Europe PMC

This website requires cookies, and the limited processing of your personal data in order to function. By using the site you are agreeing to this as outlined in our privacy notice and cookie policy.

Abstract 


Proteus mirabilis, a Gram-negative rod-shaped bacterium most noted for its swarming motility and urease activity, frequently causes catheter-associated urinary tract infections (CAUTIs) that are often polymicrobial. These infections may be accompanied by urolithiasis, the development of bladder or kidney stones due to alkalinization of urine from urease-catalyzed urea hydrolysis. Adherence of the bacterium to epithelial and catheter surfaces is mediated by 17 different fimbriae, most notably MR/P fimbriae. Repressors of motility are often encoded by these fimbrial operons. Motility is mediated by flagella encoded on a single contiguous 54-kb chromosomal sequence. On agar plates, P. mirabilis undergoes a morphological conversion to a filamentous swarmer cell expressing hundreds of flagella. When swarms from different strains meet, a line of demarcation, a "Dienes line," develops due to the killing action of each strain's type VI secretion system. During infection, histological damage is caused by cytotoxins including hemolysin and a variety of proteases, some autotransported. The pathogenesis of infection, including assessment of individual genes or global screens for virulence or fitness factors has been assessed in murine models of ascending urinary tract infections or CAUTIs using both single-species and polymicrobial models. Global gene expression studies performed in culture and in the murine model have revealed the unique metabolism of this bacterium. Vaccines, using MR/P fimbria and its adhesin, MrpH, have been shown to be efficacious in the murine model. A comprehensive review of factors associated with urinary tract infection is presented, encompassing both historical perspectives and current advances.

Free full text 


Logo of nihpaAbout Author manuscriptsSubmit a manuscriptHHS Public Access; Author Manuscript; Accepted for publication in peer reviewed journal;
EcoSal Plus. Author manuscript; available in PMC 2018 Apr 2.
Published in final edited form as:
PMCID: PMC5880328
NIHMSID: NIHMS925686
PMID: 29424333

Pathogenesis of Proteus mirabilis Infection

Abstract

Proteus mirabilis, a Gram-negative rod-shaped bacterium most noted for its swarming motility and urease activity, frequently causes catheter-associated urinary tract infections (CAUTI) that are often polymicrobial. These infections may be accompanied by urolithiasis, development of bladder or kidney stones due to alkalinization of urine from urease-catalyzed urea hydrolysis. Adherence of the bacterium to epithelial and catheter surfaces is mediated by 17 different fimbriae, most notably MR/P fimbriae. Repressors of motility are often encoded by these fimbrial operons. Motility is mediated by flagella encoded on a single contiguous 54 kb chromosomal sequence. On agar plates, P. mirabilis undergoes a morphological conversion to a filamentous swarmer cell expressing hundreds of flagella. When swarms from different strains meet, a line of demarcation, a “Dienes line”, develops due to the killing action of each strain’s type VI secretion system. During infection, histological damage is caused by cytotoxins including hemolysin and a variety of proteases, some autotransported. The pathogenesis of infection, including assessment of individual genes or global screens for virulence or fitness factors has been assessed in murine models of ascending UTI or CAUTI using both single-species and polymicrobial models. Global gene expression studies carried out in culture and in the murine model have revealed the unique metabolism of this bacterium. Vaccines, using MR/P fimbria and its adhesin, MrpH, have been shown to be efficacious in the murine model. A comprehensive review of factors associated with urinary tract infection is presented, encompassing both historical perspectives and current advances.

I. Introduction

Proteus mirabilis, a Gram-negative rod-shaped bacterium, is well-known for its urease production and distinctive ability to differentiate into elongated swarm cells and characteristic bull’s-eye pattern of motility on agar plates. P. mirabilis belongs to the class Gammaproteobacteria, and has long been recognized as a member of the order Enterobacteriales, family Enterobacteriaceae. However, one group recently created a reconstructed phylogenetic tree based on shared core proteins, ribosomal proteins, and four multilocus sequence analysis proteins, and has proposed that the order Enterobacteriales be reclassified, placing Proteus within a new Morganellaceae family (1).

P. mirabilis can be found in a wide variety of environments, including soil, water sources, and sewage, but it is predominantly a commensal of the gastrointestinal tracts of humans and animals (2, 3). While the bacterium is capable of causing a variety of human infections, including those of wounds, the eye, the gastrointestinal tract, and the urinary tract, it is most noted for infections of the catheterized urinary tract, known as catheter-associated urinary tract infections (CAUTI) (49). These infections are common in long-term catheterized patients, such as those who reside in nursing homes and chronic care facilities, and may be of particular danger to spinal cord injury patients (10). Urinary tract infections (UTIs) and CAUTIs involving P. mirabilis are typically complicated by the formation of bladder and kidney stones (urolithiasis) and permanent renal damage (1113), and may progress to bacteremia and sepsis (14, 15). Indeed, CAUTI is the most common source of bacteremia in nursing homes, bacteremia involving P. mirabilis most frequently occurs following UTI or CAUTI compared to other sources of infection, and bacteremia and sepsis due to P. mirabilis carry a high mortality rate (1417). CAUTIs are also often polymicrobial (8, 17), and P. mirabilis is one of the most common organisms present during polymicrobial urine colonization and infection (4, 9) (Table 1).

Table 1

Epidemiology of Single-Species and Dual-Species Clinically Diagnosed Catheter-Associated Urinary Tract Infection in Nursing Home Residents.

Single-SpeciesDual-Species
Microorganisma Total Urine Culturesb Urine Culturesc %d Urine Culturese %
Proteus mirabilis4828222035
Enterococcus spp.3815122340
Escherichia coli3723181425
Pseudomonas aeruginosa3418141628
Staphylococcus aureus20119916
Klebsiella pneumoniae1465814
Citrobacter spp.932611
Morganella morganii72259
Providencia stuartii73247
Yeast63235
Acinetobacter baumanii42223
Enterobacter spp.55400
Serratia marscesens11100
Corynobacterium spp.11100
Other84347
Total18212510057100
aNumber of urine cultures containing each microorganism.
bNumber of single-species urine cultures containing each microorganism.
cPercent of all single-species cultures represented by each microorganism.
dNumber of dual-species urine cultures containing each microorganism.
ePercent of all dual-species cultures represented by each microorganism Reproduced, with permission, from (9).

P. mirabilis is an agent of catheter biofilm formation, quickly fouling the surface of a newly inserted urinary catheter. Surface organelles such as fimbriae and other adhesins appear to play a significant role in this process. The enzyme urease also contributes dramatically to this process. Urea, our means of eliminating excess nitrogen, is present in high concentrations in urine (~400 mM), is the substrate of urease, and is hydrolyzed to CO2 and NH3. The liberated ammonia raises the pH of the urine and initiates the precipitation of otherwise soluble polyvalent anions and cations present in urine. The result is urolithiasis, the formation of struvite (MgNH3PO4) or apatite (CaPO4) stones. These crystals can form on and within the lumen of catheters, blocking urine flow and necessitating catheter removal and replacement. Stones may also form in the renal tubules or renal pelvis, causing inflammation and often requiring surgical removal. This bacterium is capable of invading bladder epithelial cells, and produces a variety of cytotoxins that damage the epithelium, leading to significant histopathology.

Over the last four decades, our laboratory and others have developed a variety of approaches to study the virulence of P. mirabilis. Importantly, two primary models of ascending urinary tract infections have been developed in mice, in which organisms are transurethrally delivered to the bladder and bacteria can be enumerated in the urine, bladder, and kidneys following a suitable interval, depending on the experimental question asked. One model requires no manipulation of the murine host, yet it still represents a model of complicated urinary tract infection for P. mirabilis as urolithiasis quickly occurs, offering a substrate for bacterial colonization and sometimes blockage of proper urine flow along with significant pathology. In a more recent modification of this model, a small silicone catheter segment is introduced into the bladder, providing an immediate substrate for bacterial colonization. This latter model may more closely reflect the catheterized urinary tract and CAUTI in humans. It has also been demonstrated as an excellent model for polymicrobial bacteriuria, a common feature of CAUTI.

The advent of straightforward genome sequencing and comparative genomics studies has provided a view of the heterogeneity of the species, and has allowed for a more global understanding of metabolic pathways, regulatory schemes, and virulence determinants present in the bacterium. The importance of individual genes and operons must be addressed experimentally. Nevertheless, sequencing information allows for generation of testable hypotheses. This information has also opened the door for transcriptomic studies, both under in vitro conditions and during experimental infection. While not as facile as in other genera, allelic exchange methods for specific gene mutation may be routinely conducted in P. mirabilis strains. In addition, genome-wide transposon mutagenesis is also straightforward and has facilitated a variety of screens, including signature-tagged mutagenesis (STM) and transposon insertion-site sequencing (Tn-Seq).

Clearly, P. mirabilis possesses an impressive arsenal of virulence factors (Fig. 1). Urease is a critical feature of this species, but the bacterium also expresses a startling number of fimbriae and other adhesins. The most well studied fimbria is the mannose-resistant Proteus-like (MR/P) fimbria, whose expression is phase-variable. The mrp operon also encodes a non-structural protein, MrpJ, which directly represses flagella synthesis, thus shutting down motility while the bacterium adheres, and flagella themselves contribute to pathogenesis. As well, a variety of potent toxins and proteases compound virulence. Similar to other members of the Enterobacteriaceae, P. mirabilis carries numerous secretion systems including types I, III, IV, V, and VI. To provide co-factors and regulate intracellular metabolism, P. mirabilis also carries a myriad of ion importers and exporters. Lastly, the bacterium carries an integrative and conjugative element named ICEPm1 that can self-replicate and self-transfer to other strains and species, transferring virulence genes and antibiotic resistance.

An external file that holds a picture, illustration, etc.
Object name is nihms925686f1.jpg
Concepts of Proteus mirabilis pathogenesis during urinary tract infection (UTI)

Adherence: binding catheters, host tissues, and neighboring bacteria may all contribute to disease. Adherence is mediated by chaperone-usher fimbriae and autotransporter adhesins. Urease: involved in stones, crystalline biofilms, and possibly nutrition or host sensing. Motility: P. mirabilis swarms across catheters and may ascend to the kidneys using swimming motility. Both forms of motion are mediated by flagella. Chemotaxis proteins allow the bacteria to follow chemical gradients. Metabolism: likely permits establishment of a nutritional niche, competition with other species, and response to host cues. Metal scavenging: iron and zinc uptake are essential for growth, but are sequestered by the host; therefore, specialized proteins are required for bacteria to scavenge these metals. Toxins: proteins such as HpmA and Pta may aid in nutrient accessibility, immune evasion, or provision of surfaces to colonize. Biofilm formation: Crystalline biofilms readily form on catheters, and bacterial clusters in the bladder may be a biofilm-mediated process. Immune evasion: this can include antibody and antimicrobial peptide degradation, polymyxin resistance, lipopolysaccharide (LPS) variation, and physical obstruction of phagocytosis. Virulence regulation: required to coordinate all steps of infection. Type 6 secretion system (T6SS): involved in self-recognition; unknown role during UTI. MrpJ-controlled systems in this figure are bolded. Figure adapted, with permission, from (20).

There are no currently licensed vaccines available for this organism, and multidrug-resistant isolates are becoming increasingly common. Thus, efforts to generate effective vaccines or therapeutic treatments are warranted. The majority of experimental vaccine studies have targeted P. mirabilis fimbriae, and successes have been described in the murine model for the MR/P fimbriae themselves as well as the tip adhesin of the fimbria, MrpH. However, none of the experimental vaccines have provided complete protection against infection, and additional targets remain to be explored.

In this chapter, we will focus on the current state of knowledge regarding how this fascinating bacterium is capable of infecting the urinary tract and causing disease. We will summarize studies that have made a clear connection between specific genes, gene clusters, and operons and their role in pathogenesis. While there is an enormous literature on the fascinating phenotype of swarming motility, this has been extensively summarized elsewhere (3, 1820). We will therefore address only those studies that relate to flagellum-mediated motility and virulence in the murine model.

II. Host Interactions

A. The catheterized urinary tract

Numerous bacteria colonize the periurethral area, but are generally prevented from establishing an infection of the urinary tract by the regular flushing of the urethra during micturition (the passing of urine). Urothelial cells lining the bladder also provide a barrier against bacterial adhesion and invasion, partly due to a coating of glycosaminoglycan mucin as well as their role in the innate immune response.

When an indwelling urinary catheter is inserted, it bypasses many of the natural host defenses against urinary tract infection. In contrast to normal micturition and flushing of the urethra at regular intervals, the catheter allows for continuous bladder drainage as urine accumulates, which is not of sufficient volume or force to effectively flush the urinary tract. The design of the traditional Foley catheter and draining tubing also causes retention of 10–100 ml of urine within the bladder, providing a reservoir for bacteria to replicate (21). The urothelium is only 3 to 4 cell layers thick, and is easily damaged during rushed or multiple attempts to place a catheter (22). Any rough surfaces on the catheter can also facilitate bacterial biofilm formation or cause damage to the urethral lining and bleeding during insertion or withdrawal. Furthermore, improper insertion of the catheter or inflation of the balloon can create mucosal and submucosal tears (22). All of these potential complications of catheter insertion provide P. mirabilis with additional sites of attachment, nutrients, and routes for establishing more severe infection. The presence of an indwelling urinary catheter also elicits a robust inflammatory response, both in humans and in experimental animal models, and the resulting accumulation of fibrinogen on the catheter surface can provide bacteria with an ideal substrate for attachment (2325).

Approximately 50% of individuals with long-term catheterization (>28 days) experience catheter blockage from crystalline deposits, and the urease activity of P. mirabilis is the most common cause of this blockage (5, 26, 27) (link to urease section) (Fig. 2). Urease catalyzes the hydrolysis of urea to ammonia and carbon dioxide, thereby raising urinary pH. As urine pH increases, calcium and magnesium phosphates begin to precipitate out of solution, leading to the formation of struvite (magnesium ammonium phosphate) and apatite (calcium phosphate) crystals (11, 28). The urinary pH at which precipitation occurs is referred to as the nucleation pH (pHn), and if the pH of voided urine (pHv) falls within the range of the pHn (common during colonization with urease-positive organisms), crystallization and catheter blockage are likely to occur (29).

An external file that holds a picture, illustration, etc.
Object name is nihms925686f2.jpg
Treatments, nucleation pH, voided pH, catheter changes and bacterial isolates in a patient undergoing long-term urinary catheterization

Colonization occurred with multiple species, but P. mirabilis colonization specifically led to an increase in urinary pH and repeated catheter blockage. Figure adapted, with permission, from (364).

The struvite and apatite crystals that form during P. mirabilis colonization deposit on the catheter surface, facilitating the formation of crystalline biofilms (link to biofilm section). Indeed, it has been experimentally determined that P. mirabilis typically begins establishing crystalline biofilms on a urine-bathed catheter when urinary pH reaches pHn (30). Adherence to the catheter and the formation of crystalline biofilms provides protection from the action of host cells recruited to the site of infection, particularly neutrophils. Crystalline biofilms formed by P. mirabilis can also result in blockage of the catheter lumen, which obstructs urine flow and may cause reflux of infected urine to the kidneys (31, 32). The lifestyle of P. mirabilis within the catheterized urinary tract is depicted in Fig. 3.

An external file that holds a picture, illustration, etc.
Object name is nihms925686f3.jpg
Lifestyle of Proteus mirabilis

P. mirabilis bacteria (green) form crystalline biofilms on the surface of catheters (Top). Once inside the bladder [0.5–6 hours postinfection (hpi)], this organism can invade into urothelial cells of the bladder. As early as 10–24 hpi, P. mirabilis forms intraluminal clusters that can extend the length of the bladder and are associated with urothelial cell destruction [perhaps through the production of toxins (yellow stars) or an increase in urine pH] and mineral deposition (purple rods). Host innate immune cells such as neutrophils (blue) are recruited to the site of infection and can form NETs (neutrophil extracellular traps). Figure adapted, with permission, from (19).

B. Urolithiasis

The urease of P. mirabilis is unambiguously associated with the development of infection-induced stone formation, known as urolithiasis (5, 11, 3336) (Fig. 4). Indeed, Proteus species have been isolated in 70% of cases of bacteria-induced stone formation (37). The cytoplasmic enzyme urease catalyzes the hydrolysis of urea, the nitrogenous waste product of mammals which is maintained normally at 400–500 mM in human urine. Ammonia, generated from the breakdown of urea, results in a dramatic elevation of the pH of urine, and normally soluble polyvalent anions and cations precipitate at high pH to form struvite (MgNH4PO46H2O) and carbonate apatite [Ca10(PO4)6CO3]. In vitro studies inoculating urine with P. mirabilis demonstrated that increasing the concentration of Mg2+, Ca2+, or PO43− ions intensified the magnitude of crystallization (38).

An external file that holds a picture, illustration, etc.
Object name is nihms925686f4.jpg
A particularly large urolith

(A) Reconstructed computed tomography image, showing the location and relative size of the urate cystolith (indicated by arrows). (B) Photograph of the urate cystolith, showing its absolute size. Reproduced from (365) with permission.

Experimental evidence for the involvement of urease in P. mirabilis pathogenicity was provided by infection of mice via transurethral inoculation into the bladder (39, 40). Separate groups of mice were inoculated with the wild-type strain or a urease-negative mutant of P. mirabilis, and struvite stones were found in the renal pelvis of mice infected with the wild-type strain one week after inoculation. Quantitatively, 12 of 39 mice (31%) developed stones after one week and 8 of 20 mice (40%) developed stones after two weeks, while no stones were found in the 38 mice infected with the urease-negative mutant. A caveat to the study was that the urease-negative mutant did not colonize the mouse urinary tract to the same extent as the wild-type strain. However, follow-up studies with higher inocula of the urease-negative mutant construct still failed to result in development of urolithiasis.

Because P. mirabilis is often part of a polymicrobial infection, the impact of polymicrobial colonization on urease activity and urolithiasis in a murine CAUTI model has been examined for coinfection of P. mirabilis and its common CAUTI partner Providencia stuartii (25, 41). Coinfection resulted in bacterial loads that were similar to monoinfections with either pathogen. However, there was a significantly increased incidence of urolithiasis and bacteremia during coinfection. Coinfection was also accompanied by a significant increase in urease activity that was manifested in a synergistic manner (link to urease section).

Adding to the mechanism of stone formation, it has been noted that GFP-expressing P. mirabilis could be observed within the matrix of urinary stones in mice experimentally infected with P. mirabilis (42). Genome sequences of P. mirabilis have also been detected in urinary calculi by PCR (43). Furthermore, Schaffer and colleagues (44) found, using confocal microscopy, that P. mirabilis formed extracellular clusters in the bladder lumen that served as the basis for focused mineral deposition, consistent with nascent stone formation (Fig. 5).

An external file that holds a picture, illustration, etc.
Object name is nihms925686f5.jpg
P. mirabilis extracellular clusters are precursors to stone formation

(A and B) Detection of mineral deposition using Alizarin Red staining of P. mirabilis-infected bladder sections at (A) 6, and (B) 24 hpi. (Scale bars, 100 μm.) L, bladder lumen; an asterisk indicates extracellular cluster (purple staining). (C) A representative image of a P. mirabilis cluster at 24 hpi (Scale bar, 100 μm.) Staining of bacteria (green), UPIIIa (red), and DNA (blue) show accumulation of bacterial clusters at the bacteria–bladder interface. An asterisk indicates an extracellular cluster; L, bladder lumen. The thin arrow indicates a region with increased DAPI signal, whereas thick arrows indicate areas of extensive urothelial damage. (D and E) Scanning electron micrographs of P. mirabilis urease-induced bladder stone (7 dpi). (D) One-quarter of the bladder viewed at a low magnification (bar, 500 μm). The orientation of the bladder is indicated by an arrow pointing to the inferior end of the bladder (the end leading to the urethra). (E) Higher magnification (bar, 5 μm) of the area enclosed in a box in panel D. Figure adapted, with permission, from (44) (A–C) and (42) (D and E).

C. Internalization, cytotoxicity, and histopathology

The ability of P. mirabilis to invade and lyse host cells has been explored for decades, and has been demonstrated to contribute to infection progression and severity of disease in animal models. The level of host cell invasion and cytotoxicity achieved by P. mirabilis in vitro varies dramatically by the bacterial strain being tested, growth phase, and morphology of the bacteria (for instance, log phase growth versus stationary phase and vegetative cells versus differentiated swarm cells), host cell line, multiplicity of infection, pH of the medium, and the duration of the experiment. Similarly, pathological changes in the bladder and kidneys of infected animals vary to some extent based on the bacterial strain, inoculating dose, and infection model. However, P. mirabilis has been demonstrated to be more invasive than Salmonella enterica serovar Typhimurium (Table 2) (45). Although not an exhaustive list, the invasion and cytotoxic properties of various P. mirabilis isolates are summarized in Table 3.

Table 2

Internalization of P. mirabilis strains by cultured human renal epithelial cells.

StrainGrowth conditionInternalized bacteria (CFU/ml)a
P. mirabilis WPM111Broth; aerated; 18 h22,420
E. coli HB101Broth; aerated; 18 h67
S. TyphimuriumBroth; aerated; 18 h18,130

Modified from (45).

aValues are means of three independent determinations.

Table 3

Invasive and cytotoxic properties of human P. mirabilis isolates.

Strain SourceStrain nameInvasionCytotoxicity
Cell LineRefCell LineRef
Catheterized urinary tractHI4320Human embryonic kidney (HEK293), monkey kidney (Vero), human urothelial (UMUC-3)(46)Human primary renal proximal tubular epithelial cells (HRPTEC), human embryonic kidney (HEK293)(25, 47)
BA6163NDHuman primary renal proximal tubular epithelial cells (HRPTEC), human bladder epithelial (T24), human B-cell lymphoma (Daudi and Raji), human monocyte (U-937), monkey kidney (Vero)(47, 48)
CFT295
HU2450
CFT106
SA1387
MI159
EL1131
CFT37
MA2489
DR3282
CFT403
NDHuman primary renal proximal tubular epithelial cells (HRPTEC)(47)
C6
C7
C11
C31
Human ureter (Hu 609), and human bladder epithelium (HCV 29)(49)
(49)
ND
UTIT1
31
Monkey kidney (Vero)(50)Monkey kidney (Vero)(50)
Urinary stonePM7002Human urothelial (UMUC-3)(51)ND
K8
K608
Human ureter (Hu 609)(49)ND
U6450NDHuman urothelial (EJ/28)(52)
WoundKR37Human ureter (Hu 609) and human(49)ND
KW28bladder epithelium (HCV 29)
UnspecifiedN2
P19
Human urothelial (NTUB1)(53, 54)ND

While it is clear that P. mirabilis invades host tissues during infection, it does not appear to establish a significant intracellular niche as observed for uropathogenic E. coli (44). Several virulence factors have been implicated in contributing to cell invasion and cytotoxicity in vitro, as well as histopathological changes in vivo. For instance, flagella contribute to invasion in part by allowing the bacterial cells to come into close proximity to the host cells, and mutants lacking flagella are unable to invade cells unless centrifuged directly onto the host cell monolayer (45, 52). The specifics of P. mirabilis invasion of bladder and kidney cells and the contributions of select virulence factors are detailed further below.

Bladder invasion and histopathology

Internalization of P. mirabilis by bladder epithelial cells has been directly demonstrated in vitro using a hemolysin (hpmA) mutant to avoid confounding from the effects of the cytolytic toxin. In these experiments, it was determined that P. mirabilis utilizes the AipA autotransporter for internalization into bladder cells (46, 55). In addition to internalization, P. mirabilis is capable of lysing bladder epithelial cells using a combination of the Proteus toxic agglutinin (Pta) and hemolysin (55).

In the murine model of ascending UTI, P. mirabilis invades bladder epithelial cells as early as 30 minutes post-inoculation, which may provide transient protection from the immune response and an intracellular niche for initial replication and survival (44, 50, 52, 5658). However, intracellular bacteria are uncommonly observed at later times post-inoculation, and P. mirabilis appears to instead form large, extracellular clusters within the bladder lumen and adjacent to the urothelium after this initial invasion phase rather than establishing the intracellular communities that are characteristic of uropathogenic E. coli (44) (Fig. 6). Formation of these clusters requires urease activity and the mannose-resistant Proteus-like (MR/P) fimbriae, and provides protection from infiltrating neutrophils (44).

An external file that holds a picture, illustration, etc.
Object name is nihms925686f6.jpg
P. mirabilis invades the urothelium, but infrequently forms intracellular bacterial communities (IBCs)

(A–D) Representative images of P. mirabilis (A and B) and uropathogenic E. coli (UPEC) (C and D) attachment and invasion. Bacteria (green), UPIIIa (red), and DNA (blue) show localization of the bacteria relative to the apical surface of the urothelium. Scale bars, 10 μm. (B and D) A regional view of the bladder section containing the 10 hpi IBC shown in A and C, respectively. Scale bars, 100 μm. L, bladder lumen. (E and F) Quantification of P. mirabilis (E) and UPEC (F) bladder invasion at 0.5 hpi following either ex vivo gentamicin treatment (Gent) or mock treatment (Mock) (n = 6–8). *P < 0.05. Figure adapted, with permission, from (44).

Bladder colonization by P. mirabilis most often results in mild to moderate cystitis in murine models of UTI, including transmural neutrophilic inflammation, epithelial transcytosis of neutrophils, and submucosal edema (25, 55). The pathology of P. mirabilis bladder colonization during ascending UTI is largely due to a combination of urease activity, Pta, and hemolysin (25, 55). In the murine model of CAUTI, infection with P. mirabilis results in more severe cystitis, and this appears to be largely due to urease activity and encrustation of the catheter (25).

Kidney invasion and histopathology

Infection with P. mirabilis tends to result in unique kidney pathology. For instance, P. mirabilis is the only species that causes a high incidence of kidney stone formation in a rat model of pyelonephritis (59), and it causes more kidney stones and greater kidney damage than other urease-positive organisms such as P. stuartii in murine models of ascending UTI and CAUTI (25, 41). Specifically, kidney colonization by P. mirabilis in the ascending model of UTI is most often associated with moderate pyelonephritis, including neutrophilic interstitial nephritis within the peripelvic renal cortex and occasional damage to the surrounding renal parenchyma (25, 55). The severity of pyelonephritis is more strongly influenced by the additive effect of Pta and hemolysin than urease (25, 55). The same is true in the murine model of CAUTI, with P. mirabilis infection similarly resulting in moderate pyelonephritis, largely independent of urease activity (25).

In addition to directly damaging kidney tissue and inducing inflammation, P. mirabilis proliferates within the tubular epithelium of the kidneys in both mice and rats, resulting in necrosis and nephrosis (25, 59). In contrast to pyelonephritis, the severity of nephrosis correlates with urease activity in the murine models of ascending UTI and CAUTI, indicating a role for the urease enzyme and alkaline pH in renal tubule damage (25). These foci of inflammation can also develop into inflammatory lesions, another characteristic feature of P. mirabilis pyelonephritis (25, 55, 59).

Internalization of P. mirabilis by human renal proximal tubular epithelial cells (HRPTECs) has also been directly demonstrated in vitro using a hemolysin mutant to avoid confounding by the cytotoxic effects of the toxin. Interestingly, internalization requires protein synthesis by P. mirabilis, but does not require active phagocytosis or protein synthesis by the HRPTECs (56, 58). Thus, invasion of host cells by P. mirabilis, at least in vitro, is predominantly mediated by bacterial factors. Hemolysin is also thought to be the primary virulence factor responsible for direct lysis of HRPTECs (47, 56).

D. Innate immune responses to P. mirabilis UTI

Although innate immune responses to UTI are an active area of investigation, most of these studies have focused on uropathogenic E. coli (6062). Because there are several key differences between P. mirabilis UTI and UPEC UTI, including urolithiasis, bacterial metabolism during UTI, and intracellular vs. luminal niches in the bladder, we will focus sharply here on studies that specifically examine P. mirabilis. Likewise, adaptive immune responses to vaccines are described elsewhere in this chapter (link vaccines section).

P. mirabilis induces a pro-inflammatory response during UTI; mice infected with P. mirabilis have elevated levels of CXCL1 at 72 hpi and IL-10 at 6 and 96 hpi in their urine (25). Notably, the cytokine response is further elevated during coinfection with Providencia stuartii, with increased levels of CCL2, CCL5, CXCL1, IL-6, IL-10, IL-17A, TNFα, IFNβ, and IFNγ at 48 hpi. In contrast, P. stuartii monoinfection did not lead to increased levels of any of the measured cytokines (25). Some of the elevated cytokines, such as IL-10, are anti-inflammatory (63), suggesting there are switches between pro- and anti-inflammatory responses as the infection progresses. This is consistent with changes in innate immune responses observed for other uropathogens, including UPEC and Group B streptococcus (60, 62, 64). TNF-α and IL-6 are also elevated in bladder homogenates from mice at 48 hpi (Peng 2016—it’s currently ref 308).

Neutrophils are one of the first innate responses to UTI, and are recruited by host uroepithelial cell signaling in response to conserved bacterial structures (pathogen-associated molecular patterns, or PAMPs) (65). CXCL1, mentioned above, is a potent neutrophil chemoattractant, and leukocytes are frequently detectable in urine during P. mirabilis UTI. Appropriately, there have been several studies investigating interaction of P. mirabilis with these leukocytes. By 10–24 hpi in the mouse model of UTI, P. mirabilis forms large clusters in the bladder lumen that draw a massive infiltration of neutrophils (Fig. 7) (44). Neutrophil extracellular traps (NETs) are webs of decondensed chromatin and antimicrobial proteins that are released in response to infection (66), and neutrophils in regions adjacent to the P. mirabilis bacterial clusters have been observed with extruded DNA that colocalized with extracellular H2A, suggesting development of NETs during P. mirabilis UTI (44). Supporting the mouse model data, proteomic signatures indicative of “NETosis” have been detected in urine samples from patients with P. mirabilis UTI (67, 68). Although neutrophil-mediated phagocytosis and NETosis can be effective bacterial clearance mechanisms, this line of attack may be less effective against P. mirabilis encased within clusters or urinary stones (42, 44). Neutrophils also respond to bacteremic P. mirabilis UTIs, and an increase in band neutrophils (>10% white blood cell count) is associated with an increased risk of P. mirabilis bacteremia (odds ratio 4.18) (69).

An external file that holds a picture, illustration, etc.
Object name is nihms925686f7.jpg
P. mirabilis induces neutrophil recruitment and NET formation

(A and B) Visualization of neutrophil recruitment at extracellular clusters at 24 hpi. Individual channels for the region enclosed in the dashed rectangle are shown below at the same magnification. Scale bars, 100 μm. (C) Identification of NET formation in regions of neutrophil recruitment at 24 hpi. Individual channels representing nuclear stains (H2A and DAPI) and membrane stains (Ly6G) show the overlap of DAPI and H2A distant from Ly6G staining. Scale bars, 10 μm. (D) Neutrophil phagocytosis of P. mirabilis at 24 hpi. Arrows indicate neutrophils which have phagocytosed bacteria. (E) Individual neutrophil recruitment in sections of murine bladders infected with P. mirabilis without clusters at 24 hpi and UPEC-infected sections at 10 hpi. Bacteria (green), Ly6G (red), and DNA (blue) show neutrophils adjacent to bacteria. Arrows indicate intact neutrophils. Figure adapted, with permission, from (44).

Inflammasomes are protein complexes that form in response to a variety of stimuli and act to induce inflammation (70). One particular type, the NLRP3 inflammasome, has been specifically linked to P. mirabilis in the intestine, where P. mirabilis induces a potent IL-1β response (71). In this setting, NLRP3 activation is dependent on HpmA hemolysin-induced K+ efflux (71). NLRP3 inflammasomes are also activated by crystalline material, and thus urease activity during UTI may further boost a potent inflammatory response (44, 70).

Tamm-Horsfall protein (THP, also called uromodulin) is the most abundant protein in normal human urine (72). Mice deficient in THP are susceptible to UTI (73), and P. mirabilis likewise causes more severe infection in THP-deficient mice (74). THP is heavily glycosylated and has been proposed to be a competitive ligand for UCA/NAF fimbriae (link fimbriae) (74).

Flagellin is another conserved structure that is recognized by the innate immune response, specifically toll-like receptor 5 (TLR5) (75), and flagellar motility is a fundamental feature of P. mirabilis (link flagella). In the intestine, Lypd8 protein prevents flagellated bacteria, including P. mirabilis, from invading colonic epithelium and causing inflammation (76). Installation of purified P. mirabilis flagellin into the bladder elicits leukocyte infiltration, histological changes in bladder tissue, and elevated Cxcl1, Cxcl10, and Il6 mRNA (77). Interestingly, pre-treatment with flagellin did not aid clearance of P. mirabilis in a subsequent challenge, and instead led to increased bacterial recovery from the kidneys (77).

Immune evasion

P. mirabilis has several tools for evading the innate immune response. These bacteria readily invade cultured cells (link invasion). However, although invasion has been observed during experimental UTI, P. mirabilis is typically found in the bladder lumen (44). It is possible that invasion plays a larger role during pyelonephritis. Urinary stones may also provide shelter from attack by leukocytes or antimicrobial peptides (3, 19, 42, 78). P. mirabilis is noted for its general resistance to killing by antimicrobial peptides, particularly polymyxins (79). Two components of this resistance are Zap protease, which degrades antimicrobial peptides (link), and LPS modifications to alter surface charge (80). Notably, different types of P. mirabilis LPS O-antigen elicit distinct proinflammatory cytokine IL-8 responses from cultured urothelial and renal cells (81). Furthermore, P. mirabilis produces a lysozyme inhibitor named PliC (82). Finally, at least two major antigenic proteins on the bacterial surface, MR/P fimbriae and flagella, are subject to phase variation. Thus, bacteria producing altered flagellins or phase-off MR/P fimbriae may evade both innate and adaptive immune responses.

P. mirabilis as a possible trigger of autoimmunity

Rheumatoid arthritis (RA) is a chronic autoimmune disorder that primarily attacks the joints. There is an intriguing correlation between P. mirabilis and RA, suggesting that in some populations, P. mirabilis UTI may be a triggering event for autoantibody development (83). Specifically, amino acid sequences IRRET in UreC (a component of urease) and ESRRAL in HpmB (an accessory protein which allows secretion of HpmA hemolysin) may lead to collagen autoantibodies due to cross-reactivity in individuals with particular HLA-DR4 subtypes. A correlation between P. mirabilis and RA has been observed in multiple patient studies (83); however, the urease/hemolysin hypothesis has not yet been directly tested. Importantly, other microbes have also been linked to RA, and it is possible that there are multiple infectious triggers for RA in vulnerable populations (84).

P. mirabilis is also associated with inflammatory bowel disease, which likewise has autoimmune origins. In a TRUC mouse ulcerative colitis model, P. mirabilis and Klebsiella pneumoniae correlated with colitis (85). Similar to the Lypd8 mice mentioned above, P. mirabilis bacteria were observed in the colonic mucus adjacent to epithelial cells in colitic mice, but not in wild-type mice (76, 85). While wild-type mice were negative for P. mirabilis, wild-type pups fostered on TRUC dams developed colitis and were positive for P. mirabilis and K. pneumoniae; furthermore, administration of TRUC-derived P. mirabilis and K. pneumoniae strains to specific pathogen-free wild-type mice elicited colitis (85).

III. Approaches to Studying Virulence

A. Animal models

The first murine model of UTI was described in 1967 (86), and while several different experimental methods have been explored for P. mirabilis infection studies, the most widely used model for investigating P. mirabilis pathogenicity is based on a method developed in the 1980s, referred to as the Hagberg protocol for ascending, unobstructed UTI (87). This protocol was initially developed to study E. coli virulence, particularly bacterial attachment to host bladder cells. Female CBA mice were favored as the model organism in this protocol due to several observations: 1) bacteria were better able to attach to uroepithelial cells from mice than rats, 2) CBA mice exhibited greater susceptibility to colonization and experimental UTI than BALB/c, C57BL/6, or C3H/HeN mice, and 3) the anatomy of male mice was not permissive for transurethral inoculation directly into the bladder. However, many mouse strains are susceptible to bladder colonization by P. mirabilis and ascension to the kidneys. For example, CF-1 mice have been utilized for noninvasive biophotonic imaging studies to monitor progression bacterial progression from the bladder to the kidneys and response to antibiotic treatment (88) (Fig. 8).

