”
Type 1 pili mediate binding to a variety of surfaces and host tissues in a mannose-sensitive manner. Type 1 pili are a major virulence factor of UPEC and facilitate adhesion to the bladder via specific binding to uroplakins and other mannosylated proteins that coat the bladder lumen (178, 179). Antibodies against the type 1 pilus adhesin, FimH, confer protection against UTI in both nonhuman primates and mice (180, 181). However, a definitive requirement for type 1 pili in human UTI remains elusive, likely because of the many adhesins and other virulence factors with redundant or overlapping functions in uropathogenic strains (182). Type 1 pili bind to a variety of receptors and surfaces in addition to urothelial cells, including surface glycoproteins of immune cells, extracellular matrix proteins, and abiotic surfaces (121, 183–185).
Type 1 pili perform a variety of roles in addition to initiating host cell contact and facilitating colonization. The steps by which E. coli infects the urinary tract are illustrative of these pilus-mediated functions. Introduction of UPEC into the bladder typically occurs via ascension of the urethra by fecal bacteria that contaminate the periurethral area. UPEC bind to the bladder surface via interaction of the type 1 pilus adhesin FimH with mannosylated receptors such as the uroplakins. This allows the bacteria to gain a foothold in the urinary tract and resist eradication by urine flow (178, 186). In the mouse UTI model, the bladder epithelium rapidly responds to type 1 pilus-mediated binding of E. coli through the activation of NF-κB. This upregulation occurs specifically in the terminally differentiated superficial umbrella cells of the bladder and is dependent on a dual ligand/receptor interaction: FimH and the uroplakin Ia receptor; and lipopolysaccharide and Toll-like receptor 4 (TLR4) (187).
Bacterial binding via type 1 pili also activates host cell pathways that lead to actin cytoskeletal rearrangements in the urothelium, and subsequent bacterial invasion into the host cells via a zipper-like mechanism (186, 188). Bacterial uptake by bladder epithelial cells is initiated by FimH binding to host β1- and α3-integrins (179, 188). Type 1 pili also promote adhesion to phagocytes and survival inside macrophages (185, 189–191). In addition to type 1 pilus functions mediated by the FimH adhesin, the FimA major rod subunit acts to potently inhibit Bax-mediated cytochrome C release in infected cells, thereby suppressing host cell apoptosis (192). FimA does this by binding to VDAC1 (voltage-dependent anion channel-1) and stabilizing the mitochondrial VDAC1-hexokinase complex, preventing its dissociation in response to apoptotic stimuli (192). This effect may be mediated by a soluble version of the FimA monomer, resulting from intramolecular self-DSC (193).
Following invasion, UPEC replicate within the superficial umbrella cells of the bladder and form intracellular biofilm-like communities (IBC). Type 1 pili continue to be expressed by the intracellular bacteria and are important in IBC formation, illustrating a function for pili that is distinct from initial binding and invasion (194). Bacteria expressing mutants of FimH that are functional for mannose binding, but defective for IBC formation, are attenuated for pathogenesis and rapidly cleared in the mouse UTI model, underscoring the distinct functions of type 1 pili (195). Type 1 pili also contribute to extracellular biofilm production by E. coli in a variety of environments (183).
Bladder cells respond to UPEC binding and invasion by sloughing off into the bladder lumen, a mechanism used by the host to eradicate the bacteria (186). This sloughing off is counteracted by UPEC fluxing out of the host cells and undergoing further rounds of binding to the bladder epithelium and invasion, in a process likely mediated by type 1 pili as in the initial round of infection (196, 197). Through additional iterations of this process, UPEC may gain access to underlying layers of the bladder epithelium, wherein the bacteria survive in quiescent reservoirs. The bacteria may then reseed the bladder lumen from these reservoirs and initiate a subsequent infection. Such reseeding of the bladder from these quiescent reservoirs likely accounts for a subset of recurrent UTI cases, which will also depend on type 1 pili for host cell adhesion and invasion (196, 198).
P pili
P pili are encoded by the chromosomal pap (pyelonephritis-associated pili) gene cluster papIBAHCDJKEFG (Fig. 2). The pap gene cluster resides on pathogenicity islands of UPEC strains (the cluster is absent in laboratory and many nonpathogenic strains), and is also found in E. coli strains that cause neonatal meningitis (199). More than one pap gene cluster may be present in an individual E. coli strain. The papI and papB genes form the upstream regulatory region, as described in “Chaperone/Usher Genetic Loci and Regulation.” PapA is the major pilus subunit that forms the helical pilus rod; PapK, E, and F are adapter subunits that form the distal P pilus tip fiber; and PapG is the tip-associated adhesin (Fig. 1). PapD is the chaperone and PapC is the usher for P pili. PapH is the rod-terminator subunit that determines pilus length and anchors the pilus fiber in the OM (92, 142). PapJ’s function is unknown, although it may be involved in ensuring the integrity of the pilus fiber during assembly (200).
P pili are associated with the ability of UPEC to colonize the kidney and cause pyelonephritis (112, 201, 202). P pili bind to Gal(α1-4)Gal moieties present on the globoseries of glycolipids of kidney epithelial cells. Binding of PapG to kidney epithelial cells facilitates attachment of UPEC during an ascending UTI, and is one of the initiating steps of pyelonephritis. The glycolipid receptor is also present on the P blood group antigen, allowing P pilus-mediated agglutination of human erythrocytes (and the basis of the useful laboratory hemagglutination assay to quantitate adhesive pilus assembly) in a mannose-resistant manner (10). There are three variants of the PapG adhesin: PapGI, GII, and GIII (203, 204). Each of these recognizes distinct isoreceptors, differing in carbohydrate residues distal to the Gal(α1-4)Gal core. PapGI, GII, and PapGIII bind to globosides GbO3, GbO4, and GbO5, respectively. PapGII binds preferentially to the common kidney glycolipid isoreceptor, globoside (GbO4; tetrasaccharide GalNAcβ1-3Galα1-4Galβ1-4Glc linked to ceramide), and is correlated with acute pyelonephritis in humans. In contrast, PapGIII is correlated with human bladder infection.
In support of the role of P pili in UPEC pathogenesis, vaccination of both mice and primates with P pili results in protection against pyelonephritis (205, 206). However, as found for type 1 pili, mutagenesis studies have had mixed results in establishing P pili as essential for bacterial infection (207). P-pilus-mediated adhesion of E. coli to the urinary tract stimulates cytokine production and release, and thus induces inflammatory responses that likely exacerbate kidney damage in acute pyelonephritis (208–210). P pilus binding to its globoside receptor also induces release of the second messenger ceramide. Ceramide is the membrane-anchoring portion of the receptor and a TLR4 agonist. In turn, TLR4 activation leads to the production of proinflammatory cytokines (IL-6) and chemokines (CXCL-8), and neutrophil recruitment (211, 212). Additionally, through TLR4 interaction, P pilus binding modulates the local secretory antibody immune response by downregulating polymeric immunoglobulin receptor (PIGR) expression, thereby impairing immunoglobulin A transport to the kidney lumen (213). Thus, UPEC evades this host-protective mechanism, facilitating the establishment of an ascending infection (43, 213). These findings demonstrate a link between bacterial adhesion and the induction and modulation of host innate immune pathways. UPEC binding via PapG also activates signaling pathways within the bacteria that lead to upregulation of iron acquisition processes and machinery, and may prepare UPEC for urinary tract colonization (214).
Aggregative adherence fimbriae
Aggregative adherence fimbriae (AAF) are a family of pili expressed by enteroaggregative E. coli (EAEC). Genes clusters coding for AAF are present on large plasmids in EAEC strains and comprise aafD (chaperone), aafC (usher), aafB (minor pilin subunit), and aafA (major pilin subunit) (90, 215–218). AAF belong to the FGL subfamily of CU pathways, and assemble as thin (2- to 3-nm diameter) fibers with variable morphology. The AAF/I and AAF/II pilus fibers are somewhat flexible and form bundles, wrapping around neighboring bacterial cells, whereas AAF/III and AAF/IV pilus fibers are longer and more flexible (217, 219–221).
EAEC use AAF to promote binding to the human intestinal wall (215, 222, 223). AAF bind to several different epithelial cell receptors, including decay-accelerating factor (DAF) and extracellular matrix proteins (90, 224). Once bound to the intestine, EAEC elaborate toxins, which in turn cause induction of proinflammatory cytokines from the infected epithelium and a watery diarrheal illness (225, 226).
Afa/Dr pili
Both uropathogenic and diarrhea-associated E. coli strains use afa gene clusters to encode afimbrial adhesins (95–98). Based on the commonality of binding to the decay-accelerating factor (DAF, CD55) as a pilus receptor, the Afa/Dr family includes Afa, Dr, and F1845 pili, among others (90, 91). The genes coding for the structural and assembly components of Afa/Dr pili are organizationally well conserved and ordered as follows: afaABCDE. The afaA gene encodes a regulatory protein, afaB the pilus chaperone, afaC the usher, afaD the tip-located invasin subunit, and afaE the major subunit protein. AfaE, which also functions as the pilus polyadhesin, is the least conserved in sequence among family members, accounting for receptor-binding variability. UPEC code for Dr adhesins using the draABCDE gene cluster (Fig. 2) (227–229). F1845 pili facilitate diffuse cell adherence of E. coli diarrheal isolates, and have a genetic organization similar to both afa and dra. While the genes encoding F1845 pili were first cloned from the bacterial chromosome, hybridization studies revealed that similar coding regions may be plasmid-associated in other strains (230, 231).
Afa/Dr pili belong to the FGL subfamily of CU pathways. The pili assemble as flexible, polyadhesive fibers (Fig. 1), which confer mannose-resistant hemagglutination (95, 232–235). Afa/Dr pili enable binding to the Dra blood-group antigen present on CD55/DAF, a signaling molecule involved in complement regulation (233). CD55/DAF prevents amplification of the complement cascade, via direct binding of C3b or C4b (236, 237). The Afa/Dr adhesins also bind carcinoembryonic antigen (CEA)-related cellular adhesion molecules (CEACAM). Pathogen binding to this receptor is followed by a CEACAM-associated signaling cascade, which triggers a remodeling of the host cell surface that assists bacterial colonization of the urinary tract (238, 239).
The different binding tendencies of the Afa/Dr pili give rise to different pathologies. There are correlations between Afa/Dr-expressing strains of E. coli and low birth weight and preterm delivery in mice, as well as gestational pyelonephritis (240, 241). Additionally, there is a correlation between Afa/Dr-encoding strains and increased colonization and invasive capacity (242). Afa/Dr pili have also been implicated in inducing proinflammatory responses in host cells. E. coli strains expressing Afa/Dr pili promote the basolateral secretion of IL-8 through the activation of mitogen-activated protein kinases (MAPK) (243). IL-8 then induces transmigration of neutrophils across the epithelial monolayer, which in turn induces tumor necrosis factor alpha and IL-1β, upregulating DAF and thereby increasing Afa/Dr adhesion (244). Additionally, Afa/Dr diffusely adhering E. coli induce activation of CD55/DAF signaling in an adhesin-dependent manner, which in turn activates vascular endothelial growth factor (VEGF). Elevated VEGF promotes angiogenesis in inflammatory bowel diseases, including Crohn’s disease and ulcerative colitis, providing a link between pilus-mediated bacterial adhesion and angiogenesis in disease (245).
F1C pili
F1C pili are nonhemagglutinating adherence factors present in approximately 14% of UPEC isolates and 7% of E. coli fecal isolates (246–250). Eight chromosomally encoded genes, termed foc, are necessary for F1C pilus biogenesis, and have a nearly identical organization to the type 1 pilus fim operon. F1C pili are similar in architecture to type 1 pili, with dimensions of ∼1 μm in length and 7 nm in diameter (251). FocA is the major pilus subunit, FocF and FocG are the minor pilus subunits, and FocH is the tip-located adhesin. FocC is the pilus chaperone and FocD the usher (251, 252). F1C pili adhere to the distal tubules and collecting ducts of the kidney, and to vascular endothelial cells of the kidney and bladder cells (248, 250, 253, 254). F1C pili have also been implicated in biofilm formation, and in induction of the proinflammatory cytokine IL-8 on renal epithelial cell attachment (255, 256).
S pili
S pili are virulence-associated adhesive organelles present in extraintestinal pathogenic E. coli (ExPEC) strains that cause urinary tract infections and neonatal meningitis (250, 257–259). The genes coding for S pili were first isolated from a UPEC strain (260). The sfa gene cluster is chromosomal and consists of nine genes (261). Similar to type 1 and P pili, S pili contain a proximal helical rod (7 nm in diameter) and a distal linear tip fiber. The rod is composed of the major subunit SfaA. The tip, similar to P pili, is relatively long and flexible, and contains the sialic acid-binding adhesin SfaS at its distal end (248, 262). The usher in the S pilus assembly system is SfaE, and the chaperone is SfaF (21, 263, 264).
S pili bind to sialic acid-containing oligosaccharides, facilitating the recognition of and binding to the urinary tract and also to brain microvascular endothelial cells (265, 266). S pili have also been implicated in binding to erythrocytes, human fibronectin, and the extracellular matrix protein laminin (267, 268). Furthermore, S pili immobilize human plasminogen and tissue-type plasminogen activator (tPA), thereby activating plasminogen and promoting plasmin proteolytic activity on the bacterial cell surface (269).
CS1, CFA/I, and CFA/II pili
ETEC express a large group of pili, termed colonization factor antigens (CFAs) or coli surface antigen (CS), which are part of the alternate family of CU pili that belong to the σ clade (29). The coli surface antigen 1 (CS1) and CFA/I pili are representative members of this CU family. CS1 pili are encoded by the cooBACD operon, which is present on the large pCoo plasmid found in ETEC strains (270). Expression of this operon requires the Rns transcriptional activator, which belongs to the AraC DNA-binding protein family, and is encoded on a separate plasmid (271, 272). In the absence of Rns, coo operon expression is repressed by the global regulatory protein H-NS (272). CFA pili are coded for by the cfaABCE gene cluster, present on the NTP113 ETEC plasmid (273, 274). Expression of these pili is regulated by the CfaD-positive regulator, present on the same plasmid, and homologous to Rns (275). CFA/I pili are expressed at 37°C, but remain repressed at lower temperatures, in a process mediated by H-NS, similar to the CS pili (276).
CS1 and CFA pili are large and rigid: approximately 6 to 7 nm in diameter (277). The pili are composed of two structural subunits and lack a distinct tip structure (31, 278). CFA and CS pilus fibers are composed of repeating copies of the major pilus subunit, cooA for CS1 and CfaB for CFA/I pili. CooD and CfaE are the minor subunits for CS1 and CFA/1, respectively (279, 280). Both the major and minor subunits may contribute to adhesion, with each appearing to possess distinct binding properties (31, 281–284). In CS1 pili, CooB serves as the periplasmic chaperone and CooC serves as the OM usher. In CFA/1 pili, CfaA and CfaC are the chaperone and usher, respectively.
ETEC colonizes the human small intestine and secretes enterotoxins into the gut lumen. This causes a disruption in fluid homeostasis, which leads to acute gastroenteritis (285, 286). ETEC is the predominant causative agent of gastroenteritis, otherwise known as traveler’s diarrhea, and also is a major cause of diarrhea in developing nations, particularly among children (287, 288). The CS and CFA pili, which primarily mediate adhesion to the small intestine, are critical for ETEC virulence (277).
Salmonella CU Pili
The CU pili of Salmonella are less well characterized than the E. coli pili, in part because of a general lack of pilus expression by strains grown under laboratory culture conditions (50). The S. enterica serotype Typhimurium genome contains 13 gene clusters potentially coding for surface structures—fim, agf (csg), pef, lpf, saf, bcf, stb, stc, std, sth, stf, sti, and stj (17). However, among the CU pathways, evidence for in vitro expression of the various pili is limited (50, 289). Nevertheless, findings from studies on pilus expression in vivo and on the role of pilus gene clusters during colonization of the mouse intestine suggest that most of the Salmonella CU genetic loci encode functional pili (50, 289–292).
Long polar fimbriae
The long polar fimbriae (LPF) locus consists of chromosomal genes lpfABCDE and was identified during a screen for S. Typhimurium loci not present in related Enterobacteriaceae (293, 294). Based on the homology of flanking chromosomal regions to E. coli, the lpf genes likely were acquired by horizontal gene transfer. The lpf operon has a similar organization to that of the type 1 pilus fim operon, and the Lpf protein sequences are homologous to the Fim proteins (293, 295). LpfD is the adhesin, LpfA and LpfE encode subunit proteins, LpfB serves as the pilus chaperone, and LpfC is the OM usher (21, 293). LPF facilitate bacterial adhesion to M cells of Peyer’s patches in the small intestine, as determined using a mouse model of infection (290). Deletion of the lpf locus has only a minor effect on the virulence of S. Typhimurium in the mouse infection model, but a double-deletion mutant of lpf together with the invA type III secretion system (T3SS) component produces a synergistic effect (296), suggesting that LPF assist in bacterial docking to host cells to facilitate delivery of effector proteins via the T3SS injectisome. Phase-variable lpf expression has been demonstrated, with LPF phase-ON bacteria appearing more frequently in the mouse Peyer’s patches than phase-OFF bacteria (297). Such a differential was not seen in bacteria isolated from the mesenteric lymph nodes or spleens of the infected animals. Additionally, phase-ON variants were strongly selected against upon immunization of mice with the LpfA major fimbrial subunit protein (297). These findings support a model wherein LPF are expressed in vivo and are important for an early adhesive step of infection via the oral route, but not during later stages of infection when the bacteria have disseminated to systemic sites.
