2,572
Views
67
CrossRef citations to date
0
Altmetric
Original Article

Omptin proteins: an expanding family of outer membrane proteases in Gram-negative Enterobacteriaceae (Review)

&
Pages 395-406 | Received 21 Aug 2006, Published online: 09 Jul 2009

Abstract

The Escherichia coli K-12 outer membrane protein OmpT is a prototype of a unique family of bacterial endopeptidases known as the omptins. This family includes OmpT and OmpP of E. coli, SopA of Shigella flexneri, PgtE of Salmonella enterica, and Pla of Yersinia pestis. Despite their sequence similarities, the omptins vary in their reported functions. The OmpT protease is characterized by narrow cleavage specificity defined by the extracellular loops of the β-barrel protruding above the lipid bilayer. It employs a distinct proteolytic mechanism that involves a histidine and an aspartate residue. Most of the omptin proteins have been implicated in bacterial pathogenesis. As a result, the omptins are potential targets for antimicrobial drug and vaccine development. This review summarizes recent developments in omptins structure and function, emphasizes their role in pathogenesis, proposes evolutionary relation among the existing omptins, and offers possible directions for future research.

Introduction

Omptins have been identified in several Gram-negative bacteria. These proteins are embedded in the bacterial outer membrane and folded in a conserved 10-stranded β-barrel structure Citation[1]. Their activities and specificities are determined by the long, extracellular loops that connect the strands of the β-barrel Citation[2]. Most of the known omptins are bacterial virulence factors and function as either proteases, adhesins, or invasins () Citation[3], Citation[4]. E. coli K12 expresses the temperature-regulated OmpT Citation[5] and its close homologue OmpP Citation[6]. Yersinia pestis, the causative agent of plague, possesses a highly virulent plasminogen activator Pla Citation[7]. PlaA, a close homologue of Pla, is found in the plant pathogen Erwiniapyrifoliae. Three similar forms of SopA (IcsP) are present in three oral-faecal pathogens: enteroinvasive E. coli, a major cause of diarrhea and death among children in the developing countries, Shigella flexneri and Shigella dysenteriae which cause bacterial dysentery Citation[8]. Salmonella enterica expresses the two homologous forms of PgtE (also referred to as the E protein) Citation[9].

Table I.  Characteristics of the proteins of the omptin family.

Thus, omptin proteins are present in a number of human and plant (e.g., Erwinia) bacterial pathogens. Many are important contributors to bacterial virulence. A decreased virulence of the bacterial pathogen was seen upon the loss of a gene coding for Pla omptin Citation[10–12]. Because of their pathogenic potency, the design of specific protease inhibitors against omptin proteases would offer promising leads for new antimicrobial therapeutics. Besides proteolysis, proteolytically-inactive omptins can carry functions such as adhesion and invasion Citation[13]. Therefore, vaccine development against omptins is another fascinating avenue of research. A recent x-ray crystal structure of one of the omptins Citation[1] offers an opportunity for structure-based drug design.

This review discusses the crystal structure of OmpT, which led to elucidation of its proteolytic mechanism and the characterization of its cleavage specificity. The review proceeds to the overview of diverse functional characteristics of the omptin family members. Then, the contribution of the omptins to the bacterial pathogenesis is provided. This review continues with the evolutionary analysis of the omptins. Finally, the review concludes by laying out the directions for future research.

Structural characteristics of OmpT, Pla, and PgtE

OmpT is the only member, whose structure has been elucidated Citation[1]. Its crystal structure, resolved at 2.6 Å, reveals a transmembrane β-barrel (A). The β-barrel structure is a conserved feature among the integral membrane proteins found in the outer membrane of the Gram-negative bacteria. This vase-shaped barrel consists of 10 antiparallel β-strands connected by four short periplasmic loops (T1–T4) with no known activity and five longer extracellular loops (L1–L5) Citation[1], Citation[14]. The extracellular loops (B) are responsible for protein's activity and specificity Citation[2]. The active site of OmpT is located in the grove formed by these extracellular loops (A). Like other transmembrane proteins, OmpT has two rings of aromatic amino acid residues at the water-lipid interface Citation[1]. The protein extends about 40 Å above the lipid bilayer with several extracellular loops extending just above the outer edge of LPS.

Figure 1.  Crystal structure and membrane topology of OmpT (1I78). Adapted and reproduced with permission from reference Citation[1]. (A) Front and side views of OmpT based on its crystal structure. (B) Topology model of OmpT. Amino acids in squares form β-strands. Periplasmic loops are labeled T1–T5. Loops exposed to the extracellular space are labeled L1–L5. The position of outer membrane is highlighted in gray. This Figure is reproduced in color in Molecular Membrane Biology online.

Figure 1.  Crystal structure and membrane topology of OmpT (1I78). Adapted and reproduced with permission from reference Citation[1]. (A) Front and side views of OmpT based on its crystal structure. (B) Topology model of OmpT. Amino acids in squares form β-strands. Periplasmic loops are labeled T1–T5. Loops exposed to the extracellular space are labeled L1–L5. The position of outer membrane is highlighted in gray. This Figure is reproduced in color in Molecular Membrane Biology online.

The association with LPS molecules is an absolute requirement for the proteolytic activities of OmpT, Pla, and PgtE Citation[15], Citation[16]. Activation by LPS might be a safety mechanism to ensure that the protease does not become active in the bacterial cytoplasm or periplasm, and is only active when inserted into the outer membrane Citation[15]. Circular dichroism (CD) studies Citation[15] indicate that OmpT-LPS interaction does not cause major changes in the protease structure. The interactions between Arg138, Arg175 and the 4′ phosphate of lipid A result in a proper alignment of the residues of the active site Citation[15]. To bring OmpT to its full enzymatic potential, the LPS molecule must have heptose-bound phosphates and a fully acylated lipid A Citation[14], Citation[15].