An external file that holds a picture, illustration, etc.
Object name is nihms925686f8.jpg
Real-time monitoring of P. mirabilis UTI using biophotonic imaging

Mice were inoculated with a P. mirabilis strain, Xen 44, engineered to constitutively produce luciferase. A representative animal is shown. Bacteria were initially observed in the bladder (ventral view), with kidney progression visible by day 2 (dorsal view). Figure adapted, with permission, from (88).

Two overviews of the procedures for induction of urinary tract infection in mice have been published in the Journal of Visualized Experiments (JoVE) (89, 90). Briefly, 6–9 week old mice are anesthetized and the peri-urethra and peri-anal areas are sterilized, generally with 10% povidone iodine. A 25 mm segment of gas-sterilized polypropylene tubing (0.28 mm inner diameter, 0.61 mm outer diameter) is threaded over a sterile 30-gauge hypodermic needle in such a way that approximately 15 mm of tubing extend beyond the tip of the needle. For transurethral inoculation, approximately 10 mm of the protruding polypropylene tubing is inserted through the urethra, beneath the pelvic bone, so that the bacterial inoculum is instilled directly into the bladder. This is generally performed using a mechanical syringe pump to slowly infuse a 0.05 ml inoculum over a 30-second interval to reduce the likelihood of reflux into the kidneys, so that the bacteria must actively ascend the ureters to colonize the kidneys and establish an upper urinary tract infection.

Initial studies of P. mirabilis pathogenesis using the murine model of ascending UTI were conducted using an inoculum of 2×1010 CFU/ml (1×109 CFU/mouse) and time points of 2, 7, or 14 days post-inoculation (39, 40, 91). For P. mirabilis HI4320, inoculation with this high dose resulted in significant mortality in mice at later time points (25% after 7 days, and 25% after 14 days), as well as severe pyelonephritis, renal damage, and urolithiasis (39). Inoculation with a lower dose (2×108 CFU/ml, or 1×107 CFU/mouse) was therefore determined to be ideal for studying the contribution of P. mirabilis virulence factors to pathogenicity. The contribution of specific genes to fitness and disease progression are generally assessed by one of two ways: 1) independent challenge, in which one group of mice is inoculated with the wild-type strain and the other is inoculated with the mutant of interest to see if the mutant exhibits significant differences in bacterial burden, or 2) cochallenge with a 1:1 mixture of mutant to wild-type, where the mutant must compete against the wild-type parent in the same mouse. In a cochallenge, wild-type and mutant bacteria are distinguished using a marker, such as antibiotic resistance, and a competitive index is calculated for the mutant by dividing the ratio of mutant/wild-type recovered from the mouse to the ratio of mutant/wild-type in the inoculum (Fig. 9).

An external file that holds a picture, illustration, etc.
Object name is nihms925686f9.jpg
Example of a competitive index (CI) calculated from a cochallenge experiment

Mice were challenged with a 1:1 ratio of wild-type P. mirabilis HI4320 and a lon mutant. P. mirabilis was recovered at 7 d post-inoculation (dpi). Wild type and mutant bacteria were distinguished by plating on solid media with or without antibiotic selection. (A) Cochallenge data. Solid circles represent wild-type CFU and open squares are mutant CFU recovered from each mouse. Bars show median CFU. The limit of detection in this assay is 200 CFU/g tissue. (B) In vivo CIs calculated from cochallenge data. Each dot represents the CI from an individual animal in the urine (U), bladder (B), kidneys (K), or spleen (S). Bars indicate the median CI. Significant differences in colonization (*P<0.05) were determined with the Wilcoxon signed-rank test. A CI<1 indicates a fitness defect. Figure adapted, with permission, from (129).

With respect to catheter-associated UTI, the presence of a catheter segment within the bladder provides bacteria with a new substrate for attachment and colonization and also elicits numerous changes within the bladder environment, including induction of a potent pro-inflammatory response (25, 88, 92, 93). Thus, as P. mirabilis most commonly causes complicated or catheter-associated UTI, several adaptations to the standard ascending model have been utilized to introduce and maintain a foreign body in the bladders of mice. One such method uses 6 mm segments of polyethylene tubing that have been coiled and sterilized with ethanol, pre-colonized by P. mirabilis for biofilm formation, and stretched back to an uncoiled position for transurethral insertion into the bladder, where they will spontaneously re-coil to prevent loss of the catheter during normal urination (88). Another method uses a 4 mm segment of silicone catheter tubing, sterilized by autoclaving, and introduced into the bladder at the time of transurethral inoculation (25). Both methods require reducing the bacterial density of the inoculum approximately 100-fold over the ascending UTI model to reduce mortality within 7 days of inoculation, and both models result in a greater incidence of cystitis and pyelonephritis than the ascending model. Furthermore, use of the silicone catheter segment was shown to induce a potent pro-inflammatory response during mock infection (25, 92), thereby mimicking the host response to catheterization that is observed in humans. Different catheter segment lengths may be required for different mouse strains and ages. However, for CBA/J mice and inoculation with P. mirabilis HI4320, rapid encrustation of the catheter promotes retention of the catheter segment in the vast majority of mice (25).

The predominant murine models of UTI and CAUTI rely solely on female mice due to the technical challenges of transurethral inoculation that are posed by the anatomy of male mice. A model of ultrasound-guided percutaneous catheter implantation into the bladder was therefore developed to permit studies of CAUTI in both female and male mice (94). Comparison of percutaneous catheter implantation and transurethral catheter placement revealed that both methods resulted in a similar incidence of catheter encrustation during infection with P. mirabilis. Thus, while this model may better approximate suprapubic catheterization than urethral catheterization, it will facilitate investigation of CAUTI and testing of catheter coatings in male mice in addition to female mice.

B. Comparative genomics

P. mirabilis lives in diverse environments, and has long been noted for strain variability in swarming behavior and LPS O-antigen structure (9597). Given these observations, it might seem that this species would possess great diversity in its genomic content, much like the mosaic pangenome of E. coli (98, 99). However, several lines of evidence point toward a remarkable conservation of the P. mirabilis chromosome.

A proteomic comparison of the P. mirabilis HI4320 reference strain with three other genomes, AOUC-001, CYPM1, and BB2000, shows that the overwhelming majority of predicted proteins are at least 99% identical across all four strains (Fig. 10A). Furthermore, alignment of the first two fully-annotated P. mirabilis genomes, HI4320 and BB2000, shows a high degree of synteny with only a few gaps due to insertions or deletions (Fig. 10B). The major differences that are visible are due to insertion of bacteriophage, a conjugative transposon, and the mobile genetic element ICEPm1, described later.

An external file that holds a picture, illustration, etc.
Object name is nihms925686f10.jpg
Conservation of P. mirabilis sequences

(A) Comparison of proteomes of three P. mirabilis strains against the HI4320 type strain using the PATRIC Proteome Comparison service (125). The majority of predicted proteins are ≥ 99% identical across all four genomes (blue and purple). Notable highly-variable proteins (≤ 70% identical; orange and red) are indicated with arrows. (B) Genome alignment showing synteny (red) between P. mirabilis HI4320 and BB2000. The largest gaps are ICEPm1, a conjugative transposon, and phage. Most blue lines indicate highly-repetitive transposase genes. Plot generated using the Artemis Comparison Tool (366).

Virulence factors such as urease, HpmA hemolysin, and MR/P fimbriae, are detected in the vast majority of strains, regardless of isolation source (100, 101). For example, hpmA was found by Southern blot in 63/63 isolates tested, and correlated with hemolytic activity (102); likewise, a survey of 211 isolates using a combination of PCR and dot blot found that all encoded hpmA (103). Zap protease is also widespread, including all P. mirabilis tested from a diverse O-serogroup collection (104). Similarly, ZapA-associated protease activity was detected in the urine of 16/17 P. mirabilis-infected patients (105).

Mannose-resistant hemagglutination is another defining characteristic of P. mirabilis, as both activity and the causative mrp fimbrial genes are nearly ubiquitous in this species (100, 101, 106). In addition, P. mirabilis encodes an abundance of fimbrial operons (link fimbriae section) relative to other bacterial species, and the majority of these of these are conserved across diverse isolates (106) (Fig. 11). The presence of a given fimbrial operon does not correlate with the isolate source. Of the 17 fimbrial operons encoded by P. mirabilis HI4320, only one, fim3, is detected in fewer than half of P. mirabilis isolates. A second operon, uca, is notable because it is present in the majority of isolates, but is located on a mobile element and may be found in different regions of the chromosome (106).

An external file that holds a picture, illustration, etc.
Object name is nihms925686f11.jpg
Conservation of fimbriae encoded by P. mirabilis

The percentage of isolates with each fimbria is shown. For the six sequenced strains, fimbriae were identified using BLAST using the HI4320 fimbrial major structural subunit gene as the query. For the 1980s (n = 10) and hospital isolates (n = 48), fimbrial genes were detected by PCR. Reproduced from (106), with permission.

Despite the overall conservation of the P. mirabilis chromosome, there is variability within the species. A few loci in particular are hypervariable and merit further mention (Fig. 10A).

ICEPm1

Comparative genomic hybridization of Providencia stuartii and Morganella morganii with a P. mirabilis HI4320 microarray revealed an integrative and conjugative element (ICE) common to all three species (107) (link ICE section). Further analysis of this 94 kb element, called ICEPm1, revealed genes encoding two previously-identified virulence factors: Proteus toxic agglutinin (Pta) and Nrp siderophore (108110). Carried within ICEPm1 is the Yersinia spp. high pathogenicity island (HPI), which includes the nrp operon. Consistent with a virulence role, ICEPm1 is found more often in urinary isolates than colonizing isolates (100% vs. 65%) (107). Furthermore, ICEPm1 has been observed to transmit between clinical isolates in a laboratory setting (111). Examination of sequenced genomes reveals that ICEPm1, when present, is not always inserted at the same genetic locus.

P. mirabilis strains may encode other conjugative elements, but they do not appear to be as prevalent as ICEPm1. For example, HI4320 encodes another conjugative transposon, named as either ICEPmiUSA1 or ICEPmi HI4320, that is a member of the SXT/R391 family of ICEs (112114). No genes in ICEPmiUSA1 have thus far been identified as virulence-associated. However, other ICEs in the SXT/R391 family carry multiple antibiotic resistance genes, and some of these ICEs are found in P. mirabilis (115117).

T6SS

Because the type VI secretion system (T6SS) of P. mirabilis is involved in strain self-recognition and competition between strains (link secretion section), it is not surprising that there is a high degree of variability in these genes. Specifically, the genes encoding the T6SS secretion apparatus are highly conserved, but the secreted effectors vary in sequence, type, and number. For example, P. mirabilis HI4320 and BB2000 encode T6SS effector operons adjacent to, and in the reverse orientation from, the T6SS secretion apparatus genes (118, 119). In both cases, the effector operons begin with hcp and vgrG homologs, which are required for T6SS function (120). After that, however, the operons diverge in both sequence and number of genes; for example, BB2000’s idr operon comprises five genes while HI4320’s pef operon has nine (118, 119) (Fig. 12). Furthermore, HI4320 encodes at least two more operons identified as likely T6SS effectors than BB2000 is predicted to encode (112, 121). Examination of other P. mirabilis sequenced genomes indicates that this variability continues throughout the species (Fig. 10A).

An external file that holds a picture, illustration, etc.
Object name is nihms925686f12.jpg
T6SS effector operons are highly variable

The T6SS effector operon immediately adjacent to the conserved T6SS secretion apparatus operon is depicted for P. mirabilis HI4320 (top) and BB2000 (bottom). The HI4320 locus, called pef for primary effector operon, comprises PMI0750-0758, and the BB2000 locus, called idr for identity recognition, comprises BB2000_0822-0826. The proteins encoded by the first gene (red) and the first 596 amino acids of the second gene (orange) are 99% identical, but the remaining ~130 amino acids of the encoded proteins share no homology. The rest of the operons are not homologous (black). A portion of IdrD shares identity with PMI0761, which is not part of the pef operon. The scale at the top is in nucleotides.

O-antigen

Like other gram-negative bacteria, P. mirabilis strains have diverse O-polysaccharides as part of their LPS; 76 O-serogroups have been described thus far for Proteus spp. (122). In fact, the O and H antigen designations for LPS and flagellar typing were originally named by Weil and Felix’s observation that swarming P. mirabilis had a breath-like appearance (German Hauch, breath) while non-swarming colonies did not (ohne Hauch, without breath) (95). In a survey of P. mirabilis strains with different O-serotypes, virulence factors were widespread (urease, Zap protease, swarming), but strains differed in the levels of these activities. Specifically, strains with negatively-charged O-polysaccharides displayed higher urease, protease, and swarming activities compared with positive or neutral O-polysaccharides (104). LPS is abundant on the cell surface of gram-negative bacteria, and its properties can affect bacterial interaction with their environment. Thus, some aspects of strain variability may be determined by LPS biochemistry.

Flagellin

Flagella (link flagella section) are targets of the innate immune system and also energetically expensive to produce (123). Correspondingly, although most flagellar genes are conserved in P. mirabilis, the gene encoding the whip itself is highly variable, as in other bacterial species. P. mirabilis also encodes a second, silent flagellar gene flaB, which can recombine with flaA and yield hybrid FlaAB flagella (124). The properties of FlaAB hybrids have been extensively reviewed elsewhere (19).

Phage

Like other bacteria, P. mirabilis genomes are peppered with integrated phage. HI4320 encodes three apparently complete and three degenerate prophages (112). CRISPR-Cas systems are encoded in approximately one-third of sequenced P. mirabilis genomes as determined by the Bacterial Bioinformatics Resource Center PATRIC (125); notably, CRISPR is not present in two strains commonly used for experimentation: HI4320 and BB2000 (112, 121).

Differential expression of phage genes has been observed in P. mirabilis during swarming (126, 127); specifically, one phage is inserted adjacent to the rcs operon, which encodes a regulator of flagella and swarming. Furthermore, phage have been used to transduce swarming ability into non-swarming strains (128). Thus, it is likely that phage account for some of the observed differences in swarming behavior between strains.

Plasmids

Plasmids are not a defining feature for P. mirabilis. The original P. mirabilis sequenced genome, HI4320, included a 36 kb plasmid that was designated pHI4320 (112), and a signature-tagged mutagenesis study as well as a transposon insertion-site sequencing study indicated that mutations in pHI4320 led to decreased fitness in mice (108, 129). However, several of the targeted genes were involved in plasmid stability, and pHI4320 does not encode any obvious virulence genes (112). In the National Center for Biotechnology Information (NCBI) database, a minority of P. mirabilis strains have been reported to carry distinct plasmids, at least some of which encode antibiotic resistance genes (130).

C. Transcriptomics

P. mirabilis undergoes dramatic shifts in its appearance and surroundings, and it follows that many of these responses result from changes in transcription. Global analysis of gene expression (transcriptomics) has been conducted for P. mirabilis under multiple conditions, initially using microarrays and, more recently, RNA-seq.

Microarrays

The four following studies (109, 126, 131, 132) were conducted using a microarray with 70-mer oligonucleotide probes designed to hybridize with each of the 3,719 predicted open reading frames from an early draft version of the P. mirabilis HI4320 genome (112).

Swarming microarray

Perhaps the most dramatic change P. mirabilis undergoes is the cyclic switch from elongated, hypermotile swarmer cells to short, nonmotile consolidated cells during culture on solid surfaces (19). This transformation was the target of the first P. mirabilis transcriptomic study, where (i) swarm and consolidate transcripts were shown to be more similar to each other than to RNA derived from broth culture, and (ii) while many genes are induced during consolidation, very few are upregulated during active swarming (126). The data were consistent with earlier reports of increased expression of certain virulence genes during swarming (133, 134), and swarmer cells being less metabolically active than consolidates (135). Curiously, although P. mirabilis is largely nonmotile during consolidation, flagellar transcripts were among the most highly-transcribed messages in both swarm and consolidate (126). Based on the accumulated transcription and metabolic data, the consolidation phase has been proposed to be a time of preparation for the next round of swarming, by increasing nutrient uptake systems, central metabolism, and respiration. As P. mirabilis swarming is very robust and occurs under conditions not permissive for other species of swarming bacteria, the preparation that occurs during consolidation phase may contribute to how P. mirabilis is able to swarm under a wider range of conditions that other species, including on solid surfaces such as hard agar and catheters.

Iron limitation microarray

Iron is sequestered by the host to combat bacterial infection, and pathogens activate iron acquisition systems during UTI (link metals section). P. mirabilis was previously thought to lack common iron scavenging systems in earlier studies. However, microarray analysis conducted on iron-restricted P. mirabilis revealed 21 iron-regulated systems, including two new acquisition systems which are detailed later in this chapter (109).

Ascending UTI microarray

A critical point for studying how P. mirabilis causes disease requires knowledge regarding the physiological state of this pathogen during UTI. To answer this question, urine was collected from experimentally infected mice at 1, 3, and 7 days post-inoculation (dpi) for RNA extraction and microarray hybridization (131). Iron acquisition, urease, and peptide transporters were induced during UTI, consistent with an iron-restricted, urea- and peptide-rich urine environment. The genes encoding MR/P fimbriae (link fimbriae section), which were previously shown to contribute to infection, were the most-highly upregulated of the transcriptome compared with broth culture (Fig. 13). However, some known virulence factors were either poorly expressed or downregulated during UTI, including genes encoding PMF fimbriae, Zap metalloprotease, hemolysin, and Proteus toxic agglutinin (Pta). This suggests that voided urine provides an important, but incomplete, snapshot of bacterial gene expression during UTI.

An external file that holds a picture, illustration, etc.
Object name is nihms925686f13.jpg
P. mirabilis gene expression during experimental UTI

(A) Heat map of expression data for specific virulence-associated genes, depicting the ratio of expression in LB broth versus in vivo. The legend at the left indicates the color associated with log2 fold change: red, upregulated in vivo; green, downregulated in vivo; black, not differentially regulated. (B) Adherence and motility genes are inversely regulated during UTI. Each line represents fold-change of a specific flagellar (left panel) or fimbrial (right panel) gene in vivo relative to mid-logarithmic phase culture in vitro. Genes in the mrp operon are highly induced early during infection, but expression falls by 7 days post infection. Flagellar genes are initially repressed, but expression increases late in infection. Figure adapted, with permission, from (131).

Interestingly, central metabolic pathways that were required during infection differed in several ways compared with a previous study of uropathogenic E. coli (136). Nitrogen assimilation gene expression suggested that P. mirabilis has greater access to nitrogen compared with E. coli, but follow-up experiments indicated this was not simply due to ammonia derived from urease activity (131). In contrast with E. coli, P. mirabilis induced glucose uptake and glycolysis genes, suggesting that P. mirabilis is able to access nutrient sources not available to E. coli (perhaps due to the activity of toxins on bladder epithelium). Gene expression at 1, 3, and 7 days post-inoculation was also compared. Flagellar genes were initially repressed, but this repression was relieved by day 7. In contrast, mrp genes were highly-transcribed throughout the infection but their degree of induction was lower at day 7; this topic is further detailed in the fimbrial section (link fimbriae section).

The transcriptional regulator MrpJ is encoded by the mrp fimbrial operon (link fimbriae section). Originally proposed as a switch between adherent and motile states (137), MrpJ was found to contribute to other phenotypes as well (138). Microarray experiments comparing an mrpJ in vivo mimic, where mrpJ is expressed in trans at levels comparable to those detected in mice during experimental UTI, with an empty vector control showed that MrpJ’s regulatory network extends considerably beyond repression of flagella to include, among others, regulation of additional fimbriae and the type VI secretion system (132). MrpJ and MrpJ paralogs are discussed later in this chapter (link).

RNA-seq

The next three studies (127, 139, 140) compared gene expression for targeted mutants with their isogenic parent strain using RNA-seq. In addition to the aforementioned microarray study, broth-cultured and swarmer cells have also been compared by RNA-seq using the BB2000 lab strain (127). Even though the methods for collecting and analyzing swarmer cells were different in these two studies (126, 127), similar results were obtained for the broth-swarmer comparison regarding the genes encoding flagella, lipoproteins, iron transport, and hemolysin, among others. FliL is a flagellar protein that, when mutated, leads to elongated swarmer-like cells under non-swarming conditions (pseudoswarming) (141). Comparison of a fliL mutant cultured in broth to induce pseudoswarmer cells with the wild-type broth RNA-seq dataset revealed an increase in propanediol utilization in the fliL mutant. This suggests a possible new role for these genes in swarmer cell differentiation (127).

The Rcs phosphorelay is a repressor of swarming and an activator of biofilm formation (139, 140, 142144). RNA-seq was used to compare a broth-cultured rcsB mutant with its isogenic parent strain PM7002 (139, 140). In this condition, the rcsB mutant undergoes inappropriate elongation that is similar to the fliL phenotype. These studies showed that RcsB controls pathways important for swarming, including cell division (minCDE) and motility (flagella and flagellar regulators). Furthermore, RcsB activates biofilm-associated genes, including genes encoding MR/P, PMF, and UCA fimbriae (link biofilm section). Interestingly, RcsB also represses virulence genes that have previously been shown to be activated during swarming (zapABD, hpmBA) and type VI secretion genes (tss, ids, and idr/pef operons).

D. Mutagenesis

The first targeted mutation to be made in P. mirabilis was a ureC urease mutant (40) (Fig. 14). This mutant was constructed using a single crossover, Campbell-type insertion via an R6K ori-pir suicide vector introduced by conjugation (145), an approach that is still used by multiple groups studying P. mirabilis. This initial paper noted several difficulties with methods for mutant construction in P. mirabilis, as conjugative-incompatible plasmids, phage transduction, and spontaneous curing of multicopy plasmids all resulted in failure to yield chromosomal mutations (40). Later on, sacB counterselection was employed to facilitate identification of double crossover events (146). The ability to directly construct and test mutants in the mouse UTI model was a boon for dissecting P. mirabilis virulence, and allelic exchange via mating with suicide plasmids continues to be used for mutant construction. Even so, mutant construction remained labor-intensive and prone to failure, particularly in some clinical isolates.

An external file that holds a picture, illustration, etc.
Object name is nihms925686f14.jpg
Integration of the urease suicide plasmid into the chromosome of P. mirabilis HI4320 by homologous recombination

E. coli SM10 λ pir (Kanr), carrying transfer genes integrated into its chromosome, was used to mobilize pBSJ102 (Apr) into P. mirabilis HI4320. Integration of pBDJ102 into the chromosome was selected on medium containing tetracycline and ampicillin. Transconjugants (Apr Tetr) were urease negative. Bla, β-lactamase; C, ClaI; Ei, EcoRI; Ev, EcoRV; H, HindIII; N, NruI; P, PvuI. The 1.5-kilobase HindIII fragment of ureC (solid), a 277-base-pair deletion (striped) and chromosomal urease gene sequences (open) are shown. Adapted from (40), with permission.

Subsequently, sequencing of the HI4320 genome revealed that this strain naturally carries a plasmid that is closely related to R6K and encodes a π protein (112), possibly contributing to the difficulty in constructing mutants in this background using typical suicide vectors. Thus, another method was developed for constructing mutations, using a group II intron mutated to specifically target a gene of interest (147, 148). Advantages over the prior allelic exchange method are the relatively short time required, higher rate of success, and lack of passaging or counterselection that would normally be required to resolve double crossover events. The drawbacks are that the choice of insertion sites is limited, mutations have polar effects on downstream genes, and the method is insertion-based (that is, it is not simple to delete sequences). The latter problem can be managed by use of cre-lox recombination to remove the antibiotic resistance cassette carried within the intron, resulting in an unmarked mutant (126). Additional introns can be inserted, with the potential to remove intervening sequences via an additional round of cre-lox recombination.

Genes that have been specifically mutated and tested in a mouse model of UTI, either independently or by direct cochallenge against a parental strain, are shown in Table 4. It is important to note that the studies presented in Table 4 include a range of P. mirabilis strains, mouse strains, infectious doses, and experiment lengths, but exclude mutants generated by transposon insertion and targeted mutants that were only tested for virulence in an in vitro assay, such as cell culture.

Table 4

Targeted mutations that have been tested in mice

GeneLocusaDescriptionVirulence genebReference
ureCPMI3685Urease alpha subunitY(39, 40, 42, 44)
hpmAPMI2057HemolysinN(45, 48, 55)
mrpAPMI0263Major mannose-resistant/Proteus-like fimbrial proteinY(44, 149152)
pmfAPMI1877PMF major fimbrial subunitY(153, 154) (151)
flaDPMI1621Flagellar hook-associated protein 2Y(45)
flaABPMI1619–20Flagellin 1 and flagellin 2N(155)
mrpGPMI0269MR/P fimbrial subunitY(146)
zapAcPMI0279MetalloproteaseY(134)
mrpHPMI0270MR/P fimbrial adhesinY(156)
atfAPMI2728ATF major fimbrial subunitN(157)
mrpJPMI0271Fimbrial operon regulatorY(132, 137)
mrpIPMI0262Fimbrial recombinaseY d(158)
luxSPMI0379S-ribosylhomocysteine lyaseN(159)
ureRPMI3681Urease operon transcriptional activatorY(160)
spa47PMI2692Type III secretion system ATPaseN(148)
PMI0047Secreted 5′-nucleotidaseN(161)
rafYPMI0288Putative glycoporinN(161)
PMI0842TonB-dependent receptorY(161)
fadLPMI1810Long-chain fatty acid transport proteinN(161)
PMI2596TonB-dependent receptorY(161)
ptaPMI2341Proteus toxic agglutininY(55)
znuCPMI1151High-affinity zinc uptake system ATP-binding proteinY(162)
zurPMI2748Zinc uptake regulation proteinY(162)
dppAPMI2847Dipeptide ABC transporter, substrate-binding proteinY(126)
oppBPMI1474Oligopeptide ABC transporter, permease proteinN(126)
cysJPMI2250Sulfite reductase [NADPH] flavoprotein alpha-componentY(126)
lrhAPMI0629LysR-family transcriptional regulatorN(126)
hexAPMI1764LysR-family transcriptional regulatorY(126)
pbtAPMI0232Siderophore biosynthesis proteinN(109)
nrpRPMI2599Siderophore synthaseY(109)
aipAPMI2122Adhesion and invasion autotransporterY(46)
taaPPMI2575trimeric autoagglutinin autotransporter of ProteusY(46)
gdhAPMI3008NADP-specific glutamate dehydrogenaseY(131)
glnAPMI2882L-glutamine synthetaseY(129)
ucaAPMI0536UCA major fimbrial subunitY(163)
hfqePMI3365Hfq (host factor-I protein)Y(164)
speAPMI2094Biosynthetic arginine decarboxylaseY(165)
speBFPMI2093; PMI0307Agmatinase; ornithine decarboxylase, inducibleY(165)
pfkAPMI32036-phosphofructokinaseY(166)
pgiPMI2754Glucose-6-phosphate isomeraseY(166)
tpiAPMI3205Triosephosphate isomeraseY(166)
pykAPMI1155Pyruvate kinase IIY(166)
gndPMI06556-phosphogluconate dehydrogenase, decarboxylatingY(166)
talBPMI0006Transaldolase BY(166)
eddPMI27606-phosphogluconate dehydrataseY(166)
sdhBPMI0568Succinate dehydrogenase iron-sulfur proteinY(166)
frdAPMI3588Fumarate reductase flavoprotein subunitY(166)
fumCPMI1296Fumarate hydratase, class IIY(166)
pckAPMI3015Phosphoenolpyruvate carboxykinase [ATP]N(166)
argGPMI3239Argininosuccinate synthaseN(166)
serAPMI2031D-3-phosphoglycerate dehydrogenaseY(166)
rpoEPMI1894RNA polymerase sigma-E factorY(167)
rseAPMI1893Anti-sigma E proteinN(167)
PMI1518High-affinity nickel efflux proteinY(129)
pldAPMI3344Phospholipase AN(129)
ilvDPMI3302Dihydroxy-acid dehydratase (branched chain amino acid biosynthesis protein)Yf(129)
lonPMI0117ATP-dependent proteinase, serine peptidaseY(129)
argRPMI3399Transcriptional regulator, repressor of the arginine biosynthetic pathwayY(129)
aBased on P. mirabilis HI4320 annotated genome
bSignificant in one or more sites tested (urine, bladder, kidneys, spleen)
cAlso found to contribute to virulence in a rat prostatitis model (168)
dMR/P locked-on mrpI mutants outcompete locked-off mrpI or wild-type (see fimbrial section)
eAlso found to contribute to virulence in a rat skin wound model (164)
fContributed to fitness during polymicrobial infection with Providencia stuartii, but not during single-species infection (129)

E. Genome-wide mutagenesis

Transposon mutagenesis in P. mirabilis was first described in a publication from 1991 using a Tn5 derivative (169, 170). The mini-Tn5 transposon was carried on a suicide plasmid with the π protein-dependent origin of replication from R6K, oriT for conjugal transfer from donor strain to recipient strain, and the IS50R transposase gene tnp. This approach has since been used to generate libraries for isolation of mutants defective for a given phenotype, such as swarming motility (119, 143, 171175), sensitivity to antimicrobial peptides (80, 176), and biofilm formation (177).

In 1995, this approach was revolutionized through the development of signature-tagged mutagenesis (STM) (178), in which each transposon mutant has a specific DNA sequence within the transposon that acts as a barcode. STM led to the first genome-wide pathogenesis studies by permitting identification of unique mutants within a mixed population and allowing for assessment of mutants lost during the selective process of the infection, and therefore containing transposon insertions in genes that contribute to colonization and pathogenicity. This approach was successfully utilized three times with P. mirabilis strain HI4320 in a murine model of ascending UTI, identifying a combined total of 54 genes that represented important fitness factors for colonization (summarized in Table 5) (108, 179, 180).

Table 5

Virulence genes identified by STM and Tn-Seq.

PMI NumberGene NameIsolateSTM of Ascending UTICAUTIPolymicrobial CAUTI
Reference(180)(108)(179)(129)(129)
PMI0012nhaR4B6U, K, O
PMI0020carAD4–8U, B, KB, K
PMI0030exbD7A1U, B, K, OB
PMI0283zapE29C6K
PMI0565sdhC10E3U, B, K, O
PMI0589aroGG1–38U, B, K
PMI0711serCG4–12U, OB, K
PMI0842D6–33U, B, KK
PMI10008H4U, OK
PMI1045arnAG5–30U, B, KB, K
PMI118416F1K
PMI119340A6U, B, K, O
PMI1421ppsA17A4B, K, OB, K
PMI14486F6U, O
PMI1546guaB5F1U, B, K, OB, K
PMI1598yidA12E4U
PMI1607sdaA15C3UKK
PMI1630fliF8A6B, KBB, K
PMI1651flgE21B2KB, K
37D5U, B, O
PMI1667cheWE6–43U, B, KB, K
PMI1761nuoB15A6U, B, K, O
PMI1874E5–36U, B, KK
PMI201432C4B
PMI2046aceE8B5B, K, O
PMI2259metN23D5U, K, OB, K
PMI2332surA11C4U, B, K, OKK
32E1B
PMI2342sufI38E2K, OK
PMI2345parEC6–15U, B, K
PMI2378cbbC11G3U, K, OB, K
PMI257536D2B, O
PMI2605nrpGA1–8U, K, OK
PMI2760edd11D4U, KB, K
PMI2823ppiA6D1U, K, O
PMI2828dsbAF4–37U, B, KB, K
PMI2893pstSH4–34U, B, KB, K
PMI2894pstCG1–43U, B, KB, K
PMI30014A3U
11A4U, K, O
PMI3021mrcAB6–23U, B, KK
PMI3053asnCE6–41U, B, KB, K
PMI3190cpsF11B2B, K, OK
20C4U, B, K, O
PMI3295hdfRD5–9U, OK
C1–18B, O
PMI3329hemYB2–9UB
PMI3333cyaA20E1U, B, K, OK
PMI335913D1U, O
PMI3390F6–34U, B, KK
A3–39U, B, K
PMI3441B2P
PMI3451nrdDD6–39B, K, OB, K
PMI3623dusBCF1–29U, B, KB, KB,K
PMI3645rapZB5U, B, KK
PMI3687ureFB6–40U, B, KK
PMI3692D6–26U, B, K
PMI3700D1–30U, B, KB, K
PMI37051F6U, K, OB, K
PMIP09pilX4C4–41B, K, OKK
A5–34U, K, OK
B4–38KK

P, primary screen; U, urine; B, bladder; K, kidney; O, attenuated overall

In the first P. mirabilis STM study, mice were inoculated with 96 tagged mutants as the input sample, and bacteria recovered from the bladders of 5 mice 2 days post-inoculation were used as the output samples (180). Only two mutants were reproducibly underrepresented in the output samples from all five mice: B2 (determined to be an insertion in a secreted protease, now identified as PMI3441, a U32-family peptidase), and B5 (determined to be an insertion in the rpoN operon, now identified as PMI3645, an RNase adaptor protein RapZ). The fitness defect of each mutant was assessed by direct cochallenge with wild-type P. mirabilis HI4320, and only the PMI3645 mutant remained significantly outcompeted.

The second P. mirabilis STM study used 45 pools of mutants as the input samples (2,088 mutants total), and bacteria recovered from the bladders or kidneys of 5 mice per pool 4 days post-inoculation were used as the output samples (108). 502 mutants were identified as attenuated in the primary screen. A secondary screen was conducted with these 502 mutants divided into 19 pools, and only 114 of the mutants were reproducibly attenuated. 84 of the 114 mutants were re-tested by direct cochallenge against wild-type P. mirabilis HI4320 for a longer experimental challenge of 7 days, and 37/84 (44%) of the mutants exhibited reproducible fitness defects. Nucleotide sequences were obtained from 30 of the mutants, 27 of which map to open reading frames in the current annotation of P. mirabilis HI4320. This approach identified several categories of mutants representing both known and novel fitness factors for UTI, including motility, iron acquisition, plasmid-encoded factors, transcriptional regulation, phosphate transport, urease activity, capsule synthesis, and metabolic pathways, the majority of which would have been difficult to identify as fitness factors by other methods.

The third P. mirabilis STM study screened 40 pools of mutants as the input samples (1,880 mutants total), and bacteria recovered from the bladders or kidneys of 5 mice per pool 4 days post-inoculation were used as the output samples (179). 570 mutants were identified as attenuated in the primary screen, 217 of which had a reproducible defect in the secondary screen. 93 of the most-attenuated mutants that did not exhibit competitive defects in vitro were re-tested by direct cochallenge against wild-type P. mirabilis, and 37 (40%) had reproducible fitness defects. Nucleotide sequences were obtained from 29 of the mutants, 28 of which map to open reading frames in the current annotation of P. mirabilis HI4320. This approach again identified fitness factors pertaining to cellular processes, transport, transcriptional regulation, motility, cell surface structures, and metabolism. However, there was no overlap in the genes or operons identified in this study and the prior two STM studies; as the three combined STM studies only achieved 70% theoretical genome coverage, this is likely due to incomplete saturation and differences between the mutants present in each pool for in vivo competition.

In 2009, the basic technique of STM was further improved upon and integrated with next generation sequencing techniques to allow for massively parallel sequencing and quantitation of the relative abundance of all mutants present in a given setting. Several variations of the method were developed concurrently, including insertion sequencing (INSeq) (181), transposon-directed insertion site sequencing (TraDIS) (182), transposon insertion-site sequencing (Tn-Seq) (183, 184), and high-throughput insertion tracking by deep sequencing (HITS) (185). As Tn-Seq has been applied to the study of P. mirabilis pathogenicity (129), the specifics of this method will be explained in further detail.

The first step of Tn-Seq is to generate a genome-saturating transposon mutant library, generally using a Mariner transposon. Due to the saturating nature of the library, this method can also be used to identify genes that are essential for growth in vitro, as insertion mutants within these genes will be lacking from the starting library. Infection studies are then conducted with the transposon library, and all bacteria in the target site of infection (for instance, the bladder) are collected for extraction of genomic DNA. The frequency of each insertion mutant is then determined, en masse, from the input sample (the saturated library) and the output sample (the bladder), and the fitness contribution of each gene in the genome can be assessed by the change in frequency of insertion mutants for that gene from the output sample compared to the input sample.

