Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2016 Mar 23.
Published in final edited form as: Microbiol Spectr. 2016 Feb;4(1):10.1128/microbiolspec.VMBF-0004-2015. doi: 10.1128/microbiolspec.VMBF-0004-2015

The Structure and Function of Type III Secretion Systems

Ryan Q Notti 1,2, C Erec Stebbins 1
PMCID: PMC4804468  NIHMSID: NIHMS715954  PMID: 26999392

ARTICLE SUMMARY

Type III secretion systems (T3SS) afford gram-negative bacteria a most intimate means of altering the biology of their eukaryotic hosts — the direct delivery of effector proteins from the bacterial cytoplasm to that of the eukaryote. This incredible biophysical feat is accomplished by nanosyringe “injectisomes,” which form a conduit across the three plasma membranes, peptidoglycan layer and extracellular space that form a barrier to the direct delivery of proteins from bacterium to host. The focus of this chapter is T3SS function at the structural level; we will summarize the core findings that have shaped our understanding of the structure and function of these systems and highlight recent developments in the field. In turn, we describe the T3SS secretory apparatus, consider its engagement with secretion substrates, and discuss the post-translational regulation of secretory function. Lastly, we close with a discussion of the future prospects for the interrogation of structure-function relationships in the T3SS.

INTRODUCTION

Type III secretion systems (T3SS) afford gram-negative bacteria a most intimate means of altering the biology of their eukaryotic hosts — the direct delivery of effector proteins from the bacterial cytoplasm to that of the eukaryote (1, 2). T3SS utilize a conserved set of homologous gene products to assemble the nanosyringe “injectisomes” capable of traversing the three plasma membranes, peptidoglycan layer and extracellular space that form a barrier to the direct delivery of proteins from bacterium to host. While the injectisome is architecturally similar across disparate gram-negatives, its applications are a study in diversity: T3SS are employed by both symbionts and pathogens; they target animals, plants, and protists; and they are used to manipulate a wide array of cellular activities and pathways.

T3SS have attracted intense scientific interest since the seminal work documenting their discovery was published over two decades ago (3-5). Given their role in the virulence of several human and plant pathogens (e.g. Salmonella enterica, Shigella flexneri, Yersinia spp., pathogenic Escherichia coli, Vibrio spp., Pseudomonas spp., Chlamydia spp.), T3SS are attractive targets for the discovery or design of novel anti-infective agents and vaccine approaches. Conversely, as T3SS accomplish the biophysical feat of protein transduction across multiple membranes, their re-engineering for in vivo delivery of therapeutic proteins or in vitro production of protein reagents provides an exciting prospect for future biomedical application. In either case, the manipulation of T3SS for human benefit will require highly refined mechanistic models of T3SS function. Drawing on research from multiple disciplines and employing complementary techniques, such models are beginning to emerge. In particular, the application of structural biochemical approaches to the T3SS has provided numerous insights into the assembly and function of this system.

The focus of this chapter will be on T3SS function at the structural level; we will summarize the core findings that have shaped our understanding of the structure and function of these systems and highlight recent developments in the field. In turn, we will describe the T3SS secretory apparatus, consider its engagement with secretion substrates, and discuss the post-translational regulation of secretory function. Lastly, we close with a discussion of the future prospects for the interrogation of structure-function relationships in the T3SS.

ARCHITECTURE OF A NANOSYRINGE

The genomic islands and virulence plasmids that support T3SS encode proteins of four broad classes: the components of the secretory system itself, the effector substrates, their chaperones, and transcriptional regulators. Working in concert, these components form a complete secretory system that de-chaperones and secretes substrates in a defined hierarchy and delivers them to the host cytoplasm. The repertoire of effector proteins secreted by a given T3SS is species-specific, as is the transcriptional network regulating T3SS expression (6). A discussion of these elements is beyond the scope of this review and has been expertly reviewed elsewhere (7-16). In contrast to the diverse, species-specific catalog of effector proteins and transcriptional regulators, the nanosyringe-like secretory machinery is well conserved across species, and advances in our mechanistic understanding of one species’ injectisome are often applicable to others.

The core secretion machinery of the T3SS is comprised of a homologous set of approximately two-dozen gene products. Because of the high degree of homology of some components of the system, a universal nomenclature was previously suggested to facilitate cross-species comparisons (17), and recently others in the field have endorsed this naming system (18, 19). Similarly, we will employ this nomenclature (Table 1) wherever possible. Additionally, a subset of these proteins have conserved homologues in the flagellar apparatus (Table 1), which uses its own T3SS machine to assemble the flagellar filament (20). Given that these T3SS subtypes appear to have diverged from a common ancestor some hundreds of millions of years ago (21), their conservation is noteworthy. While the focus of this review is the injectisome T3SS subtype, we will draw on the flagellar literature where it offers insights into injectisome function.

Table 1. A unified nomenclature for the homologous core components of the T3SS.

Based on the nomenclature proposed by Hueck (17), with modifications and additions from others (16, 18, 19, 90).

Notes Universal Nomenclature Salmonella SPI-1 Shigella EPEC Yersinia spp. Pseudomonas aeruginosa Flagellar Apparatus
Basal Body SctC InvG MxiD EscC YscC PscC
SctD PrgH MxiG EscD YscD PscD
SctJ PrgK MxiJ EscJ YscJ PscJ
Pilotin InvH MxiM YscW ExsB
Inner Rod SctI PrgJ MxiI EscI YscI PscI
Needle
Fliament
SctF PrgI MxiH EscF YscF PscF
Needle
Length
Regulator
SctP InvJ Spa32 EscP
(Orf16)
YscP PscP FliK
Inner
Membrane
Machinery
SctV InvA MxiA EscV YscV
(LcrD)
PcrD FlhA
SctR SpaP Spa24 EscR YscR PscR FliP
SctS SpaQ Spa9 EscS YscS PscS FliQ
SctT SpaR Spa29 EscT YscT PscT FliR
SctU SpaS Spa40 EscU YscU PscU FlhB
Needle Tip
and
Translocon
SipB IpaB EspD YopB PopB
SipC IpaC EspB YopD PopD
SipD IpaD EspA LcrV PcrV
ATPase SctN InvC Spa47 EscN YscN PscN FliI
Coiled Coil
Linker
SctO InvI Spa13 EscO
(Orf15,
EscA)
YscO PscO FliJ
Sorting
Platform
SctQ SpaO Spa33 SepQ YscQ PscQ FliM/FliN
SctK OrgA MxiK YscK
SctL OrgB MxiN EscL YscL PscL FliH
Needle
Length
Regulator
SctP InvJ Spa32 EscP
(Orf16)
YscP PscP FliK
Export
Regulator
SctW InvE MxiC SepL/SepD YopN/TyeA PopN

The core conserved proteins of the T3SS form a double-membrane-spanning syringe-like structure (22, 23), including its extracellular needle-like appendage, and the associated cytoplasmic and membrane-integral secretion machinery (Figure 1). It is these components that are collectively responsible for the delivery of effector proteins into the cytosol of the eukaryotic host cell (24, 25). While our mechanistic model of this cytoplasm-to-cytoplasm secretion system continues to evolve, (ultra-)structural and biochemical characterization of its core components has yielded significant insights.

Figure 1. Gross architecture of the T3SS.

Figure 1

Cryo-EM reconstruction of the Salmonella typhimurium injectisome basal body at subnanometer resolution reveals its overall architecture. (A) Surface representation of the highest resolution cryo-EM map (EMD 1875, contour level 0.0233) published by Marlovits and colleagues (27). Dashed lines indicate the positions of bacterial membranes in vivo. Abbreviations used: OR, outer ring; IR, inner ring; OM, outer membrane; IM, inner membrane. (B) An axial section through the map in (A). (C) Transverse sections through the map in (A) at the level of the neck (top) and IR1 (bottom).