Recent research Citation[16], Citation[17] has indicated that LPS may have a dual function: it can activate and inactivate the omptin proteins. O-antigen in the outer core of LPS sterically inhibits Pla and PgtE proteins from cleaving larger substrates. However, OmpT targets short peptides which are small enough not to encounter steric hindrance from O-antigen. Indeed, Y. pestis, a host organism of Pla, does not produce O-antigen, due to mutations in its genes coding for the biosynthesis of O-antigen pathway Citation[16]. S. enterica, a host organism of PgtE, reduces the length of O-antigen and regulates the expression of pgtE gene during its intracellular growth in macrophages Citation[18].

The FhuA-LPS crystal structure, the only outer membrane protein structure crystallized with LPS, reveals an existence of the conserved LPS-binding motif present on the surface of FhuA Citation[19]. The comparison between FhuA and OmpT structures Citation[1] lead to a hypothesis that the OmpT amino acids associating with LPS are Tyr134, Glu136, Arg138, Arg175, and Lys226 (A). The accuracy of this theoretical prediction is supported by mutation data Citation[16]. Other omptins have slightly different LPS binding regions (). Even the closest homologue of OmpT, OmpP, has His instead of Arg175 and Ile instead of Lys226. Both forms of SopA have Leu in place of Arg175. Pla and PlaA have Ser in place of Lys226. Two forms of PgtE have Glu in place of Lys226. Of all the amino acids mediating LPS association, Tyr134 and Glu136 are the most conserved and are present in nine out of 11 omptins. Arg138 is conserved in eight of the omptins. The least conserved amino acids are Arg175 and Lys226.

Table II.  Amino acids with the known functions among the omptin proteins.

Proteolytic characteristics of the omptins

OmpT as an aspartate protease

Earlier investigations Citation[20] supported the hypothesis that OmpT was a member of serine protease family because (i) it could be partially inhibited by serine protease inhibitors, such as diisopropylfluorophosphate (DFP) and phenylmethanesulphonylfluoride (PMSF) Citation[21], and (ii) a Ser99→Ala mutation decreased OmpT proteolytic activity 500 fold. However, based on the crystal structure analysis Citation[1], OmpT was reclassified as an aspartate protease. The analysis of structural and mutagenesis studies suggested that the omptin family proteases utilize a unique mechanism that combines the elements of both serine and aspartate proteases Citation[22]. On one hand, OmpT catalytic couple Asp83-Asp85 shares similarity to a catalytic site of aspartate proteases ( A, B). On the other hand, the OmpT catalytic couple Asp210-His212 is comparable to the catalytic triad of serine carboxyl proteases ( A, B). It has been proposed that the role of the catalytic site Ser, absent in omptins, is performed by a water molecule present in the catalytic site of the omptins Citation[1]. In the proposed mechanism, catalytic residues of OmpT activate a water molecule that acts as a nucleophile to attack the scissile bond of the substrate. This water molecule plays the role of Oγ of the catalytic serine in the serine proteases. The Asp83-Asp85 couple contributes to a proton translocation or a stabilization of an oxyanion intermediate Citation[1], Citation[22].

The sequence alignment analysis of omptin proteases shows that the catalytic amino acids are conserved among the omptins. However, there are several homologues, which have mutations in the catalytic site residues (). Overall, the proteolytic mechanism is highly conserved among the proteases of the omptin family.

The active site of OmpT is surrounded by the following amino acids Glu27, Ser40, and Leu42 in L1; Met81, Asp97, Ser99, and His101 in L2; Ala143, Tyr150, and Ile170 in L3; Asp208 in L4; Thr263, Ala280, and Ile282 in L5, which are well conserved (). However, different amino acids in the active site can lead to a difference in the cleavage specificity. A recent report Citation[23] states that a change of the amino acid from Ser to Arg at the position 223 alters drastically cleavage specificity of OmpT from Arg-Arg to Ala-Arg Comparison of amino acid residues at this position among the omptins shows variation () in agreement with the experimental findings.

Ionic interactions appear important in cleavage specificity because the negative environment surrounding the active site leads to a high selectivity of OmpT towards more basic targets Citation[14]. Significance of the electrostatic interaction between OmpT and its substrates is highlighted by the evidence that increased ionic strength, high pH, and acidic amino acids at the cleavage site of a substrate are able to decrease the OmpT affinity for its substrate Citation[14], Citation[24], Citation[25].

Narrow cleavage specificity of OmpT

The OmpT protease has a narrowly-defined cleavage specificity. OmpT targets sites that consist of two basic amino acids, its main substrate is proteamine P1 Citation[26]. Proteamine P1 is 50 amino acids long peptide with 24 arginines and no lysines. The high sequence homogeneity among its targets is consistent with the narrow cleavage specificity of OmpT.

Using Schechter and Berger nomenclature Citation[27], the OmpT recognition site on the surface of a substrate protein can be designated as … P3-P2-P1-P1’-P2’-P3’ … and so on, where the proteolytic cleavage occurs between P1 and P1’. OmpT has an absolute requirement for a basic amino acid at position P1 Citation[24], Citation[25], Citation[28]. On the other hand, the amino acid at the P1’ position is more flexible. Although Lys and Arg are the most preferable for cleavage, amino acids with small non-acidic side chains are also good substrates. The flexibility at the P1’ position helps to explain the capability of OmpT to cleave plasminogen at the Arg-Val site Citation[28], while D-Arg at P1’ is an inhibitor of OmpT activity Citation[24]. Selectivity of OmpT diminishes with the distance from the scissile bond. At the P2’ site, OmpT prefers hydrophobic amino acids, such as Ile or Val. OmpT prefers Ala at the P2 position, but a variety of non-acidic amino acids are tolerated. At the P3 site, Arg and Trp are the residues of choice. The P3’ site can accommodate large amino acids like Trp, Arg, and Thr. Both P4 and P6 sites prefer a basic amino acid and disallow acidic residues Citation[25]. Overall, OmpT has a strong preference for the positively charged amino acids (). Such specificity strengthens the hypothesis that the long-range electrostatic interactions between OmpT and its substrates play an important role in the recognition of substrates.

Table III.  Summary of preferable amino acids in the recognition site of OmpT substrate.