This approach was successfully used to generate the first genome-saturating library of transposon insertion mutants in P. mirabilis HI4320, and allowed for identification of 436 genes (12% of the genome) estimated to be essential for growth in Luria broth (129). As would be expected, the essential genes largely pertained to cell cycle control and division, cell wall biogenesis, replication, and ribosomal proteins, and 64% had been identified as essential genes in other bacterial species. The transposon mutant library was then used in a murine model of CAUTI to identify fitness factors for colonization of the catheterized bladder and ascension to the kidneys 4 days post-inoculation (Fig. 15). Massively parallel sequencing and statistical analysis identified 629 genes (17% of the genome) as fitness factors for CAUTI; 93 genes were specifically important for colonization of the catheterized bladder, 209 for colonization of the kidneys, and 286 that were important for colonization of both the bladder and kidneys. Of the 54 genes and operons identified as fitness factors for ascending UTI in the combined STM studies, 31 (57%) were also identified in the CAUTI model by Tn-Seq (summarized in Table 5). Eight mutants were constructed for direct cochallenge with the parental strain to confirm the fitness defects that were identified by Tn-Seq, and fitness defects were verified for 7/8 (88%). Tn-Seq therefore provided confirmation of a role for known P. mirabilis fitness factors during CAUTI, identification of CAUTI-specific fitness factors, and a model for exploring the different metabolic pathways and transport requirements between ascending UTI and CAUTI.

An external file that holds a picture, illustration, etc.
Object name is nihms925686f15.jpg
Conceptual model of single-species and polymicrobial CAUTI Tn-Seq

For each of five transposon mutant library pools, mice were infected as follows: 1) 5±10 CBA/J mice were transurethrally inoculated with 1×105 CFU of the transposon library for single-species infection, and 2) 5±10 CBA/J mice were inoculated with 1×105 CFU of a 1:1 mixture of the transposon library and wild-type P. stuartii BE2467 (purple) for coinfection. Thus, for each input pool, the single-species infections and coinfections were conducted in parallel to utilize the same input inoculum. Input and output samples were enriched for transposon-containing sequences and subjected to next generation Illumina sequencing of the transposon-chromosome junctions. The resulting reads were mapped to the P. mirabilis genome, and the abundance of reads at each insertion site from all output samples were compared to the input samples to determine a fold change for each gene. The gene in yellow represents a candidate P. mirabilis fitness factor for single-species CAUTI that is even more important during coinfection; the gene in blue represents a P. mirabilis fitness factor for single-species CAUTI that is no longer important during coinfection; the gene in red represents a factor that does not contribute to P. mirabilis CAUTI and was therefore recovered at a similar density from the infection output pools as the input pools. Reproduced from (129), with permission.

Importantly, this approach was also applied during polymicrobial infection with P. stuartii, in parallel with the above study, to determine how polymicrobial infection influences P. mirabilis fitness requirements (129) (Fig. 15). Of the 629 genes identified as P. mirabilis fitness factors for CAUTI, 217 (35%) were also important during polymicrobial infection. The fitness factors that were important for both infection types include fimbrial genes, components of the flagellar cascade, urease, and factors involved in inorganic ion transport and metabolism. Interestingly, an additional 1,353 candidate fitness factors were identified that appear to specifically contribute to P. mirabilis fitness during polymicrobial infection, including defense mechanisms (such as type III secretion and type VI secretion), signal transduction pathways, at least 10 distinct fimbrial types, iron uptake systems, respiratory nitrate reductase, the oxidative pentose phosphate pathway and the Entner-Doudoroff pathway, arginine import and biosynthesis, and branched chain amino acid (BCAA) biosynthesis. It was further determined that the requirement for BCAA biosynthesis during coinfection was due to high-affinity import of leucine by P. stuartii, indicating the utility of this technique for dissecting polymicrobial interactions during infection (129).

IV. Virulence Factors

A. Urease

P. mirabilis was the model organism for which the genetics of urease were first delineated. Urease is a nickel-metalloenzyme, the synthesis of which is induced by the presence of its substrate, urea (186, 187).

Structure and activity

Apourease is comprised of three structural subunits, UreA, UreB, and UreC, assembled as a homotrimer of individual UreABC heterotrimers (UreABC)3 (Fig. 16). To become catalytically active, the apoenzyme must acquire divalent nickel ions through a process involving four accessory proteins, UreD, UreE, UreF, and UreG. As studied by yeast two-hybrid and immunoprecipitation with monoclonal antibodies to UreC and UreD, it is clear that the accessory proteins interact with the apoenzyme to deliver Ni2+ to the active site (188). Indeed, UreE carries a natural histidine tail that allows it to carry Ni2+ and also to be purified in a single step on a Ni2+-nitrolo-acetic acid column.

An external file that holds a picture, illustration, etc.
Object name is nihms925686f16.jpg
Model of P. mirabilis urease interactions with structural and accessory proteins based on yeast two-hybrid experiments

The three-dimensional structure of urease, inferred from the closely related urease of K. aerogenes. (A) UreD associates with UreC in the context of the apourease independently of the UreA structural protein. Although UreD and UreF interact in the absence of structural proteins, UreD is still capable of associating with the apourease without coaccessory proteins such as UreF. (B) A single molecule of UreD associated with UreABC may interact with additional UreD molecules bound to adjacent UreABC heterotrimers. These interactions could stabilize the accessory protein interactions with the apourease and hypothetically coordinate nickel uptake among the three active sites of urease. A similar hypothesis applies to UreF; homomultimeric UreF interactions in vivo could occur between individual UreF molecules bound through UreD to adjacent UreABC heterotrimers. (C) The 6,500-bp P. mirabilis urease gene cluster encodes eight proteins that comprise, regulate, and assemble the urease homoenzyme. Figure adapted, with permission, from (188).

In early studies using bacterial cell lysates of four P. mirabilis isolates, the molecular size of urease was originally estimated by molecular sieve chromatography to be 281–338 kDa (189), and the molecular size of the apoenzyme was later determined to be 252,600 Da (190). Including the two Ni2+ in each of three active sites, this brings the predicted molecular size to ~253 kDa with an isoelectric point of 5.9 for P. mirabilis HI4320. The affinity for substrate as estimated by Km is relatively weak, ranging from 22–60 mM urea among the four isolates. However, with urea concentrations of ~400 mM urea in human urine, the enzyme would be fully saturated and working at the Vmax within the urinary tract and therefore extremely potent.

Genetic properties and organization of the ure operon

The urease operon was originally characterized from a DNA fragment of 7.6 kb, derived from a cosmid gene bank of P. mirabilis genomic DNA. Six open reading frames were found within a 4,952 bp region, which were predicted to encode polypeptides of 31.0 (UreD), 11.0 (UreA), 12.2 (UreB), 61.0 (UreC), 17.9 (UreE), and 23.0 (UreF) kDa (190). Reexamination of recombinant clones for urease activity revealed that a seventh gene, ureG, was also required for production of catalytically-active urease enzyme (191) Sequences that preceded these genes later determined to confer urea-inducibility upon the operon (192), resulting in identification of UreR, a 33.4 kDa helix-turn-helix AraC family member responsible for urea-induction of the operon.

Regulation by UreR

UreR is a dimer that binds to two sites within the ureR-ureD intergenic region at a consensus binding site of TA/GT/CA/TT/GC/TTA/TT/AAATTG as predicted from DNASE I protection assays (193). UreR binds to these sites in a urea-dependent manner to activate expression of the urease operon. In the absence of urea, there is no measurable binding or induction of urease expression.

Regulation by H-NS

A poly(A) tract nucleotide sequence of A6TA2CA2TGGTA5GA6TGA5 is present 16 bp upstream of the ureR promoter, which serves as a binding site for the gene repressor histone-like nucleoid structure protein (H-NS) (194, 195). Using a ureR-lacZ reporter plasmid, it was determined that H-NS represses transcription of ureR, and therefore represses urease expression (194, 196). H-NS and UreR compete for binding the ureR-ureD intergenic region, and culture conditions dictate whether the operon will be repressed or activated. H-NS binding is favored at 25°C and in the absence of urea, while UreR binding and derepression of the operon is favored at 37°C when urea is present (196).

Role in virulence

Urease is a critical virulence determinant for P. mirabilis urinary tract infection. Indeed, the first targeted mutation of this species involved the mutation of ureC, the major structural subunit of the apoenzyme (40). The ID50 of the ureC mutant is approximately 3 logs higher than the parental strain, clearly indicating the contribution of urease to disease severity. In an independent challenge with 109 CFU of the wild-type strain or the ureC mutant, the mutant was attenuated by approximately 2 logs in the bladder and kidneys 48-hours post-inoculation and up to 6 logs by 7 days post-inoculation (39, 40). Infection with the ureC mutant also resulted in significantly less pathology in the bladder and kidneys of infected mice. It was later shown that a ureR mutant lacking urea-induction of urease synthesis also resulted in attenuation by >3 logs in the bladder and kidneys (160).

The contribution of urease to pathogenicity has also been assessed in the uncomplicated murine model of UTI compared to a model of catheter-associated UTI (CAUTI), in which a catheter segment is placed in the bladder during inoculation (25). These studies determined that urease significantly contributes to cystitis in both models, but does not play as dramatic a role in pyelonephritis in the CAUTI model. However, urease activity strongly correlated with renal tubule damage and nephrosis during both UTI and CAUTI in this study.

Potentiation of urease activity during polymicrobial infection

Urine colonization in catheterized individuals is frequently polymicrobial, and P. mirabilis is one of the most common organisms present during polymicrobial colonization and CAUTI (4, 8, 9, 17). It is therefore noteworthy that co-culture with other uropathogens can enhance production of active urease enzyme by P. mirabilis (25, 41). In addition to P. mirabilis, the most common CAUTI uropathogens are Enterococcus species, Escherichia coli, Pseudomonas aeruginosa, and Klebsiella pneumoniae, and multiple isolates of all four species are capable of enhancing P. mirabilis urease activity (25). This has important implications for the pathogenesis of polymicrobial CAUTI, as enhanced urease activity resulting from co-culture of P. mirabilis and P. stuartii was determined to significantly impact disease severity, being directly associated with increased urine pH, greater incidence of urolithiasis and bacteremia, and induction of a more potent and destructive inflammatory response (25, 41).

Urease inhibitors

Due to the important role of urease during P. mirabilis pathogenesis in the urinary tract, urease inhibitors have been explored as a potential therapeutic (197). Such inhibitors include hydroxyamic acids, which bind the nickel atoms in the urease active site, and acetohydroxamic acid (AHA), a structural analog of urea. AHA is the only analog to have been tested in clinical trials and be approved by the Food and Drug Administration. However, while it showed efficacy for preventing urolithiasis (198200), it has potentially severe side effects that limit its clinical use (201). Thus, further research is necessary to identify safe and effective urease inhibitors. It is also worth noting that while the ure operon is urea-inducible, structural analogs of urea such as AHA do not induce urease expression (202).

B. Secretion systems

Proteus mirabilis secretes proteins via Type I, III, IV, V, and VI systems as well as the Sec pathway and Twin Arginine Targeting (Tat) system. These pathways are predicted from the genome sequence (112), as analyzed by KEGG pathways (203) and from the individual studies described below. While examples of secreted proteins can be found for the systems noted, only a few examples of secreted proteins that have been tested for their contribution to virulence have been documented.

Proteases

In an early study, it was found that 90–94% of 48 P. mirabilis strains secreted proteolytic enzymes, detected at pH 8 using gelatin as substrate. A correlation between the ability to swarm and protease secretion was noted. These proteases were all EDTA-sensitive metalloproteinases (204). However, the specific secretion systems involved in secretion of these proteases were not determined in this study.

Type III Secretion

Nucleotide sequencing of the genome of P. mirabilis HI4320 (112) revealed genes appearing to encode a type III secretion system on a low GC-content pathogenicity island (148). A cluster of 24 genes clearly have the potential to encode an intact T3SS with genes for an intact needle complex and at least two effector proteins and their chaperones. The genetic organization of the T3SS is similar to that of Shigella flexneri, there was no evidence of mutation of the genes resulting in inactivation of the system, and RT-PCR analysis demonstrated that these genes were expressed in P. mirabilis HI4320. However, mutation of two genes including a putative ATPase and negative regulator of T3S had no effect on the secreted protein profile when compared to the wild-type strain (148). Furthermore, in the murine model of ascending UTI, there was no difference in CFUs between the mutant and wild-type strain in urine, bladder, or kidneys during either independent infection or cochallenge. Thus, no pathogenicity phenotype for the T3SS was identified for P. mirabilis HI4320 during ascending UTI. However, two components of the T3SS (PMI2688 and PMI2696) were identified as significant fitness factors during polymicrobial CAUTI by Tn-Seq (129), indicating that the T3SS may be involved in mediating host-microbe interactions during polymicrobial infection.

Type V Secretion

There are three subclasses of type V secretion systems (Va, Vb, and Vc), and the nucleotide sequence of P. mirabilis HI4320 predicts members of all three (112, 205, 206).

Classical autotransporters (Va)

Classical autotransporters are a family of virulence proteins in Gram-negative pathogens that contain three domains: an amino-terminal leader peptide for export across the inner membrane via the Sec pathway, a surface-localizing passenger domain, and a carboxy-terminal domain for translocation across the outer membrane (Henderson 2004). One of the three predicted classical autotransporters encoded by P. mirabilis HI4320 has been studied in detail: Proteus toxic agglutinin (pta), encodes a serine protease that was previously identified as an immunogenic outer membrane protein (161) (link Toxins section). The 117-kDa protein has a 58-amino acid long signal peptide, a 75-kDa long N-terminal passenger domain, and a 30-kDa C-terminal translocator (110). The autotransported protease, either cell-associated or secreted, has cytotoxic effects on cultured bladder and kidney epithelial cells. Catalytic residues Ser366, His147, and Asp533, when mutated, abolish protease activity. The protein also has autoaggregation properties not associated with the proteolytic activity. Virulence of a protease mutant, as tested in the murine model of ascending UTI, was significantly reduced.

Two-Partner Secretion (Vb)

P. mirabilis hemolysin, encoded by hpmA, is exported by the protein product of hmpB via the type Vb two-partner secretion pathway. This requires sequential unfolding of the HpmA protein (207), for which a partial crystal structure has been solved (208). Two polypeptides, HpmB and HpmA, synthesized in that transcriptional order, are responsible for hemolysin activity of P. mirabilis (link Toxins section). A Fur-binding site upstream of hmpB overlapping the −35 site of the promoter indicates that expression is governed by iron concentration. HpmB and HpmA have 17- and 29-amino acid leader peptides, confirmed for purified HpmA. HpmB is necessary for secretion of HpmA. HpmB and HpmA are most similar with respect to amino acid sequence to the Serratia marcescens hemolysin proteins ShlB and ShlA, respectively (209).

Trimeric autotransporters (Vc)

P. mirabilis HI4320 encodes three putative trimeric autotransporter proteins (Pearson 2008). Two of these, AipA and TaaP, are annotated as “adhesin-like” and “agglutinating adhesin-like”. Based on their homology with other trimeric autotransporters, the two autotransporters would contain four antiparallel β-sheets and form homotrimers. Recombinant AipA and TaaP bind extracellular matrix proteins, produce polypeptides of 28-kDa and 78-kDa, respectively, and form high molecular weight homotrimers. A 51-amino acid invasin-like motif of AipA is necessary for function. Gly-247 in AipA and Gly708 in TaaP are required for trimerization and activity. AipA and TaaP confer an advantage at 7 dpi during cochallenge in the murine model of ascending UTI.

Type VI secretion

P. mirabilis has long been observed to distinguish between disparate strains when swarming on agar surfaces, and there is now substantial evidence that this phenomenon is mediated by type VI secretion systems (T6SS). The discovery of T6SSs in P. mirabilis stemmed from an unusual observation about this species. The Dienes phenomenon, first described in 1946 by Louis Dienes, refers to the fascinating ability of two P. mirabilis swarming colonies of a single strain to merge with each other, while swarms of different strains form a line of demarcation where they meet (210) (Fig. 17). Formation of the Dienes line requires direct cell-cell contact by living bacteria, and involves killing of at least one strain at the boundary (211) (Fig. 18).

An external file that holds a picture, illustration, etc.
Object name is nihms925686f17.jpg
Contact-dependent preemptive antagonism is dependent on the T6SS

(A) A Dienes line (black arrows) forms between two different wild-type isolates, HI4320 and BB2000 (strain A and B kill each other). Loss of the T6SS (ΔT6) in either isolate by disruption of PMI0742 does not affect the discriminatory Dienes line (strain A kills strain B or strain B kills strain A). Loss of the T6SS in both isolates allows non-identical swarms to merge and the lack of T6SS-dependent killing appears as recognition (white arrow). (B) Swarm plate of the wild-type strain BB2000, BB2000 idsABCDEF deletion mutant (Δids), and BB2000 with the idsA-gfp vector pKG100 (labeled pPr-ids-gfp) or with an empty vector. A visible boundary formed between swarms of the wild type and the deletion mutant. Swarms of the wild type merged regardless of the presence of pKG100. The bar shown is 1 cm. Figure adapted, with permission, from (119) (A) and (212) (B).

An external file that holds a picture, illustration, etc.
Object name is nihms925686f18.jpg
Incompatible Dienes types have distinctive reactions when swarming colonies meet

(A) A P. mirabilis strain expressing red fluorescent protein (DsRed) (2R) intersecting with a strain expressing green fluorescent protein (GFP) (3G). Strain 3G produces round cells, whereas strain 2R produces no round cells. The dark areas are agar with no growth. Magnification, ×400. (Inset) Intersection and rounded cells in more detail (magnification, ×800). (B) Intersection zone for strain 2 expressing either GFP or DsRed (2G and 2R) without boundary formation or rounded cells (magnification, ×1,000). Scale bars = 50 μm. Figure adapted, with permission, from (211).

The structural components of the T6SS are Hcp and VgrG, which comprise the hollow tube and puncturing tip of the apparatus, VipA and VipB (or TssBC) that form the tube-like structures or sheath of the apparatus, and ClpV to mediate ATP hydrolysis for polysheath disassembly (120). Work conducted in P. mirabilis strain BB2000 identified an operon that contributed to Dienes line formation, named Ids for identification of self (175), which was found to encode Hcp and VgrG (175) and therefore provided the first link to T6S and Dienes line formation. Further investigations in P. mirabilis strains BB2000 and HI4320 determined that formation of the Dienes line represents either T6S-mediated killing of one strain by the other, or both strains killing each other using T6SSs (118, 119).

Using a transposon screen for P. mirabilis HI4320 mutants that failed to kill opposing strains, a 33.5-kb region of chromosome, different from the Ids locus, was identified as harboring the T6SS apparatus (PMI0749-PMI0733) and the primary effector operon (pefABCDEFG) (PMI0750-PMI0758) of this strain (119). This operon was found to encode a single immunity protein, PefE, which was responsible for providing immunity from killing by the Pef effectors. However, strain HI4320 was also found to encode at least four other potential T6S effector operons that could use the T6S apparatus, including idsABCDEF (175, 212) and three additional operons (PMI0207-PMI0212, PMI1117-PMI1121, and PMI1332-PMI1324) (112, 119) (Fig. 19). Differential use of these systems or possibly even recombination between them may therefore contribute to the wide range of strain-killing activities and Dienes types exhibited by P. mirabilis isolates.

An external file that holds a picture, illustration, etc.
Object name is nihms925686f19.jpg
The P. mirabilis HI4320 genome contains one primary and four orphan hcp-vgrG effector operons that are expressed during swarming

(A) A circular representation of the P. mirabils HI4320 genome depicting the location of the primary hcp-vgrG effector operon (hcp-vgrG1), divergent T6SS, and the four orphan hcp-vgrG effector operons (hcp-vgrG2-5). (B) PMI0750–PMI0758 encode the primary hcp-vgrG1 effector operon (pef) adjacent to the T6SS operon; PMI0207–PMI0212 encode the hcp-vgrG2 effector operon; PMI1117–PMI1121 encode the hcp-vgrG3 effector operon; PMI1332-PMI1324 encode the hcp-vgrG4 effector operon; and PMI2990–PMI2996 is the ids operon (hcp-vgrG5). Genes with homology to hcp (grey), vgrG (white), and predicted T6SS effectors (blue) are shown. Reproduced from (119), with permission.

A direct role for any individual P. mirabilis T6S operon during infection has yet to be elucidated. For instance, none of the T6S operons encoded by P. mirabilis HI4320 exhibited fitness defects in the murine model of CAUTI by transposon insertion-site sequencing (Tn-Seq) of monospecies infection. However, all of the known P. mirabilis T6S operons were strikingly overrepresented as fitness factors during coinfection with Providencia. stuartii (129), indicating a potential role for T6S in mediating competitive and cooperative interactions during polymicrobial infection. It remains to be determined if this system is important for mediating microbe-microbe interactions or microbe-host interactions in this infection model.

C. Toxins

While the complete genome sequence of P. mirabilis strain HI4320 revealed numerous potential toxins, only three have been well-characterized for a role in virulence. These are hemolysin, Proteus toxic agglutinin, and the ZapA metalloprotease.

Hemolysin

Secreted pore-forming toxins are a common feature of pathogenic bacteria. In particular, hemolysins are secreted pore-forming toxins that insert into eukaryotic cell membranes, causing efflux of sodium ions and cell damage (213). Two hemolysins have been described for members of the Proteus genus (214), one that is calcium-dependent and similar to the α-hemolysin of E. coli (hlyA), and another that is calcium-independent. The calcium-independent hemolysin is encoded by two genes comprising a two-partner secretion system: hpmA, which produces a 166 kDa secreted exoprotein, and hpmB, which produces a 63 kDa translocase protein that is required for both the activation and secretion of HpmA through cleavage of an N-terminal peptide (209). HpmA is the predominant hemolysin in Proteus species and appears to be the only hemolysin encoded by P. mirabilis, present in 268/274 (98%) of P. mirabilis clinical and fecal isolates from Brazil and the United States while none of these isolates encoded hlyA (102, 103). Similarly, hpmA is present in all 7 complete P. mirabilis genome sequences available through NCBI as of May 2017. An N-terminal fragment of HpmA, called HpmA265, has been crystallized and its structure solved (Fig. 20) (208). This fragment lacks the C-terminal pore-forming domain, but can still activate full-length HpmA in the absence of HpmB when the proteins are mixed with erythrocytes. The formation of HpmA265 dimers suggests that full-length HpmA may also function as a dimer.

An external file that holds a picture, illustration, etc.
Object name is nihms925686f20.jpg
HpmA265 crystallographic dry dimer interface leads to a filamentous appearance

Solid lines represent hydrogen bonds shared between β23 strands of both subunits. Reproduced from (208), with permission.

HpmA mediates lysis of a broad range of cell types from numerous host species and appears to be the primary P. mirabilis virulence factor responsible for cytotoxicity to human renal proximal tubular epithelial cells (HRPTECs) (47, 48, 50, 56). Deletion of hpmA dramatically decreases cytotoxicity, and allows for internalization of P. mirabilis by HRPTECs in vitro (47, 56). As HRPTECs form a protective barrier for the kidney parenchyma, it was hypothesized the hemolysin may be a critical virulence factor that mediates spread of P. mirabilis into the kidneys and development of pyelonephritis (28). However, deletion of hpmA does not appear to impact tissue colonization or damage during independent challenge in the murine model of ascending UTI (45, 48, 55), indicating that either hemolysin has less of an impact during experimental infection than the in vitro cell culture studies suggest, or the activity of other virulence factors mask its contribution in vivo. In agreement with this finding, hpmBA was not identified as a fitness factor for UTI or CAUTI in any of the genome-wide transposon mutagenesis studies (108, 129, 179, 180), likely due to a combination of complementation in trans by the other hemolysin-producing transposon mutants present during infection and possibly production of other cytolysins with similar functions.

Proteus toxic agglutinin

The genome of P. mirabilis strain HI4320 encodes six putative autotransporters (112) (link secretion section), of which only the Proteus toxic agglutinin (Pta) has been fully characterized. Pta was initially identified as an outer membrane surface-expressed protein recognized by the murine immune system (161), and determined to be encoded within the integrative and conjugative element ICEPm1 in P. mirabilis strain HI4320 (107) (link to ICEPm1).

Pta is a 120 kDa protein that is catalytically processed to become a ~110 kDa active protein within the outer membrane (110). Expression of this protein in E. coli promotes autoaggregation of the bacteria, as well as lysis of bladder epithelial cells in vitro. Both properties of Pta require translocation to the outer membrane and cytotoxicity requires serine protease activity from the passenger domain, but protease activity is not required for autoaggregation. Pta is also classified as a subtilisin-like alkaline protease owing to the observation that expression of pta is induced by alkaline pH and protease activity is maximal at pH 8.5–9.0, conditions that would be encountered by P. mirabilis during urinary tract infection due to its urease activity.

Pta contributes to colonization of the bladder and kidneys in a murine cochallenge model of ascending UTI, as well as dissemination to the spleen (110). Interestingly, HpmA and Pta have an additive effect on cytotoxicity both in vitro and during experimental UTI, particularly with respect to cystitis and possibly interstitial nephritis (55). However, Pta appears to be the more potent toxin during experimental infection, as disruption of pta has a much greater impact on infectivity than loss of hpmA. Similar to hpmA, pta is present in all seven of the complete P. mirabilis genome sequences in NCBI as of May 2017, and was also encoded in eight P. mirabilis infection and fecal isolates (55).

ZapA Metalloprotease

P. mirabilis produces a metalloprotease with broad specificity, originally thought to be an IgA protease but later determined to be capable of cleaving IgA, IgG, secretory component (the heavily glycosylated protein that complexes with dimeric and polymeric IgA), antimicrobial peptides hBD1 and LL-37, complement protein C1q and C3, fibronectin, actin, collagen, laminin, casein, and gelatin (215219). In all cases, cleavage was sensitive to the metal chelator EDTA, indicating that the enzyme is a metalloprotease. This protease was identified as a 55 kDa metalloprotease of the serralysin family, encoded by zapA (220, 221).

The contribution of ZapA to pathogenicity has been explored both in vitro and in vivo. Intact IgG and IgA1 can interact with Fc receptors on neutrophils to stimulate a respiratory burst, and this process is significantly reduced when IgG has been cleaved into Fab and Fc fragments in vitro by the P. mirabilis metalloprotease (218). Similarly, cleavage of hBD1 and LL-37 by ZapA reduces their antimicrobial activity (219). Thus, ZapA may contribute to evading the innate immune response during infection. A ZapA-deficient mutant, specifically constructed by insertional inactivation of zapA in the chromosome of P. mirabilis BB2000, was used to assess virulence in the murine model of ascending UTI (134). CBA mice were transurethrally inoculated into the bladder with 106 CFU P. mirabilis BB2000 or the zapA mutant. After 7 days, quantitative cultures indicated that the zapA mutant was dramatically attenuated (by 3 logs in the urine and 5 logs in the bladder or kidney), indicating that the protease contributed strongly to virulence of P. mirabilis BB2000. The same wild-type and mutant pair were tested in a rat model of prostatitis (168). Unlike the wild-type, inoculation of the zapA mutant resulted in reduced levels of acute prostatitis as determined by lower levels of tissue damage, bacterial colonization and inflammation (Fig. 21).

An external file that holds a picture, illustration, etc.
Object name is nihms925686f21.jpg
Tissue histology of acute and chronic P. mirabilis prostate infections

(a to c) Hematoxylin and eosin-stained rat prostate sections, showing the typical appearance of saline-treated controls (a) and WT (b)- and ZapA mutant (c)-infected prostate tissue in acute infection. (d to f) Histology of chronic infection for saline controls (d), WT infection (e), and ZapA mutant infection (f). Reproduced from (168) with permission.

D. Integrative and conjugative element ICEPm1

Pathogenicity islands (PAIs) are a specific group of genomic islands that contribute to genomic variability and virulence of bacterial pathogens. PAIs carry virulence determinants and are typically present in pathogenic strains but absent in nonpathogenic strains of the same species. These islands consist of large genomic regions ranging from 10–200 kb, and often have different G+C contents as compared to the host. They also frequently contain mobile genetic elements, are flanked by direct repeat sequences, have mosaic-like structure, insert at the site of tRNA genes, and have likely been acquired by horizontal gene transfer (222, 223).

A subset of PAIs has been identified that can excise from the bacterial chromosome, following a recombination event at the site of direct repeats flanking the island, and actively transfer via a type IV secretion system to another bacterium. These are called Integrative and Conjugative Elements (ICE) and many P. mirabilis strains carry an ICE designated ICEPm1 (107) (link comparative genomics).

ICEPm1 is a 94-kb region with a G+C content of 44.84%, which differs substantially form that of the P. mirabilis genome (38.88%) (107, 112), and contains core modules and a syntenic structure consistent with prototypical ICEs (113, 224) (Fig. 22). Direct repeats flank ICEPm1, and at the left-most end the 52-bp direct repeat is located with the 5′ coding end of tRNAPhe gene (attL). The direct repeat at the right-most end (attR) is part of the 3′ end of a truncated tRNAPhe gene. In the genome of P. mirabilis HI4320, ICEPm1 is annotated as being integrated into the phenylalanine tRNA gene pheV. However, ICEPm1 can integrate into either pheV or pheU sequences (111).

An external file that holds a picture, illustration, etc.
Object name is nihms925686f22.jpg
Characteristics of ICEPm1

(A) Modular structure of ICEPm1. Modules in yellow represent core modules; variable regions are depicted in gray. Direct repeats (DRs) are represented as triangles. (B) G+C content of ICEPm1 and the flanking chromosome. Boundaries of ICEPm1 are denoted by vertical black lines. Horizontal lines represent G+C content, with the middle line representing 39% G+C (that of the P. mirabilis HI4320 genome). DRs show the modularity of the island and suggest the evolutionary history of acquisition of regions of ICEPm1. Reproduced from (107), with permission.

ICEPm1 contains 91 ORFs, with the core genes constituting putative integration, replication and conjugative modules, including an integrase, genes for a putative tyrosine-like recombinase, putative helicase that could act as a relaxase, six transposases, and five plasmid-transfer related proteins (107). A 26-gene region encodes eight putative exported proteins and nine putative membrane proteins comprising the T4SS.

ICEPm1 also encodes known “cargo” genes that contribute to virulence, interspersed between the core modules. These cargo genes include the nrp operon (PMI2596-PMI2604), which encodes genes for the synthesis, transport, and uptake of an iron siderophore (link metals section), and Proteus toxic agglutinin (Pta), which contributes to aggregation of P. mirabilis and represents a potent autotransported protease toxin (46, 109, 110, 161).

Highly conserved homologs of ICEPm1 were present in all of 39 P. mirabilis urinary isolates tested and were also found in Providencia stuartii (6 of 10 isolates) and Morganella morganii (11 of 38 isolates) from cases of catheter-associated bacteriuria (107). ICEPm1 was also found at a reduced frequency in colonizing isolates of P. mirabilis (15 of 23 isolates) that were cultured from the oropharynx, nasopharynx, wound, groin, or perianal area.

When the ICE-encoded integrase is activated, the ICE can excise from the chromosome, form a circular intermediate, and subsequently transfer to a recipient cell via a mating pore formed by the ICE-encoded T4SS (113, 225). Transfers occurs at a frequency of 1.35 × 10−5 transconjugants/donor to ICEPm1-deficient P. mirabilis (111). Insertional inactivation of a putative integrase gene (PMI2549), a specific recombinase of the tyrosine-like family, on ICEPm1 decreases transfer frequencies to below the limit of detection. Mutation of the relaxase of ICEPm1 also eliminates transfer (111).

E. Fimbriae and adhesins

Fimbriae (pili) are hair-like protein structures that extend from the bacterial surface and usually mediate adherence to surfaces. Gram-negative bacteria produce a subset, called chaperone-usher fimbriae for their method of secretion and assembly (226, 227). P. mirabilis HI4320 encodes 17 chaperone-usher fimbrial operons (112). Although the function of most of these is not yet defined, transcription has been detected from all 17 operons, and most of these fimbriae are encoded by P. mirabilis isolated from diverse sites (Fig. 11) (106). For example, a PCR screen for the predicted major structural subunit gene for each fimbria showed that 85% of clinical isolates encode at least 14 unique fimbriae (106). Of seven clades defined for classical chaperone-usher fimbriae, three are represented in P. mirabilis1, γ2, and π) (Table 6) (106, 228).

Table 6

Classes of P. mirabilis fimbriae and their contribution to virulence.

FimbriaGenesClassaImplicated in virulenceb, cProteinMrpJ paralogRef
ICCOSTMSS CAUTICO CAUTIExpIn vitro
MR/P′PMI0254-PMI0261π[check][check]PMI0261(129)
MR/PPMI0262-PMI0271π[check][check][check][check][check]*PMI0271(44, 100, 106, 129, 131, 146, 149152, 158, 229, 240245)
Fimbria 3PMI0296-PMI0304π[check]PMI0296(129)
UCAPMI0532-PMI0536γ1[check][check][check][check][check][check]PMI0532(106, 129, 131, 163, 246251)
Fimbria 5PMI1060-PMI1067π[check]PMI1060(129, 131)
Fimbria 6PMI1185-PMI1190γ1None
Fimbria 7PMI1193-PMI1197γ1None
Fimbria 8PMI1464-PMI1470γ1[check]PMI1470(106, 131)
PMFPMI1877-PMI1881π[check][check][check][check][check]*None(106, 129, 131, 151, 153, 154, 243, 252)
Fimbria 10PMI2207-PMI2214γ1[check]PMI2209, PMI2207(129)
PMPPMI2216-PMI2224π[check][check][check]PMI2224(129, 253)
Fimbria 12PMI2533-PMI2539γ2[check]None(129)
ATFPMI2728-PMI2733γ1
An external file that holds a picture, illustration, etc.
Object name is nihms925686t1.jpg
An external file that holds a picture, illustration, etc.
Object name is nihms925686t1.jpg
[check][check]PMI2733(106, 129, 157)
Fimbria 14PMI2997-PMI3003ND[check][check][check]PMI3003(106, 129, 131, 179)
Fimbria 15PMI3086-PMI3093π[check][check]None(129, 131)
Fimbria 16PMI3348-PMI3352γ1[check]None(129)
Fimbria 17PMI3435-PMI3440γ1[check]None(129)

Table adapted from (19)

aGreek classification was determined as in (228) and (106). The operon encoding Fimbria 14 lacks a chaperone, and thus cannot be classified in the Greek system.
bThe mutant was tested in: IC, independent challenge experiment; CO, cochallenge experiment; STM, signature-tagged mutagenesis experiment; SS CAUTI, Tn-Seq CAUTI model with P. mirabilis alone; CO CAUTI, Tn-Seq CAUTI model during coinfection with Providencia stuartii; Exp, expressed in vivo; in vitro, virulence attributes (e.g., cell adherence, hemagglutination) detected using in vitro assays.
c[check]positive for the trait; An external file that holds a picture, illustration, etc.
Object name is nihms925686ig1.jpgnegative for the trait; ↑ induced in vivo compared with in vitro culture; ↓repressed in vivo compared with in vitro culture.
*Immunogenic

P. mirabilis may produce multiple types of fimbriae at once. For example, at least two fimbrial morphologies have been seen on a single bacterium by transmission electron microscopy (229, 230). In a mass spectrometry study, proteins from six of the 17 fimbriae were detected in a sheared surface protein preparation from a single broth culture (MR/P, UCA, PMF, ATF, Fim8, and Fim14) (106). This ability sets P. mirabilis apart from other bacteria such as E. coli that tend to produce one type of fimbria at a time (231). Likewise, P. mirabilis devotes a larger share of its chromosome to encoding fimbriae than most other fimbriated pathogens; for example, uropathogenic E. coli typically encode 9–12 fimbrial operons (98).

Because fimbriae are often found to contribute to virulence, there are multiple studies examining the roles of P. mirabilis-encoded fimbriae to UTI (Table 6) (19, 20, 232). P. mirabilis obtained from healthy stool or diarrhea, urine, or blood can adhere to voided human uroepithelial cells regardless of the bacterial source (233), which is consistent with the conservation of fimbrial genes in this species (106). In general, mutations in individual fimbrial genes result in bacteria that are still able to colonize the urinary tract, but are less fit in direct competition cochallenge experiments (151, 153, 234). This suggests that the 17 fimbriae have overlapping or redundant functions. Furthermore, as surface-localized, abundant, immunogenic proteins, P. mirabilis fimbriae have been tested as vaccine candidates with some success (link vaccines section) (235239).