Plasmid-encoded fimbriae
Plasmid-encoded fimbriae (PEF) are expressed from the plasmid-based pefBACD locus. The pef gene cluster is responsible for formation of surface fibers in both E. coli and S. Typhimurium (298). PefD is the chaperone, PefC the usher, PefA the major structural pilus subunit, and PefB encodes a regulatory protein. PefA polymerizes as a thin and flexible fiber, with a diameter of 2 to 4 nm. PEF promote adhesion of S. Typhimurium to the villous small intestine, and are necessary for infection-associated pathology such as fluid accumulation in an infant mouse model (299). Additionally, PEF promote bacterial adherence to various human cell lines, including HEp-2, T-84, Int-407, and HeLa cells (300). The pef genes are only found in a subset of phylogenetically related strains, and were likely recently acquired by Salmonella (301, 302). The acquisition of pilus loci such as pef may provide an evolutionary mechanism for Salmonella serovars and other bacteria to expand their host range.
Salmonella enteritidis fimbriae
The S. enteritidis fimbriae (SEF) are encoded by the chromosomal sefABCD locus. SefA and SefD are subunit proteins, SefB is the periplasmic chaperone, and SefC the OM usher (303, 304). SefC belongs to the FGL subfamily of CU pilus chaperones, and assembles two distinct, thin, polyadhesive fibers composed of either SefA or SefD, termed SEF14 or SEF18, respectively. Thus, the SefB and SefC chaperone-usher pair assembles homopolymers of either SefA or SefD. As with the pef locus, the sefABCD genes were likely acquired recently by Salmonella, and are expressed only by S. enteritidis and closely related serovars (301, 302). SEF are expressed in vivo during infection, but the exact contributions of these pili to virulence are not well characterized (304, 305). Evidence suggests that SefD may function as an adhesin for S. enteritidis that facilitates invasion inside host macrophages, and that the pili may be important for invasive stages of infection following the initial intestinal colonization (305).
Type 1 pili
The type 1 pili of Salmonella are encoded by a chromosomal gene cluster containing six genes encoding structural and assembly components (fimAICDHF) together with three separately transcribed regulatory genes (fimZYW) (306, 307). These pili exhibit mannose-sensitive hemagglutination, similar to the type 1 pili of E. coli (308). The Fim proteins of Salmonella exhibit a corresponding functional relationship to their counterparts in E. coli type 1 pili, but the Salmonella fim genes and gene products exhibit little sequence similarity with the E. coli genes. FimA is the major subunit, FimF the adaptor subunit, and FimH the tip-located adhesin. FimC is the periplasmic chaperone and FimD the OM usher. FimI has an unknown role, as a fimI deletion mutant resulted in an eightfold higher hemagglutination titer, but morphologically normal pili (306). FimI in Salmonella does not appear to be related to the E. coli FimI protein (94). Type 1 pili play a role in the initial stages of Salmonella infections in a variety of animal models including mice, chicks, and hens (309–313). Type 1 pili have also been shown to limit bacterial dissemination and colonization of mice, as well as induce intestinal inflammation during Salmonella invasion (314).
CU PILI AS THERAPEUTIC TARGETS
Pili function as virulence factors for many bacterial pathogens (1, 315). Pilus-mediated adhesion is critical for the early stages of infection, allowing bacteria to establish a foothold within the host and dictating tropism toward specific tissues. Following bacterial attachment, pili also modulate host cell signaling pathways, promote or inhibit invasion inside host cells, and function in bacterial-bacterial interactions leading to the formation of community structures such as biofilms. Given their varied and central functions in pathogenesis, pili make attractive therapeutic targets (316, 317). Knowledge gained from basic science efforts, as described in the preceding sections, has generated a comprehensive understanding of the molecular biology of CU pilus assembly and adhesion. This detailed mechanistic information, in turn, has fueled translational efforts to develop pilus-focused therapeutics, such as vaccines and small-molecule inhibitors, with the goal of disrupting pilus assembly or function, and thereby treating or preventing disease caused by Gram-negative bacterial pathogens.
Therapeutics directed against virulence factors such as pili represent an alternative to traditional antibiotics. Traditional, broad-spectrum antibiotics interfere with essential biological processes, and indiscriminately target the beneficial host flora along with bacterial pathogens. In contrast, antivirulence therapeutics are designed to inhibit systems only required for pathogens to cause disease within the host (318–320). Therefore, antivirulence therapeutics should spare the commensal bacteria and also reduce the evolutionary pressure leading to antimicrobial resistance. Rates of antibiotic resistance have reached alarming levels, raising the possibility of a postantibiotic era, in which even common infections could become life threatening (321–324). This looming health crisis is compounded by the limited availability of new antibiotics and the rapid emergence of resistance once new antibiotics are introduced into use. Antivirulence therapeutics directed against targets such as bacterial pili represent one approach to address this urgent situation.
In this section, we describe strategies being taken for the therapeutic targeting of CU pili, using the UPEC type 1 and P pili as models. Effective strategies against UPEC type 1 and P pili will likely also be relevant to the many other pili assembled by the CU pathway, due to conservation of the CU assembly mechanism, pilus structure, and pilus function.
Vaccination Approaches for CU Pili
The location of pilus fibers on the bacterial cell surface exposes them to the immune system, making pili ideal targets for vaccination. However, efforts to develop vaccines using whole pili have generally been unsuccessful, because of factors such as phase-variable expression and antigenic variation. In addition, vaccination with whole pili may not generate antibodies that inhibit pilus function, particularly for pili such as the UPEC type 1 and P pili, where the adhesin subunit is present at only one copy per pilus and the immune response is biased toward the main structural subunits instead. One promising approach to address this issue is to vaccinate with the purified adhesin subunit, rather than using whole pili. In both murine and primate cystitis models, substantial protection against UPEC infection was conferred on systemic vaccination with the FimH (type 1 pili) or PapG (P pili) adhesins (180, 181, 325, 326). Similarly, a truncated FimH construct containing only the adhesin domain conferred protection against cystitis in mice vaccinated intramuscularly or intranasally, using CpG oligonucleotides as an adjuvant (327). The effectiveness of vaccination with FimH was largely due to the generation of antibodies that blocked FimH-mediated bladder colonization. However, in some cases, rather than blocking type 1 pilus-mediated adhesion, the host antibody response may instead enhance binding of FimH to its receptor by stabilizing the adhesin’s high-affinity binding state (328). An improved understanding of the catch-bond mechanism of FimH binding, and pilus adhesion under fluid flow conditions in the urinary tract, may allow tailoring of the antigen used for vaccination to generate a desired, antiadhesive immune response. Importantly, vaccination against FimH does not appear to alter the commensal E. coli in the gut (180).
Small-Molecule Inhibitors of CU Pili
An alternative approach to vaccination is to target the assembly or function of CU pili through the use of small-molecule inhibitors. Several approaches are being taken to develop such inhibitors for UPEC and other pathogens that express CU pili, taking advantage of the extensive knowledge of the pilus assembly pathway and pilus-mediated adhesion mechanisms.
One approach is to develop small molecules that competitively inhibit adhesin-receptor interactions. For the FimH adhesin of type 1 pili, which bind to mannosylated proteins, soluble receptor analogs termed mannosides are being pursued (329, 330). Mannosides bind to and occupy the FimH receptor-binding site, and act as antiadhesives by preventing bacterial colonization of the urinary tract. Indeed, analysis of a mouse model showed that such molecules act prophylactically to prevent bacterial invasion of the bladder following urinary tract infection (329). Studies have indicated that mannosides are orally bioavailable and are also effective against an established UTI, as well as against catheter-associated UTI (329–332). Mannosides have been shown to act in synergy with traditional antibiotic treatments in their capacity to reduce urinary tract UPEC titers in mice (332). Galabiose-based inhibitors are similarly being developed to inhibit bacterial adhesion mediated by P pili (333).
A second approach toward small-molecule anti-pilus therapeutics is to develop inhibitors that target the pilus assembly and secretion process, with the goal of preventing pilus fibers from being expressed on the bacterial surface. One such class of small-molecule CU pilus inhibitors, known as pilicides, consists of molecules with a 2-pyridine scaffold (334, 335). Pilicides bind to the periplasmic chaperone and interfere with chaperone-subunit interactions or the binding of chaperone-subunit complexes to the usher. In vitro studies have demonstrated that such molecules attenuate type 1 and P pilus biogenesis and pilus-mediated adhesion and biofilm formation (334, 336). The mechanism of action and effects of pilicides may be broader than initially envisioned. In a recent study, pilicide ec240 was found to disrupt type 1, P, and S pili, and to also affect flagellar motility (337). The ec240 pilicide inhibited pilus expression as well as interfered with pilus assembly. Ec240 caused downregulation of the type 1 pilus genes by switching fimS to the phase-OFF position (337). Treatment of bacteria with ec240 also increased levels of the S and P pili regulatory proteins SfaB and PapB, which both further promote fim phase-OFF variation (337). Thus, pilicides may also work by influencing pilus phase variation and thus pilus expression, in addition to disrupting protein-protein interactions required for pilus assembly. Modified pilicides have also been developed that have activity against curli in addition to type 1 pili (338). Treatment of UPEC with one such pilicide reduced biofilm formation and attenuated bacterial colonization in the mouse UTI model (338).
Additional targets in the CU pathway are being exploited for therapeutic intervention. Small-molecule “pilicides” have been developed with alternate mechanisms of action compared to the originally developed pilicides. Lo et al. used computational screening to identify a small molecule, AL1, that inhibits polymerization of the type 1 pilus fiber by disrupting the DSE reaction between the FimH and FimG subunits (339). The AL1 compound disrupted type 1 pilus biogenesis by UPEC, and reduced biofilm formation and bacterial adhesion to human bladder epithelial cells. Another small-molecule compound, nitazoxanide (NTZ), was shown to inhibit biofilm formation by EAEC, by preventing assembly of the AAF CU pili (340). Further analysis demonstrated that NTZ inhibits assembly of type 1 and P pili as well, and that the drug acts through a novel mechanism of action to prevent proper folding of the usher protein in the bacterial OM (341). Coilicides are a recently developed class of anti-pilus inhibitors (342). Coilicides act by preventing uncoiling and recoiling of the helical pilus rod. As discussed in “Architecture of Chaperone/Usher Pili,” the compliance of the pilus rod is critical for a redistribution of forces during shear stress such as caused by urine flow. In a proof-of-concept study, the PapD chaperone was shown to bind to an uncoiled P pilus rod, preventing recoiling of the rod and thus locking it in a noncompliant form (342). Only a few molecules of PapD per pilus rod (>1000 subunits) were necessary for coilicide activity, thus making recoil attenuation a viable opportunity for therapeutic intervention.
The examples described in this section highlight the potential for a new class of antimicrobial therapeutic agents that target the assembly and function of CU pili. Such agents have the potential to alleviate disease by disrupting critical host-pathogen interactions, while sparing the beneficial bacterial flora of the host and addressing the growing health threat of antibiotic resistance.
ACKNOWLEDGMENTS
G.T.W. is supported by National Research Service Award F30AI112252 and Medical Scientist Training Program Award T32GM008444 from the U.S. National Institutes of Health (NIH). Research in the Thanassi laboratory on the subject of this review is supported by NIH grants R01GM62987 and R21AI121639.
The authors declare no conflicts.
REFERENCES
- 1.Thanassi DG, Bliska JB, Christie PJ. 2012. Surface organelles assembled by secretion systems of Gram-negative bacteria: diversity in structure and function. FEMS Microbiol Rev 36:1046–1082. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Costa TR, Felisberto-Rodrigues C, Meir A, Prevost MS, Redzej A, Trokter M, Waksman G. 2015. Secretion systems in Gram-negative bacteria: structural and mechanistic insights. Nat Rev Microbiol 13:343–359. [DOI] [PubMed] [Google Scholar]
- 3.Proft T, Baker EN. 2009. Pili in Gram-negative and Gram-positive bacteria - structure, assembly and their role in disease. Cell Mol Life Sci 66:613–635. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Ottow JC. 1975. Ecology, physiology, and genetics of fimbriae and pili. Annu Rev Microbiol 29:79–108. [DOI] [PubMed] [Google Scholar]
- 5.Duguid JP, Smith IW, Dempster G, Edmunds PN. 1955. Non-flagellar filamentous appendages (fimbriae) and haemagglutinating activity in Bacterium coli. J Pathol Bacteriol 70:335–348. [DOI] [PubMed] [Google Scholar]
- 6.Brinton CC Jr. 1959. Non-flagellar appendages of bacteria. Nature 183:782–786. [DOI] [PubMed] [Google Scholar]
- 7.Berry JL, Pelicic V. 2015. Exceptionally widespread nanomachines composed of type IV pilins: the prokaryotic Swiss Army knives. FEMS Microbiol Rev 39:134–154. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Brinton CC Jr. 1965. The structure, function, synthesis and genetic control of bacterial pili and a molecular model for DNA and RNA transport in gram negative bacteria. Trans N Y Acad Sci 27(8 Series II):1003–1054. [DOI] [PubMed] [Google Scholar]
- 9.Duguid JP. 1959. Fimbriae and adhesive properties in Klebsiella strains. J Gen Microbiol 21:271–286. [DOI] [PubMed] [Google Scholar]
- 10.Kallenius G, Mollby R, Svenson SB, Windberg J, Lundblud A, Svenson S, Cedergen B. 1980. The Pk antigen as receptor for the haemagglutinin of pyelonephritogenic Escherichia coli. FEMS Microbiol Lett 7:297–302. [Google Scholar]
- 11.Leffler H, Svanborg-Edén C. 1981. Glycolipid receptors for uropathogenic Escherichia coli on human erythrocytes and uroepithelial cells. Infect Immun 34:920–929. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Nowicki B, Labigne A, Moseley S, Hull R, Hull S, Moulds J. 1990. The Dr hemagglutinin, afimbrial adhesins AFA-I and AFA-III, and F1845 fimbriae of uropathogenic and diarrhea-associated Escherichia coli belong to a family of hemagglutinins with Dr receptor recognition. Infect Immun 58:279–281. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Nowicki B, Hart A, Coyne KE, Lublin DM, Nowicki S. 1993. Short consensus repeat-3 domain of recombinant decay-accelerating factor is recognized by Escherichia coli recombinant Dr adhesin in a model of a cell-cell interaction. J Exp Med 178:2115–2121. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Duguid JP, Anderson ES, Campbell I. 1966. Fimbriae and adhesive properties in Salmonellae. J Pathol Bacteriol 92:107–138. [DOI] [PubMed] [Google Scholar]