OmpT-Pla chimeric proteins Citation[2] demonstrate that the substrate specificity varies from one omptin to another via the sequence variability in the outer loops. The substitution of the C-terminal part of L4 of OmpT with the Pla sequence confers a Pla-like ability to efficiently cleave plasminogen to OmpT. The additional substitution of L3 and L5 gives OmpT capability to inactivate α2-antiplasmin via cleavage Citation[2]. Substrate's size also influences the rate of cleavage. Consequently, OmpT prefers to cleave small peptides, while Pla targets larger proteins, and PgtE cleaves both Citation[29].

Adhesive properties of the omptins

Several omptins have been implicated in adherence to eukaryotic cells. Pla of Y. pestis, PgtE of S. enterica, and OmpT of E. coli are the best characterized omptin adhesins Citation[13], Citation[16], Citation[30–32]. The adhesive activity of Pla is distinctly separated from its proteolytic activity because proteolytically inactive Pla Ser99→Ala and Asp206→Ala mutants still retain their ability to adhere to human endothelial cell lines Citation[13]. Y. pestis binds to the cells lines derived from human epithelial and endothelial tissues through Pla. Even the cell components, such as basement membrane preparations (Matrigel) and mammalian extracellular matrix (ECM) preparations, are sufficient for Pla-expressing Y. pestis or E. coli cells to bind Citation[13], Citation[29], Citation[31]. Pla exhibits a strong adherence to a murine laminin, a medium adherence to a murine heparan sulfate proteoglycan, and a weak adherence to human collagen types IV, I, V, human cellular, and plasma fibronectins Citation[31]. Inhibition studies Citation[30] reveal a lectin-like nature of the Pla-mediated adherence to galactose-rich carbohydrates, present in the glycoproteins, laminin and collagen IV, and the glycolipid globotetraosylceramide. Adherence to laminin is significant not only because it is very strong, but also because laminin is a highly abundant component of the extracellular matrix. Laminin provides an ample opportunity for bacterial adhesion to occur during infection.

PgtE of S. enterica is also shown to promote bacterial adhesion to a human endothelial cell line, almost as efficiently as Pla Citation[16]. OmpT-directed bacterial adherence is four-five times lower than that of Pla-expressing E. coli cells but is still significantly higher than seen in E. coli cells lacking OmpT Citation[31]. Also, OmpT-mediated E. coli adhesion to basement membrane preparation is significantly higher than to albumin-coated surfaces Citation[16]. Furthermore, studies have shown that the purified OmpT binds weakly to purified human laminin and fibronectin, the two extracellular proteins commonly found coating human cells Citation[32]. From the data it appears that Pla and PgtE are strong adhesins, while OmpT exhibits the least adhesive activity among the three omptins examined Citation[16], Citation[29–31].

Invasin activity of Pla of Y. pestis

Pla of Y. pestis is the only omptin shown to be an invasin (). It has been shown that Pla is capable of promoting bacterial invasion of the human epithelial and endothelial cells Citation[16]. Gentamycin protection assays Citation[33] show that the majority of Pla-expressing Y. pestis cells invade human pneumocyte and HeLa cell lines within 1 h of infection. The presence of the Pla-encoding plasmid, pPCP1, is responsible for 90–95% of Y. pestis invasiveness measured against a strain that lacks the pPCP1 plasmid Citation[33]. The pivotal role of Pla in the bacterial invasion is highlighted by the ability of Pla to confer invasiveness to non-invasive E. coli strains. E. coli cells with cloned Pla invade HeLa monolayers in a time-dependent manner, similar to Y. pestis. Moreover, proteolytically inactive Pla mutants Ser99→Ala and Asp206→Ala retain their ability to confer invasiveness to E. coli cells Citation[13]. As in the case of its adhesion properties, Pla's ability to invade is independent of the protein's proteolytic activity.

The internalization of Y. pestis depend on actin and less on microtubules Citation[33]. Studies with Pla-expressing E. coli have begun to dissect the bacterial invasion mechanism by determining the host factors involved. Therefore, a several fold decrease in the bacterial invasion following the staurosporin, a broad-spectrum protein kinase inhibitor, treatment suggests the involvement of several classes of host protein kinases. Further work with a more specific protein kinase inhibitor, genistein, indicates that some tyrosine protein kinases are involved. Phosphatidyl-inosinol-4-kinase, small GTPase Rho, and 5-lipoxygenase have been implicated in Pla-promoted bacterial internalization. A drastic effect of Cytochalasin D, an inhibitor of actin polymerization, on invasion strongly implies the involvement of actin in Pla-promoted invasion. Furthermore, invasion is immediately preceded by a burst of cytoskeletal rearrangement inside the target human epithelial cells Citation[34]. Pla targets eukaryotic signal transduction pathways responsible for transducing extracellular signals to the eukaryotic cell and reacting to these signals by cytoskeleton rearrangement.

PgtE of S. enterica and OmpT of E. coli are the other omptins, whose invasin properties have been tested. Neither can invade endothelial cell lines, despite the ability of PgtE to mediate significant bacterial binding to human endothelial cells Citation[29].

Coagulase activity of Pla of Y. pestis

The only omptin whose coagulase activity has been investigated is Pla of Y. pestis. Pla has been reported to coagulate rabbit plasma Citation[35], Citation[36]. The positive results require a long incubation time and a significant number of Y. pestis cells, which are unlikely during a rapidly disseminating plague infection Citation[7]. Insensitivity of Pla-promoted coagulation to thrombin inhibitors suggests that Pla initiates the process not by thrombin activation, but by the direct cleavage of rabbit fibrinogen. It is likely that Pla fibrinogen cleavage sites are unique to rabbits, because Pla cannot cleave and coagulate plasma from human, rats, or mice Citation[7]. The Pla coagulase activity is unlikely to be significant to the virulence of Y. pestis in these hosts.