Mannose-resistant Proteus-like fimbriae (MR/P)

These are the most-extensively studied of the fimbriae encoded by P. mirabilis. MR/P fimbriae were originally named for their ability to agglutinate untreated erythrocytes in the presence of mannose, a trait which was frequently found in P. mirabilis isolates (229) (Fig. 23). This was in contrast to mannose-sensitive agglutination, e.g., type 1 fimbriae from E. coli, or mannose-resistant Klebsiella-like (MR/K) agglutination, which occurred in the presence of tannic acid-treated bovine erythrocytes (254). MrpA is the major structural subunit of the fimbria, and MrpH is a two-domain adhesin which is located at the tip of MR/P fimbriae. Extensive details on MR/P structure and assembly have been recently reviewed elsewhere (19).

An external file that holds a picture, illustration, etc.
Object name is nihms925686f23.jpg
Mannose-resistant hemagglutination (MRHA) patterns of MR/P fimbriae expressed in E. coli

E. coli DH5α expressing the entire mrp operon under its native promoter (mrpA-J) (pXL4206), mrp minus mrpH and mrpJ (“ΔmrpH”; pXL4401), ΔmrpH plus empty vector (pON-184), or ΔmrpH plus complemented mrpH (pXL8906) were cultured in Luria broth at 37°C and mixed with chicken erythrocytes. MRHA only occurs when MR/P tip adhesin MrpH is present. Figure adapted, with permission, from (156).

Expression

The mrp gene cluster consists of two transcripts. Genes required for fimbrial structure and assembly are encoded by mrpABCDEFGHJ (156). The last gene, mrpJ, encodes a transcriptional regulator and is discussed later. The second transcript comprises a single gene, mrpI, which is divergently transcribed from the rest of the mrp genes (241). MrpI is a recombinase that flips a 252 bp invertible element containing the mrpABCDEFGHJ promoter (241). The transcriptional start site lies within the invertible element (132), and thus the orientation of this element dictates whether or not mrp fimbrial genes are transcribed. MrpI is the sole recombinase for the invertible element, and therefore mutations in mrpI result in bacteria that either constitutively produce MR/P fimbriae (“locked on”) or are devoid of MR/P fimbriae (“locked off”) (158). The orientation of the element may be easily detected using PCR-based assays (Fig. 24A), and phase variation of MR/P fimbriae has been observed at the single cell level (Fig. 24B).

An external file that holds a picture, illustration, etc.
Object name is nihms925686f24.jpg
Phase variation of P. mirabilis MR/P fimbriae

(A) Urine, bladder and kidney samples were collected at 7 dpi from mice infected with wild-type P. mirabilis HI4320, an MR/P locked-off mutant or an MR/P locked-on mutant and subjected to the invertible element assay. (B) Electron micrograph showing the phase variation of MR/P fimbrial expression in a broth culture of P. mirabilis. MrpH, the tip adhesin of MR/P fimbriae, is immunogold labeled. Note that the top left bacterium is gold labeled, while the top right bacterium is unlabeled. Scale bar, 500 nm. (C) Correlation between MR/P fimbrial expression and bacterial colonization in the bladder. Bladders from mice challenged with P. mirabilis HI4320 were collected at 7 dpi and bacteria were both quantitatively cultured and subjected to the IE assay. A positive correlation was found between MR/P fimbrial expression (percentage of IE in the ON orientation; y axis) and bacterial colonization in the bladder (log10 CFU/g tissue; x-axis): y = 17× − 30, r2 = 0.9, n = 18, P < 0.0001. Figure adapted, with permission, from (158) (A) and (152) (B and C).

The nine genes in the mrp operon are the most highly induced genes by bacteria in urine collected from experimentally-infected mice compared with in vitro broth culture (131). Correspondingly, even when mice are infected with P. mirabilis cultured so the population is almost completely phase off, after seven days, the majority of the bacteria have the invertible element in the on orientation (241). MR/P fimbriae are produced at a relatively low level in vitro, with the highest amounts produced during 48 h static culture in broth or culture under 5% oxygenation, followed by aerated broth, and finally, growth on an agar surface (152, 249, 255). Expression of mrp genes is both induced and subject to positive selection in a 5% oxygen atmosphere, which is logical for a virulence factor given reduced oxygen availability in the bladder (152). That is, an mrp locked on mutant outcompetes locked off in broth, and, at the same time, mrpA expression increases during oxygen limitation (152). Transcription of the recombinase mrpI may also be regulated by oxygen levels (152). Furthermore, orientation of the invertible element is influenced by the bacterium’s ability to assemble fimbriae. Assembly requires disulfide bond formation, and mutation of the gene encoding thiol:disulfide interchange protein DsbA leads to an increase in element-off bacteria (152).

In vitro function

In addition to hemagglutination, MR/P fimbriae mediate in vitro phenotypes that give clues to their function during infection. Biofilm formation (link biofilm section), measured in plastic multi-well plates, on glass coverslips, as pellicles in culture tubes, or in static cultures in glass tubing, is dependent on MR/P fimbriae (156, 244, 256). Interestingly, phase variation contributes to biofilm development. MR/P locked on bacteria establish biofilms more rapidly than wild-type, but over time the wild-type biofilm is much thicker (65 μm vs 12 μm) and has a classic structured channel appearance (244). In contrast, the locked on biofilm is short, very dense, and lacks channels. MR/P locked on bacteria autoaggregate (152, 244), which is reminiscent of the bacterial clustering that happens prior to stone formation in the bladder (44) (Fig. 25A–D). Furthermore, MR/P fimbriae allow P. mirabilis to adhere to some cultured cell lines, including T24 and Vero (150, 245). There are conflicting reports about MR/P-mediated binding to shed bladder epithelial cells, where purified MR/P and MR/P-positive P. mirabilis bound cells (240), but there was no difference in binding between a different wild-type strain and its isogenic mrpA mutant (149). It is possible that these two studies were in fact not examining the same fimbria (19).

An external file that holds a picture, illustration, etc.
Object name is nihms925686f25.jpg
P. mirabilis mrpA mutant is defective in cluster formation

(A–D) Representative sections of (A and B) P. mirabilis wild-type and (C and D) mrpA-infected murine bladders at 24 hpi. (A and C) The wild-type–infected bladder shows a regional view of clusters, whereas the mutant-infected bladder shows a close-up of the urothelial surface. Bacteria are in green, UPIIIa in red, and DAPI in blue. (Scale bars, in micrometers, are as marked.) (B and D) Alizarin Red staining of bladder sections. Only wild-type P. mirabilis-infected bladders contain significant mineral deposition. L, bladder lumen; an asterisk indicates an extracellular cluster. Arrows indicate regions with increased DAPI signal. (E and F) SEM micrographs of MR/P ON and MR/P OFF cells colonizing the murine bladder at 4 dpi. MR/P ON colonized the bladder uroepithelium (E), while MR/P OFF colonized the lamina propria where bladder cells had sloughed off (F). Bars = 10 μm (E), 2 μm (F). Figure adapted, with permission, from (44) (A–D) and (244) (E and F).

Contributions to infection

MR/P fimbriae contribute to UTI as described in multiple studies (Table 6). The magnitude of deficiency of an mrp mutant in mouse model independent challenges varies from study to study, from the initial mrpA experiment, where a 6–18-fold decrease was observed in urine, bladder, and kidneys, to a four log defect in the bladder for an mrpG mutant (146, 149, 151). A mutation in mrpH, which encodes the fimbrial adhesin, does not decrease fitness in independent challenge, although the mutant was unrecoverable following a one-week cochallenge experiment (156). Subsequent cochallenge infections with an mrpA or an mrpA-D mutant likewise resulted in the mutant being significantly out-competed in both the bladder and kidneys (150, 151).

Newer evidence may explain the observed variability in the requirement for MR/P expression. Two studies have reported equal numbers of recovered bacteria from the bladders of mice infected with wild-type compared with an mrpA mutant (44) or mrp locked-off (244). However, in both studies, the mrp mutant was not found in the same location as wild-type, suggesting that MR/P fimbriae not only facilitate adherence to urinary tract tissue but also dictate the localization of P. mirabilis within the bladder. Notably, both of these studies used a relatively early endpoint (24 hpi) (44, 244) compared with other mouse independent challenges lasting 7 days (146, 149, 151, 156). Thus, the location of P. mirabilis in the urinary tract, whether extracellular, intracellular, or bound to regions where uroepithelial cells have sloughed off, may be important for its long-term survival (Fig. 25E and F).

Although prior studies suggested that MR/P fimbriae bind a receptor in renal cells and contribute to pyelonephritis (149, 240), recent data point to their importance in cystitis. First, when bacteria are isolated from the bladder or urine of infected mice, the invertible element of the mrp promoter is overwhelmingly (95–100%) in the on orientation, allowing transcription of mrp genes. In contrast, the invertible element orientation is highly variable in the kidneys, ranging from mostly on to 85% off (158, 241). Second, the percentage of bacteria with the mrp invertible element in the on orientation directly correlates with the amount of bladder colonization (Fig. 24C) (152). Third, when a locked-on mrpI mutant is competed against wild-type bacteria in a 7 d mouse cochallenge experiment, locked-on bacteria outcompete wild-type in the bladder but not the kidneys (158). Fourth, P. mirabilis locked-on mrpI mutants adhere to the luminal surface of the bladder when examined 24 hpi, while locked-off bacteria adhere to regions where umbrella cells have been shed, revealing the lamina propria (244). Fifth, wild-type P. mirabilis forms large clusters in the bladder lumen within 24 hpi, but an mrpA mutant does not form clusters and is primarily intracellular (44) (Fig. 25A–D). The precise function of MR/P fimbriae and their contributions to cystitis and/or pyelonephritis will be difficult to address until their binding target(s) are identified.

Uroepithelial cell adhesin (UCA)

A fimbrial preparation from P. mirabilis cultured to stationary phase in a minimal medium at 37 °C was used to isolate a long, 4 nm-thin protein that adhered to human uroepithelial cells obtained from voided urine (246). The importance of this protein, UCA, in uroepithelial cell adhesion was subsequently confirmed when a mutation of the gene encoding the major structural subunit of UCA (ucaA) was tested with desquamated cells (163).

A second group working on UCA renamed the protein “non-agglutinating fimbriae” (NAF) and found the protein in shear preparations obtained from diverse P. mirabilis isolates (248). UCA/NAF are readily expressed under several laboratory conditions at 37°C, including Luria broth or agar; production may be enhanced by addition of serum to Luria broth or during culture on urine agar (248, 249).

Expressing cloned uca genes in E. coli conferred the ability to bind uroepithelial cells (247). However, no fimbriae were observed on the recombinant E. coli; only ucaA was sequenced and it is not clear whether all genes in the uca operon were present in the expression construct (247). A UCA/NAF preparation purified from P. mirabilis binds to cultured cells, including EJ/28 (urinary carcinoma), HEp-2 (laryngeal carcinoma), and MDCK (canine kidney epithelium) (249, 251), although these results have not been confirmed using other methods, such as by comparing adhesion of a uca mutant to that of wild-type.

Although UCA fimbriae are encoded by most P. mirabilis isolates, their sequence is more variable compared with other fimbriae (106). Antiserum raised against purified fimbriae further indicated that UCA is widely present but variable in size (248). It is possible that variations in UCA may explain distinct phenotypes reported by different groups, who are typically using locally-obtained isolates in their experiments. Notably, the uca operon is part of a mobile element and flanked by phage genes; in the first two fully-sequenced P. mirabilis genomes, HI4320 and BB2000, the uca operon is not in the same chromosomal locus (106).

In vitro receptor experiments

A possible receptor for UCA/NAF has been identified using fimbrial protein purified from a shear preparation using differential centrifugation (248, 250). Purified UCA/NAF bound in vitro to glycolipids, including asialo-GM1, asialo-GM2, and lactosyl ceramide (250). A follow-up study suggested another molecule, galectin-3, was also a receptor for purified UCA/NAF on MDCK cells (251). The possible receptors have not been identified in healthy bladder tissue, although galectin-3 is a marker for bladder cancer (257). Thus, it is not apparent whether these molecules are true in vivo targets for UCA.

Infection studies

Given that bladder cell adherence is a defining feature of UCA, it is perhaps paradoxical that a ucaA mutant is significantly attenuated in the kidneys but not the bladder during either independent or cochallenge of the murine urinary tract (163). When this ucaA mutant was injected into the tail vein of mice, bacteria were only recoverable from 7/16 mice 7 days post-inoculation, while the wild-type parent was recovered from 15/16 mice (163). Transcription of ucaA has been detected in urine from infected mice, although the level was 2.5-fold decreased relative to in vitro broth culture (131). UCA also contribute to CAUTI, where ucaA and ucaB were both detected in a Tn-Seq study (129).

Proteus mirabilis fimbriae (PMF)

PMF fimbriae were initially identified when a P. mirabilis cosmid gene bank was screened using antiserum generated against UCA (255). These fimbriae are encoded by the pmfACDEF operon (226, 258). Despite initial speculation that PMF was the unidentified MR/K agglutinin (255), a pmfA mutant had no effect on either MR/P or MR/K-type agglutination (153). Nevertheless, PMF fimbriae contribute to UTI in mouse models, with an 83-fold decrease in recoverable CFU for a pmfA mutant during independent challenge in the bladder (153). The defect was more profound in cochallenge, where the mutant was significantly attenuated in both bladder and kidneys (154). In a hematogenous model, where P. mirabilis was introduced through the tail vein of mice, the pmfA mutant was unrecoverable from the kidneys of 55% of mice 7 days post-inoculation, while the wild-type parent was still recovered from 94% (154). There are conflicting data on whether PMF fimbriae mediate adherence to voided human uroepithelial cells (153, 154); however, a pmfA mutant had diminished adherence to cultured T24/83 bladder carcinoma cells (154). In summary, evidence exists for PMF to contribute to infection in the bladder, kidneys, or both, and the receptor(s) have not yet been identified. PMF also contribute to fitness during CAUTI (129).

Interestingly, although PMF are expressed during UTI and mice infected with P. mirabilis generate an immune response against these fimbriae, pmf genes are repressed in urine relative to in vitro culture (131, 243). This suggests that PMF fimbriae may be transiently produced during infection, or that PMF-producing bacteria are not readily voided in urine. A pmfA/mrpA-D double mutant is more attenuated during a triple challenge in mice than either single mutation or wild-type, but interestingly, the double mutant is still infective in independent challenge (albeit less than the wild-type parent strain) (151). This is consistent with overlapping roles for the P. mirabilis “pan-fimbriome” and suggests that other fimbriae also contribute to adherence in the urinary tract (Fig. 26).

An external file that holds a picture, illustration, etc.
Object name is nihms925686f26.jpg
MR/P and PMF fimbriae have additive roles in urinary tract infection

Virulence of wild‐type P. mirabilis Pr2921 (wt), single fimbrial mutants (mrpA-D or pmfA), and double mrpA-D pmfA mutant was assessed in an ascending UTI model in mice at 7 dpi. (A and B) Independent challenge. The double mutant is significantly less fit compared with either single mutant. (A) CFU recovered per kidney, (B) CFU recovered per bladder. (C and D) Mice were challenged with a 1:1:1 mixture of three strains (tri-challenge). (C) Tri-challenge with wt, mrpA-D, and mrpA-D pmfA double mutant. (D) Tri-challenge with wt, pmfA, and mrpA-D pmfA double mutant. Each dot represents the log10 CFU recovered from each organ. The median (horizontal bar) is indicated for each group. The range of detection in this assay is 102 to 109 CFU per organ. Figure adapted, with permission from (151).

Ambient temperature fimbriae (ATF)

ATF fimbriae were first identified in a preparation of sheared surface proteins derived from an mrpA mutant (234). The atf operon was originally reported to consist of three genes, atfABC (259) as isolated from a cosmid clone, but subsequent genome sequencing revealed three additional genes (atfDEF) including an mrpJ paralog (112). Although ATF are produced at 37°C (106, 234), they are optimally expressed following static culture at 23°C (234). In a similar vein, an atfA mutant is recovered at the same rate as wild-type from either independent or cochallenge mouse studies (157). However, when an mrpA mutant was examined by immunofluorescence microscopy in infected mice, AtfA was readily detected (Fig. 27) (158, 244). This suggests that ATF may play a role in infection that is masked by other fimbriae, such as MR/P, and also there may be a fimbrial hierarchy that determines gene expression. ATF may contribute to biofilm formation under some conditions (256).

An external file that holds a picture, illustration, etc.
Object name is nihms925686f27.jpg
MR/P and ATF expression during ascending UTI

(A and B) Phase-locked mutants express MR/P fimbriae (MR/P ON) (A) or do not express MR/P fimbriae (MR/P OFF) (B) during experimental UTI at 2 dpi. Green, GFP-expressing P. mirabilis; blue, DAPI-stained bladder cell nuclei; red, rabbit anti-MrpA reacted with goat anti-rabbit IgG conjugated to Alexa Fluor 568; yellow, colocalization of bacteria and MR/P. (C and D) MR/P ON does not express ATF (C), but MR/P OFF does (D). Green, GFP-expressing P. mirabilis; blue, DAPI-stained bladder cell nuclei, red, rabbit anti-ATF serum reacted with goat anti-rabbit IgG conjugated to Alexa Fluor 568; yellow, colocalization of bacteria and ATF. Bars = 10 μm. Reproduced from (244), with permission.

P. mirabilis P-like fimbriae (PMP)

This fimbria was first discovered in a sheared protein preparation from a canine UTI isolate of P. mirabilis, where PMP was found along with UCA (253). PMP fimbriae have also been implicated in both single-species and polymicrobial CAUTI, where pmpG was identified by Tn-Seq (129), and a pmpA mutant was found to be attenuated in a diabetic mouse UTI model (260). The pmp operon was also reported to be directly regulated by cyclic AMP receptor protein (Crp) (260).

Fimbria 14

Fimbria 14, so named because it is encoded by the fourteenth of the 17 chaperone-usher fimbrial operons in HI4320, seems like it should be non-functional because the fim14 operon completely lacks a chaperone and has a frameshift mutation in usher gene fim14C (112). However, putative minor fimbrial subunit fim14B was detected in an STM screen for genes that contribute to UTI (179), and putative major structural subunit Fim14A was identified in a preparation of sheared surface proteins (106). The last gene in the operon, fim14D, is induced during UTI (131), and genes from the fim14 operon were found to contribute to polymicrobial CAUTI (129). Although chaperone-usher secretion apparatuses are usually specific for a single fimbrial type, it is likely that Fimbria 14 is assembled using one or more of the other 16 chaperone-usher systems encoded by P. mirabilis.

Other fimbriae

The remaining fimbriae encoded by P. mirabilis are poorly characterized (Table 6). P. mirabilis have been reported to employ an unknown fimbria to agglutinate tannic acid-treated erythrocytes, a pattern that is called mannose-resistant Klebsiella-like (MR/K) (100, 229, 230, 261). This ability is variable for P. mirabilis, with some studies suggesting MR/K agglutination by P. mirabilis is widespread (229, 230) but others suggesting MR/K is more often seen for Proteus vulgaris or Proteus penneri (100, 252, 261). MR/K fimbriae produced by Providencia stuartii and P. penneri mediate attachment to catheter surfaces, and therefore may contribute to catheter colonization by P. mirabilis strains that possess MR/K fimbriae (261, 262). They have also been shown to mediate adherence to Bowman’s capsules in kidney glomeruli (240). Some of the 17 fimbriae are more variable in sequence than others. For example, one operon, fim3, is present in less than half the isolates in a diverse collection (106). Likewise, UCA is widespread in P. mirabilis, but its size and sequence are variable, so UCA binding properties might not be constant from strain to strain (106, 248). Thus, MR/K agglutination could be mediated by one or more of these fimbriae.

Transcriptomic and high throughput sequencing methods are beginning to shed light on the complex collection of the 17 fimbriae. Transcripts from three additional fimbriae, fim5, fim8, and fim15, were detected by microarray analysis of urine from infected mice, although both fim5 and fim8 were decreased compared with in vitro culture (131). In addition, Fim8A, E, and F protein were detected by mass spectrometry from sheared surface proteins produced during in vitro culture (106). Southern blot analysis suggested P. mirabilis strains encoded two copies of mrp genes (150). Subsequent genome sequencing revealed the mrp′ operon, which is an apparent duplication of the mrp operon; the two operons are encoded next to each other in the chromosome (112). Tn-Seq analysis of genes that contribute to single-species CAUTI suggest a role for mrp′ but not mrp, which is unexpected given the well-documented contributions of MR/P fimbriae to uncomplicated UTI (129). Most of the remaining fimbriae were also detected in polymicrobial CAUTI Tn-Seq with P. stuartii but not single-species CAUTI (mrp, atf, fim3, fim5, fim12, fim14, fim15, fim16, fim17, plus orphan fimbrial genes PMI1812, PMI1920, and PMI3023), indicating once again that the contributions of these multiple fimbriae may be subtle and much work remains to discern their functions (129).

Non-fimbrial adhesins

P. mirabilis produces other, non-fimbrial adhesins that are described elsewhere in this chapter, including autotransporter proteins Pta, AipA, and TaaP (link toxins and autotransporter sections). P. mirabilis also encodes two putative type IV pili (Pearson 2008).

Transcriptional regulation by MrpJ

MrpJ is a transcriptional regulator encoded at the end of the mrp fimbrial operon (137). Like other mrp genes, it is not expressed well in vitro, but is among the most highly-induced genes in urine collected from infected mice (131). MrpJ was originally proposed to serve as a balance between adherent and motile states (reciprocal regulation); when MR/P fimbriae are produced (adherent state), MrpJ is also produced and switches off flagella (motility) (137). That is, expression of mrpJ on a plasmid represses FlaA flagellin, in turn inhibiting both swimming and swarming motility (137) (Fig. 28). Similarly, an mrpI locked-on mutant constitutively produces MR/P fimbriae but not flagella, while disruption of mrpJ in the locked-on background results in a bacterium that produces both fimbriae and flagella (137).

An external file that holds a picture, illustration, etc.
Object name is nihms925686f28.jpg
Elevated expression of MrpJ in P.mirabilis inhibits motility due to reduced flagella production

The three strains assayed here are P.mirabilis HI4320 transformed with pLX3607 (vector), pLX3805 (+MrpJ) and pLX5401 (ΔmrpJ). (A) The three strains were assayed for swarming on 1.5% Luria agar and for swimming in 0.35% Luria agar. (B) Three overnight Luria broth cultures of each of the three strains were adjusted to the same optical density, and equal volumes processed for SDS–PAGE and subsequent western blot analyses with antiserum against MrpJ or P.mirabilis flagella (FlaA). Reproduced from (137), with permission.

MrpJ is predicted to be a 110 amino acid protein consisting of a helix-turn-helix (HTH) DNA-binding domain and a unique C-terminal tail (137, 138). It is a member of the xenobiotic response element (XRE) superfamily of transcriptional regulators. Consistent with other HTH proteins, mutation of residues predicted to form DNA contacts interferes with MrpJ’s ability to repress swimming motility (138). However, deletion of the C-terminal domain has almost no effect on motility repression, suggesting that this region is disposable for reciprocal regulation (138). Consistent with its role as a repressor of motility, MrpJ directly binds the promoter of the master flagellar regulatory genes flhDC (132, 138).

In addition to influencing motility, MrpJ also appears to be involved in autoregulation of the mrp operon. Expression of mrpJ on a plasmid results in elevated mrpA expression (132), and an mrpJ mutant produces less MrpA protein (137). The mrpJ mutant also has the mrp invertible element overwhelmingly in the off orientation, even under mrp inducing conditions (152).

Deletion of mrpJ has little effect in vitro, which may be due to either low expression of mrp genes in most laboratory conditions or masking of MrpJ function by redundancy (see MrpJ paralogs below) (137). However, an mrpJ mutant is significantly outcompeted during mouse independent and cochallenge experiments (132, 137), indication a strong contribution to fitness and pathogenesis within the urinary tract. The exact role of MrpJ during infection is complicated considering that it regulates other virulence factors, such as production of MR/P fimbriae and flagella; MrpJ could be acting by altering either of these proteins, or eliminating coordination of motility and adherence, or by acting on other gene targets.

Transcriptomic analysis confirms that MrpJ has other virulence-associated targets, in addition to MR/P fimbriae and flagella (132). Because mrp genes are transcribed at a low level in vitro, an in vivo mimic has been used to study MrpJ targets, where mrpJ is expressed in trans at levels comparable to those detected in mice during experimental UTI (132). In a comparison of the in vivo mimic with a vector control, 217 genes were differentially regulated. Classes of MrpJ target genes include flagella, MR/P and other fimbriae, type VI secretion, metabolism, LPS modifications, virulence factors such as Zap and Pta, and transporters (132). Because many MrpJ-regulated genes have known or predicted roles in disease, MrpJ has been proposed as a master regulator of virulence.

Chromatin immunoprecipitation followed by PCR (ChIP-PCR) indicated that MrpJ binds within the invertible element of the mrp operon, in agreement with detection of an MrpJ-responsive element 156–256 bp upstream of the transcriptional start site of the mrp promoter (132). This is unusual, given that bacterial regulators typically bind in the vicinity of the −35 (activator) or +0 (repressor) promoter sequences so they can interact or interfere with RNA polymerase, respectively (263). However, a distant site of action for MrpJ, relative to other transcriptional regulators, is consistent with the flhDC promoter binding data mentioned above (132, 138).

MrpJ paralogs

Genomic sequencing of P. mirabilis revealed that this species encodes 14 additional mrpJ-type genes (112, 138). All but four are part of fimbrial operons, and they are always located at the beginning or the end of the operon (Table 6). Expression of all but two in trans causes repression of motility (138). Apparently, the paralogs do not all repress motility by the same mechanism, because overexpression of individual paralogs leads to flagellin levels ranging from wild-type to near-complete repression (138) (Fig. 29A). Different paralogs also induce distinctive swarming phenotypes and uniquely aberrant differentiation into swarmer cells (138) (Fig. 29B). This suggests that although most MrpJ paralogs repress motility, they likely have other, non-overlapping functions. However, at least one paralog, UcaJ, has been shown to bind the same fragment of the flhDC promoter as MrpJ (138).

An external file that holds a picture, illustration, etc.
Object name is nihms925686f29.jpg
MrpJ paralogs exert unique control over flagella and swarming

(A) Western blot of flagellin expression in P. mirabilis HI4320 mrpJ or mrpJ paralog overexpression strains. Whole cell lysates of uninduced strains were denatured, electrophoresed on 10% SDS-PAGE, and blotted with anti-FlaA antibody, which recognizes the major subunit of the flagellum. Lysates were also blotted with anti-UreD antibody as a loading control (lower panel). Molecular weight markers are indicated on the left side in kDa. (B and C), P. mirabilis cultures were spotted onto the center of swarming agar. (B) Representative swarming phenotypes of strain HI4320 expressing mrpJ paralogs. (C), Gram stains of bacteria taken from the edge of swarm fronts. The reference bar is 50 μm. Figure adapted, with permission, from (138).

Comparison of MrpJ and its paralogs revealed a conserved SQQQFSRYE motif within the helix-turn-helix (HTH) DNA-binding domain, plus a highly-variable C-terminal tail and short, highly-variable N-terminal sequence (138). Notably, MrpJ paralogs are not restricted to P. mirabilis as they have been detected in fimbrial operons from related genera (264266). Because fimbriae allow bacteria to adhere or form biofilms in specific environments, and thus respond to niche-specific signals, MrpJ paralogs have been proposed to orchestrate genetic programs that are beneficial to those environments (132, 267).

Like MrpJ itself, the MrpJ paralogs regulate other fimbrial operons in addition to repressing motility (267). Screening of MrpJ, UcaJ, AtfJ, and Fim8J showed that some paralogs positively auto-regulate their operons. Strikingly, AtfA and Fim8A strongly induced their respective operons while having modest or no effects on other fimbrial genes (mrpA, ucaA, or pmfA); in contrast, UcaJ has no effect on the uca operon (267). Examination of the atf promoter using deletion analysis indicated that AtfJ interacts with a region 486 to 655 bp upstream of the transcriptional start site (267). Thus, other MrpJ paralogs may share a common feature of binding DNA far away from transcriptional machinery binding sites. Deletion of the AtfJ C-terminal tail had little effect on motility repression, but was required for atf autoregulation (267). Protein modeling of the AtfJ C-terminus suggests this region is involved in protein-protein interactions, and it has been proposed that MrpJ paralogs exert their unique properties via their divergent C-termini (138, 267). These potential, unidentified, binding partners may explain how MrpJ paralogs exert transcriptional control from distant DNA binding sites.

F. Biofilms

Biofilms, which are adherent microbial communities, are a notorious problem on catheter surfaces, including urinary catheters. The ability of P. mirabilis to form biofilms on catheter surfaces is well-established (5, 268). Less well-understood is the potential establishment of biofilms within the urinary tract, and to what extent these biofilms contribute to disease. Because catheterization is a major risk factor for P. mirabilis UTI, biofilms within both catheters and urinary tissue will be considered here.

P. mirabilis readily adheres to a wide variety of surfaces, including clinically-relevant materials such as silicone, latex, glass, and polystyrene. In the presence of urine, struvite and apatite minerals are deposited among the developing P. mirabilis-colonized surface, leading to a crystalline biofilm. In laboratory models, crystalline biofilms form within 6 h after inoculation with P. mirabilis (269); similarly, mineralization has been detected in experimentally-infected mice within 24 hrs (44), and stones are detectable by 4 days after inoculation (42). This process occurs as a result of urease activity (link urease section), which causes locally-increased pH and subsequent mineral precipitation. This trait is a major reason why P. mirabilis is particularly problematic in patients with indwelling urinary catheters: crystalline biofilms can completely obstruct catheters (5, 27, 270) (Fig. 30).

An external file that holds a picture, illustration, etc.
Object name is nihms925686f30.jpg
Scanning electron micrographs of crystalline biofilms encrusting catheters

(A) shows a cross-section of an all-silicone catheter removed from a patient after 8 weeks; (B) shows a longitudinal section of a blocked silver/hydrogel-coated latex catheter removed from a patient after 11 days. In both these cases extensive crystalline material can be seen occluding the catheter lumen. Figure adapted, with permission, from (367) (A) and (288) (B).

Random mutagenesis to find biofilm mutants indicates that, as in other bacterial species, the biofilm lifestyle for P. mirabilis is complex, heterogeneous, and involves the coordination of many genes (177, 271). For example, a non-saturating transposon screen identified altered biofilm formation in 575/3850 mutants (15%) (177). When a subset of these mutants were further examined, most were also deficient in swimming and swarming motility. Notably, swimming and swarming may be involved in both establishment of biofilms, as a means of rapidly colonizing a surface, and dispersal into new environments. P. mirabilis swarms on catheters, and differentiated swarmer cells have been reported on P. mirabilis biofilms (272, 273); thus, swarming may help distribute bacteria from the catheter to the urinary tract. However, motility may also interfere with establishing adherent communities (274), and it is likely that balance is required for fully-developed biofilm formation and swarming behavior.

P. mirabilis encodes an arsenal of adhesive proteins, several of which have been shown to contribute to biofilm formation (Table 7). In particular, MR/P fimbriae have been linked to biofilm formation using multiple methods, including crystal violet-stained multi-well plates (256, 275), confocal microscopy on glass coverslips (244), glass tubing (256), bacterial clustering within the bladder lumen (44), and aggregative adherence to cultured cells (275). MR/P fimbriae allow complex biofilm architecture to develop, but only if the bacteria are allowed to modulate their fimbrial production (link fimbrial section). If MR/P expression is genetically shut off, biofilm formation is greatly diminished. However, bacteria engineered to constitutively produce MR/P fimbriae produce relatively thin (12 μm compared with 65 μm for wild-type), albeit very dense, biofilms (244).

Table 7

Factors involved in biofilm formation

Biofilm FactorProposed RoleReference
UreaseNickel metalloenzyme, local increase in pH causes mineral deposition which facilitates crystalline biofilms(5, 30, 306)
MR/P fimbriaeAdherence and auto-aggregation; mutants are defective in biofilm formation(44, 177, 244, 256, 275)
Fimbria 10 (PMI2210)Adherence; mutant has increased biofilm (perhaps due to fimbrial cross-talk)(177)
UCA, PMF, and ATF fimbriaeMutations have varying effects on biofilms (CFU, matrix) depending on culture conditions(256)
RcsBCD (RsbA) phosphorelayPhosphorelay system that enhances extracellular polysaccharide production; may control biofilm formation via regulation of fimbriae(140, 177, 279)
Pst transporterHigh-affinity phosphate transporter, mutants are defective in biofilm formation(297)
Capsule or extracellular polysaccharidesFacilitates mineral aggregation into crystalline biofilms and bacterial colonization(282284)
LPSP. mirabilis with different LPS charges vary in biofilm formation; pmrI LPS modification mutant and waaE inner-core LPS biosynthetic protein mutant are biofilm-defective(53, 280, 281)
RsmARNA binding protein; expression of P. mirabilis rsmA in Escherichia coli suppresses biofilm(307)
PpkPolyphosphate kinase; may act by regulating MR/P fimbriae(308)
HfqRNA chaperone; may act by regulating MR/P fimbriae or motility(164)
GlnEGlutamate-ammonia ligase adenylyltransferase; mutant has increased biofilm(177)
LrpLeucine-responsive regulator(177)
NirBNitrate reductase; biofilm-deficient in crystal violet assay and takes longer to block catheters(177)
BcrMultidrug efflux pump; mutant has increased biofilm but takes longer to block catheters(177)
GltSSodium/glutamate symport carrier protein(177)
PMI1551Putative lipoprotein(177)
PMI1608Unknown; biofilm-deficient in crystal violet assay but blocks catheters more rapidly(177)
PMI2861Putative membrane protein(177)
PMI3402MuA-like DNA binding protein; mutant has increased biofilm in crystal violet assay and blocks catheters more rapidly(177)

Crystalline biofilm formation is thought to increase risk of bladder stones, as one study found that 62% of patients with encrusted catheters also had bladder stones (276). Notably, P. mirabilis is recovered from most (65–79%) patients with obstructed catheters (5, 276). Infection stones and crystalline biofilms are both dense, complex bacterial communities, and at least two P. mirabilis proteins contribute to both processes: urease and MR/P fimbriae. During experimental UTI in mice, P. mirabilis assembles into large clusters in the bladder lumen, where mineralization starts to occur within 24 hours. Mutation of either urease or MR/P genes results in loss of both clustering and mineralization (44). P. mirabilis bacteria become embedded within stones, and likewise, may be embedded in extracellular matrix produced during biofilm growth (42, 277). In both cases, the bacteria may be protected from immune responses (44, 278).

Other processes also contribute to P. mirabilis biofilm formation (Table 7). Logically, regulators of the above biofilm factors, such as the RcsBCD phosphorelay (RcsD is also known as RsbA) (140, 177, 279) and RNA chaperone Hfq (164), control biofilm formation. Surface hydrophobicity, mediated by lipopolysaccharide (LPS) (53, 280, 281) and capsular polysaccharides (282284), influence both initial surface colonization and biofilm maturation.

Prevention of catheter blockage

Currently, the typical approaches to avoid urinary catheter biofilms are to 1) limit the duration of catheterization and 2) replace blocked/encrusted catheters. Multiple groups are developing new approaches to prevent biofilm establishment, including alternative catheter materials, antimicrobial coatings or instillation of anti-biofilm chemicals, mechanical or electrical biofilm dispersal, bacteriophage, and control of urinary pH (278, 285287). Because P. mirabilis is the major agent of crystalline biofilm formation and catheter blockage during CAUTI, many of these efforts focus on this species. Thus far, no catheter type has been immune to P. mirabilis biofilm (270), although some materials and coatings delay blockage. In some cases, P. mirabilis colonizes antimicrobial surfaces after crystalline material has coated the catheter (288). Filling the retention balloon of urinary catheters with 10 g/L triclosan is able to inhibit P. mirabilis biofilm in an in vitro model (289, 290), although exposure of P. mirabilis to triclosan may rapidly select for triclosan resistance (291). Ultimately, a combination of these strategies may be necessary to successfully combat P. mirabilis crystalline biofilm formation in a clinical setting.

Methods for studying P. mirabilis biofilms

An important caveat for biofilm studies in general is that biofilms can be generated using many techniques, and the chosen method has a strong influence on the outcome of the study (292). For example, P. mirabilis mutants with altered biofilm in a high-throughput crystal violet screen had no correlation with time to blockage in a catheter/artificial bladder model (177). Thus, there is a need to use a combination of simple and/or high-throughput in vitro methods, and more complex in vitro models and in vivo methods to study the contributions of P. mirabilis biofilms to infection.