- 15.Kisiela D, Sapeta A, Kuczkowski M, Stefaniak T, Wieliczko A, Ugorski M. 2005. Characterization of FimH adhesins expressed by Salmonella enterica serovar Gallinarum biovars Gallinarum and Pullorum: reconstitution of mannose-binding properties by single amino acid substitution. Infect Immun 73:6187–6190. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Blattner FR, Plunkett G III, Bloch CA, Perna NT, Burland V, Riley M, Collado-Vides J, Glasner JD, Rode CK, Mayhew GF, Gregor J, Davis NW, Kirkpatrick HA, Goeden MA, Rose DJ, Mau B, Shao Y. 1997. The complete genome sequence of Escherichia coli K-12. Science 277:1453–1462. [DOI] [PubMed] [Google Scholar]
- 17.McClelland M, Sanderson KE, Spieth J, Clifton SW, Latreille P, Courtney L, Porwollik S, Ali J, Dante M, Du F, Hou S, Layman D, Leonard S, Nguyen C, Scott K, Holmes A, Grewal N, Mulvaney E, Ryan E, Sun H, Florea L, Miller W, Stoneking T, Nhan M, Waterston R, Wilson RK. 2001. Complete genome sequence of Salmonella enterica serovar Typhimurium LT2. Nature 413:852–856. [DOI] [PubMed] [Google Scholar]
- 18.Orskov I, Orskov F. 1990. Serologic classification of fimbriae. Curr Top Microbiol Immunol 151:71–90. [PubMed] [Google Scholar]
- 19.Piatek R, Zalewska B, Bury K, Kur J. 2005. The chaperone-usher pathway of bacterial adhesin biogenesis -- from molecular mechanism to strategies of anti-bacterial prevention and modern vaccine design. Acta Biochim Pol 52:639–646. [PubMed] [Google Scholar]
- 20.Zavialov A, Zav’yalova G, Korpela T, Zav’yalov V. 2007. FGL chaperone-assembled fimbrial polyadhesins: anti-immune armament of Gram-negative bacterial pathogens. FEMS Microbiol Rev 31:478–514. [DOI] [PubMed] [Google Scholar]
- 21.Zav’yalov V, Zavialov A, Zav’yalova G, Korpela T. 2010. Adhesive organelles of Gram-negative pathogens assembled with the classical chaperone/usher machinery: structure and function from a clinical standpoint. FEMS Microbiol Rev 34:317–378. [DOI] [PubMed] [Google Scholar]
- 22.Ilangovan A, Connery S, Waksman G. 2015. Structural biology of the Gram-negative bacterial conjugation systems. Trends Microbiol 23:301–310. [DOI] [PubMed] [Google Scholar]
- 23.Evans ML, Chapman MR. 2014. Curli biogenesis: order out of disorder. Biochim Biophys Acta 1843:1551–1558. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Craig L, Li J. 2008. Type IV pili: paradoxes in form and function. Curr Opin Struct Biol 18:267–277. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Pelicic V. 2008. Type IV pili: e pluribus unum? Mol Microbiol 68:827–837. [DOI] [PubMed] [Google Scholar]
- 26.Arutyunov D, Frost LS. 2013. F conjugation: back to the beginning. Plasmid 70:18–32. [DOI] [PubMed] [Google Scholar]
- 27.Thanassi DG, Saulino ET, Hultgren SJ. 1998. The chaperone/usher pathway: a major terminal branch of the general secretory pathway. Curr Opin Microbiol 1:223–231. [DOI] [PubMed] [Google Scholar]
- 28.Geibel S, Waksman G. 2014. The molecular dissection of the chaperone-usher pathway. Biochim Biophys Acta 1843:1559–1567. [DOI] [PubMed] [Google Scholar]
- 29.Nuccio SP, Bäumler AJ. 2007. Evolution of the chaperone/usher assembly pathway: fimbrial classification goes Greek. Microbiol Mol Biol Rev 71:551–575. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Sakellaris H, Scott JR. 1998. New tools in an old trade: CS1 pilus morphogenesis. Mol Microbiol 30:681–687. [DOI] [PubMed] [Google Scholar]
- 31.Anantha RP, McVeigh AL, Lee LH, Agnew MK, Cassels FJ, Scott DA, Whittam TS, Savarino SJ. 2004. Evolutionary and functional relationships of colonization factor antigen i and other class 5 adhesive fimbriae of enterotoxigenic Escherichia coli. Infect Immun 72:7190–7201. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Altboum Z, Levine MM, Galen JE, Barry EM. 2003. Genetic characterization and immunogenicity of coli surface antigen 4 from enterotoxigenic Escherichia coli when it is expressed in a Shigella live-vector strain. Infect Immun 71:1352–1360. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Froehlich BJ, Karakashian A, Sakellaris H, Scott JR. 1995. Genes for CS2 pili of enterotoxigenic Escherichia coli and their interchangeability with those for CS1 pili. Infect Immun 63:4849–4856. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Hamers AM, Pel HJ, Willshaw GA, Kusters JG, van der Zeijst BA, Gaastra W. 1989. The nucleotide sequence of the first two genes of the CFA/I fimbrial operon of human enterotoxigenic Escherichia coli. Microb Pathog 6:297–309. [DOI] [PubMed] [Google Scholar]
- 35.Froehlich B, Husmann L, Caron J, Scott JR. 1994. Regulation of rns, a positive regulatory factor for pili of enterotoxigenic Escherichia coli. J Bacteriol 176:5385–5392. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Folkesson A, Advani A, Sukupolvi S, Pfeifer JD, Normark S, Löfdahl S. 1999. Multiple insertions of fimbrial operons correlate with the evolution of Salmonella serovars responsible for human disease. Mol Microbiol 33:612–622. [DOI] [PubMed] [Google Scholar]
- 37.Low D.B. B, van der Woude M. 1996. Escherichia coli and Salmonella: Cellular and Molecular Biology, 2nd ed. ASM Press, Washington, DC. [Google Scholar]
- 38.Zav’yalov VP, Chernovskaya TV, Navolotskaya EV, Karlyshev AV, MacIntyre S, Vasiliev AM, Abramov VM. 1995. Specific high affinity binding of human interleukin 1 beta by Caf1A usher protein of Yersinia pestis. FEBS Lett 371:65–68. [DOI] [PubMed] [Google Scholar]
- 39.Kennan RM, Dhungyel OP, Whittington RJ, Egerton JR, Rood JI. 2001. The type IV fimbrial subunit gene ( fimA) of Dichelobacter nodosus is essential for virulence, protease secretion, and natural competence. J Bacteriol 183:4451–4458. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Hung DL, Knight SD, Woods RM, Pinkner JS, Hultgren SJ. 1996. Molecular basis of two subfamilies of immunoglobulin-like chaperones. EMBO J 15:3792–3805. [PMC free article] [PubMed] [Google Scholar]
- 41.Marklund BI, Tennent JM, Garcia E, Hamers A, Båga M, Lindberg F, Gaastra W, Normark S. 1992. Horizontal gene transfer of the Escherichia coli pap and prs pili operons as a mechanism for the development of tissue-specific adhesive properties. Mol Microbiol 6:2225–2242. [DOI] [PubMed] [Google Scholar]
- 42.Abraham SN, Sun D, Dale JB, Beachey EH. 1988. Conservation of the D-mannose-adhesion protein among type 1 fimbriated members of the family Enterobacteriaceae. Nature 336:682–684. [DOI] [PubMed] [Google Scholar]
- 43.Flores-Mireles AL, Walker JN, Caparon M, Hultgren SJ. 2015. Urinary tract infections: epidemiology, mechanisms of infection and treatment options. Nat Rev Microbiol 13:269–284. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Stamm WE, Norrby SR. 2001. Urinary tract infections: disease panorama and challenges. J Infect Dis 183(Suppl 1):S1–S4. [DOI] [PubMed] [Google Scholar]
- 45.Foxman B. 2010. The epidemiology of urinary tract infection. Nat Rev Urol 7:653–660. [DOI] [PubMed] [Google Scholar]
- 46.Oelschlaeger TA, Dobrindt U, Hacker J. 2002. Virulence factors of uropathogens. Curr Opin Urol 12:33–38. [DOI] [PubMed] [Google Scholar]
- 47.Wurpel DJ, Beatson SA, Totsika M, Petty NK, Schembri MA. 2013. Chaperone-usher fimbriae of Escherichia coli. PLoS One 8:e52835. 10.1371/journal.pone.0052835. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Korea CG, Badouraly R, Prevost MC, Ghigo JM, Beloin C. 2010. Escherichia coli K-12 possesses multiple cryptic but functional chaperone-usher fimbriae with distinct surface specificities. Environ Microbiol 12:1957–1977. [DOI] [PubMed] [Google Scholar]
- 49.Townsend SM, Kramer NE, Edwards R, Baker S, Hamlin N, Simmonds M, Stevens K, Maloy S, Parkhill J, Dougan G, Bäumler AJ. 2001. Salmonella enterica serovar Typhi possesses a unique repertoire of fimbrial gene sequences. Infect Immun 69:2894–2901. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Humphries A, Deridder S, Bäumler AJ. 2005. Salmonella enterica serotype Typhimurium fimbrial proteins serve as antigens during infection of mice. Infect Immun 73:5329–5338. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.McClelland M, Sanderson KE, Clifton SW, Latreille P, Porwollik S, Sabo A, Meyer R, Bieri T, Ozersky P, McLellan M, Harkins CR, Wang C, Nguyen C, Berghoff A, Elliott G, Kohlberg S, Strong C, Du F, Carter J, Kremizki C, Layman D, Leonard S, Sun H, Fulton L, Nash W, Miner T, Minx P, Delehaunty K, Fronick C, Magrini V, Nhan M, Warren W, Florea L, Spieth J, Wilson RK. 2004. Comparison of genome degradation in Paratyphi A and Typhi, human-restricted serovars of Salmonella enterica that cause typhoid. Nat Genet 36:1268–1274. [DOI] [PubMed] [Google Scholar]
- 52.Parkhill J, Dougan G, James KD, Thomson NR, Pickard D, Wain J, Churcher C, Mungall KL, Bentley SD, Holden MT, Sebaihia M, Baker S, Basham D, Brooks K, Chillingworth T, Connerton P, Cronin A, Davis P, Davies RM, Dowd L, White N, Farrar J, Feltwell T, Hamlin N, Haque A, Hien TT, Holroyd S, Jagels K, Krogh A, Larsen TS, Leather S, Moule S, O’Gaora P, Parry C, Quail M, Rutherford K, Simmonds M, Skelton J, Stevens K, Whitehead S, Barrell BG. 2001. Complete genome sequence of a multiple drug resistant Salmonella enterica serovar Typhi CT18. Nature 413:848–852. [DOI] [PubMed] [Google Scholar]
- 53.Weissman SJ, Chattopadhyay S, Aprikian P, Obata-Yasuoka M, Yarova-Yarovaya Y, Stapleton A, Ba-Thein W, Dykhuizen D, Johnson JR, Sokurenko EV. 2006. Clonal analysis reveals high rate of structural mutations in fimbrial adhesins of extraintestinal pathogenic Escherichia coli. Mol Microbiol 59:975–988. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Schwartz DJ, Kalas V, Pinkner JS, Chen SL, Spaulding CN, Dodson KW, Hultgren SJ. 2013. Positively selected FimH residues enhance virulence during urinary tract infection by altering FimH conformation. Proc Natl Acad Sci USA 110:15530–15537. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Sokurenko EV, Chesnokova V, Doyle RJ, Hasty DL. 1997. Diversity of the Escherichia coli type 1 fimbrial lectin. Differential binding to mannosides and uroepithelial cells. J Biol Chem 272:17880–17886. [DOI] [PubMed] [Google Scholar]
- 56.Clegg S, Wilson J, Johnson J. 2011. More than one way to control hair growth: regulatory mechanisms in enterobacteria that affect fimbriae assembled by the chaperone/usher pathway. J Bacteriol 193:2081–2088. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Blomfield IC. 2001. The regulation of pap and type 1 fimbriation in Escherichia coli. Adv Microb Physiol 45:1–49. [DOI] [PubMed] [Google Scholar]
- 58.van der Woude M, Braaten B, Low D. 1996. Epigenetic phase variation of the pap operon in Escherichia coli. Trends Microbiol 4:5–9. [DOI] [PubMed] [Google Scholar]
- 59.Schwan WR. 2011. Regulation of fim genes in uropathogenic Escherichia coli. World J Clin Infect Dis 1:17–25. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Xia Y, Gally D, Forsman-Semb K, Uhlin BE. 2000. Regulatory cross-talk between adhesin operons in Escherichia coli: inhibition of type 1 fimbriae expression by the PapB protein. EMBO J 19:1450–1457. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Totsika M, Beatson SA, Holden N, Gally DL. 2008. Regulatory interplay between pap operons in uropathogenic Escherichia coli. Mol Microbiol 67:996–1011. [DOI] [PubMed] [Google Scholar]
- 62.Snyder JA, Haugen BJ, Lockatell CV, Maroncle N, Hagan EC, Johnson DE, Welch RA, Mobley HL. 2005. Coordinate expression of fimbriae in uropathogenic Escherichia coli. Infect Immun 73:7588–7596. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Holden NJ, Totsika M, Mahler E, Roe AJ, Catherwood K, Lindner K, Dobrindt U, Gally DL. 2006. Demonstration of regulatory cross-talk between P fimbriae and type 1 fimbriae in uropathogenic Escherichia coli. Microbiology 152:1143–1153. [DOI] [PubMed] [Google Scholar]
- 64.Lane MC, Simms AN, Mobley HLT. 2007. complex interplay between type 1 fimbrial expression and flagellum-mediated motility of uropathogenic Escherichia coli. J Bacteriol 189:5523–5533. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Gally DL, Leathart J, Blomfield IC. 1996. Interaction of FimB and FimE with the fim switch that controls the phase variation of type 1 fimbriae in Escherichia coli K-12. Mol Microbiol 21:725–738. [DOI] [PubMed] [Google Scholar]
- 66.Olsen PB, Klemm P. 1994. Localization of promoters in the fim gene cluster and the effect of H-NS on the transcription of fimB and fimE. FEMS Microbiol Lett 116:95–100. [DOI] [PubMed] [Google Scholar]
- 67.Aberg A, Shingler V, Balsalobre C. 2008. Regulation of the fimB promoter: a case of differential regulation by ppGpp and DksA in vivo. Mol Microbiol 67:1223–1241. [DOI] [PubMed] [Google Scholar]
- 68.Adiciptaningrum AM, Blomfield IC, Tans SJ. 2009. Direct observation of type 1 fimbrial switching. EMBO Rep 10:527–532. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Eisenstein BI. 1981. Phase variation of type 1 fimbriae in Escherichia coli is under transcriptional control. Science 214:337–339. [DOI] [PubMed] [Google Scholar]
- 70.Orndorff PE, Spears PA, Schauer D, Falkow S. 1985. Two modes of control of pilA, the gene encoding type 1 pilin in Escherichia coli. J Bacteriol 164:321–330. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Wu Y, Outten FW. 2009. IscR controls iron-dependent biofilm formation in Escherichia coli by regulating type I fimbria expression. J Bacteriol 191:1248–1257. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Aberg A, Shingler V, Balsalobre C. 2006. (p)ppGpp regulates type 1 fimbriation of Escherichia coli by modulating the expression of the site-specific recombinase FimB. Mol Microbiol 60:1520–1533. [DOI] [PubMed] [Google Scholar]
- 73.Holden NJ, Uhlin BE, Gally DL. 2001. PapB paralogues and their effect on the phase variation of type 1 fimbriae in Escherichia coli. Mol Microbiol 42:319–330. [DOI] [PubMed] [Google Scholar]
- 74.Sjöström AE, Sondén B, Müller C, Rydström A, Dobrindt U, Wai SN, Uhlin BE. 2009. Analysis of the sfaX(II) locus in the Escherichia coli meningitis isolate IHE3034 reveals two novel regulatory genes within the promoter-distal region of the main S fimbrial operon. Microb Pathog 46:150–158. [DOI] [PubMed] [Google Scholar]
- 75.Corcoran CP, Dorman CJ. 2009. DNA relaxation-dependent phase biasing of the fim genetic switch in Escherichia coli depends on the interplay of H-NS, IHF and LRP. Mol Microbiol 74:1071–1082. [DOI] [PubMed] [Google Scholar]
- 76.Holden N, Blomfield IC, Uhlin BE, Totsika M, Kulasekara DH, Gally DL. 2007. Comparative analysis of FimB and FimE recombinase activity. Microbiology 153:4138–4149. [DOI] [PubMed] [Google Scholar]
- 77.Old DC, Duguid JP. 1970. Selective outgrowth of fimbriate bacteria in static liquid medium. J Bacteriol 103:447–456. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Schwan WR, Lee JL, Lenard FA, Matthews BT, Beck MT. 2002. Osmolarity and pH growth conditions regulate fim gene transcription and type 1 pilus expression in uropathogenic Escherichia coli. Infect Immun 70:1391–1402. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Floyd KA, Moore JL, Eberly AR, Good JA, Shaffer CL, Zaver H, Almqvist F, Skaar EP, Caprioli RM, Hadjifrangiskou M. 2015. Adhesive fiber stratification in uropathogenic Escherichia coli biofilms unveils oxygen-mediated control of type 1 pili. PLoS Pathog 11:e1004697. 10.1371/journal.ppat.1004697. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Floyd KA, Mitchell CA, Eberly AR, Colling SJ, Zhang EW, DePas W, Chapman MR, Conover M, Rogers BR, Hultgren SJ, Hadjifrangiskou M. 2016. The UbiI (VisC) aerobic ubiquinone synthase Is required for expression of type 1 pili, biofilm formation, and pathogenesis in uropathogenic Escherichia coli. J Bacteriol 198:2662–2672. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.van der Woude MW, Bäumler AJ. 2004. Phase and antigenic variation in bacteria. Clin Microbiol Rev 17:581–611. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Braaten BA, Blyn LB, Skinner BS, Low DA. 1991. Evidence for a methylation-blocking factor ( mbf) locus involved in pap pilus expression and phase variation in Escherichia coli. J Bacteriol 173:1789–1800. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Peterson SN, Reich NO. 2008. Competitive Lrp and Dam assembly at the pap regulatory region: implications for mechanisms of epigenetic regulation. J Mol Biol 383:92–105. [DOI] [PubMed] [Google Scholar]