The basal body

The bacterial double membrane and peptidoglycan layer are spanned by a stack of protein annuli known as the basal body (Figure 1). It is comprised of an outer membrane-anchored layer (SctC) and an inner membrane-anchored layer (SctD and SctJ) that interface at a “neck” (26, 27). In electron microscopic (EM) reconstructions of the injectisome, SctC forms two distinct outer rings (OR1 and OR2), SctD and SctJ together form the distal inner ring (IR1), and the cytoplasmic amino-terminus of SctD forms the innermost ring (IR2) (26). The highest resolution cryo-EM models of the Salmonella basal body reveal an overall three-fold rotational symmetry, with a resultant symmetry mismatch between the inner and outer layers: each basal body contains 24 SctD molecules, 24 SctJ molecules, and 15 SctC molecules (27). While the 24-fold symmetry of the inner membrane rings appears conserved across T3SS (28, 29), the stoichiometry of the SctC outer membrane ring may vary between species (12-15 molecules per basal body), such that some systems have an overall 12-fold rotational symmetry (29, 30).

SctC is homologous to the Type II secretion system secretins (5, 31), and like other secretin family members requires a pilotin lipoprotein for its optimal localization and assembly (32-34). The membrane-embedded, β-rich region at the SctC carboxy-terminus can be isolated and has been visualized by EM (30), but it has yet to be characterized at moderate or high-resolution. The periplasmic amino-terminus of SctC contains a modular domain architecture (35) that interacts with the inner membrane ring (36, 37).

SctD and SctJ form the inner membrane rings (26). Each is anchored to the membrane by a single transmembrane helix, and SctJ is additionally lipidated near its amino-terminus (38). Like the amino-terminal periplasmic region of SctC, the periplasmic domains of SctD and SctJ are comprised of a modular multidomain architecture (35). Despite differences in connectivity and little sequence homology, the mixed α/β domains of SctC, SctD, and SctJ show a similar three-dimensional structure: two α-helices pack against the same face of a three strand β-sheet (35, 39). Superhelical crystal packing of the E. coil SctJ periplasmic region provided initial insights into the mechanism of inner membrane ring assembly (38), and the modular arrangement of these domains seems to promote oligomerization (40); however, none of these domains have been shown to clearly form annuli in solution, suggesting that additional constraints (e.g. protein-protein interactions or lipid membrane planarity) are critical for ring formation. Similarly, despite their 1:1 stoichiometry in the basal body, the periplasmic domains of SctD and SctJ have not been crystalized in complex.

Might there be a specific functional advantage for the modular domain architecture common to the SctCDJ periplasmic regions? Recent in situ electron tomography of the Yersinia and Shigella injectisomes shows that the basal body has the ability to stretch in response to osmotic expansion of the periplasmic space (29). This resilience could be of potential importance for the maintenance of intact T3SS injectisomes under physiologic stresses and membrane deformations (29). Molecular dynamic simulations suggest that relative motions of the SctD periplasmic domains could account in part for this flexibility (29), but this hypothesis requires empiric support.

The amino-terminal cytoplasmic domain of SctD forms the innermost ring of the T3SS basal body. High-resolution structural analyses have determined this domain to have a forkhead-associated fold that interacts with cytoplasmic components of the T3SS (41-44). Forkhead-associated domains are β-sandwiches that typically serve as phosphothreonine binding scaffolds, suggesting a means of signal-dependent recruitment of the cytoplasmic secretory apparatus to the basal body. However, the potential phosphopeptide binding residues are not conserved among T3SS, and the precise nature of the interactions between SctD and the cytoplasmic apparatus remains to be determined.

Within the central lumen of the basal body annuli, SctI is believed to form a cylindrical “inner rod” (26) structure that may support the extracellular needle filament. Computational methods have suggested a predominantly α-helical structure for SctI similar to that of the needle filament protomer (below); however, structural analyses of Salmonella and Shigella SctI in solution showed little secondary or tertiary structure (45). It remains to be determined whether SctI can adopt a stable fold within the confines of the basal body. The functional significance of the inner rod in the regulation of needle length and secretion substrate switching will be discussed below.

Direct structural characterization of the T3SS basal body epitomizes the challenges associated with the interrogation of high-molecular weight macromolecular machines: the assembly spans two membranes and a layer of peptidoglycan, and in situ electron tomographic analyses suggest that the basal body is capable of substantial conformational dynamism (29). As direct, high-resolution structural characterization of the assembled T3SS basal body has not yet been possible, a multidisciplinary approach integrating cryo-EM maps, x-ray crystallographic domain structures, biochemical analyses and computer modeling has yielded a high-probability static model for injectisome architecture (46) (Figure 2). Such “hybrid” models will allow the testing of molecular-level hypotheses about ring assembly until the structure of the T3SS basal body has been determined at high-resolution in toto.

Figure 2. Hybrid models of basal body structure.

Figure 2

(A) Computational modeling of the neck (SctC, PDB 3J1V), IR1 (SctD, PDB 3J1X), and IR2 (SctD, PDB 3J1W) annuli of the Salmonella typhimurium basal body. No high-resolution structural information is available for the basal body above the neck. (B) In this model, complementary electrostatic surfaces support ring building, as shown for the SctD periplasmic domains. Note the modular domain architecture (enumerated 1, 2, 3) for SctDperiplasmic.

The inner membrane machinery

Five highly conserved inner membrane proteins (SctRSTUV) are necessary for the function of the pathogenic T3SS; however their individual functions are unclear. It is worth noting these proteins show a high degree of sequence homology to components of the evolutionarily related flagellar T3SS (Table 1), and may represent a functional core, serving critical chemical roles in initiating or powering protein secretion.

Among the SctRSTUV cohort, SctU and SctV possess cytoplasmic domains in addition to their transmembrane helices, and these domains have been best characterized to date. The cytoplasmic region of SctV contains a modular array of small domains (47-49) and crystallographic data suggest SctV may nonamerize (49). Intriguingly, such an oligomer is well suited to fit in a torus of SctV-associated density observed 5-10nm beneath the basal body in EM reconstructions of the T3SS (29, 49). Lea and colleagues (49) have forwarded the hypothesis that this SctV homo-oligomer may serve as a “cage” to facilitate the complete unfolding of folded or partially unfolded secretion substrates, but this possibility has not yet been experimentally validated. The cytoplasmic region of SctU is considerably smaller than that of SctV and contains an autoprotease (50); its potential role in the regulation of secretion will be discussed below.

SctRSTUV are important in the organized, stepwise assembly of the T3SS basal body. Galán and colleagues (51) employed a combination of genetic and structural approaches to show that SctRSTUV help to organize the SctDJ inner membrane rings. Subsequently, the SctC ring and cytoplasmic machinery (below) are recruited, and the inner rod and needle polymers assembled (19). While SctRSTUV are individually not strictly necessary for the formation of the SctCDJ basal body, the efficiency of basal body assembly is significantly ameliorated in their absence (51).

The needle filament

A needle-like filament tens of nanometers in length protrudes from the extracellular face of the T3SS basal body (22, 52). The needle is formed by a helical assembly of the protein SctF, with an outer diameter of 8nm and an inner pore diameter of 2.5nm (53). The apparent similarity of the T3SS basal body and needle filament to a macroscopic syringe makes it tempting to speculate that the T3SS directly injects its substrates into the host cell cytoplasm, with the needle filament serving as a conduit for the passage of partially unfolded effector proteins. Until recently, this hypothesis lacked direct empirical support, and alternative “non-injectisome” models for T3SS effector delivery had been proposed (54). Analyzing substrate-trapped injectisomes by cryo-EM, Marlovits and colleagues (55) and Kolbe and colleagues (56) demonstrated the presence of additional density in the lumen of the T3SS needle filament, consistent with the passage of partially unfolded substrate molecules through the needle.

High-resolution structures of monomeric SctF mutants (57) or chaperone-bound SctF (58, 59), have allowed the characterization of the needle protomer fold. SctF is a hairpin of alpha helices with an intervening conserved PXXP motif. The oligomeric nature of the needle filament had posed a practical barrier to high-resolution structure determination for the intact assembly. Recent hybrid approaches combining cryo-EM with solid-state nuclear magnetic resonance spectroscopy (NMR) and computational modeling have since afforded such a model for both the Salmonella (53) and Shigella (60) needle filaments. In these models, the SctF amino-terminus is oriented towards the convex needle exterior, the carboxy-terminus towards the lumen, and the apex loop connecting the two alpha helices points away from the bacterium. It should be noted that this arrangement is in contrast to prior lower resolution models (61), which oriented the SctF amino-terminus towards the needle lumen. This correction is significant: the orientation of SctF protomers in the solid state NMR models is such that the lumen walls are formed by highly conserved residues, consistent with the passage of secretion substrates through the lumen (53, 60).