A common, simple method for surveying biofilm formation involves culturing bacteria in multi-well plates, and visualizing biofilm by staining with crystal violet. The major advantage is that this assay allows for simultaneous screening of many samples; it is also easily quantifiable by dissolving the crystal violet and measuring the dye with a spectrophotometer. For this reason, the assay has been used to screen P. mirabilis clinical isolates and transposon mutants, as well as screening of targeted mutants (53, 177, 256, 275, 280, 281, 293, 294). Other variations on this assay include the addition of short catheter segments to multiwall plates and culture in static, closed tubes (283, 295). It is, however, important to note that these methods have their limitations. While urinary catheter biofilms are subject to the flow of urine, bacteria in multi-well plates are in a closed system where nutrients are depleted. Furthermore, if urine (artificial or real) is used as the medium, P. mirabilis urease activity will cause a rapid increase in alkalinity to approximately pH9. This lethal effect in static systems can be seen in a comparison of P. mirabilis biofilms cultured in the presence or absence of urea (296). Therefore, while the crystal violet method is useful, results may be less likely to lead to translational applications.

Continuous flow systems, where fresh medium is supplied at a constant rate, allow study of P. mirabilis biofilm development in catheters (30, 273, 297). The flow can be adjusted to match physiological urine rates in humans, and a flow cell can be inserted to facilitate microscopy. Urease activity still leads to crystalline biofilm formation if urine is used as the medium, but the pH increase is not as extreme as in the multi-well plate biofilm assay and may be maintained within a physiological range. A modification of continuous flow is an artificial bladder (298). Flow is maintained into a glass chamber that serves as a bladder analog, which is maintained at 37°C via a water jacket. A catheter is inserted into the base of the bladder chamber, which may be further connected to tubing and a drainage bag. The advantage of this bladder model is that it captures several aspects of the urinary tract: continuous introduction of urine from above, insertion of a catheter from below (including inflation of the retention balloon), and retention of a residual volume of urine similar to what occurs in a catheterized patient. However, this model requires specialized equipment and is limited in how many samples can be analyzed at one time.

P. mirabilis adheres to cultured cells in an aggregative pattern that has been likened to biofilms (275). This method may be used to assess aspects of biofilm development that are host-influenced, or require a biotic substrate for the developing biofilm. A drawback is that P. mirabilis produces toxins, such as hemolysin, which lead to cell death within a few hours and pose a challenge for studying longer-term biofilm development.

Animal models are also used to study biofilm development. Stones and clusters form in the mouse ascending UTI model (40, 42, 44), and P. mirabilis biofilms may be studied in vivo using the CAUTI mouse model (link animal models section) (25, 129). These models are essential for investigating the interaction of P. mirabilis biofilms with urinary tissues, the innate immune response, and the catheter-bladder interface.

Culture medium also greatly affects biofilm development. Using continuous flow coupled with confocal scanning laser microscopy, P. mirabilis biofilms were found to produce towers of biomass separated by channels in a rich medium (LB). However, in artificial urine (299), biofilms were flat, lacked channels, and had swarmer cells protruding from the surface (273). Crystalline biofilms will only develop if urine or artificial urine is used. Catheters in patients will be coated with cellular material and other material not present in artificial urine, and these coatings may further facilitate biofilm formation (268, 270).

Polymicrobial catheter-associated biofilms

P. mirabilis is frequently part of a mixed-species infection (4, 5, 9, 300), and catheter biofilms are a useful model for studying polymicrobial UTI. Catheters taken directly from patients may be studied to explore polymicrobial biofilms (300302), or continuous-flow systems (300, 303, 304) or static cultures (305) may be inoculated in the lab. Urease activity can give P. mirabilis a competitive advantage in catheter biofilms over Pseudomonas aeruginosa (303); similarly, in the presence of artificial urine, P. mirabilis inhibits Klebsiella pneumoniae biofilm formation (305). The CAUTI mouse model, in which a catheter segment is pushed into and retained in the bladder during the course of the experiment, is also shedding light on P. mirabilis as part of a mixed-species infection (25, 129) (link to urease and genome-wide mutagenesis sections). Other aspects of mixed-species bacteriuria and infection have been recently reviewed elsewhere (3, 20).

G. Flagella and motility

P. mirabilis was named for the Greek god Proteus due to its distinctive dimorphic nature, which is directly related to its motility: in the vegetative state, these bacteria exist as short motile rods (1–2 microns in length) with 6–10 peritrichous flagella, but under permissive conditions they differentiate into swarm cells by elongating 20- to 50-fold (20–80 microns in length) and expressing hundreds to thousands of flagella (309, 310) (Fig. 31). The vegetative swimmer cells exhibit normal chemotactic behavior in liquid medium, being attracted by nutrients and repelled by toxic substances or unfavorable conditions (310), while swarm cells allow for migration across solid surfaces, forming a characteristic bull’s eye pattern through sequential rounds of the differentiation process (95). Swarm cell differentiation and the mechanics and regulation of swimming and swarming motility in P. mirabilis have been extensively reviewed elsewhere (3, 18, 19, 311). We will therefore sharply focus on the direct contribution of flagella and motility to pathogenesis.

An external file that holds a picture, illustration, etc.
Object name is nihms925686f31.jpg
Swarming behavior of P. mirabilis

(A) Swarming colony of P. mirabilis HI4320. (B) The swarming migration distance of wild-type strain P19 (open circles) and a super-swarming rsbA mutant (solid circles). The periodic shift from swarming to consolidation can be seen. (C) Cartoon and TEM showing differentiated swimmer (broth-cultured) and swarmer cells. (D and F) The edge of an advancing swarm colony during consolidation (D) or swarming (F). (E and G) Gram stains showing consolidate (E) and swarmer cells (G) obtained from the edge of a growing P. mirabilis swarm. Figure adapted, with permission, from (19) (A and C), (143) (B), and (126) (C–G).

Flagella

In contrast to other motile bacteria, all of the flagellar components and chemotaxis proteins of P. mirabilis are encoded within a single locus in the chromosome that spans approximately 54 kb (112). Within this region, P. mirabilis encodes two flagellin genes, flaA and flaB (312). FlaA appears to be the predominant flagellin produced by P. mirabilis, but recombination can occur between flaA and flaB, resulting in occasional hybrids (124, 313, 314). As the flagella produced by P. mirabilis are recognized by the host immune response and can elicit a pro-inflammatory response, antigenic variation and flaAB hybrids may play a role in immune evasion during infection (77, 161).

Flagella are not an absolute requirement for establishment of UTI, as naturally-occurring nonmotile strains and flagellar mutants of P. mirabilis are still capable of colonizing the urinary tract in experimental models (155, 315). However, there is clear evidence for the contribution of flagella to pathogenesis. The importance of flagella to host cell invasion and ascending UTI was directly assessed by disrupting the gene encoding the flagellar filament (flaD) in a P. mirabilis mutant lacking hemolysin activity to avoid confounding by expression of the toxin. Production of the flagellar filament was found to facilitate invasion of human renal proximal tubular epithelial cells in vitro, partly by allowing the bacterial cells to come into close proximity to the host cells (45). Flagella may also promote internalization through other mechanisms, as the flaD mutant still exhibited reduced internalization compared to the parental strain when the bacteria were centrifuged onto the cell monolayer. This mutant also exhibited reduced bladder and kidney colonization in the murine model of ascending UTI, while loss of hemolysin alone did not significantly impact pathogenesis.

A role for chemotaxis and components of the flagellar apparatus in ascending UTI was also directly indicated by signature-tagged mutagenesis (STM) studies as a positive regulator of chemotaxis (cheW), the flagellar M-ring (fliF) and flagellar hook protein (flgE) were all identified as fitness factors for ascending UTI (108, 179). Transposon insertion in any of these genes resulted in a defect in kidney colonization, while fliF and cheW also affected bladder colonization. Transposon insertion-site sequencing (Tn-Seq) similarly indicated a role for chemotaxis and flagella during catheter-associated CAUTI (129). Specifically, a regulator of chemotaxis (cheR), the flagellar master regulator (flhDC), and a gene involved in regulation of flhDC (umoA) were identified as fitness factors for kidney colonization, while another gene involved in regulation of flhDC (umoB) and the flagellar basal-body rod protein (flgC) were fitness factors for colonization of the kidneys and the catheterized bladder, and the flagellar M-ring (fliF) and ATP synthase (fliI) were fitness factors for bladder colonization (129). Thus, the flagellar M-ring has been identified as having a significant contribution to fitness through two separate mutagenesis screens in two distinct models of infection. It is also noteworthy that flagellar genes are temporally regulated during ascending UTI, being poorly expressed in vivo relative to growth in rich medium at early time points post-inoculation but increasing in expression by 7 dpi, further indicating a role for flagella in pathogenicity (131).

Motility

Despite numerous studies, the exact contribution of flagella-mediated motility to P. mirabilis pathogenicity remains inconclusive. P. mirabilis has been observed to differentiate into swarm cells on catheter surfaces, which is thought to allow for migration and entry into the catheterized urinary tract (7, 272, 298, 316) (Fig. 32). As non-swarming mutants are only capable of migrating across hydrogel-coated catheters, swarming is likely an important aspect of initial bladder colonization in catheterized individuals. Five cues that can induce swarming in P. mirabilis are present in human urine (arginine, glutamine, histidine, malate, and ornithine), and may contribute to swarm cell differentiation and migration across urine-bathed catheters as they promote swarming on urine agar plates (165, 317, 318) (Fig. 33).

An external file that holds a picture, illustration, etc.
Object name is nihms925686f32.jpg
P. mirabilis swarms across 1 cm sections of hydrogel-coated latex catheter

Reproduced from (272), with permission.

An external file that holds a picture, illustration, etc.
Object name is nihms925686f33.jpg
Swarming on urine agar in response to five cues

(A) Urine agar is not normally permissive for swarming as the high pH resulting from urease activity inhibits swarming. A ureC mutant is capable of modest swarming on urine agar, and supplementation with any of five swarming cues to a final concentration of 20 mM dramatically enhances swarming. (B) Quantitation of swarm colony diameter on urine agar for P. mirabilis HI4320 and a ureC mutant. White lines indicate swarm diameter. Error bars represent means and standard deviations for four independent experiments with four replicates each. Statistical significance was determined by comparing the swarm diameter under each condition to the diameter on plain medium for each strain. *, P < 0.05; **, P < 0.01; ***, P < 0.001. Figure adapted, with permission, from (317).

Once within the urinary tract, in vitro experiments indicate that swimming motility may facilitate contact with uroepithelial cells, thereby promoting internalization and cytotoxicity (45). Swarm cells in particular have been postulated to contribute to host cell invasion, as these differentiated cells may be capable of invading urothelial cells faster and to a greater extent than vegetative cells (52). Numerous virulence factors are coordinately expressed during swarming, including Zap protease and hemolysin, which may cause swarm cells to be more cytotoxic to the host urothelium (52, 133) (Fig. 34).

An external file that holds a picture, illustration, etc.
Object name is nihms925686f34.jpg
Expression of the virulence genes zapA and hpmB is coordinately regulated with swarming

The upstream promoter regions of zapA and hpmB, respectively, were fused to a promoterless luxCDABE cassette harbored on a low-copy-number plasmid and transformed into wild-type P. mirabilis. Each of the resulting strains was inoculated as a 5 μl spot in the center of an L agar plate and incubated at 37°C. Gene expression was measured as luminescence (Lux), which is displayed in false color corresponding to relative Lux intensity, where blue is lowest and red is most intense. An image of the colony growth photographed in natural light is shown below its corresponding Lux image. Expression of zapA (first two rows) and hpmB (third and fourth row) is shown at 4, 8, 12, and 16 h. Figure adapted, with permission, from (141).

Swimming motility is also thought to contribute to dissemination within the urinary tract, particularly ascension from the bladder to the kidneys and spread between kidneys. For instance, immunization to stimulate production of antibodies that immobilize P. mirabilis can prevent spread to the contralateral kidney following direct inoculation into the right renal medulla (319). Swarm cells have also been implicated for contributing to kidney colonization and development of pyelonephritis, particularly during long-term infection, as swarm cells have been visualized within the kidney parenchyma (52, 320). However, elongated swarm cells were rarely observed during four-day infection studies in the (uncatheterized) murine model of ascending UTI (321). Thus, while flagella clearly contribute to P. mirabilis pathogenesis, the importance of swimming motility and swarm cell differentiation to disease progression and severity remain to be fully elucidated.

H. Metal acquisition

Bacterial pathogens compete with the host for micronutrients. One method the host uses to combat pathogens is sequestration of these nutrients so that the bacteria become starved for these ions. Bacteria that have evolved to colonize humans, therefore, have strategies to scavenge these elements, including iron and zinc (322, 323).

Iron

The most highly studied of the nutrient competition systems is iron, and much like the rest of the human body, the urinary tract is iron-restricted (324). Similar to other pathogens, P. mirabilis encodes iron scavenging and uptake systems that are induced during infection (109, 112, 131).

Bacteria typically produce iron-scavenging molecules called siderophores, which bind iron with higher affinity than host iron sequestration systems (322). Early evidence suggested that P. mirabilis lacked siderophores because it fares poorly relative to other bacteria in iron chelation experiments and tested negative in siderophore detection assays (325, 326). In addition, P. mirabilis is inhibited in its ability to establish infection in an iron-deficient rat pyelonephritis model (327). However, we now know that P. mirabilis has at least three ways to scavenge iron: proteobactin (Pbt), a non-ribosomal peptide synthesis system (Nrp), and α-keto acids. It can also use enterobactin, a siderophore produced by other enteric bacteria, although this species is not able to synthesize enterobactin on its own (109).

The nrp system was initially identified as a possible iron-regulated swarming signaling system (328). Genome annotation, microarray analysis, and a more sensitive siderophore detection method revealed that nrp and a second system, proteobactin (pbt), are siderophore biosynthesis and transport operons (109, 112). While the nrp genes are homologous to yersiniabactin genes and are encoded within a high pathogenicity island derived from Yersinia spp. (107), the products are not interchangeable (109). Although both siderophore systems must be mutated to produce a negative iron scavenging result in the laboratory (chrome azurol S, or CAS assay), only the nrp system has a significant contribution to infection (109).

P. mirabilis produces α-keto acids from amino acids, likely via two amino acid deaminases (PMI2834 aad and PMI2149) (326, 329, 330). These α-keto acids act as non-canonical siderophores and can restore growth of P. mirabilis on iron-chelated medium when supplied externally (326, 329). Amino acid deaminase activity is restricted to genera closely related to P. mirabilis (that is, Proteus, Providencia, and Morganella), as most enteric bacteria do not have Aad activity and cannot use externally-provided α-keto or α-hydroxy acids to overcome iron chelation (326, 329).

P. mirabilis encodes ferric, ferrous, and heme uptake systems that have been discovered by analysis of genes and proteins differentially regulated during iron restriction, genome annotation, and transposon mutagenesis (109, 112, 331, 332). In vivo analyses have demonstrated that iron uptake proteins are immunogenic (161, 331) and produced in a rat intraperitoneal chamber model (333); likewise, iron-related genes are induced during UTI (131), and several of these have directly been shown to contribute to UTI (108, 161, 179, 332). Iron acquisition and storage systems are even more strongly required during polymicrobial CAUTI, suggesting that competition with other species increases pressure on P. mirabilis iron-scavenging capabilities (129). Details on iron-associated genes in P. mirabilis and their contributions to infection are listed in Table 8.

Table 8

Iron-related genes in P. mirabilis

Iron-related genes from P. mirabilis were identified by homology to other iron-related genes (334, 335). Genes identified as iron-related by homology but not identified using one of the four conditions shown were excluded. A checkmark indicates that one or more of the genes in the row were identified using the condition specified. Adapted from (19).

Gene Designation(s)Proposed functionPMI number (s)Upregulated in iron limitationUpregulated in vivoImplicated in vivoAntigenic in vivoCAUTICitations
SinglePoly
Heme uptake
TonB-dependent receptorPMI0409[check][check][check](336) (335) (129)
hemin uptake proteinPMI1424[check][check](335) (337)
hmuR1R2STUVhemin uptake systemPMI1425–1430[check][check][check][check][check](336) (335) (337) (338) (331) (129)
Ferrous Iron Uptake
sitDCBAIron ABC transporterPMI1024–1027[check][check][check](335) (337) (129)
feoABferrus iron transportPMI2920–2921[check][check](337) (129)
Ferric citrate transport
exported proteasePMI3704[check][check](335) (129)
iron-related ABC transporterPMI3705[check][check][check](335) (339) (129)
iron receptorPMI3706–3707[check][check](335) (129)
extracytoplasmic function family σ factorPMI3708[check][check](335) (129)
TonB-like proteinPMI3709[check][check](335) (129)
Siderophore production
TonB-dependent receptorPMI2596[check][check][check][check][check](336) (335) (333) (129)
nrpXYRSUTABGnrp siderophorePMI2597–2605[check][check][check][check](335) (337) (340) (129)
pbtIABCDEFGHproteobact inPMI0231–0239[check][check][check](339) (129)
Other TonB-dependent receptors
irgATonB-dependent enterobact in receptorPMI0842[check][check][check][check][check](336) (335) (337) (333) (129)
ireAferric siderophore receptorPMI1945[check][check][check][check][check](336) (335) (337) (333) (129)
TonB-dependent receptorPMI0363[check][check][check](335) (337) (129)
TonB-dependent receptorPMI1548–1551[check][check][check](335) (337) (129)
hasRTonB-dependent receptorPMI3120–3121[check][check](340) (129)
ABC-transport system
iron-related ABC transporterPMI0331[check][check][check](335) (337) (129)
iron-related ABC transporterPMI2957–2960[check][check][check][check](335) (337) (340) (129)
Iron-relatedPMI0229[check][check](335) (129)
ABC transporter−0230
Iron Metabolism
iron utilization proteinPMI1437[check][check][check](335) (337) (129)
Iron Sulfur Cluster Formation/Uptake
iron sulfur clusterPMI1411–1416[check][check][check][check](335) (337) (129)
iron sulfur clusterPMI3253[check][check][check](335) (337) (129)
iron sulfur clusterPMI0176–0172[check][check][check](335) (337) (129)

Zinc

The high-affinity zinc importer genes znuACB are induced during UTI compared with laboratory culture (131), and ZnuB is recognized by the humoral immune response (161). Furthermore, a znuC mutant is outcompeted in mouse cochallenge experiments (162), and znuACB contributes to fitness in a CAUTI model (129). However, these systems are not an absolute requirement for infection, as the znuC mutant remains able to colonize during independent infection (162), and the requirement for znuACB in single-species CAUTI disappears during polymicrobial CAUTI (129). Zinc export may also contribute to infection, as exporter ppaA contributes to single-species CAUTI (129).

At least two virulence factors are linked to zinc: flagella and ZapA. ZapA is a member of the serralysin metalloprotease family of zinc metalloproteases, which have a conserved zinc-binding motif (219, 341). While this clearly suggests a zinc requirement for ZapA, it has not been directly tested. Work in E. coli showed that zinc is bound by the flagellar master regulator FlhC (342), so it is not surprising that a znuC mutant exhibits defects in flagella-mediated motility (both swimming and swarming) (162). Swarming is also inhibited in the presence of the transitional metal chelator TPEN, further supporting a role for zinc in motility (162). Interestingly, a zinc exporter ppaA mutant also exhibits aberrant swarming (343).

Nickel

Nickel is an essential component of catalytically active urease, in addition to other bacterial enzymes. However, there are few studies where P. mirabilis nickel homeostasis has been examined. Likewise, nickel sequestration by the host as a pathogen defense strategy has been proposed, but not yet conclusively demonstrated (344). Yet, there are some indications that nickel balance contributes to P. mirabilis fitness in the urinary tract. Genes from two predicted nickel import systems are induced in experimentally-infected mice compared with laboratory culture: nikAB and yntABCD (131). Curiously, in a CAUTI model, these import systems were not found to be essential, but nickel export (PMI1518) was (129). Thus, nickel homeostasis appears to be important during infection, but the balance between enzymatic requirements and metal toxicity may be affected by the presence of a catheter and/or competing nutritional requirements of other bacteria.

Phosphate

Phosphorus, another essential element for life, is most often incorporated as inorganic phosphate. Phosphate sensing and uptake has been linked to virulence in many bacterial species (345), and P. mirabilis is no exception. Several studies have pointed to the importance of the pstSCAB phosphate uptake system to P. mirabilis virulence. Transposon mutations in pstS and pstA were both identified in a signature-tagged mutagenesis study (108), and complemented in the mouse model (346). All four genes in this operon were also required in a CAUTI model (129). One way the pst system contributes to virulence may be via regulation of biofilm formation (297). Interestingly, phosphate mutants outcompete wild-type bacteria during culture in human urine (346) and also during polymicrobial UTI (129), suggesting that the energy cost of operating this system may actually confer a fitness defect if the bacteria are growing under conditions in which phosphate sensing and uptake are not required.

I. Metabolism in urine

Unlike most E. coli strains, P. mirabilis isolates generally possess two important metabolic features that make them ideally suited to growth in human urine: the ability to utilize citrate as a sole carbon source, and the ability to hydrolyze urea to produce an abundant nitrogen source. Urea is the most abundant organic metabolite in human urine, at concentrations of ~400 mM (347349). Filtration of citrate occurs in the kidney glomeruli, with approximately 10–35% of filtered citrate excreted in urine, making it the most abundant organic anion in human urine (350, 351). Thus, the combined use of citrate as a carbon source and ammonia as a nitrogen source provides P. mirabilis with a significant metabolic advantage over other uropathogens, and has implications for the carbon/nitrogen ratio that is sensed by P. mirabilis during growth in urine and the metabolic pathways favored by this organism.

Ascending UTI

Much has been learned regarding the metabolic pathways favored by P. mirabilis in the urinary tract through a combination of transcriptomics and genome-wide mutagenesis studies (Fig. 35). P. mirabilis appears to utilize glucose uptake and glycolysis during ascending UTI, as genes in these pathways are upregulated in vivo (131). The importance of glycolysis to P. mirabilis fitness was further verified in the murine model of ascending UTI using mutants in glucose-6-phopsphate isomerase (pgi), 6-phosphofructokinase transferase (pfkA), triosephosphate isomerase (tpiA), and pyruvate kinase (pykA) (166). Disruption of any single gene resulted in a dramatic fitness defect in bladder colonization, and all but pykA caused a fitness defect in kidney colonization.

An external file that holds a picture, illustration, etc.
Object name is nihms925686f35.jpg
Central metabolism during P. mirabilis UTI and CAUTI

Microarray analysis and targeted mutagenesis studies in an ascending UTI model both point to roles for glycolysis, oxidative pentose phosphate pathway (PPP), Entner-Doudoroff pathway (E–D), and both oxidative and branched/reductive TCA cycles during infection, while gluconeogenesis appears to be dispensable. Induction and requirement of gdhA suggests a low carbon-to-nitrogen ratio. Gene names in red were induced in vivo compared to broth culture as detected by microarray analysis; gene names in black were not differentially regulated while names in purple were repressed. An “X” over a pathway indicates a targeted mutant was assessed in cochallenge, with a red X indicating the mutant was outcompeted and a purple X indicating no fitness defect. Interestingly, a different picture emerges from Tn-Seq in a CAUTI model, where a catheter remains in the bladder. Here, gluconeogenesis contributes to fitness instead of glycolysis, and ammonia uptake relies on glnA instead of gdhA. During coinfection with Providencia stuartii in the CAUTI model, gluconeogenesis remains important, and PPP, E–D, and the oxidative TCA cycle are once again contributors to fitness. Figure adapted from (131), (166), and (129).

Both the oxidative and non-oxidative branches of the pentose phosphate pathway for NADPH production have also been shown to contribute to P. mirabilis fitness within the urinary tract. Transaldolase (talB) catalyzes a reversible step in the non-oxidative branch of the pentose phosphate pathway. While this gene was not identified as differentially regulated by in vivo transcriptomics and has not been detected as a fitness factor by signature-tagged mutagenesis (STM), it has been shown experimentally to contribute to P. mirabilis fitness in the bladder and kidneys during ascending UTI (166). Similarly, 6-phosphogluconate dehydrogenase (gnd) catalyzes a reaction in the oxidative branch of pentose phosphate pathway, and this gene has been shown to contribute to P. mirabilis fitness within the bladder and kidneys during ascending UTI (166). Notably, the D-glyceraldehyde 3-phosphate generated by the combined branches of the pentose phosphate pathway can feed directly into glycolysis.

P. mirabilis also appears to utilize the Entner-Doudoroff pathway during ascending UTI, as phosphogluconate dehydrogenase (edd) was upregulated during ascending UTI, identified as a fitness factor by STM, and disruption of this gene results in a fitness defect in bladder and kidney colonization (131, 166, 179). The other member of the Entner-Doudoroff pathway, a bifunctional 4-hydroxy-2-oxoglutarate aldolase/2-dehydro-3-deoxy-phosphogluconate aldolase encoded by eda, was also upregulated during ascending UTI, underscoring the importance of this metabolic pathway to P. mirabilis fitness (131). Products of the Entner-Doudoroff pathway are D-glyceraldehyde 3-phosphate, which can feed into glycolysis, and pyruvate, which can be directly catabolized to generate acetyl-CoA.

Regarding the aerobic tricarboxylic acid (TCA) cycle, subunit C of succinate dehydrogenase (sdhC) was identified as a fitness factor by STM, and subunit B (sdhB) was shown to directly contribute to fitness during kidney colonization (166, 179). Fumarate dehydratase (fumC) also contributes to P. mirabilis fitness in the bladder and kidneys (166). These results indicate a role for the aerobic TCA cycle during infection, which is further supported by the finding that cytochrome bo3 oxidase (cyoABCD), a component of the aerobic respiratory chain, is upregulated during ascending UTI (131). However, fumarate reductase (frdA), which catalyzes the reduction of fumarate to succinate during anaerobic respiration, also contributes to P. mirabilis fitness in the bladder and kidneys, which may indicate use of a branched TCA cycle during infection (166). Other portions of the TCA cycle found to be upregulated during ascending UTI include citrate synthase (gltA), citrate hydro-lyase 2 (acnB), and isocitrate dehydrogenase (icd) (131).

Additionally, transcriptomics and STM data indicate that P. mirabilis utilizes pyruvate catabolism via pyruvate dehydrogenase and dihydrolipoamide acetyltransferase (encoded by poxB and aceEF) to generate acetyl-CoA, which would feed into the TCA cycle through the action of citrate synthase (gltA) (131, 179). An abundance of factors pertaining to amino acid transport and metabolism are also upregulated during ascending UTI, including D-serine dehydratase (dsdA). D-serine is present in a relatively high concentration in human urine (352, 353). Thus, P. mirabilis may utilize DsdA as an additional pathway for generating pyruvate and ammonia.

The transcriptome of P. mirabilis in the murine urinary tract also revealed a preferential role for glutamate dehydrogenase (gdhA) over glutamine synthetase (glnA) for nitrogen assimilation during ascending infection, which was recapitulated in vitro only when citrate was supplied as the sole carbon source (131). It was further determined that gdhA provides P. mirabilis with a competitive advantage for bladder and kidney colonization (131). Glutamate dehydrogenase is generally favored by bacteria for nitrogen assimilation when energy and carbon are limited but there is an excess of ammonia and phosphate (354). Thus, the combined pattern of differential gene regulation coupled with mutagenesis studies indicates that P. mirabilis primarily utilizes pathways that feed into glycolysis, pyruvate catabolism, and citrate metabolism to fuel the TCA cycle through to production of α-ketoglutarate, which intersects with the production of L-glutamate by glutamate dehydrogenase using the excess of ammonia produced by urease (131, 179).

Catheter-Associated UTI (CAUTI)

The recent use of a genome-saturating transposon library and transposon insertion-site sequencing (Tn-Seq) has provided insight into the metabolic pathways favored by P. mirabilis in the catheterized urinary tract compared to during ascending UTI (129). A list of fitness factors in common between CAUTI and other UTI studies is found in Table 9. However, this study indicated that the pro-inflammatory environment induced by the presence of a urinary catheter dramatically impacts P. mirabilis metabolism. For instance, while glycolysis, the pentose phosphate pathway, and the Entner-Doudoroff pathway contributed to P. mirabilis fitness in ascending UTI, the only gene pertaining to these pathways that was a fitness factor in the CAUTI model was pyruvate kinase (pykF), which catalyzes the transfer of a phosphate group from phosphoenolpyruvate to yield ATP and pyruvate. For the TCA cycle enzymes, only subunit A of succinate dehydrogenase (sdhA) and malate dehydrogenase (mdh) represented fitness factors for CAUTI. However, cytochrome bo3 quinol oxidase (cyoABCDE) remained an important fitness factor for P. mirabilis in the CAUTI model, indicating a continued need for aerobic respiration.

Table 9

Genes identified as fitness factors for CAUTI by Tn-Seq that were also identified in the ascending UTI model by STM or transcriptomics.

PathwayPMI
Number
GeneFunctionUTI
Transcriptome
UTI
STM
CAUTI
Tn-Seq
Amino acid transport and metabolismPMI0145hypothetical proteinupB
PMI0187dsdAD-serine ammonia-lyaseupB, K
PMI0347lysAdiaminopimelate decarboxylaseupB, K
PMI0682artPL-arginine ABC transporter ATP-binding proteindownB
PMI0841lysPlysine:proton symporter, AAT familyupK
PMI1607sdaAL-serine ammonia-lyaseUK
PMI1827cysKcysteine synthaseupK
PMI2093speBagmatinasedownK
PMI2094speAarginine decarboxylasedownB, K
PMI2259metND-methionine ABC transporter, ATP-binding proteinupU, KB, K
PMI2736potBspermidine/putrescine ABC transporter, permease proteinupK
PMI2822pabAaminodeoxychorismate synthase, glutamine amidotransferase subunitupB
PMI2882glnAL-glutamine synthetasedownB, K
PMI3052asnAaspartate-ammonia ligaseupB, K
Carbohydrate transport and metabolismPMI0453nagAN-acetylglucosamine 6-phosphate deacetylaseupK
PMI0454nagBglucosamine-6-phosphate deaminaseupB, K
PMI0873nagZbeta-N-acetylhexosaminidaseupB
PMI1046polysaccharide deacetylaseU, B, KB, K
PMI1776PTS system IIB component, L-Asc familyupB
PMI1828hpr/ptsHPTS system phosphocarrier proteinupB, K
PMI1829ptsIPhosphoenolpyruvate–protein phosphotransferaseupB, K
PMI1830crrPTS system D-glucose-specific IIA component, Glc familyupB, K
PMI1943MFS-family transporterdownB, K
PMI2134agaAN-acetylgalactosamine 6-phosphate deacetylaseupB
PMI2141agaRtranscriptional regulator, DeoR familyupB, K
PMI2414deoBphosphopentomutaseupB, K
PMI3095N-acetylglucosamine kinaseupK
PMI3716gpmBprobable phosphoglycerate mutaseupK
Coenzyme transport and metabolismPMI0105cyoEprotoheme IX farnesyltransferaseupB, K
PMI2022visC2-octaprenyl-3-methyl-6-methoxy-1,4-benzoquinol hydroxylaseupK
PMI2353ribB3,4-dihydroxy-2-butanone 4-phosphate synthasedownK
PMI3547freNAD(P)H-flavin reductaseupB, K
Energy production and conversionPMI0107cyoBcytochrome bo3 quinol oxidase subunit 1 apoproteinupB, K
PMI0108cyoAcytochrome bo3 quinol oxidase subunit 2upB, K
PMI0354nqrCNa+-translocating NADH-quinone reductase subunit CupB, K
PMI0355nqrDNa+-translocating NADH-quinone reductase subunit DupB, K
PMI0705pflBformate acetyltransferaseupK
PMI1200pntANAD(P) transhydrogenase subunit alpha (pyridine nucleotide transhydrogenase subunit alpha)upB
PMI1753nuoKNADH dehydrogenase subunit KupB
PMI1754nuoJNADH dehydrogenase subunit JupK
PMI1760nuoC/nuoDNADH dehydrogenase subunit C (EC 1.6.5.3)/NADH dehydrogenase subunit DupB, K
PMI1771ackAacetate kinaseupK
PMI1772ptaphosphotransacetylaseupK
PMI2045aceFdihydrolipoyllysine-residue acetyltransferase component of pyruvate dehydrogenase complexupB, K
PMI2046aceEpyruvate dehydrogenase E1 componentupB, KB, K
PMI2740qorquinone oxidoreductaseupK
PMI2930glpD/glyDhomodimeric glycerol 3-phosphate dehydrogenase (quinone)downB, K
PMI3108fdoIformate dehydrogenase (quinone-dependent) cytochrome b subunitupK
PMI3300ilvMacetolactate synthase, small subunitupK

The CAUTI Tn-Seq results predominantly support a role for energy production through the catabolism of peptides and amino acids, and again implicate a role for D-serine dehydratase (dsdA) (129). Pyruvate catabolism remained important for both infection models, as pyruvate dehydrogenase (aceEF) represented an important fitness factor during CAUTI. Several genes involved in chorismate biosynthesis were also identified as CAUTI fitness factors, including aroF, aroK, and aroL. Chorismate is a branch point for biosynthesis of tryptophan, phenylalanine, and tyrosine. None of the tryptophan biosynthesis genes after the step catalyzed by ubiC (trpEDCBA) were identified as fitness factors for CAUTI, but pheA was a candidate fitness factor for bladder colonization and tyrB for both bladder and kidney colonization, indicating the potential importance of phenylalanine and tyrosine during CAUTI. It is also worth noting that pyruvate can be generated from chorismate in two ways: 1) ubiC, a fitness factor for bladder and kidney colonization, produces pyruvate from chorismate during synthesis of the ubiquinone precursor 4-hydroxybenzoate, and 2) pyruvate is generated from chorismate and glutamine during synthesis of 4-aminobenzoate by pabB, a fitness factors for bladder colonization. Thus, interplay between chorismate biosynthesis and pyruvate catabolism may be important for P. mirabilis fitness during CAUTI.

Most striking, however, was the shift in nitrogen assimilation requirements between ascending UTI and CAUTI. In contrast to the ascending UTI studies, glutamate dehydrogenase (gdhA) was not identified as a fitness factor for CAUTI, while glutamine synthetase (glnA) was identified as a fitness factor and verified to significantly contribute to urine and kidney colonization in the CAUTI model (129). In E. coli, glutamate dehydrogenase is favored when energy is limited, and glutamine synthetase is favored when energy limitation is no longer an issue (354). Thus, the preference for glutamine synthetase in the CAUTI model may reflect an excess of carbon sources in the catheterized urinary tract.

Polymicrobial Infection

P. mirabilis is one of the most common organisms present during polymicrobial urine colonization and CAUTI (4, 8, 9, 17). Considering the unique features of P. mirabilis metabolism in urine, it is likely that the presence of this bacterium could have a dramatic impact on the metabolic pathways favored by other uropathogens during infection, and possibly vice versa. For instance, in the ascending model of UTI, P. mirabilis and E. coli appear to require different but complementary central metabolism pathways as P. mirabilis utilizes glycolysis, while E. coli favors gluconeogenesis (166). P. mirabilis also appears to utilize the pentose-phosphate pathway, Entner-Doudoroff, and the aerobic TCA cycle during ascending UTI as mutations affecting any of these pathways resulted in a fitness defect (166). Interestingly, coinfection of P. mirabilis with E. coli shifted the metabolic requirements for both species by alleviating the pentose-phosphate pathway requirement for P. mirabilis but making this pathway required for E. coli fitness during the coinfection.

Considering that a urease-negative isolate of E. coli was used for these experiments, the shift in metabolic requirements could be due in part to the impact of P. mirabilis urease on the carbon/nitrogen ratio available to E. coli. However, a genome-wide exploration of P. mirabilis fitness factors conducted during coinfection with Providencia stuartii, a urease-positive organism, similarly identified a dramatic impact on metabolic pathways in a murine model of CAUTI (129). Similar to coinfection with E. coli, coinfection with P. stuartii shifted the requirements for the pentose-phosphate pathway and also the Entner-Doudoroff pathway. Coinfection also resulted in a greater need for the import and synthesis of arginine and branched-chain amino acids. The requirement for branch-chain amino acid biosynthesis was determined to be the result of high-affinity import by P. stuartii, as mutation of a gene involved in branched-chain amino acid import in P. stuartii alleviated the requirement for synthesis by P. mirabilis. Regarding energy production and respiration, coinfection appeared to alleviate the requirement for cytochrome bo3, but imposed a requirement for the respiratory nitrate reductase system (narGHJI, narXL, narP,and narK) and an Rnf redox-driven ion pump (rnfABCDGE).