- 84.Peterson SN, Reich NO. 2006. GATC flanking sequences regulate Dam activity: evidence for how Dam specificity may influence pap expression. J Mol Biol 355:459–472. [DOI] [PubMed] [Google Scholar]
- 85.Nou X, Braaten B, Kaltenbach L, Low DA. 1995. Differential binding of Lrp to two sets of pap DNA binding sites mediated by Pap I regulates Pap phase variation in Escherichia coli. EMBO J 14:5785–5797. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Forsman K, Göransson M, Uhlin BE. 1989. Autoregulation and multiple DNA interactions by a transcriptional regulatory protein in E. coli pili biogenesis. EMBO J 8:1271–1277. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Holden N, Totsika M, Dixon L, Catherwood K, Gally DL. 2007. Regulation of P-fimbrial phase variation frequencies in Escherichia coli CFT073. Infect Immun 75:3325–3334. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Hernday AD, Braaten BA, Broitman-Maduro G, Engelberts P, Low DA. 2004. Regulation of the pap epigenetic switch by CpxAR: phosphorylated CpxR inhibits transition to the phase ON state by competition with Lrp. Mol Cell 16:537–547. [DOI] [PubMed] [Google Scholar]
- 89.Khandige S, Kronborg T, Uhlin BE, Møller-Jensen J. 2015. sRNA-mediated regulation of P-fimbriae phase variation in uropathogenic Escherichia coli. PLoS Pathog 11:e1005109. 10.1371/journal.ppat.1005109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Servin AL. 2005. Pathogenesis of Afa/Dr diffusely adhering Escherichia coli. Clin Microbiol Rev 18:264–292. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Nowicki B, Selvarangan R, Nowicki S. 2001. Family of Escherichia coli Dr adhesins: decay-accelerating factor receptor recognition and invasiveness. J Infect Dis 183(Suppl 1):S24–S27. [DOI] [PubMed] [Google Scholar]
- 92.Båga M, Norgren M, Normark S. 1987. Biogenesis of E. coli Pap pili: papH, a minor pilin subunit involved in cell anchoring and length modulation. Cell 49:241–251. [DOI] [PubMed] [Google Scholar]
- 93.Rossolini GM, Muscas P, Chiesurin A, Satta G. 1993. Analysis of the Salmonella fim gene cluster: identification of a new gene ( fimI) encoding a fimbrin-like protein and located downstream from the fimA gene. FEMS Microbiol Lett 114:259–265. [DOI] [PubMed] [Google Scholar]
- 94.Valenski ML, Harris SL, Spears PA, Horton JR, Orndorff PE. 2003. The Product of the fimI gene is necessary for Escherichia coli type 1 pilus biosynthesis. J Bacteriol 185:5007–5011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Garcia MI, Gounon P, Courcoux P, Labigne A, Le Bouguenec C. 1996. The afimbrial adhesive sheath encoded by the afa-3 gene cluster of pathogenic Escherichia coli is composed of two adhesins. Mol Microbiol 19:683–693. [DOI] [PubMed] [Google Scholar]
- 96.Labigne-Roussel AF, Lark D, Schoolnik G, Falkow S. 1984. Cloning and expression of an afimbrial adhesin (AFA-I) responsible for P blood group-independent, mannose-resistant hemagglutination from a pyelonephritic Escherichia coli strain. Infect Immun 46:251–259. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Labigne-Roussel AF, Schmidt MA, Walz W, Falkow S. 1985. Genetic organization of the AFA operon and nucleotide sequence from a uropathogenic Escherichia coli gene encoding an afimbrial adhesin (AFA-1). J Bacteriol 162:1285–1292. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Walz W, Schmidt MA, Labigne-Roussel AF, Falkow S, Schoolnik G. 1985. AFA-I, a cloned afimbrial X-type adhesin from a human pyelonephritic Escherichia coli strain. Purification and chemical, functional and serologic characterization. Eur J Biochem 152:315–321. [DOI] [PubMed] [Google Scholar]
- 99.Anderson KL, Billington J, Pettigrew D, Cota E, Simpson P, Roversi P, Chen HA, Urvil P, du Merle L, Barlow PN, Medof ME, Smith RA, Nowicki B, Le Bouguénec C, Lea SM, Matthews S. 2004. An atomic resolution model for assembly, architecture, and function of the Dr adhesins. Mol Cell 15:647–657. [DOI] [PubMed] [Google Scholar]
- 100.Jouve M, Garcia MI, Courcoux P, Labigne A, Gounon P, Le Bouguénec C. 1997. Adhesion to and invasion of HeLa cells by pathogenic Escherichia coli carrying the afa-3 gene cluster are mediated by the AfaE and AfaD proteins, respectively. Infect Immun 65:4082–4089. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Gounon P, Jouve M, Le Bouguénec C. 2000. Immunocytochemistry of the AfaE adhesin and AfaD invasin produced by pathogenic Escherichia coli strains during interaction of the bacteria with HeLa cells by high-resolution scanning electron microscopy. Microbes Infect 2:359–365. [DOI] [PubMed] [Google Scholar]
- 102.Zalewska B, Piatek R, Bury K, Samet A, Nowicki B, Nowicki S, Kur J. 2005. A surface-exposed DraD protein of uropathogenic Escherichia coli bearing Dr fimbriae may be expressed and secreted independently from DraC usher and DraE adhesin. Microbiology 151:2477–2486. [DOI] [PubMed] [Google Scholar]
- 103.Pettigrew D, Anderson KL, Billington J, Cota E, Simpson P, Urvil P, Rabuzin F, Roversi P, Nowicki B, du Merle L, Le Bouguénec C, Matthews S, Lea SM. 2004. High resolution studies of the Afa/Dr adhesin DraE and its interaction with chloramphenicol. J Biol Chem 279:46851–46857. [DOI] [PubMed] [Google Scholar]
- 104.Korotkova N, Le Trong I, Samudrala R, Korotkov K, Van Loy CP, Bui AL, Moseley SL, Stenkamp RE. 2006. Crystal structure and mutational analysis of the DaaE adhesin of Escherichia coli. J Biol Chem 281:22367–22377. [DOI] [PubMed] [Google Scholar]
- 105.Bork P, Holm L, Sander C. 1994. The immunoglobulin fold. Structural classification, sequence patterns and common core. J Mol Biol 242:309–320. [DOI] [PubMed] [Google Scholar]
- 106.Choudhury D, Thompson A, Stojanoff V, Langermann S, Pinkner J, Hultgren SJ, Knight SD. 1999. X-ray structure of the FimC-FimH chaperone-adhesin complex from uropathogenic Escherichia coli. Science 285:1061–1066. [DOI] [PubMed] [Google Scholar]
- 107.Sauer FG, Fütterer K, Pinkner JS, Dodson KW, Hultgren SJ, Waksman G. 1999. Structural basis of chaperone function and pilus biogenesis. Science 285:1058–1061. [DOI] [PubMed] [Google Scholar]
- 108.Sauer FG, Pinkner JS, Waksman G, Hultgren SJ. 2002. Chaperone priming of pilus subunits facilitates a topological transition that drives fiber formation. Cell 111:543–551. [DOI] [PubMed] [Google Scholar]
- 109.Zavialov AV, Berglund J, Pudney AF, Fooks LJ, Ibrahim TM, MacIntyre S, Knight SD. 2003. Structure and biogenesis of the capsular F1 antigen from Yersinia pestis: preserved folding energy drives fiber formation. Cell 113:587–596. [DOI] [PubMed] [Google Scholar]
- 110.Westerlund-Wikström B, Korhonen TK. 2005. Molecular structure of adhesin domains in Escherichia coli fimbriae. Int J Med Microbiol 295:479–486. [DOI] [PubMed] [Google Scholar]
- 111.Hung CS, Bouckaert J, Hung D, Pinkner J, Widberg C, DeFusco A, Auguste CG, Strouse R, Langermann S, Waksman G, Hultgren SJ. 2002. Structural basis of tropism of Escherichia coli to the bladder during urinary tract infection. Mol Microbiol 44:903–915. [DOI] [PubMed] [Google Scholar]
- 112.Dodson KW, Pinkner JS, Rose T, Magnusson G, Hultgren SJ, Waksman G. 2001. Structural basis of the interaction of the pyelonephritic E. coli adhesin to its human kidney receptor. Cell 105:733–743. [DOI] [PubMed] [Google Scholar]
- 113.Sung MA, Chen HA, Matthews S. 2001. Sequential assignment and secondary structure of the triple-labelled carbohydrate-binding domain of papG from uropathogenic E. coli. J Biomol NMR 19:197–198. [DOI] [PubMed] [Google Scholar]
- 114.Piątek R, Bruździak P, Wojciechowski M, Zalewska-Piątek B, Kur J. 2010. The noncanonical disulfide bond as the important stabilizing element of the immunoglobulin fold of the Dr fimbrial DraE subunit. Biochemistry 49:1460–1468. [DOI] [PubMed] [Google Scholar]
- 115.Thomas WE, Trintchina E, Forero M, Vogel V, Sokurenko EV. 2002. Bacterial adhesion to target cells enhanced by shear force. Cell 109:913–923. [DOI] [PubMed] [Google Scholar]
- 116.Thomas WE, Nilsson LM, Forero M, Sokurenko EV, Vogel V. 2004. Shear-dependent ‘stick-and-roll’ adhesion of type 1 fimbriated Escherichia coli. Mol Microbiol 53:1545–1557. [DOI] [PubMed] [Google Scholar]
- 117.Geibel S, Procko E, Hultgren SJ, Baker D, Waksman G. 2013. Structural and energetic basis of folded-protein transport by the FimD usher. Nature 496:243–246. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Le Trong I, Aprikian P, Kidd BA, Forero-Shelton M, Tchesnokova V, Rajagopal P, Rodriguez V, Interlandi G, Klevit R, Vogel V, Stenkamp RE, Sokurenko EV, Thomas WE. 2010. Structural basis for mechanical force regulation of the adhesin FimH via finger trap-like beta sheet twisting. Cell 141:645–655. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Sauer MM, Jakob RP, Eras J, Baday S, Eriş D, Navarra G, Bernèche S, Ernst B, Maier T, Glockshuber R. 2016. Catch-bond mechanism of the bacterial adhesin FimH. Nat Commun 7:10738. 10.1038/ncomms10738. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120.Yakovenko O, Tchesnokova V, Sokurenko EV, Thomas WE. 2015. Inactive conformation enhances binding function in physiological conditions. Proc Natl Acad Sci USA 112:9884–9889. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Pak J, Pu Y, Zhang ZT, Hasty DL, Wu XR. 2001. Tamm-Horsfall protein binds to type 1 fimbriated Escherichia coli and prevents E. coli from binding to uroplakin Ia and Ib receptors. J Biol Chem 276:9924–9930. [DOI] [PubMed] [Google Scholar]
- 122.Lugmaier RA, Schedin S, Kühner F, Benoit M. 2008. Dynamic restacking of Escherichia coli P-pili. Eur Biophys J 37:111–120. [DOI] [PubMed] [Google Scholar]
- 123.Larsson A, Ohlsson J, Dodson KW, Hultgren SJ, Nilsson U, Kihlberg J. 2003. Quantitative studies of the binding of the class II PapG adhesin from uropathogenic Escherichia coli to oligosaccharides. Bioorg Med Chem 11:2255–2261. [DOI] [PubMed] [Google Scholar]
- 124.Zakrisson J, Wiklund K, Axner O, Andersson M. 2013. The shaft of the type 1 fimbriae regulates an external force to match the FimH catch bond. Biophys J 104:2137–2148. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Jass J, Schedin S, Fällman E, Ohlsson J, Nilsson UJ, Uhlin BE, Axner O. 2004. Physical properties of Escherichia coli P pili measured by optical tweezers. Biophys J 87:4271–4283. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126.Andersson M, Axner O, Almqvist F, Uhlin BE, Fällman E. 2008. Physical properties of biopolymers assessed by optical tweezers: analysis of folding and refolding of bacterial pili. ChemPhysChem 9:221–235. [DOI] [PubMed] [Google Scholar]
- 127.Fällman E, Schedin S, Jass J, Uhlin BE, Axner O. 2005. The unfolding of the P pili quaternary structure by stretching is reversible, not plastic. EMBO Rep 6:52–56. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128.Hospenthal MK, Redzej A, Dodson K, Ukleja M, Frenz B, Rodrigues C, Hultgren SJ, DiMaio F, Egelman EH, Waksman G. 2016. Structure of a chaperone-usher pilus reveals the molecular basis of rod uncoiling. Cell 164:269–278. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129.Mortezaei N, Singh B, Zakrisson J, Bullitt E, Andersson M. 2015. Biomechanical and structural features of CS2 fimbriae of enterotoxigenic Escherichia coli. Biophys J 109:49–56. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130.Jacob-Dubuisson F, Striker R, Hultgren SJ. 1994. Chaperone-assisted self-assembly of pili independent of cellular energy. J Biol Chem 269:12447–12455. [PubMed] [Google Scholar]
- 131.Thanassi DG, Stathopoulos C, Karkal A, Li H. 2005. Protein secretion in the absence of ATP: the autotransporter, two-partner secretion and chaperone/usher pathways of gram-negative bacteria (review). Mol Membr Biol 22:63–72. [DOI] [PubMed] [Google Scholar]
- 132.Lycklama A Nijeholt JA, Driessen AJ, Driessen AJ. 2012. The bacterial Sec-translocase: structure and mechanism. Philos Trans R Soc Lond B Biol Sci 367:1016–1028. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133.Totsika M, Heras B, Wurpel DJ, Schembri MA. 2009. Characterization of two homologous disulfide bond systems involved in virulence factor biogenesis in uropathogenic Escherichia coli CFT073. J Bacteriol 191:3901–3908. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134.Crespo MD, Puorger C, Schärer MA, Eidam O, Grütter MG, Capitani G, Glockshuber R. 2012. Quality control of disulfide bond formation in pilus subunits by the chaperone FimC. Nat Chem Biol 8:707–713. [DOI] [PubMed] [Google Scholar]
- 135.Waksman G, Hultgren SJ. 2009. Structural biology of the chaperone-usher pathway of pilus biogenesis. Nat Rev Microbiol 7:765–774. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 136.Zavialov AV, Tischenko VM, Fooks LJ, Brandsdal BO, Aqvist J, Zav’yalov VP, Macintyre S, Knight SD. 2005. Resolving the energy paradox of chaperone/usher-mediated fibre assembly. Biochem J 389:685–694. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137.Nishiyama M, Vetsch M, Puorger C, Jelesarov I, Glockshuber R. 2003. Identification and characterization of the chaperone-subunit complex-binding domain from the type 1 pilus assembly platform FimD. J Mol Biol 330:513–525. [DOI] [PubMed] [Google Scholar]
- 138.Bann JG, Pinkner JS, Frieden C, Hultgren SJ. 2004. Catalysis of protein folding by chaperones in pathogenic bacteria. Proc Natl Acad Sci USA 101:17389–17393. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 139.Zav’yalov VP, Chernovskaya TV, Chapman DA, Karlyshev AV, MacIntyre S, Zavialov AV, Vasiliev AM, Denesyuk AI, Zav’yalova GA, Dudich IV, Korpela T, Abramov VM. 1997. Influence of the conserved disulphide bond, exposed to the putative binding pocket, on the structure and function of the immunoglobulin-like molecular chaperone Caf1M of Yersinia pestis. Biochem J 324:571–578. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140.Eidam O, Dworkowski FS, Glockshuber R, Grütter MG, Capitani G. 2008. Crystal structure of the ternary FimC-FimF(t)-FimD(N) complex indicates conserved pilus chaperone-subunit complex recognition by the usher FimD. FEBS Lett 582:651–655. [DOI] [PubMed] [Google Scholar]
- 141.Verger D, Bullitt E, Hultgren SJ, Waksman G. 2007. Crystal structure of the P pilus rod subunit PapA. PLoS Pathog 3:e73. 10.1371/journal.ppat.0030073. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142.Verger D, Miller E, Remaut H, Waksman G, Hultgren S. 2006. Molecular mechanism of P pilus termination in uropathogenic Escherichia coli. EMBO Rep 7:1228–1232. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 143.Verger D, Rose RJ, Paci E, Costakes G, Daviter T, Hultgren S, Remaut H, Ashcroft AE, Radford SE, Waksman G. 2008. Structural determinants of polymerization reactivity of the P pilus adaptor subunit PapF. Structure 16:1724–1731. [DOI] [PubMed] [Google Scholar]
- 144.Ford B, Verger D, Dodson K, Volkan E, Kostakioti M, Elam J, Pinkner J, Waksman G, Hultgren S. 2012. The structure of the PapD-PapGII pilin complex reveals an open and flexible P5 pocket. J Bacteriol 194:6390–6397. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145.Yu XD, Fooks LJ, Moslehi-Mohebi E, Tischenko VM, Askarieh G, Knight SD, Macintyre S, Zavialov AV. 2012. Large is fast, small is tight: determinants of speed and affinity in subunit capture by a periplasmic chaperone. J Mol Biol 417:294–308. [DOI] [PubMed] [Google Scholar]