Assembly of needle filaments of a given length is necessary for the proper infectivity of T3SS-bearing pathogens, possibly matching the dimensions of host-pathogen adhesion complexes (62). How, though, is the length of the needle filament controlled? SctP regulates the length of the needle filament in several species (63-65). In Yersinia spp., the number of residues in SctP correlates with needle filament length, leading to the hypothesis that SctP functions as a “molecular ruler” (65). That is, SctP might attach at one end to the basal body or cytoplasmic apparatus and at the other end to the growing needle filament, and once SctP was stretched beyond a given length, it would signal to the secretion apparatus to change substrates.

In contrast to the molecular ruler model, work in Salmonella suggest that SctP regulates needle length through control of inner rod assembly (52). Salmonella lacking SctP show decreased density in the inner rod-supporting socket region, lack a polymerized inner rod, and generate long needles (52, 66). Accordingly, Galán and colleagues hypothesized that SctI and SctF are secreted simultaneously and completion of inner rod assembly terminates needle growth in a SctP-dependent fashion. Consistent with this “timer” model of length control (19), overexpression of SctF or SctI leads to longer or shorter needles, respectively (52). Moreover, alanine-scanning mutagenesis analyses of the inner rod protein SctI in Salmonella revealed numerous point mutations that increased needle length without compromising secretory function, perhaps by slowing the rate of inner rod polymerization (66). Intriguingly, the elongate needles generated by most of these mutants remained attached to the basal body (66), in contrast to those produced by sctP deletion mutants, which are easily sheered off (52). This observation is consistent with the hypothesis that the polymerized inner rod joins with the needle filament, anchoring it to the basal body (66).

The needle tip and translocon pore

At the tip of the T3SS needle filament is a pentameric cap formed by the hydrophilic translocator protein (67, 68). The needle tip is believed to interact directly with the host cell surface to facilitate the insertion of a multimeric pore (69, 70), thus completing the cytoplasm-to-cytoplasm protein conduit. The structure and function of the needle tip is of particular biomedical interest, as the hydrophilic translocator protein is a protective antigen in anti-Yersinia vaccine formulations (71) and is targeted by a recently developed passive immunization strategy for the treatment of Pseudomonas aeruginosa infections (72).

X-ray crystallographic analysis of the monomeric tip protein from several species reveals some conserved architectural features (73, 74): the tips of all species show an elongate coiled-coil region and a central mixed α/β subdomain. The overall structure of the tip protein shows some interspecies variation though, with Salmonella/Shigella SipD/IpaD possessing an amino-terminal autochaperoning subdomain (74) lacked by the Yersinia/Pseudomonas LcrV/PcrV tip proteins (19, 70). High-resolution models of the pentameric needle tip are not available, but low-resolution negative stain EM models have offered some insight into its gross architecture. While both appear pentameric, the Shigella tip complex is narrow and elongated relative to that of the Yersinia tip (67, 68), and fitting the Shigella IpaD monomeric crystal structure into the EM map required a significant rearrangement of the mixed α/β domain (68).

Attempts to model the tip protein-needle filament interaction at high-resolution using NMR or x-ray crystallographic data have so far proven challenging, with incompatibilities arising between the proposed models and other data sets (75-77). The crystal structure of a Salmonella SctF-SipD fusion protein identified a potential binding mode for the tip with the needle (75); however, modeling the fusion structure onto the solid state NMR model of the needle filament (53) resulted in steric clashes, suggesting that artifactual constraints imposed by the protein fusion strategy biased the architecture of the complex (77). Regardless, a synthesis of the available data shows that the needle filament interacts with the tip protein at least in part through its elongate coiled-coil, a motif observed in all T3SS tip proteins described to date.

EM and biochemical analyses have shown that the Shigella tip complex is actually a heteropentamer containing four copies of the hydrophilic translocator IpaD and one copy of the hydrophobic translocator IpaB (76). A refined tip model incorporating this insight is similar to previous models, in that the amino- and carboxy-termini of IpaD are oriented towards the needle filament, with portions of the coiled-coil region contacting the needle. However, the refined model presents an IpaD orientation consistent with antibody binding data and the heteropentameric architecture explains the transition from a helical needle filament to a nearly flat-topped tip complex (76).

In contrast to the annular, pentameric tip complexes observed for other T3SS, enteropathogenic E. coli possess a long filamentous needle accessory comprised of EspA, the SipD/IpaD/LcrV/PcrV homologue (78). EspA forms a helical filament similar to the SctF needle, with an internal diameter of approximately 2.5nm (79), suggesting that it functions to extend the T3SS transport conduit. Filling a functional niche similar to needle filament length control in other T3SS, E. coli EspA polymers may adapt the injectisome to reach the target cell membrane beneath the intestinal glycocalyx (70).

The translocon permeating the host cell membrane is formed by the hydrophobic translocators SipB and SipC in Salmonella and their homologues (Table 1). Experiments in red blood cell membranes have shown that the needle tip is crucial for the insertion of the hydrophobic translocators into the host membrane and/or the organization of inserted translocators into functional pores (69, 80). While numerous experimental approaches have been employed to characterize the pore diameter and structure of the translocon (70), direct structural interrogation of native translocons is lacking. Intriguingly, the amino-termini of Salmonella SipB and Shigella IpaB contain extended coiled-coils reminiscent of the colicin family of bacteriocins (81), which are known to function in the delivery of protein toxins across bacterial membranes. However, the precise mechanisms of host cell recognition, membrane insertion, pore formation, and protein translocation remain unclear for the T3SS translocon.

SUBSTRATE RECRUITMENT AND SECRETION

T3SS secrete only a small fraction of the proteins present in the bacterial cytosol (82), and do so in a defined hierarchy. How does the T3SS select its substrates, how is their secretion hierarchy maintained, and what is the mechanism of secretion? A combination of genetic, biochemical, and structural data provide insight into the role of cytoplasmic injectisome-associated proteins in these processes.

Secretion chaperones

The amino-terminal ~100 amino acids of T3SS substrates possess two secretion signals: an unstructured extreme amino-terminus followed by a chaperone-binding region. While the extreme amino-termini of T3SS substrates are highly variable, computational approaches have identified commonalities in the chemical composition of this secretion signal (83-85): the first ~15 amino acids in an effector sequence are enriched in serine, threonine, isoleucine, and proline. Indeed, an effector with a synthetic, amphipathic poly-serine/isoleucine secretion signal was secreted by Yersinia even in the absence of the second secretion signal (its secretion chaperone) (86). However, other studies have shown that chaperone-substrate interactions are necessary for targeting substrates specifically to the injectisome T3SS, as the extreme amino-terminal secretion signal sequence can facilitate injectisome substrate export through the flagellar apparatus in the absence of a chaperone-binding domain. This finding suggests that the extreme amino-terminal secretion signal is evolutionarily “ancient” and shared by both types of T3SS (87).

Downstream of the amino-terminal secretion signal, each secretion substrate is recognized by a specific chaperone protein that maintains the bound region of the substrate in a partially unfolded state (88). It is hypothesized that this nonglobular conformation primes the substrate for secretion through the narrow aperture of the T3SS conduit (89). Chaperones can be classified by their structure and substrate type (90), as follows: Class IA chaperones are mixed α/β homodimers that bind to one effector; Class IB are structurally similar to IA but bind multiple effectors; Class II are alpha helical tetratricopeptide repeat (TPR) proteins that bind to translocon proteins; and Class III are heterodimeric TPR proteins that bind SctF needle filament protomers. It should be noted that while the premature polymerization of SipD/IpaD tip proteins is prevented by their autochaperoning domain (74), tip proteins in Yersinia and Pseudomonas utilize a unique chaperone (LcrG/PcrG) to stabilize their tip protein monomers (91).

Like the extreme amino-terminal secretion signal, the chaperone-binding sequences of secretion substrates are variable. However, Class I chaperone recognition of a conserved β-strand(s) is a common feature of a diverse array of T3SS effectors (92), and it is conserved from animal to plant pathogens (93) (Figure 3). Indeed, the sparse sequence conservation associated with the “β-motif” (92) can be used to recognize previously unknown T3SS effector proteins (94). In contrast, Class II and III TPR chaperones can recognize substrate sequences in either extended unstructured or α-helical conformations; the commonality here is that the TPR concavity is used to bind the substrate (90) (Figure 3).