V. Vaccines

Infections associated with urolithiasis (stone formation) can be difficult to treat. Indeed, prevention of infection by vaccination may be preferable in high-risk groups such as those who are chronically catheterized in the urinary tract. To that end, there have been a number of relevant studies in murine models that suggest that vaccination is feasible, an example of which is given in Fig. 36.

An external file that holds a picture, illustration, etc.
Object name is nihms925686f36.jpg
P. mirabilis colonization in bladders and kidneys of naive mice and mice nasally immunized with MrpH using cholera toxin (CT) as an adjuvant

Immunized mice were given primary immunization on day 0 and two booster immunizations on days 14 and 24. On day 34, all mice were challenged with 5 × 107 CFU P. mirabilis. After 7 d, bacterial burden was assessed. MHT, maltose-binding protein fusion of MrpH truncate; MHT-CT, MHT covalently coupled to CT; HA2-B, MrpH23–157-CT chimera; HA2-B + CT, HA2-B mixed with CT. Each diamond represents the log10 CFU per gram of tissue from an individual mouse. Samples with undetectable colonization were given a value of 2 log10 CFU/g of tissue (the limit of detection). Horizontal bars represent the median log10 CFU per gram of tissue for each column. One-tailed P values were determined by the Mann-Whitney test, comparing the colonization levels in bladders and kidneys of the naive mice with those of the immunized mice. Reproduced from (238), with permission.

Multiple approaches have been tested that include formalin-killed whole bacteria (237) and numerous surface-exposed or secreted proteins (55, 236238, 355359). In an early study (237), CBA/J mice were vaccinated with formalin-killed P. mirabilis bacteria or purified MR/P fimbriae or MrpH, the adhesin of the fimbria. For systemic immunizations, complete Freund’s adjuvant was used as adjuvant; for intranasal administration, cholera toxin was used as adjuvant. Four routes of immunization were examined: subcutaneous, intranasal, transurethral, and oral. Antibodies to the antigens were measured in serum, urine, bladder, vaginal washes, and bile. For whole-killed bacteria, vaccination by the subcutaneous route resulted in protection for 60% of mice (defined as having fewer than 100 CFU per gram of tissue 7 days post-inoculation), and intranasal vaccination provided protection in 67% of mice, but vaccination by the oral or transurethral routes failed to provide protection. Protection correlated with high serum antibody titers at day 21. For the intranasal route, measurable antibody was also found in the urine and bladder, albeit at a much reduced titer. When purified MR/P fimbriae were used, 60–63% protection, also at 2.5–3-log reduction, was noted by the intranasal or transurethral route. Protection also correlated with high serum antibody titer to the purified MR/P fimbriae. The tip adhesin was also examined as an antigen. Maltose-binding protein (MBP) fusions with MrpH (residues 23–157) and a truncated MrpH (residues 23–157) representing the receptor-binding domain were also used for intranasal immunization and yielded significant protection, especially for the MrpH receptor-binding domain, which provided up to 5-logs of protection in the bladder.

In a follow-up study, the MrpH receptor binding domain (residues 23–157) was translationally fused with the cholera toxin A2 domain, which was co-expressed with the cholera toxin B subunit (a holotoxin-like chimera) (238). Mice immunized with MrpH receptor-binding domain-CT complex were colonized by significantly fewer bacteria in the bladder (median log10 CFU per gram of tissue, 6.43 for the naive mice versus 2.00 for the immunized mice; P = 0.02) and the kidneys (median log10 CFU per gram of tissue, 5.28 for the naive mice versus 3.43 for the immunized mice; P= 0.046).

As well, two additional studies demonstrated that vaccination, either intranasally or transurethrally, with fusion proteins that contain both MrpH of P. mirabilis and FimH, the adhesin of type 1 fimbriae of E. coli, conferred protection against P. mirabilis UTI (359, 360).

In another fimbrial vaccination study (236), the main structural subunits (MrpA, UcaA, and PmfA) of three P. mirabilis fimbriae (MR/P, UCA, and PMF) were administered as vaccine antigens by the subcutaneous route. Mice were challenged with P. mirabilis either transurethrally or intravenously. MrpA protected against UTI and hematogenous challenge. UcaA protected only against hematogenous challenge. PmfA did not protect by either route of P. mirabilis challenge. In studies where flagellin was included with MrpA as antigens in intranasal vaccination, neither flagellin alone nor flagellin with MrpA led to protection from P. mirabilis challenge in the urinary tract (358). Administration of adjuvant cholera toxin in the urinary tract enhanced humoral and cytokine response but did not influence the degree of protection against UTI provided by MrpA (356).

MrpA expressed in heterologous hosts has also been tested as a vaccine antigen. MrpA was expressed as either a cell wall-anchored or secreted protein in Lactococcus lactis (355). Protection assays against challenge by P. mirabilis were tested in the mouse model, and significant induction of specific serum IgG and IgA was found in mice immunized with L. lactis expressing the MrpA antigen. A significant reduction of renal bacterial colonization was observed using both constructs. MrpA was also expressed as a MrpA-TetC (tetanus toxin fragment) in Salmonella Typhimurium (357) and used for intranasal vaccination. After two immunization doses, intranasally vaccinated mice showed a significant increase in specific serum IgG against MrpA and against the LPS of the Salmonella strain. Significant decreases in both kidney and bladder colonization by P. mirabilis after transurethral challenge were noted.

A non-fimbrial secreted protein, Proteus toxic agglutinin, proved to be a potent protective antigen for the upper urinary tract (55). Mice intranasally vaccinated with a site-directed mutant (S366A) of Proteus toxic agglutinin, which inactivated the protease active site, conjugated to cholera toxin, had significantly lower bacterial counts in their kidneys (P=.001) and spleens (P=.002) than mice that received cholera toxin alone. Bladders were not protected. Serum IgG levels correlated with protection.

Lastly, it is interesting to note that a vaccine used in Europe, Solco-Urovac, contains heat-killed enteric bacteria including P. mirabilis (361, 362). The vaccine has been shown to have demonstrable, though limited, efficacy against recurrent UTI (363). Further research into P. mirabilis vaccines is necessary, but heat-killed bacteria and surface-exposed or secreted proteins may have efficacy as vaccine antigens, and an intranasal route may be effective for vaccination. One issue is the identification of the target population for vaccination again P. mirabilis. For example, the population that might benefit most from vaccination against this species is elderly catheterized patients in chronic care facilities, but they may have diminished capacity to respond immunologically to vaccination. This poses an additional hurdle for development of an effective vaccine.

VI. Summary/Conclusions

When surveying the extensive literature on P. mirabilis infection and considering the relatively few labs that have undertaken studying this pathogen, there has been remarkable progress made in identifying virulence factors and mechanisms of pathogenesis for this uropathogen. Clearly, the catheterized urinary tract is a preferred niche for P. mirabilis and often the bacterium coexists with several other members of a polymicrobial community. However, monoinfection has been used for the most part to identify and characterize the virulence arsenal that includes urease, 17 different fimbriae, secreted cytolethal and cytolytic toxins and proteases, extensive networks for iron and other metal ion acquisition, flagella, and numerous secretion systems. We are just beginning to exploit relatively new tools and screens such as Tn-Seq and RNA-Seq to investigate mechanisms of pathogenesis in models of polymicrobial bacteriuria. This added layer of complexity is daunting, but ultimately manageable and will open the door to the development of treatment strategies.

In the early days, investigators worked with simple molecular tools to identify single genes or operons involved in pathogenesis. Frankly, victories were hard fought although targets of discovery were plentiful. With the advent of facile sequencing technologies, PCR, microarrays, fruitful screens including signature-tagged mutagenesis and transposon-directed insertion site sequencing and RNA-Seq, numerous new candidate genes that contribute to colonization, establishment of infection, and damage to the host have been highlighted for P. mirabilis.

While often tedious, such screens are straightforward and yield a staggering number of genes that represent virulence or fitness factors. What lies ahead is the hard work of probing the biology of these new-found genes. Thus, with the groundwork presented here, we can renew our efforts to refine a view of pathogenesis that includes the interactions and temporal expression of the battery of virulence factors and metabolic cycles discussed in this review.

Acknowledgments

This work was supported by Public Health Service grants from the National Institutes of Health (K99 DK105205 to C.E.A.; AI059722 and DK094777 to H.L.T.M.).

We are grateful to the many Proteus mirabilis researchers whose work contributed to this review.