Role of the omptins in bacterial pathogenesis

OmpT of E. coli

OmpT is an outer membrane protein T (temperature-regulated), previously known as protein A and protease VII Citation[21] whose expression is upregulated at the higher temperatures. OmpT is detected in both nonpathogenic (e.g., K12), and human uropathogenic E. coli strains (UPEC). The ompT gene is detected in 83.1–88.5% of E. coli isolates from urinary tract infections (UTI) and in 67.7% of faecal E. coli isolates Citation[37]. All UTI E. coli isolates had a statistically higher prevalence of ompT gene compared to the faecal E. coli isolates, based on a pairwise z-test with a 5% margin of error. The majority (>50%) of human E. coli isolates have the ompT gene. However, the proteolytic activity of OmpT is detected in E. coli samples from urine (13%) and wounds (19%) Citation[38]. A comprehensive study Citation[39] of E. coli isolates from neonatal bacterial meningitis in the Netherlands finds ompT DNA in 96% of the isolates. In the UPEC strains, OmpT is reported to cleave proteamine P1, an antimicrobial peptide is secreted by epithelial cells of the urinary tract. Protamine associates with DNA in the nuclei of sperm of various animal species and is a part of innate immunity. This peptide is secreted by host cells in the effort to kill the invading bacterial infection. It is thought that it acts on the bacterial cytoplasmic membrane and targets phospholipids in the membrane. However, the details of its mechanism are still unknown. Studies have shown that it disrupts bacterial protein synthesis, diminishes cellular ATP content, inhibits amino acid transport, disrupts oxygen consumption, and causes leakage of intracellular components Citation[40], Citation[41]. OmpT-mediated deactivation of proteamine P1 contributes to bacterial survival inside the host and classifies OmpT as a virulence factor () Citation[26].

As an outer membrane protease, OmpT processes some of the autotransporters that are shown to be virulence factors Citation[42–44]. Reported examples of such autotransporters are IgAp, IcsA, PrtS Citation[42], and YapA Citation[45]. When autotransporters, with the OmpT recognition site, are introduced into the OmpT-expressing E. coli strains, they are processed and released into the extracellular space ().

Purified OmpT can slowly cleave plasminogen Citation[28], releasing active host serine protease plasmin Citation[46]. However, the plasmin activated by OmpT is quickly inhibited by α2-antiplasmin, a natural plasmin inhibitor Citation[2]. The limited plasmin activation might be sufficient for the UPEC cells to penetrate beneath the upper layer of cells of the urinary tract to establish an infection Citation[38], but insufficient to promote an invasive infection.

The ability of OmpT to adhere to basement membrane preparations Citation[16], extracellular matrix proteins Citation[32], and human epithelial cells Citation[13] may contribute to bacterial cell attachment to host epithelial tissues (e.g., urinary tract) and the establishment of a persisting bacterial infection. Moreover, it has been recently shown that the purified OmpT induces the production of cytokine TNFα in mononuclear cells Citation[47]. Identity of this receptor on the surface of mononuclear cells has not been achieved.

Pla of Y. pestis

Plasminogen activator (Pla) is a 292 amino acid protein encoded by the pPCP1 plasmid of Y. pestisCitation[13]. Despite its small size (9.6 kbp), this plasmid is crucial for a rapid dissemination of bacteria throughout a host body from a subcutaneous infection. Bacteria that express Pla are highly virulent yielding an LD50 increase of 106 fold in the absence of Pla Citation[7], Citation[10], Citation[11]. It has been recently shown that the expression of Pla is needed for establishment of pneumonic plague Citation[12] and is necessary to initiate bubonic plague Citation[11]. Furthermore, pla has been found to be one of the most up-regulated genes of Y. pestis isolated from the bubo Citation[48].

The proteolytic activity of Pla may help Y. pestis to overcome barriers, such as fibrin clots set up by a host to limit the spread of the infection. Pla is not able to dissolve fibrin clots, cleave laminin, nor cause a degradation of ECM human preparations and basal membrane murine preparations. At the same time, Pla can induce these processes by activating plasmin. Pla activates the human plasmin from a plasminogen precursor via proteolytic cleavage Citation[31]. Pla cleaves plasminogen at the same site and with the same efficiency as the host plasmin activators (e.g., urokinase) Citation[7]. The host has to tightly control the amount of active plasmin since plasmin has a broad proteolytic specificity. Host α2-antiplasmin and aprotinin deactivate free plasmin, however, to extend plasmin activity, Pla cleaves and inactivates α2-antiplasmin Citation[31]. Furthermore, the plasmin is likely to remain bound to the bacterial surface to ensure its activity. This suggests that Y. pestis uses Pla to overcome host barriers and to facilitate bacterial spread from a subcutaneous infection site Citation[31]. Pla protease is unregulated by host protease inhibitors. As a result, it induces uncontrollable activation of proteolysis leading to detrimental effects of infection ().

Adhesive properties of Pla contribute to the bacterial colonization in damaged tissue, such as epithelium after a flea-bite. This protein also helps the bacteria invade human endothelial cells Citation[13]. Lesions at the subcutaneous infection sites of Pla-expressing Y. pestis cells have much lower counts of inflammatory cells than the sites infected with Y. pestis cells lacking Pla. Pla cleaves the C3 component of the complement system on the bacterial surface. Since C3 component is one of the key factors in the complement activation, its deactivation leads to a reduced production of chemoattractants and subsequent suppression of the local inflammation Citation[7].

SopA of S. flexneri

SopA (Shigella outer membrane protease) is encoded by S. flexneri virulence plasmid. The main function of SopA (IcsP) protease is to cleave autotransporter IcsA (VirG) between Arg758 and Arg759 and release the passenger domain (IcsAα) from the bacterial surface Citation[8]. The localization of accessible IcsAα at one pole of the bacterial cell is essential for the formation of actin ‘comet’ tails on the bacterial surface Citation[49]. These tails propel the bacterial pathogen within the host cell as well as enable the bacilli to spread to neighboring cells Citation[50]. Interestingly, in the presence of OmpT, IcsA is degraded and the virulence of S. flexneri is suppressed Citation[44]. IcsA is encoded by the same virulence plasmid as SopA. Moreover the expression of both genes is controlled by VirF. The expression of SopA is additionally controlled by VirB. VirB and VirF proteins regulate the virulence phenotype in S. flexneriCitation[51]. SopA is an important virulence factor () since it plays an important role in the exclusive polar localization of IcsA, Citation[52].