Figure 3. Chaperone-substrate interactions.

Figure 3

Structural distinctions between effector-chaperone and translocator-chaperone complexes. (A) The structure of the Salmonella effector SipA chaperone-binding domain (CBD, red and yellow) in complex with the Class IB chaperone InvB (dark grey, light grey). PDB 2FM8 (47). The structurally conserved β-motif is highlighted in yellow. (B) The SipA β-motif is bound by a hydrophobic (grey) patch on the InvB surface (blue/grey). (C) Superposition of the CBDs from effectors from multiple species shows a common binding mode marked by the structurally conserved β-motif. The prototypical Class I chaperone SicP is shown in place of the various chaperones. PDB codes: YopN, 1XKP (153); YopE, 1L2W (170); YscM2, 1TTW (171); SptP-SicP, 1JYO (88); SipA, 2FM8 (47); HopA1, 4G6T (93). (D) The Yersinia translocator YopD CBD (red) lacks secondary structure and is bound by the concave cleft of the Class II chaperone SycD (grey). PDB 4AM9 (172).

The aforementioned substrate secretion signals are not only necessary for protein secretion through the T3SS, they are sufficient. Fusion of the secretion signal and chaperone-binding domain from endogenous T3SS substrates to heterologously expressed proteins results in their secretion through the T3SS (95), provided they can be properly unfolded for transit through the needle (55, 56). While this observation is noteworthy for its mechanistic insight into the targeting of virulence factors for secretion, it has allowed the benevolent re-engineering of the system to deliver protective antigens in vaccine design (96) and the large scale production of challenging protein reagents, like spider silk (97).

The ATPase

Both the injectisome and flagellar T3SS include an ATPase with notable sequence homology to the β subunit of the F0F1 ATPase (98). For the injectisome, this ATPase is SctN; for the flagellar apparatus, FliI (Table 1). High-resolution structural analysis of the E. coli SctN catalytic domain showed similarities to V- and F-type ATPases and confirmed that SctN hexamerization would be required for efficient ATP hydrolysis (99). Both FliI (100, 101) and SctN (98, 102) form such oligomers, but neither has been characterized structurally as a catalytically active hexamer.

Interaction of chaperone-substrate complexes with the T3SS ATPase SctN causes the dechaperoning and unfolding of the substrate in an ATP hydrolysis-dependent fashion (103). Given that disruption of tertiary structure is necessary to fit protein substrates in the 2.5nm conduit of the needle filament (55, 56), it is not surprising that loss of function mutations in SctN cause the near complete abrogation of T3SS function (98, 103). Similarly, genomic deletion of fliI severely compromises flagellar filament assembly (104-107). Consistent with the role of SctN in preparing substrates for export, SctN/FliI-dependent density is observed directly beneath the T3SS basal body by EM (108, 109). The partial structural similarity of SctO/FliJ with the γ-subunit of F-type ATPases and the ability of FliJ to stimulate FliI hexamerization (101) has led to the hypothesis that this coiled-coil containing protein (110) might connect oligomeric SctN to the SctV export gate, thus linking the sub-basal body toruses of electron density (49). However, the mere presence of a coiled-coil is insufficient evidence to ascribe a γ-like function to SctO (19), and its precise role remains to be determined.

Given the ATPase activity of SctN/FliI, it is tempting to speculate that ATP hydrolysis provides the free energy for protein secretion; however, chemical and genetic analyses show that SctN/FliI is not the sole energizer of the T3SS. Experiments with the protonophore CCCP have shown that the inner membrane proton motive force is necessary for secretion by both injectisome (111) and flagellar (104, 105) T3SS. Additionally, the flagellar assembly defect of ATPase deletion mutants can be at least partially corrected by mutations that alter the export apparatus, increase substrate levels, or increase the magnitude of the proton motive force (104-106). A similar result was recently reported for SPI-1 T3SS in Salmonella (106). These results suggest that under sufficiently permissive conditions, the actual transit of substrates into and/or through the conduit is powered by the proton motive force. However, one must interpret these results with some caution, as ionophores can significantly perturb cellular physiology (112) and SctN-independent injectisome secretion of a substrate requiring dechaperoning has not yet been demonstrated (106). A reasonable synthesis of the available data might surmise that both ATP hydrolysis and the proton motive force are important for energizing T3SS: the SctN/FliI ATPase functions to dechaperone and begin unfolding the secretion substrates with optimal efficiency under (non-ideal) physiologic conditions, while the proton motive force is responsible for the apical transit of the nonglobular substrate (106).

Regardless of the quantitative contributions of either energy source, the mechanics of secretion remain poorly understood. There are no available high-resolution structural models of the interaction between chaperone-substrate complexes and the SctN ATPase. One computational model suggests a mode of ATPase-chaperone interaction based on structural similarities between Class I chaperones and the F-type ATPase γ-subunit (113). While this model and the accompanying biochemical data are consistent with the observation that relatively carboxy-terminal residues of SctN interact with chaperones (113), its structural accuracy lacks empiric support. Recent SAXS data suggest an alternative model for the interaction of substrates, chaperones, and the ATPase. Complexes of the Salmonella effector-chaperone pair SopB-SigE are able to hexamerize in a concentration dependent manner with dimensions comparable to the hexameric models of the ATPase (114). While it is too early to say whether other chaperone-effector complexes can oligomerize (115), whether these oligomers can interact with SctN, or whether such interactions — even if possible — are physiologically relevant, these results raise the possibility of alternate ATPase-cargo stoichiometries.

Recent solution NMR analyses suggest an interesting role for chaperone structure in the targeting of substrates to the T3SS. In solution, the E. coli chaperone CesAB is a partially folded molten globule (116) that does not interact with the hexameric SctN (117). However, upon binding to its substrate — the EspA tip filament protein — CesAB becomes fully structured and is able to bind SctN (117). The binding site for SctN was mapped onto the CesAB-EspA heterodimer, where it covered regions of CesAB unstructured in the absence of substrate (117). Consistent with the hypothesis that substrate-induced folding of the chaperone allows for targeting to the ATPase, a mutation-stabilized, structured CesAB homodimer was able to bind SctN in the absence of substrate (117). While it remains to be determined whether similar disorder-to-order transitions effect SctN binding for other chaperone classes, these results are consistent with the role of chaperone-substrate interactions in targeting substrates for secretion. Additionally, the observation that substrate-chaperone complexes are recognized by hexameric SctN, but not its amino-terminally truncated monomeric form, suggests that ATPase hexamerization is critical for both hydrolytic catalysis and substrate recognition (117).

The sorting platform

Located at the peripheral cytoplasmic face of the flagellar basal body is a ring of robust density in EM reconstructions (109). Known as the “C-ring,” this annulus is composed of the flagellar proteins FliM, FliN, and FliG, and it plays a role in flagellar motor function (torque generation) and rotational switching (20). FliM and FliN have an injectisome homologue with some conserved domains (SctQ), but torque generation by the injectisome is controversial (118) and a robust C-ring is absent in tomographic reconstructions of the injectisome basal body (29, 109). However, immuno-EM analysis of purified Shigella injectisomes shows localization of SctQ to the cytoplasmic face (119), suggesting that it plays some role in protein secretion or the regulation of the secretory process. Indeed, recent cryo-electron tomographic analyses of the Shigella injectisome have identified six SctQ-dependent “pods” of density proximal to the cytoplasmic face of the basal body, forming a structure distinct from that of the flagellar C-ring (120).

Seminal biochemical and genetics work by Galán and colleagues revealed that SctQ forms a critical “sorting platform” for the T3SS (121). Affinity purification of SctQ from secretion-competent Salmonella produces high molecular weight complexes containing the SctN ATPase, regulatory proteins, chaperones and secretion substrates (121). Most notably, the sorting platform plays a role in the hierarchical secretion of substrates, queuing substrates in their appropriate order. For example, in Salmonella with assembled injectisomes, the sorting platform is predominantly occupied by translocon proteins, but genomic deletion of the translocators allows the next tier of substrates (effector proteins) to access the sorting platform (121).