References

1. Adeolu M, Alnajar S, Naushad S, Gupta SR. Genome-based phylogeny and taxonomy of the ‘Enterobacteriales’: proposal for Enterobacterales ord. nov. divided into the families Enterobacteriaceae, Erwiniaceae fam. nov., Pectobacteriaceae fam. nov., Yersiniaceae fam. nov., Hafniaceae fam. nov., Morganellaceae fam. nov., and Budviciaceae fam. nov. International Journal of Systematic and Evolutionary Microbiology. 2016;66:5575–5599. [Abstract] [Google Scholar]
2. Wenner JJ, Rettger LF. A Systematic Study of the Proteus Group of Bacteria. J Bacteriol. 1919;4:331–353. [Europe PMC free article] [Abstract] [Google Scholar]
3. Armbruster CE, Mobley HLT. Merging mythology and morphology: the multifaceted lifestyle of Proteus mirabilis. Nat Rev Micro. 2012;10:743–754. [Europe PMC free article] [Abstract] [Google Scholar]
4. Warren JW, Tenney JH, Hoopes JM, Muncie HL, Anthony WC. A Prospective Microbiologic Study of Bacteriuria in Patients with Chronic Indwelling Urethral Catheters. Journal of Infectious Diseases. 1982;146:719–723. [Abstract] [Google Scholar]
5. Mobley HLT, Warren JW. Urease-Positive Bacteriuria and Obstruction of Long-Term Urinary Catheters. Journal of Clinical Microbiology. 1987;25:2216–2217. [Europe PMC free article] [Abstract] [Google Scholar]
6. Breitenbucher RB. Bacterial changes in the urine samples of patients with long-term indwelling catheters. Arch Intern Med. 1984;144:1585–1588. [Abstract] [Google Scholar]
7. Jacobsen SM, Stickler DJ, Mobley HLT, Shirtliff ME. Complicated Catheter-Associated Urinary Tract Infections Due to Escherichia coli and Proteus mirabilis. Clin Microbiol Rev. 2008;21:26–59. [Europe PMC free article] [Abstract] [Google Scholar]
8. Nicolle LE. Catheter-Related Urinary Tract Infection. Drugs & Aging. 2005;22:627–639. [Abstract] [Google Scholar]
9. Armbruster CE, Prenovost K, Mobley HLT, Mody L. How Often Do Clinically Diagnosed Catheter-Associated Urinary Tract Infections in Nursing Home Residents Meet Standardized Criteria? J Am Geriatr Soc. 2016 10.1111/jgs.14533. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
10. Hung EW, Darouiche RO, Trautner BW. Proteus bacteriuria is associated with significant morbidity in spinal cord injury. Spinal Cord. 2007;45:616–620. [Europe PMC free article] [Abstract] [Google Scholar]
11. Griffith DP, Musher DM, Itin C. Urease. The primary cause of infection-induced urinary stones. Invest Urol. 1976;13:346–350. [Abstract] [Google Scholar]
12. Li X, Zhao H, Lockatell CV, Drachenberg CB, Johnson DE, Mobley HLT. Visualization of Proteus mirabilis within the Matrix of Urease-Induced Bladder Stones during Experimental Urinary Tract Infection. Infect Immun. 2002;70:389–394. [Europe PMC free article] [Abstract] [Google Scholar]
13. Foxman B, Brown P. Epidemiology of urinary tract infections: transmission and risk factors, incidence, and costs. Infect Dis Clin North Am. 2003;17:227–241. [Abstract] [Google Scholar]
14. Kim BN, Kim NJ, Kim MN, Kim YS, Woo JH, Ryu J. Bacteraemia due to tribe Proteeae: a review of 132 cases during a decade (1991–2000) Scand J Infect Dis. 2003;35:98–103. [Abstract] [Google Scholar]
15. Watanakunakorn C, Perni SC. Proteus mirabilis bacteremia: a review of 176 cases during 1980–1992. Scand J Infect Dis. 1994;26:361–367. [Abstract] [Google Scholar]
16. Daniels KR, Lee GC, Frei CR. Trends in catheter-associated urinary tract infections among a national cohort of hospitalized adults, 2001–2010. American Journal of Infection Control. 2014;42:17–22. [Abstract] [Google Scholar]
17. Hooton TM, Bradley SF, Cardenas DD, Colgan R, Geerlings SE, Rice JC, Saint S, Schaeffer AJ, Tambayh PA, Tenke P, Nicolle LE. Diagnosis, Prevention, and Treatment of Catheter-Associated Urinary Tract Infection in Adults: 2009 International Clinical Practice Guidelines from the Infectious Diseases Society of America. Clinical Infectious Diseases. 2010;50:625–663. [Abstract] [Google Scholar]
18. Morgenstein RM, Szostek B, Rather PN. Regulation of gene expression during swarmer cell differentiation in Proteus mirabilis. FEMS Microbiology Reviews. 2010;34:753–763. [Abstract] [Google Scholar]
19. Schaffer JN, Pearson MM. Proteus mirabilis and Urinary Tract Infections. Microbiol Spectr. 2015;3 [Europe PMC free article] [Abstract] [Google Scholar]
20. Norsworthy AN, Pearson MM. From Catheter to Kidney Stone: The Uropathogenic Lifestyle of Proteus mirabilis. Trends Microbiol. 2016 10.1016/j.tim.2016.11.015. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
21. Garcia MM, Gulati S, Liepmann D, Stackhouse GB, Greene K, Stoller ML. Traditional Foley Drainage Systems—Do They Drain the Bladder? The Journal of Urology. 2007;177:203–207. [Abstract] [Google Scholar]
22. Willette PA, Coffield SK. Current Trends in the Management of Difficult Urinary Catheterizations. Western Journal of Emergency Medicine. 2012;13 [Europe PMC free article] [Abstract] [Google Scholar]
23. Hofseth LJ, Dunn BP, Rosin MP. Micronucleus frequencies in urothelial cells of catheterized patients with chronic bladder inflammation. Mutation Research/Fundamental and Molecular Mechanisms of Mutagenesis. 1996;352:65–72. [Abstract] [Google Scholar]
24. Anderson RU. Response of bladder and urethral mucosa to catheterization. JAMA. 1979;242:451–453. [Abstract] [Google Scholar]
25. Armbruster CE, Smith SN, Johnson AO, DeOrnellas V, Eaton KA, Yep A, Mody L, Wu W, Mobley HLT. The Pathogenic Potential of Proteus mirabilis is Enhanced by Other Uropathogens During Polymicrobial Urinary Tract Infection. Infection and Immunity. 2017;85:e00808–00816. [Europe PMC free article] [Abstract] [Google Scholar]
26. Kunin CM. Blockage of urinary catheters: Role of microorganisms and constituents of the urine on formation of encrustations. Journal of Clinical Epidemiology. 1989;42:835–842. [Abstract] [Google Scholar]
27. Stickler D, Ganderton L, King J, Nettleton J, Winters C. Proteus mirabilis biofilms and the encrustation of urethral catheters. Urol Res. 1993;21:407–411. [Abstract] [Google Scholar]
28. Coker C, Poore CA, Li X, Mobley HLT. Pathogenesis of Proteus mirabilis urinary tract infection. Microbes and Infection. 2000;2:1497–1505. [Abstract] [Google Scholar]
29. Choong S, Wood S, Fry C, Whitfield H. Catheter associated urinary tract infection and encrustation. International Journal of Antimicrobial Agents. 2001;17:305–310. [Abstract] [Google Scholar]
30. Stickler DJ, Lear JC, Morris NS, Macleod SM, Downer A, Cadd DH, Feast WJ. Observations on the adherence of Proteus mirabilis onto polymer surfaces. J Appl Microbiol. 2006;100:1028–1033. [Abstract] [Google Scholar]
31. Sabbuba NA, Mahenthiralingam E, Stickler DJ. Molecular Epidemiology of Proteus mirabilis Infections of the Catheterized Urinary Tract. J Clin Microbiol. 2003;41:4961–4965. [Europe PMC free article] [Abstract] [Google Scholar]
32. Kunin CM. Urinary tract infections : detection, prevention, and management. 5th. Williams & Wilkins; Baltimore: 1997. [Google Scholar]
33. Griffith DP, Osborne CA. Infection (urease) stones. Miner Electrolyte Metab. 1987;13:278–285. [Abstract] [Google Scholar]
34. Fowler JE., Jr Bacteriology of branched renal calculi and accompanying urinary tract infection. J Urol. 1984;131:213–215. [Abstract] [Google Scholar]
35. Lerner SP, Gleeson MJ, Griffith DP. Infection stones. J Urol. 1989;141:753–758. [Abstract] [Google Scholar]
36. McLean RJ, Cheng KJ, Gould WD, Nickel JC, Costerton JW. Histochemical and biochemical urease localization in the periplasm and outer membrane of two Proteus mirabilis strains. Can J Microbiol. 1986;32:772–778. [Abstract] [Google Scholar]
37. Prywer J, Olszynski M. Bacterially Induced Formation of Infectious Urinary Stones: Recent Developments and Future Challenges. Curr Med Chem. 2017;24:292–311. [Abstract] [Google Scholar]
38. Torzewska A, Rozalski A. Various intensity of Proteus mirabilis-induced crystallization resulting from the changes in the mineral composition of urine. Acta Biochim Pol. 2015;62:127–132. [Abstract] [Google Scholar]
39. Johnson DE, Russell RG, Lockatell CV, Zulty JC, Warren JW, Mobley HL. Contribution of Proteus mirabilis urease to persistence, urolithiasis, and acute pyelonephritis in a mouse model of ascending urinary tract infection. Infect Immun. 1993;61:2748–2754. [Europe PMC free article] [Abstract] [Google Scholar]
40. Jones BD, Lockatell CV, Johnson DE, Warren JW, Mobley HL. Construction of a urease-negative mutant of Proteus mirabilis: analysis of virulence in a mouse model of ascending urinary tract infection. Infect Immun. 1990;58:1120–1123. [Europe PMC free article] [Abstract] [Google Scholar]
41. Armbruster CE, Smith SN, Yep A, Mobley HLT. Increased Incidence of Urolithiasis and Bacteremia During Proteus mirabilis and Providencia stuartii Coinfection Due to Synergistic Induction of Urease Activity. Journal of Infectious Diseases. 2014;209:1524–1532. [Europe PMC free article] [Abstract] [Google Scholar]
42. Li X, Zhao H, Lockatell CV, Drachenberg CB, Johnson DE, Mobley HL. Visualization of Proteus mirabilis within the matrix of urease-induced bladder stones during experimental urinary tract infection. Infect Immun. 2002;70:389–394. [Europe PMC free article] [Abstract] [Google Scholar]
43. Huang HS, Chen J, Teng LJ, Lai MK. Use of polymerase chain reaction to detect Proteus mirabilis and Ureaplasma urealyticum in urinary calculi. J Formos Med Assoc. 1999;98:844–850. [Abstract] [Google Scholar]
44. Schaffer JN, Norsworthy AN, Sun T-T, Pearson MM. Proteus mirabilis fimbriae- and urease-dependent clusters assemble in an extracellular niche to initiate bladder stone formation. Proceedings of the National Academy of Sciences. 2016;113:4494–4499. [Europe PMC free article] [Abstract] [Google Scholar]
45. Mobley H, Belas R, Lockatell V, Chippendale G, Trifillis A, Johnson D, Warren J. Construction of a flagellum-negative mutant of Proteus mirabilis: effect on internalization by human renal epithelial cells and virulence in a mouse model of ascending urinary tract infection. Infect Immun. 1996;64:5332–5340. [Europe PMC free article] [Abstract] [Google Scholar]
46. Alamuri P, Lower M, Hiss JA, Himpsl SD, Schneider G, Mobley HLT. Adhesion, Invasion, and Agglutination Mediated by Two Trimeric Autotransporters in the Human Uropathogen Proteus mirabilis. Infect Immun. 2010;78:4882–4894. [Europe PMC free article] [Abstract] [Google Scholar]
47. Mobley HL, Chippendale GR, Swihart KG, Welch RA. Cytotoxicity of the HpmA hemolysin and urease of Proteus mirabilis and Proteus vulgaris against cultured human renal proximal tubular epithelial cells. Infect Immun. 1991;59:2036–2042. [Europe PMC free article] [Abstract] [Google Scholar]
48. Swihart KG, Welch RA. Cytotoxic activity of the Proteus hemolysin HpmA. Infect Immun. 1990;58:1861–1869. [Europe PMC free article] [Abstract] [Google Scholar]
49. Torzewska A, Budzynska A, Bialczak-Kokot M, Rozalski A. In vitro studies of epithelium-associated crystallization caused by uropathogens during urinary calculi development. Microb Pathog. 2014;71–72:25–31. [Abstract] [Google Scholar]
50. Peerbooms PG, Verweij AM, MacLaren DM. Vero cell invasiveness of Proteus mirabilis. Infect Immun. 1984;43:1068–1071. [Europe PMC free article] [Abstract] [Google Scholar]
51. Kurihara S, Sakai Y, Suzuki H, Muth A, Phanstiel O, Rather PN. Putrescine Importer PlaP Contributes to Swarming Motility and Urothelial Cell Invasion in Proteus mirabilis. Journal of Biological Chemistry. 2013;288:15668–15676. [Europe PMC free article] [Abstract] [Google Scholar]
52. Allison C, Coleman N, Jones PL, Hughes C. Ability of Proteus mirabilis to invade human urothelial cells is coupled to motility and swarming differentiation. Infect Immun. 1992;60:4740–4746. [Europe PMC free article] [Abstract] [Google Scholar]
53. Jiang SS, Liu MC, Teng LJ, Wang WB, Hsueh PR, Liaw SJ. Proteus mirabilis pmrI, an RppA-regulated gene necessary for polymyxin B resistance, biofilm formation, and urothelial cell invasion. Antimicrob Agents Chemother. 2010;54:1564–1571. [Europe PMC free article] [Abstract] [Google Scholar]
54. Wang WB, Lai HC, Hsueh PR, Chiou RY, Lin SB, Liaw SJ. Inhibition of swarming and virulence factor expression in Proteus mirabilis by resveratrol. J Med Microbiol. 2006;55:1313–1321. [Abstract] [Google Scholar]
55. Alamuri P, Eaton KA, Himpsl SD, Smith SN, Mobley HLT. Vaccination with Proteus Toxic Agglutinin, a Hemolysin-Independent Cytotoxin In Vivo, Protects against Proteus mirabilis Urinary Tract Infection. Infect Immun. 2009;77:632–641. [Europe PMC free article] [Abstract] [Google Scholar]
56. Chippendale GR, Warren JW, Trifillis AL, Mobley HL. Internalization of Proteus mirabilis by human renal epithelial cells. Infect Immun. 1994;62:3115–3121. [Europe PMC free article] [Abstract] [Google Scholar]
57. Mathoera RB, Kok DJ, Verduin CM, Nijman RJM. Pathological and Therapeutic Significance of Cellular Invasion by Proteus mirabilis in an Enterocystoplasty Infection Stone Model. Infect Immun. 2002;70:7022–7032. [Europe PMC free article] [Abstract] [Google Scholar]
58. Oelschlaeger TA, Tall BD. Uptake pathways of clinical isolates of Proteus mirabilis into human epithelial cell lines. Microb Pathog. 1996;21:1–16. [Abstract] [Google Scholar]
59. Braude AI, Siemienski J. ROLE OF BACTERIAL UREASE IN EXPERIMENTAL PYELONEPHRITIS. J Bacteriol. 1960;80:171–179. [Europe PMC free article] [Abstract] [Google Scholar]
60. Ingersoll MA, Albert ML. From infection to immunotherapy: host immune responses to bacteria at the bladder mucosa. Mucosal Immunol. 2013;6:1041–1053. [Abstract] [Google Scholar]
61. Ambite I, Nagy K, Godaly G, Puthia M, Wullt B, Svanborg C. Susceptibility to Urinary Tract Infection: Benefits and Hazards of the Antibacterial Host Response. Microbiol Spectr. 2016;4 [Abstract] [Google Scholar]
62. Hayes BW, Abraham SN. Innate Immune Responses to Bladder Infection. Microbiol Spectr. 2016;4 [Europe PMC free article] [Abstract] [Google Scholar]
63. Kolaczkowska E, Kubes P. Neutrophil recruitment and function in health and inflammation. Nat Rev Immunol. 2013;13:159–175. [Abstract] [Google Scholar]
64. Kline KA, Schwartz DJ, Lewis WG, Hultgren SJ, Lewis AL. Immune activation and suppression by group B streptococcus in a murine model of urinary tract infection. Infect Immun. 2011;79:3588–3595. [Europe PMC free article] [Abstract] [Google Scholar]
65. Weichhart T, Haidinger M, Horl WH, Saemann MD. Current concepts of molecular defence mechanisms operative during urinary tract infection. Eur J Clin Invest. 2008;38(Suppl 2):29–38. [Abstract] [Google Scholar]
66. Papayannopoulos V, Zychlinsky A. NETs: a new strategy for using old weapons. Trends Immunol. 2009;30:513–521. [Abstract] [Google Scholar]
67. Yu Y, Sikorski P, Bowman-Gholston C, Cacciabeve N, Nelson KE, Pieper R. Diagnosing inflammation and infection in the urinary system via proteomics. J Transl Med. 2015;13:111. [Europe PMC free article] [Abstract] [Google Scholar]
68. Yu Y, Kwon K, Tsitrin T, Bekele S, Sikorski P, Nelson KE, Pieper R. Characterization of Early-Phase Neutrophil Extracellular Traps in Urinary Tract Infections. PLoS Pathog. 2017;13:e1006151. [Europe PMC free article] [Abstract] [Google Scholar]
69. Chen CY, Chen YH, Lu PL, Lin WR, Chen TC, Lin CY. Proteus mirabilis urinary tract infection and bacteremia: risk factors, clinical presentation, and outcomes. J Microbiol Immunol Infect. 2012;45:228–236. [Abstract] [Google Scholar]
70. Guo H, Callaway JB, Ting JP. Inflammasomes: mechanism of action, role in disease, and therapeutics. Nat Med. 2015;21:677–687. [Europe PMC free article] [Abstract] [Google Scholar]
71. Seo SU, Kamada N, Munoz-Planillo R, Kim YG, Kim D, Koizumi Y, Hasegawa M, Himpsl SD, Browne HP, Lawley TD, Mobley HL, Inohara N, Nunez G. Distinct Commensals Induce Interleukin-1beta via NLRP3 Inflammasome in Inflammatory Monocytes to Promote Intestinal Inflammation in Response to Injury. Immunity. 2015;42:744–755. [Europe PMC free article] [Abstract] [Google Scholar]
72. Devuyst O, Olinger E, Rampoldi L. Uromodulin: from physiology to rare and complex kidney disorders. Nat Rev Nephrol. 2017;13:525–544. [Abstract] [Google Scholar]
73. Bates JM, Raffi HM, Prasadan K, Mascarenhas R, Laszik Z, Maeda N, Hultgren SJ, Kumar S. Tamm-Horsfall protein knockout mice are more prone to urinary tract infection: rapid communication. Kidney Int. 2004;65:791–797. [Abstract] [Google Scholar]
74. Raffi HS, Bates JM, Jr, Laszik Z, Kumar S. Tamm-horsfall protein protects against urinary tract infection by proteus mirabilis. J Urol. 2009;181:2332–2338. [Europe PMC free article] [Abstract] [Google Scholar]
75. Steiner TS. How flagellin and toll-like receptor 5 contribute to enteric infection. Infect Immun. 2007;75:545–552. [Europe PMC free article] [Abstract] [Google Scholar]
76. Okumura R, Kurakawa T, Nakano T, Kayama H, Kinoshita M, Motooka D, Gotoh K, Kimura T, Kamiyama N, Kusu T, Ueda Y, Wu H, Iijima H, Barman S, Osawa H, Matsuno H, Nishimura J, Ohba Y, Nakamura S, Iida T, Yamamoto M, Umemoto E, Sano K, Takeda K. Lypd8 promotes the segregation of flagellated microbiota and colonic epithelia. Nature. 2016;532:117–121. [Abstract] [Google Scholar]
77. Umpiérrez A, Scavone P, Romanin D, Marqués JM, Chabalgoity JA, Rumbo M, Zunino P. Innate immune responses to Proteus mirabilis flagellin in the urinary tract. Microbes and Infection. 2013;15:688–696. [Abstract] [Google Scholar]
78. Mobley HLT, Belas R. Swarming and pathogenicity of Proteus mirabilis in the urinary tract. Trends in Microbiology. 1995;3:280–284. [Abstract] [Google Scholar]
79. Olaitan AO, Morand S, Rolain JM. Mechanisms of polymyxin resistance: acquired and intrinsic resistance in bacteria. Front Microbiol. 2014;5:643. [Europe PMC free article] [Abstract] [Google Scholar]
80. McCoy AJ, Liu H, Falla TJ, Gunn JS. Identification of Proteus mirabilis mutants with increased sensitivity to antimicrobial peptides. Antimicrob Agents Chemother. 2001;45:2030–2037. [Europe PMC free article] [Abstract] [Google Scholar]
81. Chromek M, Stankowska D, Dadfar E, Kaca W, Rabbani H, Brauner A. Interleukin-8 response in cells from the human urinary tract induced by lipopolysaccharides of Proteus mirabilis O3 and O18. J Urol. 2005;173:1381–1384. [Abstract] [Google Scholar]
82. Callewaert L, Vanderkelen L, Deckers D, Aertsen A, Robben J, Michiels CW. Detection of a lysozyme inhibitor in Proteus mirabilis by a new reverse zymogram method. Appl Environ Microbiol. 2008;74:4978–4981. [Europe PMC free article] [Abstract] [Google Scholar]
83. Rashid T, Ebringer A, Wilson C. The link between Proteus mirabilis, environmental factors and autoantibodies in rheumatoid arthritis. Clin Exp Rheumatol 2017 [Abstract] [Google Scholar]
84. Pretorius E, Akeredolu OO, Soma P, Kell DB. Major involvement of bacterial components in rheumatoid arthritis and its accompanying oxidative stress, systemic inflammation and hypercoagulability. Exp Biol Med (Maywood) 2017;242:355–373. [Europe PMC free article] [Abstract] [Google Scholar]
85. Garrett WS, Gallini CA, Yatsunenko T, Michaud M, DuBois A, Delaney ML, Punit S, Karlsson M, Bry L, Glickman JN, Gordon JI, Onderdonk AB, Glimcher LH. Enterobacteriaceae act in concert with the gut microbiota to induce spontaneous and maternally transmitted colitis. Cell Host Microbe. 2010;8:292–300. [Europe PMC free article] [Abstract] [Google Scholar]
86. Keane WF, Freedman LR. Experimental pyelonephritis. XIV. Pyelonephritis in normal mice produced by inoculation of E. coli into the bladder lumen during water diuresis. Yale J Biol Med. 1967;40:231–237. [Europe PMC free article] [Abstract] [Google Scholar]
87. Hagberg L, Engberg I, Freter R, Lam J, Olling S, Svanborg Eden C. Ascending, unobstructed urinary tract infection in mice caused by pyelonephritogenic Escherichia coli of human origin. Infect Immun. 1983;40:273–283. [Europe PMC free article] [Abstract] [Google Scholar]
88. Kadurugamuwa JL, Modi K, Yu J, Francis KP, Purchio T, Contag PR. Noninvasive Biophotonic Imaging for Monitoring of Catheter-Associated Urinary Tract Infections and Therapy in Mice. Infect Immun. 2005;73:3878–3887. [Europe PMC free article] [Abstract] [Google Scholar]
89. Conover MS, Flores-Mireles AL, Hibbing ME, Dodson K, Hultgren SJ. Establishment and Characterization of UTI and CAUTI in a Mouse Model. J Vis Exp. 2015 10.3791/52892. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
90. Thai KH, Thathireddy A, Hsieh MH. Transurethral induction of mouse urinary tract infection. J Vis Exp. 2010 10.3791/2070. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
91. Johnson DE, Lockatell CV, Hall-Craigs M, Mobley HL, Warren JW. Uropathogenicity in rats and mice of Providencia stuartii from long-term catheterized patients. J Urol. 1987;138:632–635. [Abstract] [Google Scholar]
92. Guiton PS, Hung CS, Hancock LE, Caparon MG, Hultgren SJ. Enterococcal Biofilm Formation and Virulence in an Optimized Murine Model of Foreign Body-Associated Urinary Tract Infections. Infection and Immunity. 2010;78:4166–4175. [Europe PMC free article] [Abstract] [Google Scholar]
93. Kurosaka Y, Ishida Y, Yamamura E, Takase H, Otani T, Kumon H. A non-surgical rat model of foreign body-associated urinary tract infection with Pseudomonas aeruginosa. Microbiol Immunol. 2001;45:9–15. [Abstract] [Google Scholar]
94. Janssen C, Lo J, Jager W, Moskalev I, Law A, Chew BH, Lange D. A high throughput, minimally invasive, ultrasound guided model for the study of catheter associated urinary tract infections and device encrustation in mice. J Urol. 2014;192:1856–1863. [Abstract] [Google Scholar]
95. Williams FD, Schwarzhoff RH. Nature of the swarming phenomenon in Proteus. Annu Rev Microbiol. 1978;32:101–122. [Abstract] [Google Scholar]
96. Rozalski A, Sidorczyk Z, Kotelko K. Potential virulence factors of Proteus bacilli. Microbiol Mol Biol Rev. 1997;61:65–89. [Europe PMC free article] [Abstract] [Google Scholar]
97. O’Hara CM, Brenner FW, Miller JM. Classification, identification, and clinical significance of Proteus, Providencia, and Morganella. Clin Microbiol Rev. 2000;13:534–546. [Europe PMC free article] [Abstract] [Google Scholar]
98. Welch RA, Burland V, Plunkett G, 3rd, Redford P, Roesch P, Rasko D, Buckles EL, Liou SR, Boutin A, Hackett J, Stroud D, Mayhew GF, Rose DJ, Zhou S, Schwartz DC, Perna NT, Mobley HL, Donnenberg MS, Blattner FR. Extensive mosaic structure revealed by the complete genome sequence of uropathogenic Escherichia coli. Proc Natl Acad Sci U S A. 2002;99:17020–17024. [Europe PMC free article] [Abstract] [Google Scholar]
99. Touchon M, Hoede C, Tenaillon O, Barbe V, Baeriswyl S, Bidet P, Bingen E, Bonacorsi S, Bouchier C, Bouvet O, Calteau A, Chiapello H, Clermont O, Cruveiller S, Danchin A, Diard M, Dossat C, Karoui ME, Frapy E, Garry L, Ghigo JM, Gilles AM, Johnson J, Le Bouguenec C, Lescat M, Mangenot S, Martinez-Jehanne V, Matic I, Nassif X, Oztas S, Petit MA, Pichon C, Rouy Z, Ruf CS, Schneider D, Tourret J, Vacherie B, Vallenet D, Medigue C, Rocha EP, Denamur E. Organised genome dynamics in the Escherichia coli species results in highly diverse adaptive paths. PLoS Genet. 2009;5:e1000344. [Europe PMC free article] [Abstract] [Google Scholar]
100. Mobley HL, Chippendale GR. Hemagglutinin, urease, and hemolysin production by Proteus mirabilis from clinical sources. J Infect Dis. 1990;161:525–530. [Abstract] [Google Scholar]
101. Sosa V, Schlapp G, Zunino P. Proteus mirabilis isolates of different origins do not show correlation with virulence attributes and can colonize the urinary tract of mice. Microbiology. 2006;152:2149–2157. [Abstract] [Google Scholar]
102. Swihart KG, Welch RA. The HpmA hemolysin is more common than HlyA among Proteus isolates. Infect Immun. 1990;58:1853–1860. [Europe PMC free article] [Abstract] [Google Scholar]
103. Cestari SE, Ludovico MS, Martins FH, da Rocha SP, Elias WP, Pelayo JS. Molecular detection of HpmA and HlyA hemolysin of uropathogenic Proteus mirabilis. Curr Microbiol. 2013;67:703–707. [Abstract] [Google Scholar]
104. Stankowska D, Kwinkowski M, Kaca W. Quantification of Proteus mirabilis virulence factors and modulation by acylated homoserine lactones. J Microbiol Immunol Infect. 2008;41:243–253. [Abstract] [Google Scholar]
105. Loomes LM, Senior BW, Kerr MA. Proteinases of Proteus spp.: purification, properties, and detection in urine of infected patients. Infect Immun. 1992;60:2267–2273. [Europe PMC free article] [Abstract] [Google Scholar]
106. Kuan L, Schaffer JN, Zouzias CD, Pearson MM. Characterization of 17 chaperone-usher fimbriae encoded by Proteus mirabilis reveals strong conservation. J Med Microbiol. 2014;63:911–922. [Europe PMC free article] [Abstract] [Google Scholar]
107. Flannery EL, Mody L, Mobley HLT. Identification of a Modular Pathogenicity Island That Is Widespread among Urease-Producing Uropathogens and Shares Features with a Diverse Group of Mobile Elements. Infect Immun. 2009;77:4887–4894. [Europe PMC free article] [Abstract] [Google Scholar]
108. Burall LS, Harro JM, Li X, Lockatell CV, Himpsl SD, Hebel JR, Johnson DE, Mobley HLT. Proteus mirabilis Genes That Contribute to Pathogenesis of Urinary Tract Infection: Identification of 25 Signature-Tagged Mutants Attenuated at Least 100-Fold. Infect Immun. 2004;72:2922–2938. [Europe PMC free article] [Abstract] [Google Scholar]
109. Himpsl SD, Pearson MM, Arewång CJ, Nusca TD, Sherman DH, Mobley HLT. Proteobactin and a yersiniabactin-related siderophore mediate iron acquisition in Proteus mirabilis. Molecular Microbiology. 2010;78:138–157. [Europe PMC free article] [Abstract] [Google Scholar]
110. Alamuri P, Mobley HLT. A novel autotransporter of uropathogenic Proteus mirabilis is both a cytotoxin and an agglutinin. Molecular Microbiology. 2008;68:997–1017. [Abstract] [Google Scholar]
111. Flannery EL, Antczak SM, Mobley HLT. Self-Transmissibility of the Integrative and Conjugative Element ICEPm1 between Clinical Isolates Requires a Functional Integrase, Relaxase, and Type IV Secretion System. J Bacteriol. 2011;193:4104–4112. [Europe PMC free article] [Abstract] [Google Scholar]
112. Pearson MM, Sebaihia M, Churcher C, Quail MA, Seshasayee AS, Luscombe NM, Abdellah Z, Arrosmith C, Atkin B, Chillingworth T, Hauser H, Jagels K, Moule S, Mungall K, Norbertczak H, Rabbinowitsch E, Walker D, Whithead S, Thomson NR, Rather PN, Parkhill J, Mobley HLT. Complete Genome Sequence of Uropathogenic Proteus mirabilis, a Master of both Adherence and Motility. J Bacteriol. 2008;190:4027–4037. [Europe PMC free article] [Abstract] [Google Scholar]
113. Wozniak RA, Waldor MK. Integrative and conjugative elements: mosaic mobile genetic elements enabling dynamic lateral gene flow. Nat Rev Microbiol. 2010;8:552–563. [Abstract] [Google Scholar]
114. Ceccarelli D, Daccord A, Rene M, Burrus V. Identification of the origin of transfer (oriT) and a new gene required for mobilization of the SXT/R391 family of integrating conjugative elements. J Bacteriol. 2008;190:5328–5338. [Europe PMC free article] [Abstract] [Google Scholar]
115. Harada S, Ishii Y, Saga T, Tateda K, Yamaguchi K. Chromosomally encoded blaCMY-2 located on a novel SXT/R391-related integrating conjugative element in a Proteus mirabilis clinical isolate. Antimicrob Agents Chemother. 2010;54:3545–3550. [Europe PMC free article] [Abstract] [Google Scholar]
116. Li X, Du Y, Du P, Dai H, Fang Y, Li Z, Lv N, Zhu B, Kan B, Wang D. SXT/R391 integrative and conjugative elements in Proteus species reveal abundant genetic diversity and multidrug resistance. Sci Rep. 2016;6:37372. [Europe PMC free article] [Abstract] [Google Scholar]
117. Bie L, Wu H, Wang XH, Wang M, Xu H. Identification and characterization of new members of the SXT/R391 family of integrative and conjugative elements (ICEs) in Proteus mirabilis. Int J Antimicrob Agents. 2017 10.1016/j.ijantimicag.2017.01.045. [Abstract] [CrossRef] [Google Scholar]
118. Wenren LM, Sullivan NL, Cardarelli L, Septer AN, Gibbs KA. Two Independent Pathways for Self-Recognition in Proteus mirabilis Are Linked by Type VI-Dependent Export. mBio. 2013;4:e00374–00313. [Europe PMC free article] [Abstract] [Google Scholar]
119. Alteri CJ, Himpsl SD, Pickens SR, Lindner JR, Zora JS, Miller JE, Arno PD, Straight SW, Mobley HLT. Multicellular Bacteria Deploy the Type VI Secretion System to Preemptively Strike Neighboring Cells. PLoS Pathogens. 2013;9:e1003608. [Europe PMC free article] [Abstract] [Google Scholar]
120. Alteri CJ, Mobley HL. The Versatile Type VI Secretion System. Microbiol Spectr. 2016;4 [Europe PMC free article] [Abstract] [Google Scholar]
121. Sullivan NL, Septer AN, Fields AT, Wenren LM, Gibbs KA. The Complete Genome Sequence of Proteus mirabilis Strain BB2000 Reveals Differences from the P. mirabilis Reference Strain. Genome Announcements. 2013;1 [Europe PMC free article] [Abstract] [Google Scholar]
122. Knirel YA, Perepelov AV, Kondakova AN, Senchenkova SN, Sidorczyk Z, Rozalski A, Kaca W. Structure and serology of O-antigens as the basis for classification of Proteus strains. Innate Immun. 2011;17:70–96. [Abstract] [Google Scholar]
123. Ramos HC, Rumbo M, Sirard JC. Bacterial flagellins: mediators of pathogenicity and host immune responses in mucosa. Trends Microbiol. 2004;12:509–517. [Abstract] [Google Scholar]
124. Murphy CA, Belas R. Genomic rearrangements in the flagellin genes of Proteus mirabilis. Molecular Microbiology. 1999;31:679–690. [Abstract] [Google Scholar]
125. Wattam AR, Davis JJ, Assaf R, Boisvert S, Brettin T, Bun C, Conrad N, Dietrich EM, Disz T, Gabbard JL, Gerdes S, Henry CS, Kenyon RW, Machi D, Mao C, Nordberg EK, Olsen GJ, Murphy-Olson DE, Olson R, Overbeek R, Parrello B, Pusch GD, Shukla M, Vonstein V, Warren A, Xia F, Yoo H, Stevens RL. Improvements to PATRIC, the all-bacterial Bioinformatics Database and Analysis Resource Center. Nucleic Acids Res. 2017;45:D535–D542. [Europe PMC free article] [Abstract] [Google Scholar]
126. Pearson MM, Rasko DA, Smith SN, Mobley HLT. Transcriptome of Swarming Proteus mirabilis. Infect Immun. 2010;78:2834–2845. [Europe PMC free article] [Abstract] [Google Scholar]
127. Cusick K, Lee YY, Youchak B, Belas R. Perturbation of FliL interferes with Proteus mirabilis swarmer cell gene expression and differentiation. J Bacteriol. 2012;194:437–447. [Europe PMC free article] [Abstract] [Google Scholar]
128. Coetzee JN. Transduction of Swarming in Proteus Mirabilis. J Gen Microbiol. 1963;33:1–7. [Abstract] [Google Scholar]
129. Armbruster CE, Forsyth-DeOrnellas V, Johnson AO, Smith SN, Zhao L, Wu W, Mobley HLT. Genome-wide transposon mutagenesis of Proteus mirabilis: Essential genes, fitness factors for catheter-associated urinary tract infection, and the impact of polymicrobial infection on fitness requirements. PLOS Pathogens. 2017;13:e1006434. [Europe PMC free article] [Abstract] [Google Scholar]
130. Tatusova T. Update on Genomic Databases and Resources at the National Center for Biotechnology Information. Methods Mol Biol. 2016;1415:3–30. [Abstract] [Google Scholar]
131. Pearson MM, Yep A, Smith SN, Mobley HLT. Transcriptome of Proteus mirabilis in the Murine Urinary Tract: Virulence and Nitrogen Assimilation Gene Expression. Infect Immun. 2011 10.1128/iai.05152-11. IAI.05152-05111. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
132. Bode NJ, Debnath I, Kuan L, Schulfer A, Ty M, Pearson MM. Transcriptional analysis of the MrpJ network: modulation of diverse virulence-associated genes and direct regulation of mrp fimbrial and flhDC flagellar operons in Proteus mirabilis. Infect Immun. 2015;83:2542–2556. [Europe PMC free article] [Abstract] [Google Scholar]
133. Allison C, Lai H-C, Hughes C. Co-ordinate expression of virulence genes during swarm-cell differentiation and population migration of Proteus mirabilis. Molecular Microbiology. 1992;6:1583–1591. [Abstract] [Google Scholar]
134. Walker KE, Moghaddame-Jafari S, Lockatell CV, Johnson D, Belas R. ZapA, the IgA-degrading metalloprotease of Proteus mirabilis, is a virulence factor expressed specifically in swarmer cells. Molecular Microbiology. 1999;32:825–836. [Abstract] [Google Scholar]
135. Armitage JP. Changes in metabolic activity of Proteus mirabilis during swarming. J Gen Microbiol. 1981;125:445–450. [Abstract] [Google Scholar]
136. Snyder JA, Haugen BJ, Buckles EL, Lockatell CV, Johnson DE, Donnenberg MS, Welch RA, Mobley HL. Transcriptome of uropathogenic Escherichia coli during urinary tract infection. Infect Immun. 2004;72:6373–6381. [Europe PMC free article] [Abstract] [Google Scholar]
137. Li X, Rasko DA, Lockatell CV, Johnson DE, Mobley HLT. Repression of bacterial motility by a novel fimbrial gene product. EMBO J. 2001;20:4854–4862. [Europe PMC free article] [Abstract] [Google Scholar]
138. Pearson MM, Mobley HLT. Repression of motility during fimbrial expression: identification of 14 mrpJ gene paralogues in Proteus mirabilis. Molecular Microbiology. 2008;69:548–558. [Europe PMC free article] [Abstract] [Google Scholar]
139. Howery KE, Clemmer KM, Simsek E, Kim M, Rather PN. Regulation of the Min Cell Division Inhibition Complex by the Rcs Phosphorelay in Proteus mirabilis. J Bacteriol. 2015;197:2499–2507. [Europe PMC free article] [Abstract] [Google Scholar]
140. Howery KE, Clemmer KM, Rather PN. The Rcs regulon in Proteus mirabilis: implications for motility, biofilm formation, and virulence. Current Genetics. 2016;62:775–789. [Abstract] [Google Scholar]
141. Belas R, Suvanasuthi R. The Ability of Proteus mirabilis To Sense Surfaces and Regulate Virulence Gene Expression Involves FliL, a Flagellar Basal Body Protein. J Bacteriol. 2005;187:6789–6803. [Europe PMC free article] [Abstract] [Google Scholar]
142. Belas R, Schneider R, Melch M. Characterization of Proteus mirabilis Precocious Swarming Mutants: Identification of rsbA, Encoding a Regulator of Swarming Behavior. J Bacteriol. 1998;180:6126–6139. [Europe PMC free article] [Abstract] [Google Scholar]
143. LIAW S-J, LAI H-C, HO S-W, LUH K-T, WANG W-B. Characterisation of p-nitrophenylglycerol-resistant Proteus mirabilis super-swarming mutants. Journal of Medical Microbiology. 2001;50:1039–1048. [Abstract] [Google Scholar]
144. Clemmer KM, Rather PN. The Lon protease regulates swarming motility and virulence gene expression in Proteus mirabilis. J Med Microbiol. 2008;57:931–937. [Abstract] [Google Scholar]
145. Miller VL, Mekalanos JJ. A novel suicide vector and its use in construction of insertion mutations: osmoregulation of outer membrane proteins and virulence determinants in Vibrio cholerae requires toxR. J Bacteriol. 1988;170:2575–2583. [Europe PMC free article] [Abstract] [Google Scholar]
146. Li X, Zhao H, Geymonat L, Bahrani F, Johnson D, Mobley H. Proteus mirabilis mannose-resistant, Proteus-like fimbriae: MrpG is located at the fimbrial tip and is required for fimbrial assembly. Infect Immun. 1997;65:1327–1334. [Europe PMC free article] [Abstract] [Google Scholar]
147. Karberg M, Guo H, Zhong J, Coon R, Perutka J, Lambowitz AM. Group II introns as controllable gene targeting vectors for genetic manipulation of bacteria. Nat Biotechnol. 2001;19:1162–1167. [Abstract] [Google Scholar]
148. Pearson MM, Mobley HLT. The type III secretion system of Proteus mirabilis HI4320 does not contribute to virulence in the mouse model of ascending urinary tract infection. J Med Microbiol. 2007;56:1277–1283. [Abstract] [Google Scholar]
149. Bahrani FK, Massad G, Lockatell CV, Johnson DE, Russell RG, Warren JW, Mobley HL. Construction of an MR/P fimbrial mutant of Proteus mirabilis: role in virulence in a mouse model of ascending urinary tract infection. Infect Immun. 1994;62:3363–3371. [Europe PMC free article] [Abstract] [Google Scholar]
150. Zunino P, Geymonat L, Allen AG, Preston A, Sosa V, Maskell DJ. New aspects of the role of MR/P fimbriae in Proteus mirabilis urinary tract infection. FEMS Immunology & Medical Microbiology. 2001;31:113–120. [Abstract] [Google Scholar]
151. Zunino P, Sosa V, Schlapp G, Allen AG, Preston A, Maskell DJ. Mannose-resistant Proteus-like and P. mirabilis fimbriae have specific and additive roles in P. mirabilis urinary tract infections. FEMS Immunology & Medical Microbiology. 2007;51:125–133. [Abstract] [Google Scholar]
152. Lane MC, Li X, Pearson MM, Simms AN, Mobley HLT. Oxygen-Limiting Conditions Enrich for Fimbriate Cells of Uropathogenic Proteus mirabilis and Escherichia coli. J Bacteriol. 2009;191:1382–1392. [Europe PMC free article] [Abstract] [Google Scholar]
153. Massad G, Lockatell CV, Johnson DE, Mobley HL. Proteus mirabilis fimbriae: construction of an isogenic pmfA mutant and analysis of virulence in a CBA mouse model of ascending urinary tract infection. Infect Immun. 1994;62:536–542. [Europe PMC free article] [Abstract] [Google Scholar]
154. Zunino P, Sosa V, Allen AG, Preston A, Schlapp G, Maskell DJ. Proteus mirabilis fimbriae (PMF) are important for both bladder and kidney colonization in mice. Microbiology. 2003;149:3231–3237. [Abstract] [Google Scholar]
155. Legnani-Fajardo C, Zunino P, Piccini C, Allen A, Maskell D. Defined mutants of Proteus mirabilis lacking flagella cause ascending urinary tract infection in mice. Microbial Pathogenesis. 1996;21:395–405. [Abstract] [Google Scholar]
156. Li X, Johnson DE, Mobley HLT. Requirement of MrpH for Mannose-Resistant Proteus-Like Fimbria-Mediated Hemagglutination by Proteus mirabilis. Infect Immun. 1999;67:2822–2833. [Europe PMC free article] [Abstract] [Google Scholar]
157. Zunino P, Geymonat L, Allen AG, Legnani-Fajardo C, Maskell DJ. Virulence of a Proteus mirabilis ATF isogenic mutant is not impaired in a mouse model of ascending urinary tract infection. FEMS Immunology & Medical Microbiology. 2000;29:137–143. [Abstract] [Google Scholar]
158. Li X, Lockatell CV, Johnson DE, Mobley HLT. Identification of MrpI as the sole recombinase that regulates the phase variation of MR/P fimbria, a bladder colonization factor of uropathogenic Proteus mirabilis. Molecular Microbiology. 2002;45:865–874. [Abstract] [Google Scholar]
159. Schneider R, Lockatell CV, Johnson D, Belas R. Detection and mutation of a luxS-encoded autoinducer in Proteus mirabilis. Microbiology. 2002;148:773–782. [Abstract] [Google Scholar]
160. Dattelbaum JD, Lockatell CV, Johnson DE, Mobley HLT. UreR, the Transcriptional Activator of the Proteus mirabilis Urease Gene Cluster, Is Required for Urease Activity and Virulence in Experimental Urinary Tract Infections. Infect Immun. 2003;71:1026–1030. [Europe PMC free article] [Abstract] [Google Scholar]
161. Nielubowicz GR, Smith SN, Mobley HLT. Outer Membrane Antigens of the Uropathogen Proteus mirabilis Recognized by the Humoral Response during Experimental Murine Urinary Tract Infection. Infect Immun. 2008;76:4222–4231. [Europe PMC free article] [Abstract] [Google Scholar]
162. Nielubowicz GR, Smith SN, Mobley HLT. Zinc Uptake Contributes to Motility and Provides a Competitive Advantage to Proteus mirabilis during Experimental Urinary Tract Infection. Infect Immun. 2010;78:2823–2833. [Europe PMC free article] [Abstract] [Google Scholar]
163. Pellegrino R, Scavone P, Umpierrez A, Maskell DJ, Zunino P. Proteus mirabilis uroepithelial cell adhesin (UCA) fimbria plays a role in the colonization of the urinary tract. Pathog Dis. 2013;67:104–107. [Abstract] [Google Scholar]
164. Wang MC, Chien HF, Tsai YL, Liu MC, Liaw SJ. The RNA chaperone Hfq is involved in stress tolerance and virulence in uropathogenic Proteus mirabilis. PLoS One. 2014;9:e85626. [Europe PMC free article] [Abstract] [Google Scholar]
165. Armbruster CE, Hodges SA, Smith SN, Alteri CJ, Mobley HL. Arginine promotes Proteus mirabilis motility and fitness by contributing to conservation of the proton gradient and proton motive force. Microbiologyopen. 2014;3:630–641. [Europe PMC free article] [Abstract] [Google Scholar]
166. Alteri CJ, Himpsl SD, Mobley HL. Preferential use of central metabolism in vivo reveals a nutritional basis for polymicrobial infection. PLoS Pathog. 2015;11:e1004601. [Europe PMC free article] [Abstract] [Google Scholar]
167. Liu MC, Kuo KT, Chien HF, Tsai YL, Liaw SJ. New aspects of RpoE in uropathogenic Proteus mirabilis. Infect Immun. 2015;83:966–977. [Europe PMC free article] [Abstract] [Google Scholar]
168. Phan V, Belas R, Gilmore BF, Ceri H. ZapA, a virulence factor in a rat model of Proteus mirabilis-induced acute and chronic prostatitis. Infect Immun. 2008;76:4859–4864. [Europe PMC free article] [Abstract] [Google Scholar]
169. Belas R, Erskine D, Flaherty D. Transposon mutagenesis in Proteus mirabilis. J Bacteriol. 1991;173:6289–6293. [Europe PMC free article] [Abstract] [Google Scholar]
170. de Lorenzo V, Herrero M, Jakubzik U, Timmis KN. Mini-Tn5 transposon derivatives for insertion mutagenesis, promoter probing, and chromosomal insertion of cloned DNA in gram-negative eubacteria. J Bacteriol. 1990;172:6568–6572. [Europe PMC free article] [Abstract] [Google Scholar]
171. Belas R, Erskine D, Flaherty D. Proteus mirabilis mutants defective in swarmer cell differentiation and multicellular behavior. J Bacteriol. 1991;173:6279–6288. [Europe PMC free article] [Abstract] [Google Scholar]
172. Belas R, Goldman M, Ashliman K. Genetic analysis of Proteus mirabilis mutants defective in swarmer cell elongation. J Bacteriol. 1995;177:823–828. [Europe PMC free article] [Abstract] [Google Scholar]
173. Hay NA, Tipper DJ, Gygi D, Hughes C. A nonswarming mutant of Proteus mirabilis lacks the Lrp global transcriptional regulator. J Bacteriol. 1997;179:4741–4746. [Europe PMC free article] [Abstract] [Google Scholar]
174. Alteri CJ, Himpsl SD, Engstrom MD, Mobley HL. Anaerobic Respiration Using a Complete Oxidative TCA Cycle Drives Multicellular Swarming in Proteus mirabilis. MBio. 2012;3 [Europe PMC free article] [Abstract] [Google Scholar]
175. Gibbs KA, Urbanowski ML, Greenberg EP. Genetic Determinants of Self Identity and Social Recognition in Bacteria. Science. 2008;321:256–259. [Europe PMC free article] [Abstract] [Google Scholar]
176. Jiang S-S, Lin T-Y, Wang W-B, Liu M-C, Hsueh P-R, Liaw S-J. Characterization of UDP-Glucose Dehydrogenase and UDP-Glucose Pyrophosphorylase Mutants of Proteus mirabilis: Defectiveness in Polymyxin B Resistance, Swarming, and Virulence. Antimicrob Agents Chemother. 2010;54:2000–2009. [Europe PMC free article] [Abstract] [Google Scholar]
177. Holling N, Lednor D, Tsang S, Bissell A, Campbell L, Nzakizwanayo J, Dedi C, Hawthorne JA, Hanlon G, Ogilvie LA, Salvage JP, Patel BA, Barnes LM, Jones BV. Elucidating the Genetic Basis of Crystalline Biofilm Formation in Proteus mirabilis. Infection and Immunity. 2014;82:1616–1626. [Europe PMC free article] [Abstract] [Google Scholar]
178. Hensel M, Shea JE, Gleeson C, Jones MD, Dalton E, Holden DW. Simultaneous identification of bacterial virulence genes by negative selection. Science. 1995;269:400–403. [Abstract] [Google Scholar]
179. Himpsl SD, Lockatell CV, Hebel JR, Johnson DE, Mobley HLT. Identification of virulence determinants in uropathogenic Proteus mirabilis using signature-tagged mutagenesis. J Med Microbiol. 2008;57:1068–1078. [Abstract] [Google Scholar]
180. Zhao H, Li X, Johnson DE, Mobley HLT. Identification of protease and rpoN-associated genes of uropathogenic Proteus mirabilis by negative selection in a mouse model of ascending urinary tract infection. Microbiology. 1999;145:185–195. [Abstract] [Google Scholar]