PgtE of S. enterica

PgtE is a chromosomally-encoded protein of S. enterica. The activity of PgtE is controlled by the SlyA and PhoP/PhoQ mechanisms known to partially control some of Salmonella virulence and resistance to antimicrobial peptides Citation[3], Citation[53], Citation[54].

Like Pla, PgtE can activate plasminogen to plasmin, which degrades fibrin clots and extracellular matrices of eukaryotic tissues. Just like Pla, PgtE can be deactivated by long O-antigen chains of LPS. Like OmpT, PgtE degrades small peptides. PgtE can cleave alpha-helical cationic antimicrobial peptides, such as C18G (). Thus PgtE contributes to bacterial survival inside the host during infection Citation[53]. PgtE expression is upregulated in S. enterica inside vacuoles of murine macrophages and the length of O-antigen chain is reduced, thus increasing PgtE protease activity Citation[18], Citation[55]. Therefore it might contribute to the bacterial survival inside the host cells.

Evolutionary aspects

Recently new members of the omptin family have been identified, as summarized in MEROPS database under family A26 Citation[56]. Plasmids encode genes for most of the identified omptins. Notable exceptions are ompT of E. coli, pgtE of S. enterica, Q8ZGQ6 of Y. pestis, and CAH20506.1 of Y. pseudotuberculosis which are found on chromosomes. The ompT gene is flanked by repeats, suggesting that it was encoded by a lambiod prophage Citation[44]. Based on the proteins’ similarity, an evolutionary tree of the omptin family proteases was constructed (). Known omptins fall into two major groups, one group contains PgtE, Pla and PlaA, while the other group contains OmpT, OmpP, and SopA Citation[3]. Each group is characterized by 60 to 84% sequence similarity among its members. Moreover, members in each group are functionally similar. PgtE and Pla are both strong adhesins Citation[16], Citation[31] able to cleave large substrates Citation[29], while the O-antigen of LPS inhibits the proteolytic activities of both Pla and PgtE Citation[16]. PlaA of Erwinia spp. is likely to play an important role in bacteria-plant interaction.

Figure 2.  Phylogenetic tree of the omptin family members from theA26 family from the MEROPS database Citation[56]. Scale bar refers to time in million years. MEGA3 software and UPGMA method were used to contract the tree Citation[68].

Figure 2.  Phylogenetic tree of the omptin family members from theA26 family from the MEROPS database Citation[56]. Scale bar refers to time in million years. MEGA3 software and UPGMA method were used to contract the tree Citation[68].

Y. pestis evolved from a mild Y. pseudotuberculosis some time between 1,500 to 20,000 years ago Citation[57]. Both species have chromosome-encoded omptin homologs (Q8ZGQ6 and CAH20506.1) with Asn instead of Asp in their catalytic sites. This mutation probably renders them proteolytically inactive. Nevertheless, studies have shown that Q8ZGQ6 of Y. pestis contributes to the establishment of infection in the soil nematode, Caenorhabditis elegansCitation[58]. Y. pestis pla gene most likely evolved from an ancestral S. enterica chromosomal pgtE gene, found in two homologous forms, one in S. enterica serovar Typhimurium and the other one in serovar Typhi Citation[7]. Functional similarities between Pla and PgtE support their common ancestry Citation[2], Citation[16]. There are several homologues of Pla found in plant pathogens Erwinia spp. PlaA and AAN04526.1 have a conserved catalytic site and are likely to be functional proteases and may play a role in establishment of Erwinia infection in plants. Another Erwinia omptin homolog is AAG31039, most likely a nonsecreted nonproteolytic omptin homologue since all four catalytic site residues are mutated and SignalP software found no signal sequence Citation[59]. Another omptin belonging to this group is lpl2311 of Legionella pneumophila, a cause of Legionnaires’ disease. Based on SignalP predictions and sequence analysis, lpl2311 contains a signal peptide, necessary to the secretion and conserved catalytic site residues. Lpl2311 may play a role in bacterial survival inside the phagocytic cells because lpl2311 sequence resembles that of PgtE and L. pneumophila like S. enterica, proliferates inside macrophages. Alternatively lpl2311 can also contribute to L. pneumophila survival inside protozoa.

The second group includes OmpT and SopA, which process autotransporter proteins Citation[8], Citation[42] important in virulence. SopA1 of S. flexneri, SopA2 of enteroinvasive E. coli, and SopA3 of S. dysenteriae are all located on a virulence plasmid that arose in ancestral E. coli that subsequently gave rise to Shigella and enteroinvasive E. coliCitation[60]. Interestingly, enteroinvasive E. coli strains have lost the ompT gene, through the prophage excision, but gained the sopA2 gene, possibly, through the horizontal gene transfer Citation[44]. If ompT gene is introduced in either enteroinvasive E. coli strains or Shigella, these bacteria lose the ability of intercellular spreading Citation[44].

E. coli OmpT is a housekeeping protein that plays a role in terminating SOS response to UV radiation. OmpT cleaves and deactivates UvrB, preventing UvrABC endonuclease from hydrolyzing undamaged DNA Citation[61]. OmpT has been observed to cleave bound colicins Citation[62] and it is thought to contribute to bacterial survival in the presence of colicins. Also, OmpT has an affinity for denatured substrates such as over-expressed proteins in inclusion bodies under denaturing conditions (up to 5 M urea) Citation[63]. This property of OmpT plays a role as a housekeeping protein in denatured protein turn-over at elevated temperatures and processing of proteins in the membrane. This omptin also contributes to the pathogenicity of UPEC strains by cleaving protamine Citation[26] and by processing certain autotransporters.