In addition to SctQ, formation of the sorting platform requires the proteins SctK and SctL (121). While the role of SctK is at present unclear, biochemical analyses of the flagellar apparatus shed light on the potential function of SctL. SctL is a homologue of the flagellar protein FliH (Table 1). The SctL/FliH family is predicted to have a conserved domain architecture: an amino-terminal disordered region is followed by a coiled-coil and then a mixed α/β domain (122). The carboxy-terminus of FliH has interacts with the amino-terminal oligomerization domain of FliI (122), inhibiting its ATPase activity (123). While sequence similarities between the FliH carboxy-terminal domain and the F-type ATPase δ-subunit suggest a role for FliH as a “stator” (124), it is not obvious that FliH interacts with oligomeric FliI, and the structural details of this interaction are not yet known. The amino-terminus of FliH interacts with FliN (125), and this FliH-FliN interaction is important (126) — if not absolutely necessary (127) — for the recruitment of FliI to the export apparatus. Given that the homologous injectisome complex (SctQ-SctL-SctN) forms a portion of the sorting platform and that chaperone-substrate complexes interact with the ATPase, these data suggest that one function of the SctQ sorting platform could be to localize chaperone-effector-ATPase complexes to the injectisome export apparatus. Indeed, Minamino, Namba and colleagues have hypothesized that the FliI ATPase exists in two forms: an ATP-hydrolyzing hexamer and a dynamic substrate-carrying monomer bound to FliH and the C-ring (128). Similarly, SctQ-injectisome interactions are dynamic in Yersinia, as injectisome-associated SctQ exchanges with a cytoplasmic pool with a half-time of approximately one minute (129).

Structural models of SctQ are a work in progress, and have focused to date on the carboxy-terminus of the molecule. In Pseudomonas syringae, SctQ is spread over two open reading frames (hrcQA and hrcQB), much like FliM and FliN in the flagellar system. The structure of the carboxy-terminal domain of HrcQB is quite similar to that of the carboxy-terminal domain of FliN (130, 131); both domains are homodimers of the “surface presentation of antigens” (SpoA) fold. The folded core of each protomer is an antiparallel β-sheet, and a loop from each protomer containing a β-strand and α-helix wraps around the β-sheet core of the other protomer. In Yersinia, SctQ is the product of a single open reading frame (as in most injectisomes), but the carboxy-terminal SpoA domain is duplicitously translated from an internal translation start site (132). The homodimer produced by this translation product is able to interact with full-length SctQ and, at least in the Yersinia system, is necessary for secretion in vivo. In both the flagellar (133, 134) and injectisome (135) systems, this SpoA domain tetramerizes as a dimer of dimers, but appears to do so in different orientations in each system. Cross-linking analyses suggest that the FliN SpoA tetramers form a “doughnut” at the base of the C-ring (133), but high-resolution support for this arrangement is lacking.

Despite the progress that has been made, numerous structural questions remain unanswered for the SctQ sorting platform. The function of the SpoA domain is unclear, SctQ-SctQ(SpoA) interactions have yet to be structurally characterized, and the structural basis for the interaction of SctL with SctQ has not yet been determined. Moreover, while the amino-terminal domains of FliM have well characterized functions in the regulation of flagellar rotation switching (136, 137), these motor functions are likely flagella-specific and involve interactions with partners not conserved from the flagellar apparatus to the injectisome (e.g. FliG, CheY). Thus, the function of the SctQ amino-terminus is also unclear. Lastly, how and when SctQ or its soluble interaction partners interface with the basal body or export apparatus remains to be determined.

Substrate switching

T3SS substrates are secreted in a defined order that is necessary for the proper assembly and function of the system (138, 139): secretion of the needle filament (SctF) and inner rod (SctI) is followed by secretion of the needle tip protein and translocon pore proteins, which is followed by the secretion of effector proteins. Thus, it seems that there are several sequential substrate “switching” events that must occur for the hierarchical secretion of substrates to be maintained (139).

The first such switching event halts the extension of the growing needle filament upon completion of the inner rod and allows for secretion of the needle tip protein and translocators. As discussed above, the length of the needle filament is controlled by the assembly of the inner rod in a SctP-dependent fashion (52, 66). Full deletion of sctP locks the T3SS into a mode of exclusive SctF filament secretion; that is, deletion of sctP results in not only elongate needles, but a lack of translocon and effector secretion (63, 65, 140, 141). Indeed, deletion of sctP results in the absence of translocon components from the SctQ sorting platform (121). However, small deletions in the amino-terminal regions of SctP alter needle length without compromising translocon secretion, suggesting that some portion of SctP performs a crucial switching function (141). Deletions within the conserved mixed α/β region at the carboxy-terminus of the protein compromise translocon secretion (in addition to disrupting needle length regulation), and this presumptive domain has been termed the “type III secretion substrate specificity switch” (T3S4) domain (141).

The three-dimensional structure of an injectisome T3S4 domain has not yet been determined, but the flagellar FliK T3S4 domain has been solved by NMR (142). The carboxy-terminal domain of FliK possesses two α-helices folded against a four-strand β-sheet (142), and the predicted structural conservation of these secondary structural elements in SctP suggests that this model may be generalizable to the injectisome. While it is still unclear at the molecular level how SctP functions to promote specificity switching, its interaction partners suggest some viable hypotheses. For example, the SctP T3S4 domain interacts with the SctO protein (143), suggesting that it may be able to transmit regulatory information to the SctN ATPase or the SctV export gate. Moreover, the T3S4 domain interacts with the cytoplasmic autoprotease domain of SctU (144). Like SctP, SctU regulates the secretion of the inner rod protein (145). The interaction between these two proteins is intriguing given that SctU interacts with components of the SctQ sorting platform and the SctV export apparatus (146), again suggesting mechanisms for the relay of switching information throughout the secretory apparatus.

The second major switching event distinguishes between translocon components and effector proteins (121). Deletion of translocon components allows for the localization of effector proteins to the sorting platform, consistent with a model where a gradient of substrate and/or chaperone affinities for the sorting platform controls the hierarchy of secretion (121). The identification of several classes of secretion apparatus mutants that can secrete effectors but not translocon proteins offers some insights into the establishment of secretion hierarchy. Deletion of sctW in Salmonella results in the specific loss of translocon component secretion (138). SctW binds the translocon proteins and their chaperone in Salmonella (SicA) (138), and it is necessary for translocator binding to the SctQ sorting platform (121). These observations are consistent with SctW enhancing the affinity of translocon-containing complexes for the sorting platform. However, recent genetics data suggest the mechanism of hierarchy control for SctW may be more complex. A subset of the SctI alanine mutants identified by Lefebre and Galán (66) have normal needle lengths but phenocopy sctW deletion, and an interaction between SctW and SctI was recently reported in Shigella (147). Together, these data raise the possibility of SctW binding not only the sorting platform but also portions of the basal body.

Further clouding the role of SctW in T3SS is the observation of species-specific effects of sctW mutation. In Yersinia and Shigella, SctW is secreted and sctW deletion does not specifically impair translocon protein secretion (148, 149). Moreover, the Yersinia SctW protein pair YopN/TyeA is part of a complex calcium response apparatus in the bacterial cytosol (150) that involves several Yersinia-specific proteins (151, 152). While the structures of the Shigella and Yersinia SctW homologues have been determined (153, 154), a fuller understanding of SctW function (and its species-specific nuances) will require structural characterization in complex with other injectisome components.

In addition to its role in the first switching event, SctU is also involved in the second switch. The cytoplasmic domain of the SctU family autocatalyzes cleavage between the asparagine and proline residues of its conserved NPTH cleavage site (139). Alanine mutations on either side of the cleavage site cause aberrant specificity switching: translocon proteins are no longer secreted but effector secretion remains intact (50, 148). An amphipathic linker connects the SctU transmembrane region to the cytoplasmic autoprotease, and this linker undergoes a disorder-to-order transition in the presence of anionic lipids (155). Introducing charge-altering mutations in the linker impaired T3SS function, suggesting that the ordering of the SctU linker against the bacterial inner membrane is crucial, perhaps favorably orienting the autoprotease domain for interactions with other members of the export apparatus (155). As mentioned above, SctU interacts with multiple members of the sorting platform, but the bases for these interactions — and the mechanisms by which they would effect specificity switching — are unclear.