- 146.Remaut H, Rose RJ, Hannan TJ, Hultgren SJ, Radford SE, Ashcroft AE, Waksman G. 2006. Donor-strand exchange in chaperone-assisted pilus assembly proceeds through a concerted beta strand displacement mechanism. Mol Cell 22:831–842. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147.Piątek R, Zalewska B, Kolaj O, Ferens M, Nowicki B, Kur J. 2005. Molecular aspects of biogenesis of Escherichia coli Dr Fimbriae: characterization of DraB-DraE complexes. Infect Immun 73:135–145. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148.Puorger C, Eidam O, Capitani G, Erilov D, Grütter MG, Glockshuber R. 2008. Infinite kinetic stability against dissociation of supramolecular protein complexes through donor strand complementation. Structure 16:631–642. [DOI] [PubMed] [Google Scholar]
- 149.Nishiyama M, Ishikawa T, Rechsteiner H, Glockshuber R. 2008. Reconstitution of pilus assembly reveals a bacterial outer membrane catalyst. Science 320:376–379. [DOI] [PubMed] [Google Scholar]
- 150.Nishiyama M, Horst R, Eidam O, Herrmann T, Ignatov O, Vetsch M, Bettendorff P, Jelesarov I, Grütter MG, Wüthrich K, Glockshuber R, Capitani G. 2005. Structural basis of chaperone-subunit complex recognition by the type 1 pilus assembly platform FimD. EMBO J 24:2075–2086. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151.Phan G, Remaut H, Wang T, Allen WJ, Pirker KF, Lebedev A, Henderson NS, Geibel S, Volkan E, Yan J, Kunze MB, Pinkner JS, Ford B, Kay CW, Li H, Hultgren SJ, Thanassi DG, Waksman G. 2011. Crystal structure of the FimD usher bound to its cognate FimC-FimH substrate. Nature 474:49–53. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 152.Remaut H, Tang C, Henderson NS, Pinkner JS, Wang T, Hultgren SJ, Thanassi DG, Waksman G, Li H. 2008. Fiber formation across the bacterial outer membrane by the chaperone/usher pathway. Cell 133:640–652. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 153.Huang Y, Smith BS, Chen LX, Baxter RH, Deisenhofer J. 2009. Insights into pilus assembly and secretion from the structure and functional characterization of usher PapC. Proc Natl Acad Sci USA 106:7403–7407. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154.Di Yu X, Dubnovitsky A, Pudney AF, Macintyre S, Knight SD, Zavialov AV. 2012. Allosteric mechanism controls traffic in the chaperone/usher pathway. Structure 20:1861–1871. [DOI] [PubMed] [Google Scholar]
- 155.Werneburg GT, Henderson NS, Portnoy EB, Sarowar S, Hultgren SJ, Li H, Thanassi DG. 2015. The pilus usher controls protein interactions via domain masking and is functional as an oligomer. Nat Struct Mol Biol 22:540–546. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156.Volkan E, Ford BA, Pinkner JS, Dodson KW, Henderson NS, Thanassi DG, Waksman G, Hultgren SJ. 2012. Domain activities of PapC usher reveal the mechanism of action of an Escherichia coli molecular machine. Proc Natl Acad Sci USA 109:9563–9568. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157.Thanassi DG, Saulino ET, Lombardo MJ, Roth R, Heuser J, Hultgren SJ. 1998. The PapC usher forms an oligomeric channel: implications for pilus biogenesis across the outer membrane. Proc Natl Acad Sci USA 95:3146–3151. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 158.Li H, Qian L, Chen Z, Thibault D, Liu G, Liu T, Thanassi DG. 2004. The outer membrane usher forms a twin-pore secretion complex. J Mol Biol 344:1397–1407. [DOI] [PubMed] [Google Scholar]
- 159.So SS, Thanassi DG. 2006. Analysis of the requirements for pilus biogenesis at the outer membrane usher and the function of the usher C-terminus. Mol Microbiol 60:364–375. [DOI] [PubMed] [Google Scholar]
- 160.Nishiyama M, Glockshuber R. 2010. The outer membrane usher guarantees the formation of functional pili by selectively catalyzing donor-strand exchange between subunits that are adjacent in the mature pilus. J Mol Biol 396:1–8. [DOI] [PubMed] [Google Scholar]
- 161.Li Q, Ng TW, Dodson KW, So SS, Bayle KM, Pinkner JS, Scarlata S, Hultgren SJ, Thanassi DG. 2010. The differential affinity of the usher for chaperone-subunit complexes is required for assembly of complete pili. Mol Microbiol 76:159–172. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 162.Saulino ET, Thanassi DG, Pinkner JS, Hultgren SJ. 1998. Ramifications of kinetic partitioning on usher-mediated pilus biogenesis. EMBO J 17:2177–2185. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 163.Allen WJ, Phan G, Hultgren SJ, Waksman G. 2013. Dissection of pilus tip assembly by the FimD usher monomer. J Mol Biol 425:958–967. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 164.Durno C, Soni R, Sherman P. 1989. Adherence of vero cytotoxin-producing Escherichia coli serotype O157:H7 to isolated epithelial cells and brush border membranes in vitro: role of type 1 fimbriae (pili) as a bacterial adhesin expressed by strain CL-49. Clin Invest Med 12:194–200. [PubMed] [Google Scholar]
- 165.Sherman P, Soni R, Petric M, Karmali M. 1987. Surface properties of the Vero cytotoxin-producing Escherichia coli O157:H7. Infect Immun 55:1824–1829. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166.Kim SH, Kim YH. 2004. Escherichia coli O157:H7 adherence to HEp-2 cells is implicated with curli expression and outer membrane integrity. J Vet Sci 5:119–124. [PubMed] [Google Scholar]
- 167.Brunder W, Khan AS, Hacker J, Karch H. 2001. Novel type of fimbriae encoded by the large plasmid of sorbitol-fermenting enterohemorrhagic Escherichia coli O157:H(-). Infect Immun 69:4447–4457. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 168.Karch H, Heesemann J, Laufs R, O’Brien AD, Tacket CO, Levine MM. 1987. A plasmid of enterohemorrhagic Escherichia coli O157:H7 is required for expression of a new fimbrial antigen and for adhesion to epithelial cells. Infect Immun 55:455–461. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 169.Hayashi T, Makino K, Ohnishi M, Kurokawa K, Ishii K, Yokoyama K, Han CG, Ohtsubo E, Nakayama K, Murata T, Tanaka M, Tobe T, Iida T, Takami H, Honda T, Sasakawa C, Ogasawara N, Yasunaga T, Kuhara S, Shiba T, Hattori M, Shinagawa H. 2001. Complete genome sequence of enterohemorrhagic Escherichia coli O157:H7 and genomic comparison with a laboratory strain K-12. DNA Res 8:11–22. [DOI] [PubMed] [Google Scholar]
- 170.Perna NT, Plunkett G III, Burland V, Mau B, Glasner JD, Rose DJ, Mayhew GF, Evans PS, Gregor J, Kirkpatrick HA, Pósfai G, Hackett J, Klink S, Boutin A, Shao Y, Miller L, Grotbeck EJ, Davis NW, Lim A, Dimalanta ET, Potamousis KD, Apodaca J, Anantharaman TS, Lin J, Yen G, Schwartz DC, Welch RA, Blattner FR. 2001. Genome sequence of enterohaemorrhagic Escherichia coli O157:H7. Nature 409:529–533. [DOI] [PubMed] [Google Scholar]
- 171.Low AS, Holden N, Rosser T, Roe AJ, Constantinidou C, Hobman JL, Smith DG, Low JC, Gally DL. 2006. Analysis of fimbrial gene clusters and their expression in enterohaemorrhagic Escherichia coli O157:H7. Environ Microbiol 8:1033–1047. [DOI] [PubMed] [Google Scholar]
- 172.Roe AJ, Currie C, Smith DG, Gally DL. 2001. Analysis of type 1 fimbriae expression in verotoxigenic Escherichia coli: a comparison between serotypes O157 and O26. Microbiology 147:145–152. [DOI] [PubMed] [Google Scholar]
- 173.Dziva F, van Diemen PM, Stevens MP, Smith AJ, Wallis TS. 2004. Identification of Escherichia coli O157 : H7 genes influencing colonization of the bovine gastrointestinal tract using signature-tagged mutagenesis. Microbiology 150:3631–3645. [DOI] [PubMed] [Google Scholar]
- 174.Jordan DM, Cornick N, Torres AG, Dean-Nystrom EA, Kaper JB, Moon HW. 2004. Long polar fimbriae contribute to colonization by Escherichia coli O157:H7 in vivo. Infect Immun 72:6168–6171. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 175.Low AS, Dziva F, Torres AG, Martinez JL, Rosser T, Naylor S, Spears K, Holden N, Mahajan A, Findlay J, Sales J, Smith DG, Low JC, Stevens MP, Gally DL. 2006. Cloning, expression, and characterization of fimbrial operon F9 from enterohemorrhagic Escherichia coli O157:H7. Infect Immun 74:2233–2244. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 176.Torres AG, Giron JA, Perna NT, Burland V, Blattner FR, Avelino-Flores F, Kaper JB. 2002. Identification and characterization of lpfABCC’DE, a fimbrial operon of enterohemorrhagic Escherichia coli O157:H7. Infect Immun 70:5416–5427. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 177.Torres AG, Kanack KJ, Tutt CB, Popov V, Kaper JB. 2004. Characterization of the second long polar (LP) fimbriae of Escherichia coli O157:H7 and distribution of LP fimbriae in other pathogenic E. coli strains. FEMS Microbiol Lett 238:333–344. [DOI] [PubMed] [Google Scholar]
- 178.Zhou G, Mo WJ, Sebbel P, Min G, Neubert TA, Glockshuber R, Wu XR, Sun TT, Kong XP. 2001. Uroplakin Ia is the urothelial receptor for uropathogenic Escherichia coli: evidence from in vitro FimH binding. J Cell Sci 114:4095–4103. [DOI] [PubMed] [Google Scholar]
- 179.Eto DS, Jones TA, Sundsbak JL, Mulvey MA. 2007. Integrin-mediated host cell invasion by type 1-piliated uropathogenic Escherichia coli. PLoS Pathog 3:e100. 10.1371/journal.ppat.0030100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 180.Langermann S, Möllby R, Burlein JE, Palaszynski SR, Auguste CG, DeFusco A, Strouse R, Schenerman MA, Hultgren SJ, Pinkner JS, Winberg J, Guldevall L, Söderhäll M, Ishikawa K, Normark S, Koenig S. 2000. Vaccination with FimH adhesin protects cynomolgus monkeys from colonization and infection by uropathogenic Escherichia coli. J Infect Dis 181:774–778. [DOI] [PubMed] [Google Scholar]
- 181.Langermann S, Palaszynski S, Barnhart M, Auguste G, Pinkner JS, Burlein J, Barren P, Koenig S, Leath S, Jones CH, Hultgren SJ. 1997. Prevention of mucosal Escherichia coli infection by FimH-adhesin-based systemic vaccination. Science 276:607–611. [DOI] [PubMed] [Google Scholar]
- 182.Hannan TJ, Totsika M, Mansfield KJ, Moore KH, Schembri MA, Hultgren SJ. 2012. Host-pathogen checkpoints and population bottlenecks in persistent and intracellular uropathogenic Escherichia coli bladder infection. FEMS Microbiol Rev 36:616–648. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 183.Pratt LA, Kolter R. 1998. Genetic analysis of Escherichia coli biofilm formation: roles of flagella, motility, chemotaxis and type I pili. Mol Microbiol 30:285–293. [DOI] [PubMed] [Google Scholar]
- 184.Kukkonen M, Raunio T, Virkola R, Lähteenmäki K, Mäkelä PH, Klemm P, Clegg S, Korhonen TK. 1993. Basement membrane carbohydrate as a target for bacterial adhesion: binding of type I fimbriae of Salmonella enterica and Escherichia coli to laminin. Mol Microbiol 7:229–237. [DOI] [PubMed] [Google Scholar]
- 185.Baorto DM, Gao Z, Malaviya R, Dustin ML, van der Merwe A, Lublin DM, Abraham SN. 1997. Survival of FimH-expressing enterobacteria in macrophages relies on glycolipid traffic. Nature 389:636–639. [DOI] [PubMed] [Google Scholar]
- 186.Mulvey MA, Lopez-Boado YS, Wilson CL, Roth R, Parks WC, Heuser J, Hultgren SJ. 1998. Induction and evasion of host defenses by type 1-piliated uropathogenic Escherichia coli. Science 282:1494–1497. [DOI] [PubMed] [Google Scholar]
- 187.Liu Y, Mémet S, Saban R, Kong X, Aprikian P, Sokurenko E, Sun TT, Wu XR. 2015. Dual ligand/receptor interactions activate urothelial defenses against uropathogenic E. coli. Sci Rep 5:16234. 10.1038/srep16234. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 188.Martinez JJ, Mulvey MA, Schilling JD, Pinkner JS, Hultgren SJ. 2000. Type 1 pilus-mediated bacterial invasion of bladder epithelial cells. EMBO J 19:2803–2812. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 189.Keith BR, Harris SL, Russell PW, Orndorff PE. 1990. Effect of type 1 piliation on in vitro killing of Escherichia coli by mouse peritoneal macrophages. Infect Immun 58:3448–3454. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 190.Avalos Vizcarra I, Hosseini V, Kollmannsberger P, Meier S, Weber SS, Arnoldini M, Ackermann M, Vogel V. 2016. How type 1 fimbriae help Escherichia coli to evade extracellular antibiotics. Sci Rep 6:18109. 10.1038/srep18109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 191.Glasser AL, Boudeau J, Barnich N, Perruchot MH, Colombel JF, Darfeuille-Michaud A. 2001. Adherent invasive Escherichia coli strains from patients with Crohn’s disease survive and replicate within macrophages without inducing host cell death. Infect Immun 69:5529–5537. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 192.Sukumaran SK, Fu NY, Tin CB, Wan KF, Lee SS, Yu VC. 2010. A soluble form of the pilus protein FimA targets the VDAC-hexokinase complex at mitochondria to suppress host cell apoptosis. Mol Cell 37:768–783. [DOI] [PubMed] [Google Scholar]
- 193.Walczak MJ, Puorger C, Glockshuber R, Wider G. 2014. Intramolecular donor strand complementation in the E. coli type 1 pilus subunit FimA explains the existence of FimA monomers as off-pathway products of pilus assembly that inhibit host cell apoptosis. J Mol Biol 426:542–549. [DOI] [PubMed] [Google Scholar]
- 194.Wright KJ, Seed PC, Hultgren SJ. 2007. Development of intracellular bacterial communities of uropathogenic Escherichia coli depends on type 1 pili. Cell Microbiol 9:2230–2241. [DOI] [PubMed] [Google Scholar]
- 195.Chen SL, Hung CS, Pinkner JS, Walker JN, Cusumano CK, Li Z, Bouckaert J, Gordon JI, Hultgren SJ. 2009. Positive selection identifies an in vivo role for FimH during urinary tract infection in addition to mannose binding. Proc Natl Acad Sci USA 106:22439–22444. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 196.Mulvey MA, Schilling JD, Hultgren SJ. 2001. Establishment of a persistent Escherichia coli reservoir during the acute phase of a bladder infection. Infect Immun 69:4572–4579. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 197.Justice SS, Hung C, Theriot JA, Fletcher DA, Anderson GG, Footer MJ, Hultgren SJ. 2004. Differentiation and developmental pathways of uropathogenic Escherichia coli in urinary tract pathogenesis. Proc Natl Acad Sci USA 101:1333–1338. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 198.Mysorekar IU, Hultgren SJ. 2006. Mechanisms of uropathogenic Escherichia coli persistence and eradication from the urinary tract. Proc Natl Acad Sci USA 103:14170–14175. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 199.Ewers C, Li G, Wilking H, Kiessling S, Alt K, Antáo EM, Laturnus C, Diehl I, Glodde S, Homeier T, Böhnke U, Steinrück H, Philipp HC, Wieler LH. 2007. Avian pathogenic, uropathogenic, and newborn meningitis-causing Escherichia coli: how closely related are they? Int J Med Microbiol 297:163–176. [DOI] [PubMed] [Google Scholar]
- 200.Tennent JM, Lindberg F, Normark S. 1990. Integrity of Escherichia coli P pili during biogenesis: properties and role of PapJ. Mol Microbiol 4:747–758. [DOI] [PubMed] [Google Scholar]
- 201.Roberts JA, Marklund B-I, Ilver D, Haslam D, Kaack MB, Baskin G, Louis M, Möllby R, Winberg J, Normark S. 1994. The Gal(alpha 1-4)Gal-specific tip adhesin of Escherichia coli P-fimbriae is needed for pyelonephritis to occur in the normal urinary tract. Proc Natl Acad Sci USA 91:11889–11893. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 202.Bock K, Breimer ME, Brignole A, Hansson GC, Karlsson K-A, Larson G, Leffler H, Samuelsson BE, Strömberg N, Edén CS, Thurin J. 1985. Specificity of binding of a strain of uropathogenic Escherichia coli to Gal alpha 1----4Gal-containing glycosphingolipids. J Biol Chem 260:8545–8551. [PubMed] [Google Scholar]
- 203.Strömberg N, Marklund BI, Lund B, Ilver D, Hamers A, Gaastra W, Karlsson KA, Normark S. 1990. Host-specificity of uropathogenic Escherichia coli depends on differences in binding specificity to Gal alpha 1-4Gal-containing isoreceptors. EMBO J 9:2001–2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 204.Johnson JR, Russo TA, Brown JJ, Stapleton A. 1998. papG alleles of Escherichia coli strains causing first-episode or recurrent acute cystitis in adult women. J Infect Dis 177:97–101. [DOI] [PubMed] [Google Scholar]
- 205.Roberts JA, Hardaway K, Kaack B, Fussell EN, Baskin G. 1984. Prevention of pyelonephritis by immunization with P-fimbriae. J Urol 131:602–607. [DOI] [PubMed] [Google Scholar]