Other omptins do not fall into either of the two groups described above, such as Vibrio fischeri AW85741 a nonproteolytic omptin homologue and CAG37431 from Desulfatalea psychrophila, a sulfate-reducing bacterium from Arctic permafrost, contains a conserved catalytic site Citation[64]. Other omptin family members are the outer membrane proteases (Q98A65) of Mesorhizobium loti and EAN05038 of Mesorhisobium sp. BNC1 the symbiotic plant bacteria. Computational analysis with SignalP software identified a cleavable signal peptide Citation[59], Citation[65]. Analysis of the region upstream of the structural genes revealed a Shine-Dalgarno sequence implying that these genes are likely to be expressed. The most conserved amino acid in the active site is Ser99, a crucial active site residue for proteolytic activity (). This characteristic is shared by both proteolytic and nonproteolytic omptins, and initially even lead researchers classify OmpT as a serine protease. Ser40 and Tyr150 are the other residues that are conserved among all the known omptins. High conservation of serines and omptins’ partial susceptibility to serine inhibitors Citation[20] suggest that omptins might have evolved from serine proteases.

Concluding remarks and future research

There are many bacterial infections without effective treatment or prevention. Death from an infectious disease reached the top five causes of death in the United States in the mid-1990s Citation[66]. One of the several reasons is the increasing population of the elderly, who are particularly susceptible to infections. Another reason is that the achievements in cancer therapy and organ transplantation increased the population of immunosuppressed patients. Moreover, old infectious diseases (e.g., plague) are now classified as re-emerging infections. The omptins provide hope as new potential targets for drug and vaccine development.

Several proteolytic studies with OmpT Citation[24], Citation[25], Citation[28], Citation[67] provide numerous clues about the proteolytic mechanism of the OmpT protease and its inhibition. This could be used as a starting point of computer-based search for OmpT-inhibiting drugs. Narrow cleavage specificity of OmpT could be exploited. Modified OmpT could be a specific protease to release engineered proteins Citation[23], Citation[25]. Besides the E. coli OmpT, there are other omptins whose proteolytic properties need to be explored. One of them is a potent protease Pla. The inhibition of this protein is crucial in preventing a septisemic plague infection. Recently identified omptins provide many other directions of research. It would be interesting to see if PlaA contributes to the pathogenesis of E. pyrifoliae, and if the protease of M. loti plays a role in bacterium-plant symbiosis. Also the ability of the chromosomal Y. pestis proteolytically inactive omptin to contribute to pathogenesis needs to be explored. It is still unknown if the chromosomal omptins of Y. tuberculosis or V. fischeri have any activities. The function of lpl2311 and its possible role in L. pneumophila infection in humans or protozoa remains to be explored. The relation between the protein's activity and the length of O-antigen would be interesting to explore.

Acknowledgements

Authors would like to thank Dr Mohammed Shayib for his help with statistical analysis and Dr William Widger for reviewing the manuscript. Authors are grateful to the anonymous reviewers for their critical comments which greatly improved the quality of the manuscript.