Control of secretion

The T3SS can assemble a basal body, needle and tip, and then pause in a “primed” state until the relevant stimulus arrives and secretion resumes. This strategy prevents the wanton waste of translocon and effector proteins. Interrogating this additional level of complexity is important to our full understanding of the pathobiology of T3SS, and may suggest routes to anti-virulence compounds that prevent the activation of otherwise structurally competent injectisomes.

Bile salts play a regulatory role in the T3SS used by several enteric pathogens. The interaction of bile salts with the Shigella tip complex promotes IpaB recruitment to the tip, forming the heteropentameric tip complex described above (76, 156). In contrast, bile salts suppress Salmonella SPI-1 T3SS function (157). These observations provide an intriguing correlation between host gastrointestinal physiology and pathogen virulence that ties environmental factors to the species-specific adaptation of the T3SS. Despite reports describing the interaction of bile salts with monomeric Shigella IpaD (158) and Salmonella SipD (75, 159), the structural basis for bile salt interaction with the intact tip complex has yet to be determined in either species, and so the mechanism of its regulatory activity remains unclear.

Contact with host cells stimulates the activation of T3SS in several species (160-162). In Salmonella, contact with target cells stimulates the secretion of the translocon proteins SipB and SipC (163), and in Shigella, interaction of the IpaD-IpaB tip with liposomes resembling host cell membranes induces IpaC secretion (164). It is tempting to speculate that contact of the needle tip with the host cell sends a mechanical signal to the basal body and/or export apparatus that reinitiates secretion (6, 19, 165). As the connecting factor between the host cell surface and the basal body, the needle filament itself is a promising candidate for force transduction. Specific needle filament protein mutations can trap the Shigella T3SS in a constitutively active secretion mode, and one might hypothesize that these mutations stabilize needle filaments in a post-contact activated conformation (166). However, the filaments formed by these mutants do not exhibit the gross conformational changes one might expect if the needle filament architecture were transducing this signal (166). Alternatively, local changes in the tip environment may permit the secretion of substrates trapped within the needle by a closed tip, restarting secretion without requiring signal transduction to the bacterial cytoplasm or basal body (165).

Work from the Salmonella SPI-2 T3SS suggests a tantalizing third (and non-mutually exclusive) possibility, that the needle is not only a conduit for protein secretion, but a passageway for the diffusion of chemical signals (167). Salmonella makes use of two T3SS: broadly, the SPI-1 T3SS promotes cell invasion and subsequently the SPI-2 T3SS facilitates the formation of the Salmonella Containing Vacuole (SCV), an intracellular environment for Salmonella survival and replication (168). Holden and colleagues noted that priming of the SPI-2 T3SS requires exposure of the bacteria to low pH (as would be experienced in the endosomal compartment), but that triggering of effector secretion required a return to neutral pH (167). It is noteworthy that the SPI-2 SctW protein was required for this transition (167), consistent with the apparent role of SctW in translocon-to-effector specificity switching in other systems (above). However, it is most intriguing that this switch required intact translocon components, suggesting that the neutral pH signal may be transduced from the host cell cytosol, through the translocon and needle, to the basal body and/or export apparatus (167).

In light of these findings from Salmonella SPI-2, one wonders whether the needle conduit could serve as a channel for other small molecule signals from the host cytosol: perhaps the needle of Yersinia spp. transmits the decreased calcium ion concentration of the eukaryotic cytosol, resulting in T3SS reactivation. Intriguingly, Plano and colleagues have shown that point mutations in the Yersinia needle filament protein result in calcium-independent, constitutive T3SS (169); however, the mechanism of this altered calcium response is unclear. One might hypothesize that these mutant needles are unable to associate with tip or translocator proteins, preventing access to — and transmission of — the low calcium environment of the host cytosol. Consistent with this hypothesis, a subset of the constitutively secreting mutants were deficient in delivery of T3SS substrates to the host (169). Alternatively, extracellular calcium might regulate T3SS by altering the conformation of the needle filament protein (akin to the aforementioned force transduction hypothesis), and these mutants might simply be unable to assume some putative calcium-dependent conformations.

FUTURE PROSPECTS

In the past 20 years, our models of T3SS structure and function have evolved substantially, from the first visualization of the injectisome architecture, to the high-resolution structural interrogation of many of its individual components. Combining these insights with a plethora of genetic and biochemical data, the molecular mechanics of this astounding secretory nanomachine are coming into focus. However, numerous questions remain — the answers to which are critical to our understanding of bacterial virulence, the design of new therapeutics, and the imaginative re-engineering of the system.

Despite the improvements in cryo-EM models of the injectisome, the precise architecture of the membrane embedded components of the T3SS is still unclear, as is the structural basis for their interactions with the soluble components of the system. The native structures of the filament-bound needle tip and the translocon in the host membrane must be determined to understand how the extracellular environment regulates secretion, how proteins penetrate to the host cytosol, and how to rationally design secretion-blocking vaccines. Although the constituents of the cytoplasmic sorting platform have been identified, the structural bases for their interactions are unknown: how the sorting platform assembles, how substrate-chaperone complexes engage the system, and how the numerous regulatory elements interact to govern a secretory hierarchy all remain to be determined.

Answering these questions is likely to require a hybrid approach, characterizing local interactions and large assemblies alike, and employing a range of structural and molecular techniques. However, it is clear that high-resolution models of intact macromolecular assemblies (e.g. basal body, needle tip, translocon pore) would greatly advance the field. Much like the role that atomic models of the ribosome have played in the interrogation of its multiple functional states, one can imagine the watershed of insight that would come from successful visualization of the injectisome or sorting platform in each of their several forms: needle-assembling, translocator-secreting, and effector-secreting. Ideally, these mechanistic insights will allow the uncoupling of some pathogenic gram-negative bacteria from virulence and the re-engineering of the nanosyringe for the benefit of biotechnology.

ACKNOWLEDGMENTS

Work in the laboratory of C.E.S. is funded in part by the NIH and research funds from the Rockefeller University. R.Q.N. was supported by the Hearst Foundation and by a Medical Scientist Training Program grant from the National Institute of General Medical Sciences of the National Institutes of Health under award number T32GM07739 to the Weill Cornell/Rockefeller/Sloan-Kettering Tri-Institutional MD-PhD Program.