- 206.O’Hanley P, Lark D, Falkow S, Schoolnik G. 1985. Molecular basis of Escherichia coli colonization of the upper urinary tract in BALB/c mice. Gal-Gal pili immunization prevents Escherichia coli pyelonephritis in the BALB/c mouse model of human pyelonephritis. J Clin Invest 75:347–360. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 207.Lane MC, Mobley HLT. 2007. Role of P-fimbrial-mediated adherence in pyelonephritis and persistence of uropathogenic Escherichia coli (UPEC) in the mammalian kidney. Kidney Int 72:19–25. [DOI] [PubMed] [Google Scholar]
- 208.Hedlund M, Svensson M, Nilsson A, Duan RD, Svanborg C. 1996. Role of the ceramide-signaling pathway in cytokine responses to P-fimbriated Escherichia coli. J Exp Med 183:1037–1044. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 209.Hedlund M, Wachtler C, Johansson E, Hang L, Somerville JE, Darveau RP, Svanborg C. 1999. P fimbriae-dependent, lipopolysaccharide-independent activation of epithelial cytokine responses. Mol Microbiol 33:693–703. [DOI] [PubMed] [Google Scholar]
- 210.Bergsten G, Samuelsson M, Wullt B, Leijonhufvud I, Fischer H, Svanborg C. 2004. PapG-dependent adherence breaks mucosal inertia and triggers the innate host response. J Infect Dis 189:1734–1742. [DOI] [PubMed] [Google Scholar]
- 211.Bergsten G, Wullt B, Svanborg C. 2005. Escherichia coli, fimbriae, bacterial persistence and host response induction in the human urinary tract. Int J Med Microbiol 295:487–502. [DOI] [PubMed] [Google Scholar]
- 212.Fischer H, Ellström P, Ekström K, Gustafsson L, Gustafsson M, Svanborg C. 2007. Ceramide as a TLR4 agonist; a putative signalling intermediate between sphingolipid receptors for microbial ligands and TLR4. Cell Microbiol 9:1239–1251. [DOI] [PubMed] [Google Scholar]
- 213.Rice JC, Peng T, Spence JS, Wang HQ, Goldblum RM, Corthésy B, Nowicki BJ. 2005. Pyelonephritic Escherichia coli expressing P fimbriae decrease immune response of the mouse kidney. J Am Soc Nephrol 16:3583–3591. [DOI] [PubMed] [Google Scholar]
- 214.Zhang JP, Normark S. 1996. Induction of gene expression in Escherichia coli after pilus-mediated adherence. Science 273:1234–1236. [DOI] [PubMed] [Google Scholar]
- 215.Bernier C, Gounon P, Le Bouguénec C. 2002. Identification of an aggregative adhesion fimbria (AAF) type III-encoding operon in enteroaggregative Escherichia coli as a sensitive probe for detecting the AAF-encoding operon family. Infect Immun 70:4302–4311. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 216.Boisen N, Struve C, Scheutz F, Krogfelt KA, Nataro JP. 2008. New adhesin of enteroaggregative Escherichia coli related to the Afa/Dr/AAF family. Infect Immun 76:3281–3292. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 217.Czeczulin JR, Balepur S, Hicks S, Phillips A, Hall R, Kothary MH, Navarro-Garcia F, Nataro JP. 1997. Aggregative adherence fimbria II, a second fimbrial antigen mediating aggregative adherence in enteroaggregative Escherichia coli. Infect Immun 65:4135–4145. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 218.Nataro JP, Yikang D, Giron JA, Savarino SJ, Kothary MH, Hall R. 1993. Aggregative adherence fimbria I expression in enteroaggregative Escherichia coli requires two unlinked plasmid regions. Infect Immun 61:1126–1131. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 219.Donnenberg M. 2013. Escherichia coli: Pathotypes and Principles of Pathogenesis. Academic Press, San Diego, CA. [Google Scholar]
- 220.Sheikh J, Czeczulin JR, Harrington S, Hicks S, Henderson IR, Le Bouguénec C, Gounon P, Phillips A, Nataro JP. 2002. A novel dispersin protein in enteroaggregative Escherichia coli. J Clin Invest 110:1329–1337. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 221.Nataro JP, Deng Y, Maneval DR, German AL, Martin WC, Levine MM. 1992. Aggregative adherence fimbriae I of enteroaggregative Escherichia coli mediate adherence to HEp-2 cells and hemagglutination of human erythrocytes. Infect Immun 60:2297–2304. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 222.Torres AG, Zhou X, Kaper JB. 2005. Adherence of diarrheagenic Escherichia coli strains to epithelial cells. Infect Immun 73:18–29. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 223.Elias WP Jr, Czeczulin JR, Henderson IR, Trabulsi LR, Nataro JP. 1999. Organization of biogenesis genes for aggregative adherence fimbria II defines a virulence gene cluster in enteroaggregative Escherichia coli. J Bacteriol 181:1779–1785. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 224.Farfan MJ, Inman KG, Nataro JP. 2008. The major pilin subunit of the AAF/II fimbriae from enteroaggregative Escherichia coli mediates binding to extracellular matrix proteins. Infect Immun 76:4378–4384. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 225.Harrington SM, Dudley EG, Nataro JP. 2006. Pathogenesis of enteroaggregative Escherichia coli infection. FEMS Microbiol Lett 254:12–18. [DOI] [PubMed] [Google Scholar]
- 226.Steiner TS, Lima AA, Nataro JP, Guerrant RL. 1998. Enteroaggregative Escherichia coli produce intestinal inflammation and growth impairment and cause interleukin-8 release from intestinal epithelial cells. J Infect Dis 177:88–96. [DOI] [PubMed] [Google Scholar]
- 227.Väisänen-Rhen V. 1984. Fimbria-like hemagglutinin of Escherichia coli O75 strains. Infect Immun 46:401–407. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 228.Nowicki B, Svanborg-Edén C, Hull R, Hull S. 1989. Molecular analysis and epidemiology of the Dr hemagglutinin of uropathogenic Escherichia coli. Infect Immun 57:446–451. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 229.Nowicki B, Barrish JP, Korhonen T, Hull RA, Hull SI. 1987. Molecular cloning of the Escherichia coli O75X adhesin. Infect Immun 55:3168–3173. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 230.Bilge SS, Apostol JM Jr, Fullner KJ, Moseley SL. 1993. Transcriptional organization of the F1845 fimbrial adhesin determinant of Escherichia coli. Mol Microbiol 7:993–1006. [DOI] [PubMed] [Google Scholar]
- 231.Bilge SS, Clausen CR, Lau W, Moseley SL. 1989. Molecular characterization of a fimbrial adhesin, F1845, mediating diffuse adherence of diarrhea-associated Escherichia coli to HEp-2 cells. J Bacteriol 171:4281–4289. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 232.Pham TQ, Goluszko P, Popov V, Nowicki S, Nowicki BJ. 1997. Molecular cloning and characterization of Dr-II, a nonfimbrial adhesin-I-like adhesin isolated from gestational pyelonephritis-associated Escherichia coli that binds to decay-accelerating factor. Infect Immun 65:4309–4318. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 233.Nowicki B, Moulds J, Hull R, Hull S. 1988. A hemagglutinin of uropathogenic Escherichia coli recognizes the Dr blood group antigen. Infect Immun 56:1057–1060. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 234.Westerlund B, Kuusela P, Risteli J, Risteli L, Vartio T, Rauvala H, Virkola R, Korhonen TK. 1989. The O75X adhesin of uropathogenic Escherichia coli is a type IV collagen-binding protein. Mol Microbiol 3:329–337. [DOI] [PubMed] [Google Scholar]
- 235.Westerlund B, Korhonen TK. 1993. Bacterial proteins binding to the mammalian extracellular matrix. Mol Microbiol 9:687–694. [DOI] [PubMed] [Google Scholar]
- 236.Fujita T, Inoue T, Ogawa K, Iida K, Tamura N. 1987. The mechanism of action of decay-accelerating factor (DAF). DAF inhibits the assembly of C3 convertases by dissociating C2a and Bb. J Exp Med 166:1221–1228. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 237.Lublin DM, Atkinson JP. 1989. Decay-accelerating factor: biochemistry, molecular biology, and function. Annu Rev Immunol 7:35–58. [DOI] [PubMed] [Google Scholar]
- 238.Berger CN, Billker O, Meyer TF, Servin AL, Kansau I. 2004. Differential recognition of members of the carcinoembryonic antigen family by Afa/Dr adhesins of diffusely adhering Escherichia coli (Afa/Dr DAEC). Mol Microbiol 52:963–983. [DOI] [PubMed] [Google Scholar]
- 239.Muenzner P, Kengmo Tchoupa A, Klauser B, Brunner T, Putze J, Dobrindt U, Hauck CR. 2016. Uropathogenic E. coli exploit CEA to promote colonization of the urogenital tract mucosa. PLoS Pathog 12:e1005608. 10.1371/journal.ppat.1005608. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 240.Kaul AK, Khan S, Martens MG, Crosson JT, Lupo VR, Kaul R. 1999. Experimental gestational pyelonephritis induces preterm births and low birth weights in C3H/HeJ mice. Infect Immun 67:5958–5966. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 241.Goluszko P, Niesel D, Nowicki B, Selvarangan R, Nowicki S, Hart A, Pawelczyk E, Das M, Urvil P, Hasan R. 2001. Dr operon-associated invasiveness of Escherichia coli from pregnant patients with pyelonephritis. Infect Immun 69:4678–4680. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 242.Hart A, Nowicki BJ, Reisner B, Pawelczyk E, Goluszko P, Urvil P, Anderson G, Nowicki S. 2001. Ampicillin-resistant Escherichia coli in gestational pyelonephritis: increased occurrence and association with the colonization factor Dr adhesin. J Infect Dis 183:1526–1529. [DOI] [PubMed] [Google Scholar]
- 243.Bétis F, Brest P, Hofman V, Guignot J, Bernet-Camard MF, Rossi B, Servin A, Hofman P. 2003. The Afa/Dr adhesins of diffusely adhering Escherichia coli stimulate interleukin-8 secretion, activate mitogen-activated protein kinases, and promote polymorphonuclear transepithelial migration in T84 polarized epithelial cells. Infect Immun 71:1068–1074. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 244.Bétis F, Brest P, Hofman V, Guignot J, Kansau I, Rossi B, Servin A, Hofman P. 2003. Afa/Dr diffusely adhering Escherichia coli infection in T84 cell monolayers induces increased neutrophil transepithelial migration, which in turn promotes cytokine-dependent upregulation of decay-accelerating factor (CD55), the receptor for Afa/Dr adhesins. Infect Immun 71:1774–1783. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 245.Cane G, Moal VL, Pagès G, Servin AL, Hofman P, Vouret-Craviari V. 2007. Up-regulation of intestinal vascular endothelial growth factor by Afa/Dr diffusely adhering Escherichia coli. PLoS One 2:e1359. 10.1371/journal.pone.0001359. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 246.Hacker J, Morschhauser J. 1994. S and F1C Fimbriae, p 27–36. In Klemm P (ed), Fimbriae: Adhesion, Genetics, Biogenesis, and Vaccines. CRC Press, Boca Raton. [Google Scholar]
- 247.Pere A, Leinonen M, Väisänen-Rhen V, Rhen M, Korhonen TK. 1985. Occurrence of type-1C fimbriae on Escherichia coli strains isolated from human extraintestinal infections. J Gen Microbiol 131:1705–1711. [DOI] [PubMed] [Google Scholar]
- 248.Khan AS, Kniep B, Oelschlaeger TA, Van Die I, Korhonen T, Hacker J. 2000. Receptor structure for F1C fimbriae of uropathogenic Escherichia coli. Infect Immun 68:3541–3547. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 249.van Die I, van Megen I, Hoekstra W, Bergmans H. 1984. Molecular organisation of the genes involved in the production of F7(2) fimbriae, causing mannose-resistant haemagglutination, of a uropathogenic Escherichia coli 06:K2:H1:F7 strain. Mol Gen Genet 194:528–533. [DOI] [PubMed] [Google Scholar]
- 250.Virkola R, Westerlund B, Holthöfer H, Parkkinen J, Kekomäki M, Korhonen TK. 1988. Binding characteristics of Escherichia coli adhesins in human urinary bladder. Infect Immun 56:2615–2622. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 251.Klemm P, Christiansen G, Kreft B, Marre R, Bergmans H. 1994. Reciprocal exchange of minor components of type 1 and F1C fimbriae results in hybrid organelles with changed receptor specificities. J Bacteriol 176:2227–2234. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 252.Klemm P, Jørgensen BJ, Kreft B, Christiansen G. 1995. The export systems of type 1 and F1C fimbriae are interchangeable but work in parental pairs. J Bacteriol 177:621–627. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 253.Kreft B, Carstensen O, Straube E, Bohnet S, Hacker J, Marre R. 1992. Adherence to and cytotoxicity of Escherichia coli for eucaryotic cell lines quantified by MTT (3-[4,5-dimethylthiazol-2-yl]-2,5-diphenyltetrazolium bromide). Zentralbl Bakteriol 276:231–242. [DOI] [PubMed] [Google Scholar]
- 254.Marre R, Kreft B, Hacker J. 1990. Genetically engineered S and F1C fimbriae differ in their contribution to adherence of Escherichia coli to cultured renal tubular cells. Infect Immun 58:3434–3437. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 255.Lasaro MA, Salinger N, Zhang J, Wang Y, Zhong Z, Goulian M, Zhu J. 2009. F1C fimbriae play an important role in biofilm formation and intestinal colonization by the Escherichia coli commensal strain Nissle 1917. Appl Environ Microbiol 75:246–251. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 256.Bäckhed F, Alsén B, Roche N, Angström J, von Euler A, Breimer ME, Westerlund-Wikström B, Teneberg S, Richter-Dahlfors A. 2002. Identification of target tissue glycosphingolipid receptors for uropathogenic, F1C-fimbriated Escherichia coli and its role in mucosal inflammation. J Biol Chem 277:18198–18205. [DOI] [PubMed] [Google Scholar]
- 257.Korhonen TK, Valtonen MV, Parkkinen J, Väisänen-Rhen V, Finne J, Orskov F, Orskov I, Svenson SB, Mäkelä PH. 1985. Serotypes, hemolysin production, and receptor recognition of Escherichia coli strains associated with neonatal sepsis and meningitis. Infect Immun 48:486–491. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 258.Hacker J. 1992. Role of fimbrial adhesins in the pathogenesis of Escherichia coli infections. Can J Microbiol 38:720–727. [DOI] [PubMed] [Google Scholar]
- 259.Korhonen TK, Parkkinen J, Hacker J, Finne J, Pere A, Rhen M, Holthöfer H. 1986. Binding of Escherichia coli S fimbriae to human kidney epithelium. Infect Immun 54:322–327. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 260.Hacker J, Schmidt G, Hughes C, Knapp S, Marget M, Goebel W. 1985. Cloning and characterization of genes involved in production of mannose-resistant, neuraminidase-susceptible (X) fimbriae from a uropathogenic O6:K15:H31 Escherichia coli strain. Infect Immun 47:434–440. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 261.Schmoll T, Morschhäuser J, Ott M, Ludwig B, van Die I, Hacker J. 1990. Complete genetic organization and functional aspects of the Escherichia coli S fimbrial adhesion determinant: nucleotide sequence of the genes sfa B, C, D, E, F. Microb Pathog 9:331–343. [DOI] [PubMed] [Google Scholar]
- 262.Morschhäuser J, Vetter V, Korhonen T, Uhlin BE, Hacker J. 1993. Regulation and binding properties of S fimbriae cloned from E. coli strains causing urinary tract infection and meningitis. Zentralbl Bakteriol 278:165–176. [DOI] [PubMed] [Google Scholar]
- 263.Dobrindt U, Blum-Oehler G, Hartsch T, Gottschalk G, Ron EZ, Fünfstück R, Hacker J. 2001. S-Fimbria-encoding determinant sfa(I) is located on pathogenicity island III(536) of uropathogenic Escherichia coli strain 536. Infect Immun 69:4248–4256. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 264.Knight SD, Choudhury D, Hultgren S, Pinkner J, Stojanoff V, Thompson A. 2002. Structure of the S pilus periplasmic chaperone SfaE at 2.2 A resolution. Acta Crystallogr D Biol Crystallogr 58:1016–1022. [DOI] [PubMed] [Google Scholar]
- 265.Korhonen TK, Väisänen-Rhen V, Rhen M, Pere A, Parkkinen J, Finne J. 1984. Escherichia coli fimbriae recognizing sialyl galactosides. J Bacteriol 159:762–766. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 266.Wang Y, Wen ZG, Kim KS. 2004. Role of S fimbriae in Escherichia coli K1 binding to brain microvascular endothelial cells in vitro and penetration into the central nervous system in vivo. Microb Pathog 37:287–293. [DOI] [PubMed] [Google Scholar]
- 267.Sarén A, Virkola R, Hacker J, Korhonen TK. 1999. The cellular form of human fibronectin as an adhesion target for the S fimbriae of meningitis-associated Escherichia coli. Infect Immun 67:2671–2676. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 268.Virkola R, Parkkinen J, Hacker J, Korhonen TK. 1993. Sialyloligosaccharide chains of laminin as an extracellular matrix target for S fimbriae of Escherichia coli. Infect Immun 61:4480–4484. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 269.Parkkinen J, Hacker J, Korhonen TK. 1991. Enhancement of tissue plasminogen activator-catalyzed plasminogen activation by Escherichia coli S fimbriae associated with neonatal septicaemia and meningitis. Thromb Haemost 65:483–486. [PubMed] [Google Scholar]