References

  • Vandeputte-Rutten L, Kramer RA, Kroon J, Dekker N, Egmond MR, Gros P. Crystal structure of the outer membrane protease OmpT from Escherichia coli suggests a novel catalytic site. Embo J 2001; 20: 5033–5039
  • Kukkonen M, Lahteenmaki K, Suomalainen M, Kalkkinen N, Emody L, Lang H, Korhonen TK. Protein regions important for plasminogen activation and inactivation of alpha2-antiplasmin in the surface protease Pla of Yersinia pestis. Mol Microbiol 2001; 40: 1097–1111
  • Kukkonen M, Korhonen TK. The omptin family of enterobacterial surface proteases/adhesins: from housekeeping in Escherichia, coli to systemic spread of Yersinia pestis. Int J Med Microbiol 2004; 294: 7–14
  • Stathopoulos C. Structural features, physiological roles, and biotechnological applications of the membrane proteases of the OmpT bacterial endopeptidase family: a micro-review. Membr Cell Biol 1998; 12: 1–8
  • Grodberg J, Dunn JJ. ompT encodes the Escherichia coli outer membrane protease that cleaves T7 RNA polymerase during purification. J Bacteriol 1988; 170: 1245–1253
  • Kaufmann A, Stierhof YD, Henning U. New outer membrane-associated protease of Escherichia coli K-12. J Bacteriol 1994; 176: 359–367
  • Sodeinde OA, Subrahmanyam YV, Stark K, Quan T, Bao Y, Goguen JD. A surface protease and the invasive character of plague. Science 1992; 258: 1004–1007
  • Egile C, d'Hauteville H, Parsot C, Sansonetti PJ. SopA, the outer membrane protease responsible for polar localization of IcsA in Shigella flexneri. Mol Microbiol 1997; 23: 1063–1073
  • Grodberg J, Dunn JJ. Comparison of Escherichia coli K-12 outer membrane protease OmpT and Salmonella typhimurium E protein. J Bacteriol 1989; 171: 2903–2905
  • Brubaker RR, Beesley ED, Surgalla MJ. Pasturella pestis: role of pesticin I and iron in experimental plague. Science 1965; 149: 422–424
  • Sebbane F, Jarrett CO, Gardner D, Long D, Hinnebusch BJ. Role of the Yersinia pestis plasminogen activator in the incidence of distinct septicemic and bubonic forms of flea-borne plague. Proc Natl Acad Sci USA 2006; 103: 5526–5530
  • Lathem WW, Price PA, Miller VL, Goldman WE. A plasminogen-activating protease specifically controls the development of primary pneumonic plague. Science 2007; 315: 509–513
  • Lahteenmaki K, Kukkonen M, Korhonen TK. The Pla surface protease/adhesin of Yersinia pestis mediates bacterial invasion into human endothelial cells. FEBS Lett 2001; 504: 69–72
  • Kramer RA, Zandwijken D, Egmond MR, Dekker N. In vitro folding, purification and characterization of Escherichia coli outer membrane protease ompT. Eur J Biochem 2000; 267: 885–893
  • Kramer RA, Brandenburg K, Vandeputte-Rutten L, Werkhoven M, Gros P, Dekker N, Egmond MR. Lipopolysaccharide regions involved in the activation of Escherichia coli outer membrane protease OmpT. Eur J Biochem 2002; 269: 1746–1752
  • Kukkonen M, Suomalainen M, Kyllonen P, Lahteenmaki K, Lang H, Virkola R, Helander IM, Holst O, Korhonen TK. Lack of O-antigen is essential for plasminogen activation by Yersinia pestis and Salmonella enterica. Mol Microbiol 2004; 51: 215–225
  • Pouillot F, Derbise A, Kukkonen M, Foulon J, Korhonen TK, Carniel E. Evaluation of O-antigen inactivation on Pla activity and virulence of Yersinia pseudotuberculosis harbouring the pPla plasmid. Microbiology 2005; 151: 3759–3768
  • Lahteenmaki K, Kyllonen P, Partanen L, Korhonen TK. Antiprotease inactivation by Salmonella enterica released from infected macrophages. Cell Microbiol 2005; 7: 529–538
  • Ferguson AD, Welte W, Hofmann E, Lindner B, Holst O, Coulton JW, Diederichs K. A conserved structural motif for lipopolysaccharide recognition by procaryotic and eucaryotic proteins. Structure Fold Des 2000; 8: 585–592
  • Kramer RA, Dekker N, Egmond MR. Identification of active site serine and histidine residues in Escherichia coli outer membrane protease OmpT. FEBS Lett 2000; 468: 220–224
  • Sugimura K, Nishihara T. Purification, characterization, and primary structure of Escherichia coli protease VII with specificity for paired basic residues: identity of protease VII and OmpT. J Bacteriol 1988; 170: 5625–5632
  • Vandeputte-Rutten L, Gros P. Novel proteases: common themes and surprising features. Curr Opin Struct Biol 2002; 12: 704–708
  • Varadarajan N, Gam J, Olsen MJ, Georgiou G, Iverson BL. Engineering of protease variants exhibiting high catalytic activity and exquisite substrate selectivity. Proc Natl Acad Sci USA 2005; 102: 6855–6860
  • Dekker N, Cox RC, Kramer RA, Egmond MR. Substrate specificity of the integral membrane protease OmpT determined by spatially addressed peptide libraries. Biochemistry 2001; 40: 1694–1701
  • Okuno K, Yabuta M, Ohsuye K, Ooi T, Kinoshita S. An analysis of target preferences of Escherichia coli outer-membrane endoprotease OmpT for use in therapeutic peptide production: efficient cleavage of substrates with basic amino acids at the P4 and P6 positions. Biotechnol Appl Biochem 2002; 36: 77–84
  • Stumpe S, Schmid R, Stephens DL, Georgiou G, Bakker EP. Identification of OmpT as the protease that hydrolyzes the antimicrobial peptide protamine before it enters growing cells of Escherichia coli. J Bacteriol 1998; 180: 4002–4006
  • Schechter I, Berger A. On the size of the active site in proteases. I. Papain. Biochem Biophys Res Commun 1967; 27: 157–162
  • McCarter JD, Stephens D, Shoemaker K, Rosenberg S, Kirsch JF, Georgiou G. Substrate specificity of the Escherichia coli outer membrane protease OmpT. J Bacteriol 2004; 186: 5919–5925
  • Lahteenmaki K, Kukkonen M, Jaatinen S, Suomalainen M, Soranummi H, Virkola R, Lang H, Korhonen TK. Yersinia pestis Pla has multiple virulence-associated functions. Adv Exp Med Biol 2003; 529: 141–145
  • Kienle Z, Emody L, Svanborg C, O'Toole PW. Adhesive properties conferred by the plasminogen activator of Yersinia pestis. J Gen Microbiol 1992; 138 (Pt 8): 1679–1687
  • Lahteenmaki K, Virkola R, Saren A, Emody L, Korhonen TK. Expression of plasminogen activator pla of Yersinia pestis enhances bacterial attachment to the mammalian extracellular matrix. Infect Immun 1998; 66: 5755–5762
  • Hritonenko V. Bacterial outer membrane proteins and infectious diseases: computational and biochemical analyses. Int J Ecol Div 2007; 6: 78–86
  • Cowan C, Jones HA, Kaya YH, Perry RD, Straley SC. Invasion of epithelial cells by Yersinia pestis: evidence for a Y. pestis-specific invasin. Infect Immun 2000; 68: 4523–4530
  • Benedek O, Nagy G, Emody L. Intracellular signalling and cytoskeletal rearrangement involved in Yersinia pestis plasminogen activator (Pla) mediated HeLa cell invasion. Microb Pathog 2004; 37: 47–54