REFERENCES

  • 1.Galan JE, Collmer A. Type III secretion machines: bacterial devices for protein delivery into host cells. Science. 1999;284:1322–1328. doi: 10.1126/science.284.5418.1322. [DOI] [PubMed] [Google Scholar]
  • 2.Gauthier A, Thomas NA, Finlay BB. Bacterial injection machines. J Biol Chem. 2003;278:25273–25276. doi: 10.1074/jbc.R300012200. [DOI] [PubMed] [Google Scholar]
  • 3.Galan JE, Curtiss R., 3rd Cloning and molecular characterization of genes whose products allow Salmonella typhimurium to penetrate tissue culture cells. Proc Natl Acad Sci U S A. 1989;86:6383–6387. doi: 10.1073/pnas.86.16.6383. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Michiels T, Wattiau P, Brasseur R, Ruysschaert JM, Cornelis G. Secretion of Yop proteins by Yersiniae. Infect Immun. 1990;58:2840–2849. doi: 10.1128/iai.58.9.2840-2849.1990. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Michiels T, Vanooteghem JC, Lambert de Rouvroit C, China B, Gustin A, Boudry P, Cornelis GR. Analysis of virC, an operon involved in the secretion of Yop proteins by Yersinia enterocolitica. J Bacteriol. 1991;173:4994–5009. doi: 10.1128/jb.173.16.4994-5009.1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Galan JE, Wolf-Watz H. Protein delivery into eukaryotic cells by type III secretion machines. Nature. 2006;444:567–573. doi: 10.1038/nature05272. [DOI] [PubMed] [Google Scholar]
  • 7.Cornelis GR. Yersinia type III secretion: send in the effectors. J Cell Biol. 2002;158:401–408. doi: 10.1083/jcb.200205077. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Francis MS, Wolf-Watz H, Forsberg A. Regulation of type III secretion systems. Curr Opin Microbiol. 2002;5:166–172. doi: 10.1016/s1369-5274(02)00301-6. [DOI] [PubMed] [Google Scholar]
  • 9.Waterman SR, Holden DW. Functions and effectors of the Salmonella pathogenicity island 2 type III secretion system. Cell Microbiol. 2003;5:501–511. doi: 10.1046/j.1462-5822.2003.00294.x. [DOI] [PubMed] [Google Scholar]
  • 10.Stebbins CE. Structural insights into bacterial modulation of the host cytoskeleton. Curr Opin Struct Biol. 2004;14:731–740. doi: 10.1016/j.sbi.2004.09.011. [DOI] [PubMed] [Google Scholar]
  • 11.Yahr TL, Wolfgang MC. Transcriptional regulation of the Pseudomonas aeruginosa type III secretion system. Mol Microbiol. 2006;62:631–640. doi: 10.1111/j.1365-2958.2006.05412.x. [DOI] [PubMed] [Google Scholar]
  • 12.Bhavsar AP, Guttman JA, Finlay BB. Manipulation of host-cell pathways by bacterial pathogens. Nature. 2007;449:827–834. doi: 10.1038/nature06247. [DOI] [PubMed] [Google Scholar]
  • 13.Ellermeier JR, Slauch JM. Adaptation to the host environment: regulation of the SPI1 type III secretion system in Salmonella enterica serovar Typhimurium. Curr Opin Microbiol. 2007;10:24–29. doi: 10.1016/j.mib.2006.12.002. [DOI] [PubMed] [Google Scholar]
  • 14.Galan JE. SnapShot: effector proteins of type III secretion systems. Cell. 2007;130:192. doi: 10.1016/j.cell.2007.06.042. [DOI] [PubMed] [Google Scholar]
  • 15.Buttner D. Protein export according to schedule: architecture, assembly, and regulation of type III secretion systems from plant- and animal-pathogenic bacteria. Microbiol Mol Biol Rev. 2012;76:262–310. doi: 10.1128/MMBR.05017-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Burkinshaw BJ, Strynadka NC. Assembly and structure of the T3SS. Biochim Biophys Acta. 2014;1843:1649–1663. doi: 10.1016/j.bbamcr.2014.01.035. [DOI] [PubMed] [Google Scholar]
  • 17.Hueck CJ. Type III protein secretion systems in bacterial pathogens of animals and plants. Microbiol Mol Biol Rev. 1998;62:379–433. doi: 10.1128/mmbr.62.2.379-433.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Abrusci P, McDowell MA, Lea SM, Johnson S. Building a secreting nanomachine: a structural overview of the T3SS. Curr Opin Struct Biol. 2014;25:111–117. doi: 10.1016/j.sbi.2013.11.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Galan JE, Lara-Tejero M, Marlovits TC, Wagner S. Bacterial type III secretion systems: specialized nanomachines for protein delivery into target cells. Annu Rev Microbiol. 2014;68:415–438. doi: 10.1146/annurev-micro-092412-155725. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Macnab RM. How bacteria assemble flagella. Annu Rev Microbiol. 2003;57:77–100. doi: 10.1146/annurev.micro.57.030502.090832. [DOI] [PubMed] [Google Scholar]
  • 21.Gophna U, Ron EZ, Graur D. Bacterial type III secretion systems are ancient and evolved by multiple horizontal-transfer events. Gene. 2003;312:151–163. doi: 10.1016/s0378-1119(03)00612-7. [DOI] [PubMed] [Google Scholar]
  • 22.Kubori T, Matsushima Y, Nakamura D, Uralil J, Lara-Tejero M, Sukhan A, Galan JE, Aizawa SI. Supramolecular structure of the Salmonella typhimurium type III protein secretion system. Science. 1998;280:602–605. doi: 10.1126/science.280.5363.602. [DOI] [PubMed] [Google Scholar]
  • 23.Blocker A, Jouihri N, Larquet E, Gounon P, Ebel F, Parsot C, Sansonetti P, Allaoui A. Structure and composition of the Shigella flexneri “needle complex”, a part of its type III secreton. Mol Microbiol. 2001;39:652–663. doi: 10.1046/j.1365-2958.2001.02200.x. [DOI] [PubMed] [Google Scholar]
  • 24.Schlumberger MC, Muller AJ, Ehrbar K, Winnen B, Duss I, Stecher B, Hardt WD. Real-time imaging of type III secretion: Salmonella SipA injection into host cells. Proc Natl Acad Sci U S A. 2005;102:12548–12553. doi: 10.1073/pnas.0503407102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Van Engelenburg SB, Palmer AE. Imaging type-III secretion reveals dynamics and spatial segregation of Salmonella effectors. Nat Methods. 2010;7:325–330. doi: 10.1038/nmeth.1437. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Marlovits TC, Kubori T, Sukhan A, Thomas DR, Galan JE, Unger VM. Structural insights into the assembly of the type III secretion needle complex. Science. 2004;306:1040–1042. doi: 10.1126/science.1102610. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Schraidt O, Marlovits TC. Three-dimensional model of Salmonella's needle complex at subnanometer resolution. Science. 2011;331:1192–1195. doi: 10.1126/science.1199358. [DOI] [PubMed] [Google Scholar]
  • 28.Hodgkinson JL, Horsley A, Stabat D, Simon M, Johnson S, da Fonseca PC, Morris EP, Wall JS, Lea SM, Blocker AJ. Three-dimensional reconstruction of the Shigella T3SS transmembrane regions reveals 12-fold symmetry and novel features throughout. Nat Struct Mol Biol. 2009;16:477–485. doi: 10.1038/nsmb.1599. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Kudryashev M, Stenta M, Schmelz S, Amstutz M, Wiesand U, Castano-Diez D, Degiacomi MT, Munnich S, Bleck CK, Kowal J, Diepold A, Heinz DW, Dal Peraro M, Cornelis GR, Stahlberg H. In situ structural analysis of the Yersinia enterocolitica injectisome. Elife. 2013;2:e00792. doi: 10.7554/eLife.00792. doi: 10.7554/eLife.00792. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Kowal J, Chami M, Ringler P, Muller SA, Kudryashev M, Castano-Diez D, Amstutz M, Cornelis GR, Stahlberg H, Engel A. Structure of the dodecameric Yersinia enterocolitica secretin YscC and its trypsin-resistant core. Structure. 2013;21:2152–2161. doi: 10.1016/j.str.2013.09.012. [DOI] [PubMed] [Google Scholar]
  • 31.Kaniga K, Bossio JC, Galan JE. The Salmonella typhimurium invasion genes invF and invG encode homologues of the AraC and PulD family of proteins. Mol Microbiol. 1994;13:555–568. doi: 10.1111/j.1365-2958.1994.tb00450.x. [DOI] [PubMed] [Google Scholar]
  • 32.Crago AM, Koronakis V. Salmonella InvG forms a ring-like multimer that requires the InvH lipoprotein for outer membrane localization. Mol Microbiol. 1998;30:47–56. doi: 10.1046/j.1365-2958.1998.01036.x. [DOI] [PubMed] [Google Scholar]
  • 33.Daefler S, Russel M. The Salmonella typhimurium InvH protein is an outer membrane lipoprotein required for the proper localization of InvG. Mol Microbiol. 1998;28:1367–1380. doi: 10.1046/j.1365-2958.1998.00908.x. [DOI] [PubMed] [Google Scholar]