- 270.Perez-Casal J, Swartley JS, Scott JR. 1990. Gene encoding the major subunit of CS1 pili of human enterotoxigenic Escherichia coli. Infect Immun 58:3594–3600. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 271.Caron J, Coffield LM, Scott JR. 1989. A plasmid-encoded regulatory gene, rns, required for expression of the CS1 and CS2 adhesins of enterotoxigenic Escherichia coli. Proc Natl Acad Sci USA 86:963–967. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 272.Murphree D, Froehlich B, Scott JR. 1997. Transcriptional control of genes encoding CS1 pili: negative regulation by a silencer and positive regulation by Rns. J Bacteriol 179:5736–5743. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 273.Jordi BJ, Willshaw GA, van der Zeijst BA, Gaastra W. 1992. The complete nucleotide sequence of region 1 of the CFA/I fimbrial operon of human enterotoxigenic Escherichia coli. DNA Seq 2:257–263. [DOI] [PubMed] [Google Scholar]
- 274.Smith HR, Willshaw GA, Rowe B. 1982. Mapping of a plasmid, coding for colonization, factor antigen I and heat-stable enterotoxin production, isolated from an enterotoxigenic strain of Escherichia coli. J Bacteriol 149:264–275. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 275.Savelkoul PH, Willshaw GA, McConnell MM, Smith HR, Hamers AM, van der Zeijst BA, Gaastra W. 1990. Expression of CFA/I fimbriae is positively regulated. Microb Pathog 8:91–99. [DOI] [PubMed] [Google Scholar]
- 276.Jordi BJ, Dagberg B, de Haan LA, Hamers AM, van der Zeijst BA, Gaastra W, Uhlin BE. 1992. The positive regulator CfaD overcomes the repression mediated by histone-like protein H-NS (H1) in the CFA/I fimbrial operon of Escherichia coli. EMBO J 11:2627–2632. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 277.Gaastra W, Svennerholm AM. 1996. Colonization factors of human enterotoxigenic Escherichia coli (ETEC). Trends Microbiol 4:444–452. [DOI] [PubMed] [Google Scholar]
- 278.Li YF, Poole S, Rasulova F, McVeigh AL, Savarino SJ, Xia D. 2007. A receptor-binding site as revealed by the crystal structure of CfaE, the colonization factor antigen I fimbrial adhesin of enterotoxigenic Escherichia coli. J Biol Chem 282:23970–23980. [DOI] [PubMed] [Google Scholar]
- 279.Sakellaris H, Balding DP, Scott JR. 1996. Assembly proteins of CS1 pili of enterotoxigenic Escherichia coli. Mol Microbiol 21:529–541. [DOI] [PubMed] [Google Scholar]
- 280.Poole ST, McVeigh AL, Anantha RP, Lee LH, Akay YM, Pontzer EA, Scott DA, Bullitt E, Savarino SJ. 2007. Donor strand complementation governs intersubunit interaction of fimbriae of the alternate chaperone pathway. Mol Microbiol 63:1372–1384. [DOI] [PubMed] [Google Scholar]
- 281.Bühler T, Hoschützky H, Jann K. 1991. Analysis of colonization factor antigen I, an adhesin of enterotoxigenic Escherichia coli O78:H11: fimbrial morphology and location of the receptor-binding site. Infect Immun 59:3876–3882. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 282.Jansson L, Tobias J, Lebens M, Svennerholm AM, Teneberg S. 2006. The major subunit, CfaB, of colonization factor antigen i from enterotoxigenic Escherichia coli is a glycosphingolipid binding protein. Infect Immun 74:3488–3497. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 283.Marron MB, Smyth CJ. 1995. Molecular analysis of the cso operon of enterotoxigenic Escherichia coli reveals that CsoA is the adhesin of CS1 fimbriae and that the accessory genes are interchangeable with those of the cfa operon. Microbiology 141:2849–2859. [DOI] [PubMed] [Google Scholar]
- 284.Sakellaris H, Penumalli VR, Scott JR. 1999. The level of expression of the minor pilin subunit, CooD, determines the number of CS1 pili assembled on the cell surface of Escherichia coli. J Bacteriol 181:1694–1697. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 285.Levine MM, Kaper JB, Black RE, Clements ML. 1983. New knowledge on pathogenesis of bacterial enteric infections as applied to vaccine development. Microbiol Rev 47:510–550. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 286.Nagy B, Fekete PZ. 1999. Enterotoxigenic Escherichia coli (ETEC) in farm animals. Vet Res 30:259–284. [PubMed] [Google Scholar]
- 287.Boedeker EC. 2005. Enteric infections. Curr Opin Gastroenterol 21:1–3. [PubMed] [Google Scholar]
- 288.Qadri F, Svennerholm AM, Faruque AS, Sack RB. 2005. Enterotoxigenic Escherichia coli in developing countries: epidemiology, microbiology, clinical features, treatment, and prevention. Clin Microbiol Rev 18:465–483. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 289.Humphries AD, Raffatellu M, Winter S, Weening EH, Kingsley RA, Droleskey R, Zhang S, Figueiredo J, Khare S, Nunes J, Adams LG, Tsolis RM, Bäumler AJ. 2003. The use of flow cytometry to detect expression of subunits encoded by 11 Salmonella enterica serotype Typhimurium fimbrial operons. Mol Microbiol 48:1357–1376. [DOI] [PubMed] [Google Scholar]
- 290.Bäumler AJ, Tsolis RM, Heffron F. 1996. The lpf fimbrial operon mediates adhesion of Salmonella typhimurium to murine Peyer’s patches. Proc Natl Acad Sci USA 93:279–283. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 291.Tsolis RM, Townsend SM, Miao EA, Miller SI, Ficht TA, Adams LG, Bäumler AJ. 1999. Identification of a putative Salmonella enterica serotype typhimurium host range factor with homology to IpaH and YopM by signature-tagged mutagenesis. Infect Immun 67:6385–6393. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 292.Weening EH, Barker JD, Laarakker MC, Humphries AD, Tsolis RM, Bäumler AJ. 2005. The Salmonella enterica serotype Typhimurium lpf, bcf, stb, stc, std, and sth fimbrial operons are required for intestinal persistence in mice. Infect Immun 73:3358–3366. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 293.Bäumler AJ, Heffron F. 1995. Identification and sequence analysis of lpfABCDE, a putative fimbrial operon of Salmonella typhimurium. J Bacteriol 177:2087–2097. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 294.Darwin KH, Miller VL. 1999. Molecular basis of the interaction of Salmonella with the intestinal mucosa. Clin Microbiol Rev 12:405–428. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 295.Bäumler AJ, Tsolis RM, Heffron F. 1997. Fimbrial adhesins of Salmonella typhimurium. Role in bacterial interactions with epithelial cells. Adv Exp Med Biol 412:149–158. [PubMed] [Google Scholar]
- 296.Bäumler AJ, Tsolis RM, Valentine PJ, Ficht TA, Heffron F. 1997. Synergistic effect of mutations in invA and lpfC on the ability of Salmonella typhimurium to cause murine typhoid. Infect Immun 65:2254–2259. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 297.Norris TL, Kingsley RA, Bümler AJ. 1998. Expression and transcriptional control of the Salmonella typhimurium Ipf fimbrial operon by phase variation. Mol Microbiol 29:311–320. [DOI] [PubMed] [Google Scholar]
- 298.Friedrich MJ, Kinsey NE, Vila J, Kadner RJ. 1993. Nucleotide sequence of a 13.9 kb segment of the 90 kb virulence plasmid of Salmonella typhimurium: the presence of fimbrial biosynthetic genes. Mol Microbiol 8:543–558. [DOI] [PubMed] [Google Scholar]
- 299.Bäumler AJ, Tsolis RM, Bowe FA, Kusters JG, Hoffmann S, Heffron F. 1996. The pef fimbrial operon of Salmonella typhimurium mediates adhesion to murine small intestine and is necessary for fluid accumulation in the infant mouse. Infect Immun 64:61–68. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 300.Bäumler AJ, Tsolis RM, Heffron F. 1996. Contribution of fimbrial operons to attachment to and invasion of epithelial cell lines by Salmonella typhimurium. Infect Immun 64:1862–1865. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 301.Bäumler AJ, Gilde AJ, Tsolis RM, van der Velden AW, Ahmer BM, Heffron F. 1997. Contribution of horizontal gene transfer and deletion events to development of distinctive patterns of fimbrial operons during evolution of Salmonella serotypes. J Bacteriol 179:317–322. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 302.Rotger R, Casadesús J. 1999. The virulence plasmids of Salmonella. Int Microbiol 2:177–184. [PubMed] [Google Scholar]
- 303.Clouthier SC, Müller KH, Doran JL, Collinson SK, Kay WW. 1993. Characterization of three fimbrial genes, sefABC, of Salmonella enteritidis. J Bacteriol 175:2523–2533. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 304.Clouthier SC, Collinson SK, Kay WW. 1994. Unique fimbriae-like structures encoded by sefD of the SEF14 fimbrial gene cluster of Salmonella enteritidis. Mol Microbiol 12:893–901. [DOI] [PubMed] [Google Scholar]
- 305.Edwards RA, Schifferli DM, Maloy SR. 2000. A role for Salmonella fimbriae in intraperitoneal infections. Proc Natl Acad Sci USA 97:1258–1262. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 306.Zeiner SA, Dwyer BE, Clegg S. 2012. FimA, FimF, and FimH are necessary for assembly of type 1 fimbriae on Salmonella enterica serovar Typhimurium. Infect Immun 80:3289–3296. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 307.Boyd EF, Hartl DL. 1999. Analysis of the type 1 pilin gene cluster fim in Salmonella: its distinct evolutionary histories in the 5′ and 3′ regions. J Bacteriol 181:1301–1308. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 308.Duguid JP, Campbell I. 1969. Antigens of the type-1 fimbriae of salmonellae and other enterobacteria. J Med Microbiol 2:535–553. [DOI] [PubMed] [Google Scholar]
- 309.Lockman HA, Curtiss R III. 1992. Virulence of non-type 1-fimbriated and nonfimbriated nonflagellated Salmonella typhimurium mutants in murine typhoid fever. Infect Immun 60:491–496. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 310.van der Velden AW, Bäumler AJ, Tsolis RM, Heffron F. 1998. Multiple fimbrial adhesins are required for full virulence of Salmonella typhimurium in mice. Infect Immun 66:2803–2808. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 311.Dibb-Fuller MP, Allen-Vercoe E, Thorns CJ, Woodward MJ. 1999. Fimbriae- and flagella-mediated association with and invasion of cultured epithelial cells by Salmonella enteritidis. Microbiology 145:1023–1031. [DOI] [PubMed] [Google Scholar]
- 312.Dibb-Fuller MP, Woodward MJ. 2000. Contribution of fimbriae and flagella of Salmonella enteritidis to colonization and invasion of chicks. Avian Pathol 29:295–304. [DOI] [PubMed] [Google Scholar]
- 313.De Buck J, Van Immerseel F, Haesebrouck F, Ducatelle R. 2004. Effect of type 1 fimbriae of Salmonella enterica serotype Enteritidis on bacteraemia and reproductive tract infection in laying hens. Avian Pathol 33:314–320. [DOI] [PubMed] [Google Scholar]
- 314.Kuźmińska-Bajor M, Grzymajło K, Ugorski M. 2015. Type 1 fimbriae are important factors limiting the dissemination and colonization of mice by Salmonella Enteritidis and contribute to the induction of intestinal inflammation during Salmonella invasion. Front Microbiol 6:276. 10.3389/fmicb.2015.00276. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 315.Kline KA, Fälker S, Dahlberg S, Normark S, Henriques-Normark B. 2009. Bacterial adhesins in host-microbe interactions. Cell Host Microbe 5:580–592. [DOI] [PubMed] [Google Scholar]
- 316.Cusumano CK, Hultgren SJ. 2009. Bacterial adhesion--a source of alternate antibiotic targets. IDrugs 12:699–705. [PubMed] [Google Scholar]
- 317.Steadman D, Lo A, Waksman G, Remaut H. 2014. Bacterial surface appendages as targets for novel antibacterial therapeutics. Future Microbiol 9:887–900. [DOI] [PubMed] [Google Scholar]
- 318.Allen RC, Popat R, Diggle SP, Brown SP. 2014. Targeting virulence: can we make evolution-proof drugs? Nat Rev Microbiol 12:300–308. [DOI] [PubMed] [Google Scholar]
- 319.Zambelloni R, Marquez R, Roe AJ. 2015. Development of antivirulence compounds: a biochemical review. Chem Biol Drug Des 85:43–55. [DOI] [PubMed] [Google Scholar]
- 320.Cegelski L, Marshall GR, Eldridge GR, Hultgren SJ. 2008. The biology and future prospects of antivirulence therapies. Nat Rev Microbiol 6:17–27. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 321.World Health Organization. 2014. Antimcribial Resistance: Global Report on Surveillance. WHO Press. [Google Scholar]
- 322.Centers for Disease Control and Prevention. 2013. Antibiotic Resistance Threats in the United States, 2013. https://www.cdc.gov/drugresistance/pdf/ar-threats-2013-508.pdf
- 323.Fauci AS, Morens DM. 2012. The perpetual challenge of infectious diseases. N Engl J Med 366:454–461. [DOI] [PubMed] [Google Scholar]
- 324.Spellberg B, Blaser M, Guidos RJ, Boucher HW, Bradley JS, Eisenstein BI, Gerding D, Lynfield R, Reller LB, Rex J, Schwartz D, Septimus E, Tenover FC, Gilbert DN, Infectious Diseases Society of America (IDSA). 2011. Combating antimicrobial resistance: policy recommendations to save lives. Clin Infect Dis 52(Suppl 5):S397–S428. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 325.Thankavel K, Madison B, Ikeda T, Malaviya R, Shah AH, Arumugam PM, Abraham SN. 1997. Localization of a domain in the FimH adhesin of Escherichia coli type 1 fimbriae capable of receptor recognition and use of a domain-specific antibody to confer protection against experimental urinary tract infection. J Clin Invest 100:1123–1136. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 326.Roberts JA, Kaack MB, Baskin G, Chapman MR, Hunstad DA, Pinkner JS, Hultgren SJ. 2004. Antibody responses and protection from pyelonephritis following vaccination with purified Escherichia coli PapDG protein. J Urol 171:1682–1685. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 327.Poggio TV, La Torre JL, Scodeller EA. 2006. Intranasal immunization with a recombinant truncated FimH adhesin adjuvanted with CpG oligodeoxynucleotides protects mice against uropathogenic Escherichia coli challenge. Can J Microbiol 52:1093–1102. [DOI] [PubMed] [Google Scholar]
- 328.Tchesnokova V, Aprikian P, Kisiela D, Gowey S, Korotkova N, Thomas W, Sokurenko E. 2011. Type 1 fimbrial adhesin FimH elicits an immune response that enhances cell adhesion of Escherichia coli. Infect Immun 79:3895–3904. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 329.Cusumano CK, Pinkner JS, Han Z, Greene SE, Ford BA, Crowley JR, Henderson JP, Janetka JW, Hultgren SJ. 2011. Treatment and prevention of urinary tract infection with orally active FimH inhibitors. Sci Transl Med 3:109ra115. 10.1126/scitranslmed.3003021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 330.Klein T, Abgottspon D, Wittwer M, Rabbani S, Herold J, Jiang X, Kleeb S, Lüthi C, Scharenberg M, Bezençon J, Gubler E, Pang L, Smiesko M, Cutting B, Schwardt O, Ernst B. 2010. FimH antagonists for the oral treatment of urinary tract infections: from design and synthesis to in vitro and in vivo evaluation. J Med Chem 53:8627–8641. [DOI] [PubMed] [Google Scholar]
- 331.Totsika M, Kostakioti M, Hannan TJ, Upton M, Beatson SA, Janetka JW, Hultgren SJ, Schembri MA. 2013. A FimH inhibitor prevents acute bladder infection and treats chronic cystitis caused by multidrug-resistant uropathogenic Escherichia coli ST131. J Infect Dis 208:921–928. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 332.Guiton PS, Cusumano CK, Kline KA, Dodson KW, Han Z, Janetka JW, Henderson JP, Caparon MG, Hultgren SJ. 2012. Combinatorial small-molecule therapy prevents uropathogenic Escherichia coli catheter-associated urinary tract infections in mice. Antimicrob Agents Chemother 56:4738–4745. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 333.Salminen A, Loimaranta V, Joosten JA, Khan AS, Hacker J, Pieters RJ, Finne J. 2007. Inhibition of P-fimbriated Escherichia coli adhesion by multivalent galabiose derivatives studied by a live-bacteria application of surface plasmon resonance. J Antimicrob Chemother 60:495–501. [DOI] [PubMed] [Google Scholar]