  • Beesley ED, Brubaker RR, Janssen WA, Surgalla MJ. Pesticins. 3. Expression of coagulase and mechanism of fibrinolysis. J Bacteriol 1967; 94: 19–26
  • Sodeinde OA, Goguen JD. Genetic analysis of the 9.5-kilobase virulence plasmid of Yersinia pestis. Infect Immun 1988; 56: 2743–2748
  • Marrs CF, Zhang L, Tallman P, Manning SD, Somsel P, Raz P, Colodner R, Jantunen ME, Siitonen A, Saxen H, Foxman B. Variations in 10 putative uropathogen virulence genes among urinary, faecal and peri-urethral Escherichia coli. J Med Microbiol 2002; 51: 138–142
  • Lundrigan MD, Webb RM. Prevalence of ompT among Escherichia coli isolates of human origin. FEMS Microbiol Lett 1992; 76: 51–56
  • Johnson JR, Oswald E, O'Bryan TT, Kuskowski MA, Spanjaard L. Phylogenetic distribution of virulence-associated genes among Escherichia coli isolates associated with neonatal bacterial meningitis in the Netherlands. J Infect Dis 2002; 185: 774–784
  • Aspedon A, Groisman EA. The antibacterial action of protamine: evidence for disruption of cytoplasmic membrane energization in Salmonella typhimurium. Microbiology 1996; 142 (Pt 12): 3389–3397
  • Johansen C, Verheul A, Gram L, Gill T, Abee T. Protamine-induced permeabilization of cell envelopes of gram-positive and gram-negative bacteria. Appl Environ Microbiol 1997; 63: 1155–1159
  • Jacob-Dubuisson F, Fernandez R, Coutte L. Protein secretion through autotransporter and two-partner pathways. Biochim Biophys Acta 2004; 1694: 235–257
  • Ruiz-Olvera P, Ruiz-Perez F, Sepulveda NV, Santiago-Machuca A, Maldonado-Rodriguez R, Garcia-Elorriaga G, Gonzalez-Bonilla C. Display and release of the Plasmodium falciparum circumsporozoite protein using the autotransporter MisL of Salmonella enterica. Plasmid 2003; 50: 12–27
  • Nakata N, Tobe T, Fukuda I, Suzuki T, Komatsu K, Yoshikawa M, Sasakawa C. The absence of a surface protease, OmpT, determines the intercellular spreading ability of Shigella: the relationship between the ompT and kcpA loci. Mol Microbiol 1993; 9: 459–468
  • Yen YT, Karkal A, Bhattacharya M, Fernandez R, Stathopoulos C. Identification and characterization of autotransporter proteins of Yersinia pestis KIM. Mol Membr Biol 2007; 24: 28–40
  • Mangel WF, Toledo DL, Brown MT, Worzalla K, Lee M, Dunn JJ. Omptin: an Escherichia coli outer membrane proteinase that activates plasminogen. Methods Enzymol 1994; 244: 384–399
  • Brandenburg K, Garidel P, Schromm AB, Andra J, Kramer A, Egmond M, Wiese A. Investigation into the interaction of the bacterial protease OmpT with outer membrane lipids and biological activity of OmpT:lipopolysaccharide complexes. Eur Biophys J. 2005; 34: 28–41
  • Sebbane F, Lemaitre N, Sturdevant DE, Rebeil R, Virtaneva K, Porcella SF, Hinnebusch BJ. Adaptive response of Yersinia pestis to extracellular effectors of innate immunity during bubonic plague. Proc Natl Acad Sci USA 2006; 103: 11766–11771
  • Morona R, Daniels C, Van Den Bosch L. Genetic modulation of Shigella flexneri 2a lipopolysaccharide O antigen modal chain length reveals that it has been optimized for virulence. Microbiology 2003; 149: 925–939
  • Cossart P, Sansonetti PJ. Bacterial invasion: the paradigms of enteroinvasive pathogens. Science 2004; 304: 242–248
  • Wing HJ, Yan AW, Goldman SR, Goldberg MB. Regulation of IcsP, the outer membrane protease of the Shigella actin tail assembly protein IcsA, by virulence plasmid regulators VirF and VirB. J Bacteriol 2004; 186: 699–705
  • Monack DM, Theriot JA. Actin-based motility is sufficient for bacterial membrane protrusion formation and host cell uptake. Cell Microbiol 2001; 3: 633–647
  • Guina T, Yi EC, Wang H, Hackett M, Miller SI. A PhoP-regulated outer membrane protease of Salmonella enterica serovar typhimurium promotes resistance to alpha-helical antimicrobial peptides. J Bacteriol 2000; 182: 4077–4086
  • Navarre WW, Halsey TA, Walthers D, Frye J, McClelland M, Potter JL, Kenney LJ, Gunn JS, Fang FC, Libby SJ. Co-regulation of Salmonella enterica genes required for virulence and resistance to antimicrobial peptides by SlyA and PhoP/PhoQ. Mol Microbiol 2005; 56: 492–508
  • Eriksson S, Lucchini S, Thompson A, Rhen M, Hinton JC. Unravelling the biology of macrophage infection by gene expression profiling of intracellular Salmonella enterica. Mol Microbiol 2003; 47: 103–118
  • Rawlings ND, Morton FR, Barrett AJ. MEROPS: the peptidase database. Nucleic Acids Res 2006; 34: D270–272
  • Achtman M, Zurth K, Morelli G, Torrea G, Guiyoule A, Carniel E. Yersinia pestis, the cause of plague, is a recently emerged clone of Yersinia pseudotuberculosis. Proc Natl Acad Sci USA 1999; 96: 14043–14048
  • Styer KL, Hopkins GW, Bartra SS, Plano GV, Frothingham R, Aballay A. Yersinia pestis kills Caenorhabditis elegans by a biofilm-independent process that involves novel virulence factors. EMBO Rep 2005; 6: 992–997
  • Bendtsen JD, Nielsen H, von Heijne G, Brunak S. Improved prediction of signal peptides: SignalP 3.0. J Mol Biol 2004; 340: 783–795
  • Escobar-Paramo P, Giudicelli C, Parsot C, Denamur E. The evolutionary history of Shigella and enteroinvasive Escherichia coli revised. J Mol Evol 2003; 57: 140–148
  • Sedliakova M. A non-excision uvr-dependent DNA repair pathway of Escherichia coli (involvement of stress proteins). J Photochem Photobiol B 1998; 45: 75–81
  • Cavard D, Lazdunski C. Colicin cleavage by OmpT protease during both entry into and release from Escherichia coli cells. J Bacteriol 1990; 172: 648–652
  • White CB, Chen Q, Kenyon GL, Babbitt PC. A novel activity of OmpT. Proteolysis under extreme denaturing conditions. J Biol Chem 1995; 270: 12990–12994
  • Rabus R, Ruepp A, Frickey T, Rattei T, Fartmann B, Stark M, Bauer M, Zibat A, Lombardot T, Becker I, Amann J, Gellner K, Teeling H, Leuschner WD, Glockner FO, Lupas AN, Amann R, Klenk HP. The genome of Desulfotalea psychrophila, a sulfate-reducing bacterium from permanently cold Arctic sediments. Environ Microbiol 2004; 6: 887–902
  • Nielsen H, Krogh A. Prediction of signal peptides and signal anchors by a hidden Markov model. Proc Int Conf Intell Syst Mol Biol 1998; 6: 122–130
  • MMWR. 1999. Mortality Patterns – United States, 1997. pp. 664–668. Atlanta, GA: Centers for Disease Control and Prevention.
  • Okuno K, Yabuta M, Kawanishi K, Ohsuye K, Ooi T, Kinoshita S. Substrate specificity at the P1′ site of Escherichia coli OmpT under denaturing conditions. Biosci Biotechnol Biochem 2002; 66: 127–134
  • Kumar S, Tamura K, Nei M. MEGA3: Integrated software for molecular evolutionary genetics analysis and sequence alignment. Brief Bioinform 2004; 5: 150–163

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.