181. Goodman AL, McNulty NP, Zhao Y, Leip D, Mitra RD, Lozupone CA, Knight R, Gordon JI. Identifying genetic determinants needed to establish a human gut symbiont in its habitat. Cell Host Microbe. 2009;6:279–289. [Europe PMC free article] [Abstract] [Google Scholar]
182. Langridge GC, Phan MD, Turner DJ, Perkins TT, Parts L, Haase J, Charles I, Maskell DJ, Peters SE, Dougan G, Wain J, Parkhill J, Turner AK. Simultaneous assay of every Salmonella Typhi gene using one million transposon mutants. Genome Res. 2009;19:2308–2316. [Europe PMC free article] [Abstract] [Google Scholar]
183. Gallagher LA, Shendure J, Manoil C. Genome-Scale Identification of Resistance Functions in Pseudomonas aeruginosa Using Tn-seq. mBio. 2011;2 [Europe PMC free article] [Abstract] [Google Scholar]
184. van Opijnen T, Bodi KL, Camilli A. Tn-seq: high-throughput parallel sequencing for fitness and genetic interaction studies in microorganisms. Nat Methods. 2009;6:767–772. [Europe PMC free article] [Abstract] [Google Scholar]
185. Gawronski JD, Wong SMS, Giannoukos G, Ward DV, Akerley BJ. Tracking insertion mutants within libraries by deep sequencing and a genome-wide screen for Haemophilus genes required in the lung. Proceedings of the National Academy of Sciences of the United States of America. 2009;106:16422–16427. [Europe PMC free article] [Abstract] [Google Scholar]
186. Mobley HL, Island MD, Hausinger RP. Molecular biology of microbial ureases. Microbiol Rev. 1995;59:451–480. [Europe PMC free article] [Abstract] [Google Scholar]
187. Mobley HL, Hausinger RP. Microbial ureases: significance, regulation, and molecular characterization. Microbiol Rev. 1989;53:85–108. [Europe PMC free article] [Abstract] [Google Scholar]
188. Heimer SR, Mobley HLT. Interaction of Proteus mirabilis Urease Apoenzyme and Accessory Proteins Identified with Yeast Two-Hybrid Technology. J Bacteriol. 2001;183:1423–1433. [Europe PMC free article] [Abstract] [Google Scholar]
189. Jones BD, Mobley HL. Genetic and biochemical diversity of ureases of Proteus, Providencia, and Morganella species isolated from urinary tract infection. Infect Immun. 1987;55:2198–2203. [Europe PMC free article] [Abstract] [Google Scholar]
190. Jones BD, Mobley HL. Proteus mirabilis urease: nucleotide sequence determination and comparison with jack bean urease. J Bacteriol. 1989;171:6414–6422. [Europe PMC free article] [Abstract] [Google Scholar]
191. Sriwanthana B, Island MD, Mobley HL. Sequence of the Proteus mirabilis urease accessory gene ureG. Gene. 1993;129:103–106. [Abstract] [Google Scholar]
192. Nicholson EB, Concaugh EA, Foxall PA, Island MD, Mobley HL. Proteus mirabilis urease: transcriptional regulation by UreR. J Bacteriol. 1993;175:465–473. [Europe PMC free article] [Abstract] [Google Scholar]
193. Thomas VJ, Collins CM. Identification of UreR binding sites in the Enterobacteriaceae plasmid-encoded and Proteus mirabilis urease gene operons. Mol Microbiol. 1999;31:1417–1428. [Abstract] [Google Scholar]
194. Coker C, Bakare OO, Mobley HL. H-NS is a repressor of the Proteus mirabilis urease transcriptional activator gene ureR. J Bacteriol. 2000;182:2649–2653. [Europe PMC free article] [Abstract] [Google Scholar]
195. D’Orazio SE, Thomas V, Collins CM. Activation of transcription at divergent urea-dependent promoters by the urease gene regulator UreR. Mol Microbiol. 1996;21:643–655. [Abstract] [Google Scholar]
196. Poore CA, Mobley HLT. Differential regulation of the Proteus mirabilis urease gene cluster by UreR and H-NS. Microbiology. 2003;149:3383–3394. [Abstract] [Google Scholar]
197. Follmer C. Ureases as a target for the treatment of gastric and urinary infections. Journal of Clinical Pathology. 2010;63:424–430. [Abstract] [Google Scholar]
198. Griffith DP, Gleeson MJ, Lee H, Longuet R, Deman E, Earle N. Randomized, double-blind trial of Lithostat (acetohydroxamic acid) in the palliative treatment of infection-induced urinary calculi. Eur Urol. 1991;20:243–247. [Abstract] [Google Scholar]
199. Williams JJ, Rodman JS, Peterson CM. A randomized double-blind study of acetohydroxamic acid in struvite nephrolithiasis. N Engl J Med. 1984;311:760–764. [Abstract] [Google Scholar]
200. Griffith DP, Khonsari F, Skurnick JH, James KE. A randomized trial of acetohydroxamic acid for the treatment and prevention of infection-induced urinary stones in spinal cord injury patients. J Urol. 1988;140:318–324. [Abstract] [Google Scholar]
201. Zisman AL. Effectiveness of Treatment Modalities on Kidney Stone Recurrence. Clinical Journal of the American Society of Nephrology. 2017 10.2215/cjn.11201016. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
202. Nicholson EB, Concaugh EA, Mobley HL. Proteus mirabilis urease: use of a ureA-lacZ fusion demonstrates that induction is highly specific for urea. Infect Immun. 1991;59:3360–3365. [Europe PMC free article] [Abstract] [Google Scholar]
203. Kanehisa M, Furumichi M, Tanabe M, Sato Y, Morishima K. KEGG: new perspectives on genomes, pathways, diseases and drugs. Nucleic Acids Res. 2017;45:D353–D361. [Europe PMC free article] [Abstract] [Google Scholar]
204. Senior BW. Investigation of the types and characteristics of the proteolytic enzymes formed by diverse strains of Proteus species. J Med Microbiol. 1999;48:623–628. [Abstract] [Google Scholar]
205. Henderson IR, Navarro-Garcia F, Desvaux M, Fernandez RC, Ala’Aldeen D. Type V protein secretion pathway: the autotransporter story. Microbiol Mol Biol Rev. 2004;68:692–744. [Europe PMC free article] [Abstract] [Google Scholar]
206. Leyton DL, Rossiter AE, Henderson IR. From self sufficiency to dependence: mechanisms and factors important for autotransporter biogenesis. Nat Rev Microbiol. 2012;10:213–225. [Abstract] [Google Scholar]
207. Wimmer MR, Woods CN, Adamczak KJ, Glasgow EM, Novak WR, Grilley DP, Weaver TM. Sequential unfolding of the hemolysin two-partner secretion domain from Proteus mirabilis. Protein Sci. 2015;24:1841–1855. [Europe PMC free article] [Abstract] [Google Scholar]
208. Weaver TM, Smith JA, Hocking JM, Bailey LJ, Wawrzyn GT, Howard DR, Sikkink LA, Ramirez-Alvarado M, Thompson JR. Structural and functional studies of truncated hemolysin A from Proteus mirabilis. J Biol Chem. 2009;284:22297–22309. [Europe PMC free article] [Abstract] [Google Scholar]
209. Uphoff TS, Welch RA. Nucleotide sequencing of the Proteus mirabilis calcium-independent hemolysin genes (hpmA and hpmB) reveals sequence similarity with the Serratia marcescens hemolysin genes (shlA and shlB) J Bacteriol. 1990;172:1206–1216. [Europe PMC free article] [Abstract] [Google Scholar]
210. Dienes L. Reproductive processes in Proteus cultures. Proc Soc Exp Biol Med. 1946;63:265–270. [Abstract] [Google Scholar]
211. Budding AE, Ingham CJ, Bitter W, Vandenbroucke-Grauls CM, Schneeberger PM. The Dienes Phenomenon: Competition and Territoriality in Swarming Proteus mirabilis. J Bacteriol. 2009;191:3892–3900. [Europe PMC free article] [Abstract] [Google Scholar]
212. Gibbs KA, Wenren LM, Greenberg EP. Identity gene expression in Proteus mirabilis. J Bacteriol. 2011 10.1128/jb.01167-10. JB.01167-01110. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
213. Braun V, Focareta T. Pore-forming bacterial protein hemolysins (cytolysins) Crit Rev Microbiol. 1991;18:115–158. [Abstract] [Google Scholar]
214. Welch RA. Identification of two different hemolysin determinants in uropathogenic Proteus isolates. Infect Immun. 1987;55:2183–2190. [Europe PMC free article] [Abstract] [Google Scholar]
215. Senior BW, Albrechtsen M, Kerr MA. Proteus mirabilis strains of diverse type have IgA protease activity. J Med Microbiol. 1987;24:175–180. [Abstract] [Google Scholar]
216. Senior BW, Albrechtsen M, Kerr MA. A survey of IgA protease production among clinical isolates of Proteeae. J Med Microbiol. 1988;25:27–31. [Abstract] [Google Scholar]
217. Loomes LM, Senior BW, Kerr MA. A proteolytic enzyme secreted by Proteus mirabilis degrades immunoglobulins of the immunoglobulin A1 (IgA1), IgA2, and IgG isotypes. Infect Immun. 1990;58:1979–1985. [Europe PMC free article] [Abstract] [Google Scholar]
218. Almogren A, Senior BW, Loomes LM, Kerr MA. Structural and functional consequences of cleavage of human secretory and human serum immunoglobulin A1 by proteinases from Proteus mirabilis and Neisseria meningitidis. Infect Immun. 2003;71:3349–3356. [Europe PMC free article] [Abstract] [Google Scholar]
219. Belas R, Manos J, Suvanasuthi R. Proteus mirabilis ZapA Metalloprotease Degrades a Broad Spectrum of Substrates, Including Antimicrobial Peptides. Infect Immun. 2004;72:5159–5167. [Europe PMC free article] [Abstract] [Google Scholar]
220. Wassif C, Cheek D, Belas R. Molecular analysis of a metalloprotease from Proteus mirabilis. J Bacteriol. 1995;177:5790–5798. [Europe PMC free article] [Abstract] [Google Scholar]
221. Inouye SK, van Dyck CH, Alessi CA, Balkin S, Siegal AP, Horwitz RI. Clarifying confusion: the confusion assessment method. A new method for detection of delirium. Ann Intern Med. 1990;113:941–948. [Abstract] [Google Scholar]
222. Hacker J, Knapp S, Goebel W. Spontaneous deletions and flanking regions of the chromosomally inherited hemolysin determinant of an Escherichia coli O6 strain. Journal of Bacteriology. 1983;154:1145–1152. [Europe PMC free article] [Abstract] [Google Scholar]
223. Knapp S, Then I, Wels W, Michel G, Tschape H, Hacker J, Goebel W. Analysis of the flanking regions from different haemolysin determinants of Escherichia coli. Mol Gen Genet. 1985;200:385–392. [Abstract] [Google Scholar]
224. Burrus V, Waldor MK. Shaping bacterial genomes with integrative and conjugative elements. Res Microbiol. 2004;155:376–386. [Abstract] [Google Scholar]
225. Seth-Smith H, Croucher NJ. Genome watch: breaking the ICE. Nat Rev Microbiol. 2009;7:328–329. [Abstract] [Google Scholar]
226. Hospenthal MK, Costa TRD, Waksman G. A comprehensive guide to pilus biogenesis in Gram-negative bacteria. Nat Rev Microbiol. 2017;15:365–379. [Abstract] [Google Scholar]
227. Thanassi DG, Nuccio SP, Shu Kin So S, Baumler AJ. Fimbriae: Classification and Biochemistry. EcoSal Plus. 2007;2 [Abstract] [Google Scholar]
228. Nuccio SP, Baumler AJ. Evolution of the chaperone/usher assembly pathway: fimbrial classification goes Greek. Microbiol Mol Biol Rev. 2007;71:551–575. [Europe PMC free article] [Abstract] [Google Scholar]
229. Old DC, Adegbola RA. Haemagglutinins and fimbriae of Morganella, Proteus and Providencia. J Med Microbiol. 1982;15:551–564. [Abstract] [Google Scholar]
230. Adegbola RA, Old DC, Senior BW. The adhesins and fimbriae of Proteus mirabilis strains associated with high and low affinity for the urinary tract. J Med Microbiol. 1983;16:427–431. [Abstract] [Google Scholar]
231. Snyder JA, Haugen BJ, Lockatell CV, Maroncle N, Hagan EC, Johnson DE, Welch RA, Mobley HL. Coordinate expression of fimbriae in uropathogenic Escherichia coli. Infect Immun. 2005;73:7588–7596. [Europe PMC free article] [Abstract] [Google Scholar]
232. Rocha SPD, Pelayo JS, Elias WP. Fimbriae of uropathogenic Proteus mirabilis. FEMS Immunology & Medical Microbiology. 2007;51:1–7. [Abstract] [Google Scholar]
233. Eden CS, Larsson P, Lomberg H. Attachment of Proteus mirabilis to human urinary sediment epithelial cells in vitro is different from that of Escherichia coli. Infect Immun. 1980;27:804–807. [Europe PMC free article] [Abstract] [Google Scholar]
234. Massad G, Bahrani FK, Mobley HL. Proteus mirabilis fimbriae: identification, isolation, and characterization of a new ambient-temperature fimbria. Infect Immun. 1994;62:1989–1994. [Europe PMC free article] [Abstract] [Google Scholar]
235. Li X, Mobley HLT. Vaccines for Proteus mirabilis in urinary tract infection. International Journal of Antimicrobial Agents. 2002;19:461–465. [Abstract] [Google Scholar]
236. Pellegrino R, Galvalisi U, Scavone P, Sosa V, Zunino P. Evaluation of Proteus mirabilis structural fimbrial proteins as antigens against urinary tract infections. FEMS Immunology & Medical Microbiology. 2003;36:103–110. [Abstract] [Google Scholar]
237. Li X, Lockatell CV, Johnson DE, Lane MC, Warren JW, Mobley HLT. Development of an Intranasal Vaccine To Prevent Urinary Tract Infection by Proteus mirabilis. Infect Immun. 2004;72:66–75. [Europe PMC free article] [Abstract] [Google Scholar]
238. Li X, Erbe JL, Lockatell CV, Johnson DE, Jobling MG, Holmes RK, Mobley HLT. Use of Translational Fusion of the MrpH Fimbrial Adhesin-Binding Domain with the Cholera Toxin A2 Domain, Coexpressed with the Cholera Toxin B Subunit, as an Intranasal Vaccine To Prevent Experimental Urinary Tract Infection by Proteus mirabilis. Infect Immun. 2004;72:7306–7310. [Europe PMC free article] [Abstract] [Google Scholar]
239. Scavone P, Sosa V, Pellegrino R, Galvalisi U, Zunino P. Mucosal vaccination of mice with recombinant Proteus mirabilis structural fimbrial proteins. Microbes and Infection. 2004;6:853–860. [Abstract] [Google Scholar]
240. Sareneva T, Holthofer H, Korhonen TK. Tissue-binding affinity of Proteus mirabilis fimbriae in the human urinary tract. Infect Immun. 1990;58:3330–3336. [Europe PMC free article] [Abstract] [Google Scholar]
241. Zhao H, Li X, Johnson DE, Blomfield I, Mobley HL. In vivo phase variation of MR/P fimbrial gene expression in Proteus mirabilis infecting the urinary tract. Mol Microbiol. 1997;23:1009–1019. [Abstract] [Google Scholar]
242. Li X, Mobley HL. MrpB functions as the terminator for assembly of Proteus mirabilis mannose-resistant Proteus-like fimbriae. Infect Immun. 1998;66:1759–1763. [Europe PMC free article] [Abstract] [Google Scholar]
243. Johnson DE, Bahrani FK, Lockatell CV, Drachenberg CB, Hebel JR, Belas R, Warren JW, Mobley HLT. Serum Immunoglobulin Response and Protection from Homologous Challenge by Proteus mirabilis in a Mouse Model of Ascending Urinary Tract Infection. Infect Immun. 1999;67:6683–6687. [Europe PMC free article] [Abstract] [Google Scholar]
244. Jansen AM, Lockatell V, Johnson DE, Mobley HLT. Mannose-Resistant Proteus-Like Fimbriae Are Produced by Most Proteus mirabilis Strains Infecting the Urinary Tract, Dictate the In Vivo Localization of Bacteria, and Contribute to Biofilm Formation. Infect Immun. 2004;72:7294–7305. [Europe PMC free article] [Abstract] [Google Scholar]
245. Scavone P, Villar S, Umpierrez A, Zunino P. Role of Proteus mirabilis MR/P fimbriae and flagella in adhesion, cytotoxicity and genotoxicity induction in T24 and Vero cells. Pathog Dis. 2015;73 [Abstract] [Google Scholar]
246. Wray SK, Hull SI, Cook RG, Barrish J, Hull RA. Identification and characterization of a uroepithelial cell adhesin from a uropathogenic isolate of Proteus mirabilis. Infect Immun. 1986;54:43–49. [Europe PMC free article] [Abstract] [Google Scholar]
247. Cook SW, Mody N, Valle J, Hull R. Molecular cloning of Proteus mirabilis uroepithelial cell adherence (uca) genes. Infect Immun. 1995;63:2082–2086. [Europe PMC free article] [Abstract] [Google Scholar]
248. Tolson DL, Barrigar DL, McLean RJ, Altman E. Expression of a nonagglutinating fimbria by Proteus mirabilis. Infect Immun. 1995;63:1127–1129. [Europe PMC free article] [Abstract] [Google Scholar]
249. Latta RK, Schur MJ, Tolson DL, Altman E. The effect of growth conditions on in vitro adherence, invasion, and NAF expression by Proteus mirabilis 7570. Can J Microbiol. 1998;44:896–904. [Abstract] [Google Scholar]
250. Lee KK, Harrison BA, Latta R, Altman E. The binding of Proteus mirabilis nonagglutinating fimbriae to ganglio-series asialoglycolipids and lactosyl ceramide. Can J Microbiol. 2000;46:961–966. [Abstract] [Google Scholar]
251. Altman E, Harrison BA, Latta RK, Lee KK, Kelly JF, Thibault P. Galectin-3-mediated adherence of Proteus mirabilis to Madin-Darby canine kidney cells. Biochem Cell Biol. 2001;79:783–788. [Abstract] [Google Scholar]
252. Bahrani FK, Cook S, Hull RA, Massad G, Mobley HL. Proteus mirabilis fimbriae: N-terminal amino acid sequence of a major fimbrial subunit and nucleotide sequences of the genes from two strains. Infect Immun. 1993;61:884–891. [Europe PMC free article] [Abstract] [Google Scholar]
253. Bijlsma IG, van Dijk L, Kusters JG, Gaastra W. Nucleotide sequences of two fimbrial major subunit genes, pmpA and ucaA, from canine-uropathogenic Proteus mirabilis strains. Microbiology. 1995;141(Pt 6):1349–1357. [Abstract] [Google Scholar]
254. Duguid JP. Fimbriae and adhesive properties in Klebsiella strains. J Gen Microbiol. 1959;21:271–286. [Abstract] [Google Scholar]
255. Bahrani FK, Mobley HL. Proteus mirabilis MR/P fimbriae: molecular cloning, expression, and nucleotide sequence of the major fimbrial subunit gene. J Bacteriol. 1993;175:457–464. [Europe PMC free article] [Abstract] [Google Scholar]
256. Scavone P, Iribarnegaray V, Caetano AL, Schlapp G, Hartel S, Zunino P. Fimbriae have distinguishable roles in Proteus mirabilis biofilm formation. Pathog Dis. 2016;74 [Abstract] [Google Scholar]
257. Canesin G, Gonzalez-Peramato P, Palou J, Urrutia M, Cordon-Cardo C, Sanchez-Carbayo M. Galectin-3 expression is associated with bladder cancer progression and clinical outcome. Tumour Biol. 2010;31:277–285. [Abstract] [Google Scholar]
258. Massad G, Mobley HL. Genetic organization and complete sequence of the Proteus mirabilis pmf fimbrial operon. Gene. 1994;150:101–104. [Abstract] [Google Scholar]
259. Massad G, Fulkerson JF, Jr, Watson DC, Mobley HL. Proteus mirabilis ambient-temperature fimbriae: cloning and nucleotide sequence of the aft gene cluster. Infect Immun. 1996;64:4390–4395. [Europe PMC free article] [Abstract] [Google Scholar]
260. Tsai YL, Chien HF, Huang KT, Lin WY, Liaw SJ. cAMP receptor protein regulates mouse colonization, motility, fimbria-mediated adhesion, and stress tolerance in uropathogenic Proteus mirabilis. Sci Rep. 2017;7:7282. [Europe PMC free article] [Abstract] [Google Scholar]
261. Yakubu DE, Old DC, Senior BW. The haemagglutinins and fimbriae of Proteus penneri. J Med Microbiol. 1989;30:279–284. [Abstract] [Google Scholar]
262. Mobley HL, Chippendale GR, Tenney JH, Mayrer AR, Crisp LJ, Penner JL, Warren JW. MR/K hemagglutination of Providencia stuartii correlates with adherence to catheters and with persistence in catheter-associated bacteriuria. J Infect Dis. 1988;157:264–271. [Abstract] [Google Scholar]
263. Collado-Vides J, Magasanik B, Gralla JD. Control site location and transcriptional regulation in Escherichia coli. Microbiol Rev. 1991;55:371–394. [Europe PMC free article] [Abstract] [Google Scholar]
264. Meslet-Cladiere LM, Pimenta A, Duchaud E, Holland IB, Blight MA. In vivo expression of the mannose-resistant fimbriae of Photorhabdus temperata K122 during insect infection. J Bacteriol. 2004;186:611–622. [Europe PMC free article] [Abstract] [Google Scholar]
265. He H, Snyder HA, Forst S. Unique organization and regulation of the mrx fimbrial operon in Xenorhabdus nematophila. Microbiology. 2004;150:1439–1446. [Abstract] [Google Scholar]
266. Chen YT, Peng HL, Shia WC, Hsu FR, Ken CF, Tsao YM, Chen CH, Liu CE, Hsieh MF, Chen HC, Tang CY, Ku TH. Whole-genome sequencing and identification of Morganella morganii KT pathogenicity-related genes. BMC Genomics. 2012;13(Suppl 7):S4. [Europe PMC free article] [Abstract] [Google Scholar]
267. Bode NJ, Chan KW, Kong XP, Pearson MM. Distinct Residues Contribute to Motility Repression and Autoregulation in the Proteus mirabilis Fimbria-Associated Transcriptional Regulator AtfJ. J Bacteriol. 2016;198:2100–2112. [Europe PMC free article] [Abstract] [Google Scholar]
268. Morris NS, Stickler DJ, McLean RJ. The development of bacterial biofilms on indwelling urethral catheters. World J Urol. 1999;17:345–350. [Abstract] [Google Scholar]
269. Stickler D, Young R, Jones G, Sabbuba N, Morris N. Why are Foley catheters so vulnerable to encrustation and blockage by crystalline bacterial biofilm? Urol Res. 2003;31:306–311. [Abstract] [Google Scholar]
270. Stickler DJ. Clinical complications of urinary catheters caused by crystalline biofilms: something needs to be done. J Intern Med. 2014;276:120–129. [Abstract] [Google Scholar]
271. Davey ME, O’Toole GA. Microbial biofilms: from ecology to molecular genetics. Microbiol Mol Biol Rev. 2000;64:847–867. [Europe PMC free article] [Abstract] [Google Scholar]
272. Jones BV, Young R, Mahenthiralingam E, Stickler DJ. Ultrastructure of Proteus mirabilis Swarmer Cell Rafts and Role of Swarming in Catheter-Associated Urinary Tract Infection. Infect Immun. 2004;72:3941–3950. [Europe PMC free article] [Abstract] [Google Scholar]
273. Jones SM, Yerly J, Hu Y, Ceri H, Martinuzzi R. Structure of Proteus mirabilis biofilms grown in artificial urine and standard laboratory media. FEMS Microbiol Lett. 2007;268:16–21. [Abstract] [Google Scholar]
274. Jones BV, Mahenthiralingam E, Sabbuba NA, Stickler DJ. Role of swarming in the formation of crystalline Proteus mirabilis biofilms on urinary catheters. J Med Microbiol. 2005;54:807–813. [Abstract] [Google Scholar]
275. Rocha SPD, Elias WP, Cianciarullo AM, Menezes MA, Nara JM, Piazza RMF, Silva MRL, Moreira CG, Pelayo JS. Aggregative adherence of uropathogenic Proteus mirabilis to cultured epithelial cells. FEMS Immunology & Medical Microbiology. 2007;51:319–326. [Abstract] [Google Scholar]
276. Sabbuba NA, Stickler DJ, Mahenthiralingam E, Painter DJ, Parkin J, Feneley RC. Genotyping demonstrates that the strains of Proteus mirabilis from bladder stones and catheter encrustations of patients undergoing long-term bladder catheterization are identical. J Urol. 2004;171:1925–1928. [Abstract] [Google Scholar]
277. Winters C, Stickler DJ, Howe TJ, Wilkinson N, Buckley CJ. Some observations on the structure of encrusting biofilms of Proteus mirabilis on urethral catheters. Cells and Materials. 1995;5:245–253. [Google Scholar]
278. Jacobsen SM, Shirtliff ME. Proteus mirabilis biofilms and catheter-associated urinary tract infections. Virulence. 2011;2:460–465. [Abstract] [Google Scholar]
279. Liaw S-J, Lai H-C, Wang W-B. Modulation of Swarming and Virulence by Fatty Acids through the RsbA Protein in Proteus mirabilis. Infect Immun. 2004;72:6836–6845. [Europe PMC free article] [Abstract] [Google Scholar]
280. Izquierdo L, Abitiu N, Coderch N, Hita B, Merino S, Gavin R, Tomas JM, Regue M. The inner-core lipopolysaccharide biosynthetic waaE gene: function and genetic distribution among some Enterobacteriaceae. Microbiology. 2002;148:3485–3496. [Abstract] [Google Scholar]
281. Czerwonka G, Guzy A, Kaluza K, Grosicka M, Danczuk M, Lechowicz L, Gmiter D, Kowalczyk P, Kaca W. The role of Proteus mirabilis cell wall features in biofilm formation. Arch Microbiol. 2016;198:877–884. [Europe PMC free article] [Abstract] [Google Scholar]
282. Dumanski AJ, Hedelin H, Edin-Liljegren A, Beauchemin D, McLean RJ. Unique ability of the Proteus mirabilis capsule to enhance mineral growth in infectious urinary calculi. Infect Immun. 1994;62:2998–3003. [Europe PMC free article] [Abstract] [Google Scholar]
283. Wilks SA, Fader MJ, Keevil CW. Novel Insights into the Proteus mirabilis Crystalline Biofilm Using Real-Time Imaging. PLoS One. 2015;10:e0141711. [Europe PMC free article] [Abstract] [Google Scholar]
284. Moryl M, Kaleta A, Strzelecki K, Rozalska S, Rozalski A. Effect of nutrient and stress factors on polysaccharides synthesis in Proteus mirabilis biofilm. Acta Biochim Pol. 2014;61:133–139. [Abstract] [Google Scholar]
285. Hamill TM, Gilmore BF, Jones DS, Gorman SP. Strategies for the development of the urinary catheter. Expert Rev Med Devices. 2007;4:215–225. [Abstract] [Google Scholar]
286. Soto SM. Importance of Biofilms in Urinary Tract Infections: New Therapeutic Approaches. Advances in Biology. 2014;2014:13. [Google Scholar]
287. Levering V, Wang Q, Shivapooja P, Zhao X, Lopez GP. Soft robotic concepts in catheter design: an on-demand fouling-release urinary catheter. Adv Healthc Mater. 2014;3:1588–1596. [Europe PMC free article] [Abstract] [Google Scholar]
288. Morgan SD, Rigby D, Stickler DJ. A study of the structure of the crystalline bacterial biofilms that can encrust and block silver Foley catheters. Urol Res. 2009;37:89–93. [Abstract] [Google Scholar]
289. Stickler DJ, Jones GL, Russell AD. Control of encrustation and blockage of Foley catheters. Lancet. 2003;361:1435–1437. [Abstract] [Google Scholar]
290. Jones GL, Muller CT, O’Reilly M, Stickler DJ. Effect of triclosan on the development of bacterial biofilms by urinary tract pathogens on urinary catheters. Journal of Antimicrobial Chemotherapy. 2006;57:266–272. [Abstract] [Google Scholar]
291. Stickler DJ, Jones GL. Reduced Susceptibility of Proteus mirabilis to triclosan. Antimicrob Agents Chemother. 2008;52:991–994. [Europe PMC free article] [Abstract] [Google Scholar]
292. Azeredo J, Azevedo NF, Briandet R, Cerca N, Coenye T, Costa AR, Desvaux M, Di Bonaventura G, Hebraud M, Jaglic Z, Kacaniova M, Knochel S, Lourenco A, Mergulhao F, Meyer RL, Nychas G, Simoes M, Tresse O, Sternberg C. Critical review on biofilm methods. Crit Rev Microbiol. 2017;43:313–351. [Abstract] [Google Scholar]
293. Hola V, Peroutkova T, Ruzicka F. Virulence factors in Proteus bacteria from biofilm communities of catheter-associated urinary tract infections. FEMS Immunol Med Microbiol. 2012;65:343–349. [Abstract] [Google Scholar]
294. Kwiecinska-Pirog J, Bogiel T, Skowron K, Wieckowska E, Gospodarek E. Proteus mirabilis biofilm - qualitative and quantitative colorimetric methods-based evaluation. Braz J Microbiol. 2014;45:1423–1431. [Europe PMC free article] [Abstract] [Google Scholar]
295. Schlapp G, Scavone P, Zunino P, Hartel S. Development of 3D architecture of uropathogenic Proteus mirabilis batch culture biofilms-A quantitative confocal microscopy approach. J Microbiol Methods. 2011;87:234–240. [Abstract] [Google Scholar]
296. Czerwonka G, Arabski M, Wąsik S, Jabłońska-Wawrzycka A, Rogala P, Kaca W. Morphological changes in Proteus mirabilis O18 biofilm under the influence of a urease inhibitor and a homoserine lactone derivative. Archives of Microbiology. 2014;196:169–177. [Europe PMC free article] [Abstract] [Google Scholar]
297. O’May GA, Jacobsen SM, Longwell M, Stoodley P, Mobley HLT, Shirtliff ME. The high-affinity phosphate transporter Pst in Proteus mirabilis HI4320 and its importance in biofilm formation. Microbiology. 2009;155:1523–1535. [Europe PMC free article] [Abstract] [Google Scholar]
298. Stickler D, Hughes G. Ability of Proteus mirabilis to swarm over urethral catheters. Eur J Clin Microbiol Infect Dis. 1999;18:206–208. [Abstract] [Google Scholar]
299. Brooks T, Keevil CW. A simple artificial urine for the growth of urinary pathogens. Lett Appl Microbiol. 1997;24:203–206. [Abstract] [Google Scholar]
300. Macleod SM, Stickler DJ. Species interactions in mixed-community crystalline biofilms on urinary catheters. J Med Microbiol. 2007;56:1549–1557. [Abstract] [Google Scholar]
301. Moryl M, Torzewska A, Jalmuzna P, Rozalski A. Analysis of Proteus mirabilis distribution in multi-species biofilms on urinary catheters and determination of bacteria resistance to antimicrobial agents. Pol J Microbiol. 2013;62:377–384. [Abstract] [Google Scholar]
302. Hola V, Ruzicka F, Horka M. Microbial diversity in biofilm infections of the urinary tract with the use of sonication techniques. FEMS Immunol Med Microbiol. 2010;59:525–528. [Abstract] [Google Scholar]
303. Li X, Lu N, Brady HR, Packman AI. Biomineralization strongly modulates the formation of Proteus mirabilis and Pseudomonas aeruginosa dual-species biofilms FEMS. Microbiol Ecol. 2016 10.1093/femsec/fiw189. [Abstract] [CrossRef] [Google Scholar]
304. Williams Gareth JP, Stickler David JD. Some Observations on the Migration of Proteus mirabilis and Other Urinary Tract Pathogens Over Foley Catheters. Infection Control and Hospital Epidemiology. 2008;29:443–445. [Abstract] [Google Scholar]
305. Galvan EM, Mateyca C, Ielpi L. Role of interspecies interactions in dual-species biofilms developed in vitro by uropathogens isolated from polymicrobial urinary catheter-associated bacteriuria. Biofouling. 2016;32:1067–1077. [Abstract] [Google Scholar]
306. Stickler D, Morris N, Moreno MC, Sabbuba N. Studies on the formation of crystalline bacterial biofilms on urethral catheters. Eur J Clin Microbiol Infect Dis. 1998;17:649–652. [Abstract] [Google Scholar]
307. Liaw S-J, Lai H-C, Ho S-W, Luh K-T, Wang W-B. Role of RsmA in the regulation of swarming motility and virulence factor expression in Proteus mirabilis. Journal of Medical Microbiology. 2003;52:19–28. [Abstract] [Google Scholar]
308. Peng L, Jiang Q, Pan JY, Deng C, Yu JY, Wu XM, Huang SH, Deng XY. Involvement of polyphosphate kinase in virulence and stress tolerance of uropathogenic Proteus mirabilis. Med Microbiol Immunol. 2016;205:97–109. [Europe PMC free article] [Abstract] [Google Scholar]
309. Hoeniger JFM. Development of Flagella by Proteus mirabilis. J Gen Microbiol. 1965;40:29–42. [Google Scholar]
310. Williams FD, Anderson DM, Hoffman PS, Schwarzhoff RH, Leonard S. Evidence against the involvement of chemotaxis in swarming of Proteus mirabilis. J Bacteriol. 1976;127:237–248. [Europe PMC free article] [Abstract] [Google Scholar]
311. Harshey RM, Partridge JD. Shelter in a Swarm. Journal of Molecular Biology. 2015;427:3683–3694. [Europe PMC free article] [Abstract] [Google Scholar]
312. Belas R, Flaherty D. Sequence and genetic analysis of multiple flagellin-encoding genes from Proteus mirabilis. Gene. 1994;148:33–41. [Abstract] [Google Scholar]
313. Belas R. Expression of multiple flagellin-encoding genes of Proteus mirabilis. J Bacteriol. 1994;176:7169–7181. [Europe PMC free article] [Abstract] [Google Scholar]
314. Manos J, Belas R. Transcription of Proteus mirabilis flaAB. Microbiology. 2004;150:2857–2863. [Abstract] [Google Scholar]
315. Zunino P, Piccini C, Legnani-Fajardo C. Flagellate and non-flagellate Proteus mirabilis in the development of experimental urinary tract infection. Microbial Pathogenesis. 1994;16:379–385. [Abstract] [Google Scholar]
316. Sabbuba N, Hughes G, Stickler DJ. The migration of Proteus mirabilis and other urinary tract pathogens over Foley catheters. BJU Int. 2002;89:55–60. [Abstract] [Google Scholar]
317. Armbruster CE, Hodges SA, Mobley HL. Initiation of swarming motility by Proteus mirabilis occurs in response to specific cues present in urine and requires excess L-glutamine. J Bacteriol. 2013;195:1305–1319. [Europe PMC free article] [Abstract] [Google Scholar]
318. Allison C, Lai HC, Gygi D, Hughes C. Cell differentiation of Proteus mirabilis is initiated by glutamine, a specific chemoattractant for swarming cells. Mol Microbiol. 1993;8:53–60. [Abstract] [Google Scholar]
319. Pazin GJ, Braude AI. Immobilizing antibodies in urine. II. Prevention of ascending spread of Proteus mirabilis. Invest Urol. 1974;12:129–133. [Abstract] [Google Scholar]
320. Allison C, Emödy L, Coleman N, Hughes C. The Role of Swarm Cell Differentiation and Multicellular Migration in the Uropathogenicity of Proteus mirabilis. The Journal of Infectious Diseases. 1994;169:1155–1158. [Abstract] [Google Scholar]
321. Jansen AM, Lockatell CV, Johnson DE, Mobley HLT. Visualization of Proteus mirabilis Morphotypes in the Urinary Tract: the Elongated Swarmer Cell Is Rarely Observed in Ascending Urinary Tract Infection. Infect Immun. 2003;71:3607–3613. [Europe PMC free article] [Abstract] [Google Scholar]
322. Cassat JE, Skaar EP. Iron in infection and immunity. Cell Host Microbe. 2013;13:509–519. [Europe PMC free article] [Abstract] [Google Scholar]
323. Subashchandrabose S, Mobley HL. Back to the metal age: battle for metals at the host-pathogen interface during urinary tract infection. Metallomics. 2015;7:935–942. [Europe PMC free article] [Abstract] [Google Scholar]
324. Shand GH, Anwar H, Kadurugamuwa J, Brown MR, Silverman SH, Melling J. In vivo evidence that bacteria in urinary tract infection grow under iron-restricted conditions. Infect Immun. 1985;48:35–39. [Europe PMC free article] [Abstract] [Google Scholar]
325. Miles AA, Khimji PL. Enterobacterial chelators of iron: their occurrence, detection, and relation to pathogenicity. J Med Microbiol. 1975;8:477–490. [Abstract] [Google Scholar]
326. Evanylo LP, Kadis S, Maudsley JR. Siderophore production by Proteus mirabilis. Can J Microbiol. 1984;30:1046–1051. [Abstract] [Google Scholar]
327. Hart RC, Kadis S, Chapman WL., Jr Nutritional iron status and susceptibility to Proteus mirabilis pyelonephritis in the rat. Can J Microbiol. 1982;28:713–717. [Abstract] [Google Scholar]
328. Gaisser S, Hughes C. A locus coding for putative non-ribosomal peptide/polyketide synthase functions is mutated in a swarming-defective Proteus mirabilis strain. Mol Gen Genet. 1997;253:415–427. [Abstract] [Google Scholar]
329. Drechsel H, Thieken A, Reissbrodt R, Jung G, Winkelmann G. Alpha-keto acids are novel siderophores in the genera Proteus, Providencia, and Morganella and are produced by amino acid deaminases. J Bacteriol. 1993;175:2727–2733. [Europe PMC free article] [Abstract] [Google Scholar]
330. Massad G, Zhao H, Mobley HL. Proteus mirabilis amino acid deaminase: cloning, nucleotide sequence, and characterization of aad. J Bacteriol. 1995;177:5878–5883. [Europe PMC free article] [Abstract] [Google Scholar]
331. Piccini CD, Barbe FM, Legnani-Fajardo CL. Identification of iron-regulated outer membrane proteins in uropathogenic Proteus mirabilis and its relationship with heme uptake. FEMS Microbiol Lett. 1998;166:243–248. [Abstract] [Google Scholar]
332. Lima A, Zunino P, D’Alessandro B, Piccini C. An iron-regulated outer-membrane protein of Proteus mirabilis is a haem receptor that plays an important role in urinary tract infection and in in vivo growth. J Med Microbiol. 2007;56:1600–1607. [Abstract] [Google Scholar]
333. D’Alessandro B, Lery LM, Kruger WM, Lima A, Piccini C, Zunino P, Oswald E. Proteomic analysis of Proteus mirabilis outer membrane proteins reveals differential expression in vivo vs. in vitro conditions. FEMS Immunol Med Microbiol. 2011;63:174–182. [Abstract] [Google Scholar]
334. Pearson MM, Sebaihia M, Churcher C, Quail MA, Seshasayee AS, Luscombe NM, Abdellah Z, Arrosmith C, Atkin B, Chillingworth T, Hauser H, Jagels K, Moule S, Mungall K, Norbertczak H, Rabbinowitsch E, Walker D, Whithead S, Thomson NR, Rather PN, Parkhill J, Mobley HLT. Complete genome sequence of uropathogenic Proteus mirabilis, a master of both adherence and motility. Journal of Bacteriology. 2008;190:4027–4037. [Europe PMC free article] [Abstract] [Google Scholar]
335. Himpsl SD, Pearson MM, Arewång CJ, Nusca TD, Sherman DH, Mobley HLT. Proteobactin and a yersiniabactin-related siderophore mediate iron acquisition in Proteus mirabilis. Molecular Microbiology. 2010;78:138–157. [Europe PMC free article] [Abstract] [Google Scholar]
336. Nielubowicz GR, Smith SN, Mobley HLT. Outer membrane antigens of the uropathogen Proteus mirabilis recognized by the humoral response during experimental murine urinary tract infection. Infection and Immunity. 2008;76:4222–4231. [Europe PMC free article] [Abstract] [Google Scholar]
337. Pearson MM, Yep A, Smith SN, Mobley HLT. Transcriptome of Proteus mirabilis in the murine urinary tract: virulence and nitrogen assimilation gene expression. Infection and Immunity. 2011;79:2619–2631. [Europe PMC free article] [Abstract] [Google Scholar]
338. Lima A, Zunino P, D’Alessandro B, Piccini C. An iron-regulated outer-membrane protein of Proteus mirabilis is a haem receptor that plays an important role in urinary tract infection and in in vivo growth. Journal of Medical Microbiology. 2007;56:1600–1607. [Abstract] [Google Scholar]
339. Himpsl SD, Lockatell CV, Hebel JR, Johnson DE, Mobley HLT. Identification of virulence determinants in uropathogenic Proteus mirabilis using signature-tagged mutagenesis. Journal of Medical Microbiology. 2008;57:1068–1078. [Abstract] [Google Scholar]
340. Burall LS, Harro JM, Li X, Lockatell CV, Himpsl SD, Hebel JR, Johnson DE, Mobley HLT. Proteus mirabilis genes that contribute to pathogenesis of urinary tract infection: identification of 25 signature-tagged mutants attenuated at least 100-fold. Infection and Immunity. 2004;72:2922–2938. [Europe PMC free article] [Abstract] [Google Scholar]
341. Bode W, Gomis-Ruth FX, Stockler W. Astacins, serralysins, snake venom and matrix metalloproteinases exhibit identical zinc-binding environments (HEXXHXXGXXH and Met-turn) and topologies and should be grouped into a common family, the ‘metzincins’ FEBS Lett. 1993;331:134–140. [Abstract] [Google Scholar]
342. Wang S, Fleming RT, Westbrook EM, Matsumura P, McKay DB. Structure of the Escherichia coli FlhDC complex, a prokaryotic heteromeric regulator of transcription. J Mol Biol. 2006;355:798–808. [Abstract] [Google Scholar]
343. Lai H-C, Gygi D, Fraser GM, Hughes C. A swarming-defective mutant of Proteus mirabilis lacking a putative cation-transporting membrane P-type ATPase. Microbiology. 1998;144:1957–1961. [Abstract] [Google Scholar]
344. Nakashige TG, Zygiel EM, Drennan CL, Nolan EM. Nickel Sequestration by the Host-Defense Protein Human Calprotectin. J Am Chem Soc. 2017;139:8828–8836. [Europe PMC free article] [Abstract] [Google Scholar]
345. Lamarche MG, Wanner BL, Crepin S, Harel J. The phosphate regulon and bacterial virulence: a regulatory network connecting phosphate homeostasis and pathogenesis. FEMS Microbiol Rev. 2008;32:461–473. [Abstract] [Google Scholar]
346. Jacobsen SM, Lane MC, Harro JM, Shirtliff ME, Mobley HLT. The high-affinity phosphate transporter Pst is a virulence factor for Proteus mirabilis during complicated urinary tract infection. FEMS Immunology & Medical Microbiology. 2008;52:180–193. [Abstract] [Google Scholar]
347. Liu L, Mo H, Wei S, Raftery D. Quantitative analysis of urea in human urine and serum by (1)H nuclear magnetic resonance() The Analyst. 2012;137:595–600. [Europe PMC free article] [Abstract] [Google Scholar]
348. Shaykhutdinov RA, MacInnis GD, Dowlatabadi R, Weljie AM, Vogel HJ. Quantitative analysis of metabolite concentrations in human urine samples using 13C{1H} NMR spectroscopy. Metabolomics. 2009;5:307–317. [Google Scholar]
349. Bouatra S, Aziat F, Mandal R, Guo AC, Wilson MR, Knox C, Bjorndahl TC, Krishnamurthy R, Saleem F, Liu P, Dame ZT, Poelzer J, Huynh J, Yallou FS, Psychogios N, Dong E, Bogumil R, Roehring C, Wishart DS. The Human Urine Metabolome. PLOS ONE. 2013;8:e73076. [Europe PMC free article] [Abstract] [Google Scholar]
350. Simpson DP. Citrate excretion: a window on renal metabolism. Am J Physiol. 1983;244:F223–234. [Abstract] [Google Scholar]
351. Hess B. Urinary Citrate and Citrate Metabolism. In: Rao NP, Preminger GM, Kavanagh JP, editors. Urinary Tract Stone Disease. Springer; London, London: 2011. pp. 181–184. [CrossRef] [Google Scholar]
352. Huang Y, Nishikawa T, Satoh K, Iwata T, Fukushima T, Santa T, Homma H, Imai K. Urinary excretion of D-serine in human: comparison of different ages and species. Biol Pharm Bull. 1998;21:156–162. [Abstract] [Google Scholar]
353. Roesch PL, Redford P, Batchelet S, Moritz RL, Pellett S, Haugen BJ, Blattner FR, Welch RA. Uropathogenic Escherichia coli use d-serine deaminase to modulate infection of the murine urinary tract. Mol Microbiol. 2003;49:55–67. [Abstract] [Google Scholar]
354. Helling RB. Why does Escherichia coli have two primary pathways for synthesis of glutamate? Journal of Bacteriology. 1994;176:4664–4668. [Europe PMC free article] [Abstract] [Google Scholar]
355. Scavone P, Miyoshi A, Rial A, Chabalgoity A, Langella P, Azevedo V, Zunino P. Intranasal immunisation with recombinant Lactococcus lactis displaying either anchored or secreted forms of Proteus mirabilis MrpA fimbrial protein confers specific immune response and induces a significant reduction of kidney bacterial colonisation in mice. Microbes Infect. 2007;9:821–828. [Abstract] [Google Scholar]
356. Scavone P, Rial A, Umpierrez A, Chabalgoity A, Zunino P. Effects of the administration of cholera toxin as a mucosal adjuvant on the immune and protective response induced by Proteus mirabilis MrpA fimbrial protein in the urinary tract. Microbiol Immunol. 2009;53:233–240. [Abstract] [Google Scholar]
357. Scavone P, Umpierrez A, Maskell DJ, Zunino P. Nasal immunization with attenuated Salmonella Typhimurium expressing an MrpA-TetC fusion protein significantly reduces Proteus mirabilis colonization in the mouse urinary tract. J Med Microbiol. 2011;60:899–904. [Abstract] [Google Scholar]
358. Scavone P, Umpierrez A, Rial A, Chabalgoity JA, Zunino P. Native flagellin does not protect mice against an experimental Proteus mirabilis ascending urinary tract infection and neutralizes the protective effect of MrpA fimbrial protein. Antonie Van Leeuwenhoek. 2014;105:1139–1148. [Abstract] [Google Scholar]
359. Habibi M, Asadi Karam MR, Bouzari S. Transurethral instillation with fusion protein MrpH.FimH induces protective innate immune responses against uropathogenic Escherichia coli and Proteus mirabilis. APMIS. 2016;124:444–452. [Abstract] [Google Scholar]
360. Habibi M, Asadi Karam MR, Shokrgozar MA, Oloomi M, Jafari A, Bouzari S. Intranasal immunization with fusion protein MrpH.FimH and MPL adjuvant confers protection against urinary tract infections caused by uropathogenic Escherichia coli and Proteus mirabilis. Mol Immunol. 2015;64:285–294. [Abstract] [Google Scholar]
361. Uehling DT, Hopkins WJ, Beierle LM, Kryger JV, Heisey DM. Vaginal Mucosal Immunization for Recurrent Urinary Tract Infection: Extended Phase II Clinical Trial. The Journal of Infectious Diseases. 2001;183:S81–S83. [Abstract] [Google Scholar]
362. Hopkins WJ, Elkahwaji J, Beierle LM, Leverson GE, Uehling DT. Vaginal mucosal vaccine for recurrent urinary tract infections in women: results of a phase 2 clinical trial. J Urol. 2007;177:1349–1353. ; quiz 1591. [Abstract] [Google Scholar]
363. Kochiashvili D, Khuskivadze A, Kochiashvili G, Koberidze G, Kvakhajelidze V. Role of the bacterial vaccine Solco-Urovac(R) in treatment and prevention of recurrent urinary tract infections of bacterial origin. Georgian Med News. 2014:11–16. [Abstract] [Google Scholar]
364. Mathur S, Suller MT, Stickler DJ, Feneley RC. Prospective study of individuals with long-term urinary catheters colonized with Proteus species. BJU Int. 2006;97:121–128. [Abstract] [Google Scholar]
365. Chew R, Thomas S, Mantha ML, Killen JP, Cho Y, Baer RA. Large urate cystolith associated with Proteus urinary tract infection. Kidney Int. 2012;81:802. [Abstract] [Google Scholar]
366. Carver TJ, Rutherford KM, Berriman M, Rajandream MA, Barrell BG, Parkhill J. ACT: the Artemis Comparison Tool. Bioinformatics. 2005;21:3422–3423. [Abstract] [Google Scholar]
367. Stickler DJ. Bacterial biofilms in patients with indwelling urinary catheters. Nat Clin Pract Urol. 2008;5:598–608. [Abstract] [Google Scholar]

Citations & impact 


Impact metrics

Jump to Citations

Citations of article over time

Alternative metrics

Altmetric item for https://www.altmetric.com/details/33034726
Altmetric
Discover the attention surrounding your research
https://www.altmetric.com/details/33034726

Article citations


Go to all (121) article citations

Other citations

Lay summaries 


Funding 


Funders who supported this work.

NIAID NIH HHS (1)

NIDDK NIH HHS (2)