- 334.Pinkner JS, Remaut H, Buelens F, Miller E, Aberg V, Pemberton N, Hedenström M, Larsson A, Seed P, Waksman G, Hultgren SJ, Almqvist F. 2006. Rationally designed small compounds inhibit pilus biogenesis in uropathogenic bacteria. Proc Natl Acad Sci USA 103:17897–17902. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 335.Svensson A, Larsson A, Emtenäs H, Hedenström M, Fex T, Hultgren SJ, Pinkner JS, Almqvist F, Kihlberg J. 2001. Design and evaluation of pilicides: potential novel antibacterial agents directed against uropathogenic Escherichia coli. ChemBioChem 2:915–918. [DOI] [PubMed] [Google Scholar]
- 336.Chorell E, Pinkner JS, Bengtsson C, Banchelin TS, Edvinsson S, Linusson A, Hultgren SJ, Almqvist F. 2012. Mapping pilicide anti-virulence effect in Escherichia coli, a comprehensive structure-activity study. Bioorg Med Chem 20:3128–3142. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 337.Greene SE, Pinkner JS, Chorell E, Dodson KW, Shaffer CL, Conover MS, Livny J, Hadjifrangiskou M, Almqvist F, Hultgren SJ. 2014. Pilicide ec240 disrupts virulence circuits in uropathogenic Escherichia coli. MBio 5:e02038. 10.1128/mBio.02038-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 338.Cegelski L, Pinkner JS, Hammer ND, Cusumano CK, Hung CS, Chorell E, Aberg V, Walker JN, Seed PC, Almqvist F, Chapman MR, Hultgren SJ. 2009. Small-molecule inhibitors target Escherichia coli amyloid biogenesis and biofilm formation. Nat Chem Biol 5:913–919. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 339.Lo AW, Van de Water K, Gane PJ, Chan AW, Steadman D, Stevens K, Selwood DL, Waksman G, Remaut H. 2014. Suppression of type 1 pilus assembly in uropathogenic Escherichia coli by chemical inhibition of subunit polymerization. J Antimicrob Chemother 69:1017–1026. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 340.Shamir ER, Warthan M, Brown SP, Nataro JP, Guerrant RL, Hoffman PS. 2010. Nitazoxanide inhibits biofilm production and hemagglutination by enteroaggregative Escherichia coli strains by blocking assembly of AafA fimbriae. Antimicrob Agents Chemother 54:1526–1533. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 341.Chahales P, Hoffman PS, Thanassi DG. 2016. Nitazoxanide Inhibits Pilus Biogenesis by Interfering with Folding of the Usher Protein in the Outer Membrane. Antimicrob Agents Chemother 60:2028–2038. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 342.Klinth JE, Pinkner JS, Hultgren SJ, Almqvist F, Uhlin BE, Axner O. 2012. Impairment of the biomechanical compliance of P pili: a novel means of inhibiting uropathogenic bacterial infections? Eur Biophys J 41:285–295. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 343.Van Loy CP, Sokurenko EV, Moseley SL. 2002. The major structural subunits of Dr and F1845 fimbriae are adhesins. Infect Immun 70:1694–1702. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 344.Savarino SJ, Fox P, Deng Y, Nataro JP. 1994. Identification and characterization of a gene cluster mediating enteroaggregative Escherichia coli aggregative adherence fimbria I biogenesis. J Bacteriol 176:4949–4957. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 345.Garcia M-I, Labigne A, Le Bouguenec C. 1994. Nucleotide sequence of the afimbrial-adhesin-encoding afa-3 gene cluster and its translocation via flanking IS1 insertion sequences. J Bacteriol 176:7601–7613. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 346.Lalioui L, Le Bouguénec C. 2001. afa-8 Gene cluster is carried by a pathogenicity island inserted into the tRNA(Phe) of human and bovine pathogenic Escherichia coli isolates. Infect Immun 69:937–948. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 347.Cantey JR, Blake RK, Williford JR, Moseley SL. 1999. Characterization of the Escherichia coli AF/R1 pilus operon: novel genes necessary for transcriptional regulation and for pilus-mediated adherence. Infect Immun 67:2292–2298. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 348.Buckles EL, Bahrani-Mougeot FK, Molina A, Lockatell CV, Johnson DE, Drachenberg CB, Burland V, Blattner FR, Donnenberg MS. 2004. Identification and characterization of a novel uropathogenic Escherichia coli-associated fimbrial gene cluster. Infect Immun 72:3890–3901. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 349.Korea CG, Ghigo JM, Beloin C. 2011. The sweet connection: Solving the riddle of multiple sugar-binding fimbrial adhesins in Escherichia coli: Multiple E. coli fimbriae form a versatile arsenal of sugar-binding lectins potentially involved in surface-colonisation and tissue tropism. BioEssays 33:300–311. [DOI] [PubMed] [Google Scholar]
- 350.Jalajakumari MB, Thomas CJ, Halter R, Manning PA. 1989. Genes for biosynthesis and assembly of CS3 pili of CFA/II enterotoxigenic Escherichia coli: novel regulation of pilus production by bypassing an amber codon. Mol Microbiol 3:1685–1695. [DOI] [PubMed] [Google Scholar]
- 351.Duthy TG, Manning PA, Heuzenroeder MW. 2002. Identification and characterization of assembly proteins of CS5 pili from enterotoxigenic Escherichia coli. J Bacteriol 184:1065–1077. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 352.Wolf MK, de Haan LA, Cassels FJ, Willshaw GA, Warren R, Boedeker EC, Gaastra W. 1997. The CS6 colonization factor of human enterotoxigenic Escherichia coli contains two heterologous major subunits. FEMS Microbiol Lett 148:35–42. [DOI] [PubMed] [Google Scholar]
- 353.Del Canto F, Botkin DJ, Valenzuela P, Popov V, Ruiz-Perez F, Nataro JP, Levine MM, Stine OC, Pop M, Torres AG, Vidal R. 2012. Identification of coli surface antigen 23, a novel adhesin of enterotoxigenic Escherichia coli. Infect Immun 80:2791–2801. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 354.Gaastra W, Sommerfelt H, van Dijk L, Kusters JG, Svennerholm A-M, Grewal HMS. 2002. Antigenic variation within the subunit protein of members of the colonization factor antigen I group of fimbrial proteins in human enterotoxigenic Escherichia coli. Int J Med Microbiol 292:43–50. [DOI] [PubMed] [Google Scholar]
- 355.Honarvar S, Choi BK, Schifferli DM. 2003. Phase variation of the 987P-like CS18 fimbriae of human enterotoxigenic Escherichia coli is regulated by site-specific recombinases. Mol Microbiol 48:157–171. [DOI] [PubMed] [Google Scholar]
- 356.Grewal HM, Valvatne H, Bhan MK, van Dijk L, Gaastra W, Sommerfelt H. 1997. A new putative fimbrial colonization factor, CS19, of human enterotoxigenic Escherichia coli. Infect Immun 65:507–513. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 357.Valvatne H, Sommerfelt H, Gaastra W, Bhan MK, Grewal HM. 1996. Identification and characterization of CS20, a new putative colonization factor of enterotoxigenic Escherichia coli. Infect Immun 64:2635–2642. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 358.Bertin Y, Girardeau J-P, Der Vartanian M, Martin C. 1993. The ClpE protein involved in biogenesis of the CS31A capsule-like antigen is a member of a periplasmic chaperone family in gram-negative bacteria. FEMS Microbiol Lett 108:59–67. [DOI] [PubMed] [Google Scholar]
- 359.Keller R, Ordoñez JG, de Oliveira RR, Trabulsi LR, Baldwin TJ, Knutton S. 2002. Afa, a diffuse adherence fibrillar adhesin associated with enteropathogenic Escherichia coli. Infect Immun 70:2681–2689. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 360.Rendón MA, Saldaña Z, Erdem AL, Monteiro-Neto V, Vázquez A, Kaper JB, Puente JL, Girón JA. 2007. Commensal and pathogenic Escherichia coli use a common pilus adherence factor for epithelial cell colonization. Proc Natl Acad Sci USA 104:10637–10642. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 361.Riegman N, van Die I, Leunissen J, Hoekstra W, Bergmans H. 1988. Biogenesis of F71 and F72 fimbriae of uropathogenic Escherichia coli: influence of the FsoF and FstFG proteins and localization of the Fso/FstE protein. Mol Microbiol 2:73–80. [DOI] [PubMed] [Google Scholar]
- 362.van Die I, Spierings G, van Megen I, Zuidweg E, Hoekstra W, Bergmans H. 1985. Cloning and genetic organization of the gene cluster encoding F71 fimbriae of a uropathogenic Escherichia coli and comparison with the F72 gene cluster. FEMS Microbiol Lett 28:329–334. [Google Scholar]
- 363.Riegman N, Hoschützky H, van Die I, Hoekstra W, Jann K, Bergmans H. 1990. Immunocytochemical analysis of P-fimbrial structure: localization of minor subunits and the influence of the minor subunit FsoE on the biogenesis of the adhesin. Mol Microbiol 4:1193–1198. [DOI] [PubMed] [Google Scholar]
- 364.Ulett GC, Mabbett AN, Fung KC, Webb RI, Schembri MA. 2007. The role of F9 fimbriae of uropathogenic Escherichia coli in biofilm formation. Microbiology 153:2321–2331. [DOI] [PubMed] [Google Scholar]
- 365.Conover MS, Ruer S, Taganna J, Kalas V, De Greve H, Pinkner JS, Dodson KW, Remaut H, Hultgren SJ. 2016. Inflammation-induced adhesin-receptor interaction provides a fitness advantage to uropathogenic E. coli during chronic infection. Cell Host Microbe 20:482–492. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 366.Lintermans P, Pohl P, Deboeck F, Bertels A, Schlicker C, Vandekerckhove J, Van Damme J, Van Montagu M, De Greve H. 1988. Isolation and nucleotide sequence of the F17-A gene encoding the structural protein of the F17 fimbriae in bovine enterotoxigenic Escherichia coli. Infect Immun 56:1475–1484. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 367.Imberechts H, et al. 1992. Characterization of F107 fimbriae of Escherichia coli 107/86, which causes edema disease in pigs, and nucleotide sequence of the F107 major fimbrial subunit gene, fedA. Infect Immun 60:1963–1971. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 368.Fairbrother JM, Lallier R, Leblanc L, Jacques M, Larivière S. 1988. Production and purification of Escherichia coli fimbrial antigen F165. FEMS Microbiol Lett 56:247–252. [Google Scholar]
- 369.Bakker D, Vader CE, Roosendaal B, Mooi FR, Oudega B, de Graaf FK. 1991. Structure and function of periplasmic chaperone-like proteins involved in the biosynthesis of K88 and K99 fimbriae in enterotoxigenic Escherichia coli. Mol Microbiol 5:875–886. [DOI] [PubMed] [Google Scholar]
- 370.Scaletsky ICA, Michalski J, Torres AG, Dulguer MV, Kaper JB. 2005. Identification and characterization of the locus for diffuse adherence, which encodes a novel afimbrial adhesin found in atypical enteropathogenic Escherichia coli. Infect Immun 73:4753–4765. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 371.Ahrens R, Ott M, Ritter A, Hoschützky H, Bühler T, Lottspeich F, Boulnois GJ, Jann K, Hacker J. 1993. Genetic analysis of the gene cluster encoding nonfimbrial adhesin I from an Escherichia coli uropathogen. Infect Immun 61:2505–2512. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 372.Oh KH, Kim DW, Jung SM, Cho SH. 2014. Molecular characterization of Enterotoxigenic Escherichia coli strains isolated from diarrheal patients in Korea during 2003–2011. PLoS One 9:e96896. 10.1371/journal.pone.0096896. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 373.Klemm P, Roos V, Ulett GC, Svanborg C, Schembri MA. 2006. Molecular characterization of the Escherichia coli asymptomatic bacteriuria strain 83972: the taming of a pathogen. Infect Immun 74:781–785. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 374.Adams LM, Simmons CP, Rezmann L, Strugnell RA, Robins-Browne RM. 1997. Identification and characterization of a K88- and CS31A-like operon of a rabbit enteropathogenic Escherichia coli strain which encodes fimbriae involved in the colonization of rabbit intestine. Infect Immun 65:5222–5230. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 375.Lymberopoulos MH, Houle S, Daigle F, Léveillé S, Brée A, Moulin-Schouleur M, Johnson JR, Dozois CM. 2006. Characterization of Stg fimbriae from an avian pathogenic Escherichia coli O78:K80 strain and assessment of their contribution to colonization of the chicken respiratory tract. J Bacteriol 188:6449–6459. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 376.Spurbeck RR, Stapleton AE, Johnson JR, Walk ST, Hooton TM, Mobley HL. 2011. Fimbrial profiles predict virulence of uropathogenic Escherichia coli strains: contribution of ygi and yad fimbriae. Infect Immun 79:4753–4763. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 377.Schifferli DM, Beachey EH, Taylor RK. 1991. Genetic analysis of 987P adhesion and fimbriation of Escherichia coli: the fas genes link both phenotypes. J Bacteriol 173:1230–1240. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 378.Dean EA. 1990. Comparison of receptors for 987P pili of enterotoxigenic Escherichia coli in the small intestines of neonatal and older pig. Infect Immun 58:4030–4035. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 379.Yue M, Rankin SC, Blanchet RT, Nulton JD, Edwards RA, Schifferli DM. 2012. Diversification of the Salmonella fimbriae: a model of macro- and microevolution. PLoS One 7:e38596. 10.1371/journal.pone.0038596. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 380.Silva CA, Blondel CJ, Quezada CP, Porwollik S, Andrews-Polymenis HL, Toro CS, Zaldívar M, Contreras I, McClelland M, Santiviago CA. 2012. Infection of mice by Salmonella enterica serovar Enteritidis involves additional genes that are absent in the genome of serovar Typhimurium. Infect Immun 80:839–849. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 381.Thomson NR, Clayton DJ, Windhorst D, Vernikos G, Davidson S, Churcher C, Quail MA, Stevens M, Jones MA, Watson M, Barron A, Layton A, Pickard D, Kingsley RA, Bignell A, Clark L, Harris B, Ormond D, Abdellah Z, Brooks K, Cherevach I, Chillingworth T, Woodward J, Norberczak H, Lord A, Arrowsmith C, Jagels K, Moule S, Mungall K, Sanders M, Whitehead S, Chabalgoity JA, Maskell D, Humphrey T, Roberts M, Barrow PA, Dougan G, Parkhill J. 2008. Comparative genome analysis of Salmonella Enteritidis PT4 and Salmonella Gallinarum 287/91 provides insights into evolutionary and host adaptation pathways. Genome Res 18:1624–1637. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 382.Carnell SC, Bowen A, Morgan E, Maskell DJ, Wallis TS, Stevens MP. 2007. Role in virulence and protective efficacy in pigs of Salmonella enterica serovar Typhimurium secreted components identified by signature-tagged mutagenesis. Microbiology 153:1940–1952. [DOI] [PubMed] [Google Scholar]
- 383.Dufresne K, Daigle F. 2017. Salmonella Fimbriae: What is the Clue to Their Hairdo? Current Topics in Salmonella and Salmonellosis. InTech. [Google Scholar]
- 384.Seth-Smith HMB, Fookes MC, Okoro CK, Baker S, Harris SR, Scott P, Pickard D, Quail MA, Churcher C, Sanders M, Harmse J, Dougan G, Parkhill J, Thomson NR. 2012. Structure, diversity, and mobility of the Salmonella pathogenicity island 7 family of integrative and conjugative elements within Enterobacteriaceae. J Bacteriol 194:1494–1504. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 385.Porwollik S, Wong RM, McClelland M. 2002. Evolutionary genomics of Salmonella: gene acquisitions revealed by microarray analysis. Proc Natl Acad Sci USA 99:8956–8961. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 386.Thorns CJ, Turcotte C, Gemmell CG, Woodward MJ. 1996. Studies into the role of the SEF14 fimbrial antigen in the pathogenesis of Salmonella enteritidis. Microb Pathog 20:235–246. [DOI] [PubMed] [Google Scholar]
- 387.De Masi L, Yue M, Hu C, Rakov AV, Rankin SC, Schifferli DM. 2017. Cooperation of adhesin alleles in Salmonella-host tropism. MSphere 2:e00066–e00017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 388.Berrocal L, Fuentes JA, Trombert AN, Jofré MR, Villagra NA, Valenzuela LM, Mora GC. 2015. stg fimbrial operon from S. Typhi STH2370 contributes to association and cell disruption of epithelial and macrophage-like cells. Biol Res 48:34. 10.1186/s40659-015-0024-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 389.Ernst RK, Dombroski DM, Merrick JM. 1990. Anaerobiosis, type 1 fimbriae, and growth phase are factors that affect invasion of HEp-2 cells by Salmonella typhimurium. Infect Immun 58:2014–2016. [DOI] [PMC free article] [PubMed] [Google Scholar]