  • 34.Burghout P, Beckers F, de Wit E, van Boxtel R, Cornelis GR, Tommassen J, Koster M. Role of the pilot protein YscW in the biogenesis of the YscC secretin in Yersinia enterocolitica. J Bacteriol. 2004;186:5366–5375. doi: 10.1128/JB.186.16.5366-5375.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Spreter T, Yip CK, Sanowar S, Andre I, Kimbrough TG, Vuckovic M, Pfuetzner RA, Deng W, Yu AC, Finlay BB, Baker D, Miller SI, Strynadka NC. A conserved structural motif mediates formation of the periplasmic rings in the type III secretion system. Nat Struct Mol Biol. 2009;16:468–476. doi: 10.1038/nsmb.1603. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Sanowar S, Singh P, Pfuetzner RA, Andre I, Zheng H, Spreter T, Strynadka NC, Gonen T, Baker D, Goodlett DR, Miller SI. Interactions of the transmembrane polymeric rings of the Salmonella enterica serovar Typhimurium type III secretion system. MBio. 2010;1:e00158–10. doi: 10.1128/mBio.00158-10. doi: 10.1128/mBio.00158-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Schraidt O, Lefebre MD, Brunner MJ, Schmied WH, Schmidt A, Radics J, Mechtler K, Galan JE, Marlovits TC. Topology and organization of the Salmonella typhimurium type III secretion needle complex components. PLoS Pathog. 2010;6:e1000824. doi: 10.1371/journal.ppat.1000824. doi: 10.1371/journal.ppat.1000824. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Yip CK, Kimbrough TG, Felise HB, Vuckovic M, Thomas NA, Pfuetzner RA, Frey EA, Finlay BB, Miller SI, Strynadka NC. Structural characterization of the molecular platform for type III secretion system assembly. Nature. 2005;435:702–707. doi: 10.1038/nature03554. [DOI] [PubMed] [Google Scholar]
  • 39.Marlovits TC, Stebbins CE. Type III secretion systems shape up as they ship out. Curr Opin Microbiol. 2010;13:47–52. doi: 10.1016/j.mib.2009.11.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Bergeron JR, Worrall LJ, De S, Sgourakis NG, Cheung AH, Lameignere E, Okon M, Wasney GA, Baker D, McIntosh LP, Strynadka NC. The modular structure of the inner-membrane ring component PrgK facilitates assembly of the type III secretion system basal body. Structure. 2015;23:161–172. doi: 10.1016/j.str.2014.10.021. [DOI] [PubMed] [Google Scholar]
  • 41.McDowell MA, Johnson S, Deane JE, Cheung M, Roehrich AD, Blocker AJ, McDonnell JM, Lea SM. Structural and functional studies on the N-terminal domain of the Shigella type III secretion protein MxiG. J Biol Chem. 2011;286:30606–30614. doi: 10.1074/jbc.M111.243865. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Barison N, Lambers J, Hurwitz R, Kolbe M. Interaction of MxiG with the cytosolic complex of the type III secretion system controls Shigella virulence. FASEB J. 2012;26:1717–1726. doi: 10.1096/fj.11-197160. [DOI] [PubMed] [Google Scholar]
  • 43.Gamez A, Mukerjea R, Alayyoubi M, Ghassemian M, Ghosh P. Structure and interactions of the cytoplasmic domain of the Yersinia type III secretion protein YscD. J Bacteriol. 2012;194:5949–5958. doi: 10.1128/JB.00513-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Lountos GT, Tropea JE, Waugh DS. Structure of the cytoplasmic domain of Yersinia pestis YscD, an essential component of the type III secretion system. Acta Crystallogr D Biol Crystallogr. 2012;68:201–209. doi: 10.1107/S0907444911054308. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Zhong D, Lefebre M, Kaur K, McDowell MA, Gdowski C, Jo S, Wang Y, Benedict SH, Lea SM, Galan JE, De Guzman RN. The Salmonella type III secretion system inner rod protein PrgJ is partially folded. J Biol Chem. 2012;287:25303–25311. doi: 10.1074/jbc.M112.381574. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Bergeron JR, Worrall LJ, Sgourakis NG, DiMaio F, Pfuetzner RA, Felise HB, Vuckovic M, Yu AC, Miller SI, Baker D, Strynadka NC. A refined model of the prototypical Salmonella SPI-1 T3SS basal body reveals the molecular basis for its assembly. PLoS Pathog. 2013;9:e1003307. doi: 10.1371/journal.ppat.1003307. doi: 10.1371/journal.ppat.1003307. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Lilic M, Quezada CM, Stebbins CE. A conserved domain in type III secretion links the cytoplasmic domain of InvA to elements of the basal body. Acta Crystallogr D Biol Crystallogr. 2010;66:709–713. doi: 10.1107/S0907444910010796. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Worrall LJ, Vuckovic M, Strynadka NC. Crystal structure of the C-terminal domain of the Salmonella type III secretion system export apparatus protein InvA. Protein Sci. 2010;19:1091–1096. doi: 10.1002/pro.382. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Abrusci P, Vergara-Irigaray M, Johnson S, Beeby MD, Hendrixson DR, Roversi P, Friede ME, Deane JE, Jensen GJ, Tang CM, Lea SM. Architecture of the major component of the type III secretion system export apparatus. Nat Struct Mol Biol. 2013;20:99–104. doi: 10.1038/nsmb.2452. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Zarivach R, Deng W, Vuckovic M, Felise HB, Nguyen HV, Miller SI, Finlay BB, Strynadka NC. Structural analysis of the essential self-cleaving type III secretion proteins EscU and SpaS. Nature. 2008;453:124–127. doi: 10.1038/nature06832. [DOI] [PubMed] [Google Scholar]
  • 51.Wagner S, Konigsmaier L, Lara-Tejero M, Lefebre M, Marlovits TC, Galan JE. Organization and coordinated assembly of the type III secretion export apparatus. Proc Natl Acad Sci U S A. 2010;107:17745–17750. doi: 10.1073/pnas.1008053107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Marlovits TC, Kubori T, Lara-Tejero M, Thomas D, Unger VM, Galan JE. Assembly of the inner rod determines needle length in the type III secretion injectisome. Nature. 2006;441:637–640. doi: 10.1038/nature04822. [DOI] [PubMed] [Google Scholar]
  • 53.Loquet A, Sgourakis NG, Gupta R, Giller K, Riedel D, Goosmann C, Griesinger C, Kolbe M, Baker D, Becker S, Lange A. Atomic model of the type III secretion system needle. Nature. 2012;486:276–279. doi: 10.1038/nature11079. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Edgren T, Forsberg A, Rosqvist R, Wolf-Watz H. Type III secretion in Yersinia: injectisome or not? PLoS Pathog. 2012;8:e1002669. doi: 10.1371/journal.ppat.1002669. doi: 10.1371/journal.ppat.1002669. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Radics J, Konigsmaier L, Marlovits TC. Structure of a pathogenic type 3 secretion system in action. Nat Struct Mol Biol. 2014;21:82–87. doi: 10.1038/nsmb.2722. [DOI] [PubMed] [Google Scholar]
  • 56.Dohlich K, Zumsteg AB, Goosmann C, Kolbe M. A substrate-fusion protein is trapped inside the Type III Secretion System channel in Shigella flexneri. PLoS Pathog. 2014;10:e1003881. doi: 10.1371/journal.ppat.1003881. doi: 10.1371/journal.ppat.1003881. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Deane JE, Roversi P, Cordes FS, Johnson S, Kenjale R, Daniell S, Booy F, Picking WD, Picking WL, Blocker AJ, Lea SM. Molecular model of a type III secretion system needle: Implications for host-cell sensing. Proc Natl Acad Sci U S A. 2006;103:12529–12533. doi: 10.1073/pnas.0602689103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Cordes FS, Komoriya K, Larquet E, Yang S, Egelman EH, Blocker A, Lea SM. Helical structure of the needle of the type III secretion system of Shigella flexneri. J Biol Chem. 2003;278:17103–17107. doi: 10.1074/jbc.M300091200. [DOI] [PubMed] [Google Scholar]
  • 59.Poyraz O, Schmidt H, Seidel K, Delissen F, Ader C, Tenenboim H, Goosmann C, Laube B, Thunemann AF, Zychlinsky A, Baldus M, Lange A, Griesinger C, Kolbe M. Protein refolding is required for assembly of the type three secretion needle. Nat Struct Mol Biol. 2010;17:788–792. doi: 10.1038/nsmb.1822. [DOI] [PubMed] [Google Scholar]
  • 60.Demers JP, Habenstein B, Loquet A, Kumar Vasa S, Giller K, Becker S, Baker D, Lange A, Sgourakis NG. High-resolution structure of the Shigella type-III secretion needle by solid-state NMR and cryo-electron microscopy. Nat Commun. 2014;5:4976. doi: 10.1038/ncomms5976. doi: 10.1038/ncomms5976. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Fujii T, Cheung M, Blanco A, Kato T, Blocker AJ, Namba K. Structure of a type III secretion needle at 7-A resolution provides insights into its assembly and signaling mechanisms. Proc Natl Acad Sci U S A. 2012;109:4461–4466. doi: 10.1073/pnas.1116126109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Mota LJ, Journet L, Sorg I, Agrain C, Cornelis GR. Bacterial injectisomes: needle length does matter. Science. 2005;307:1278. doi: 10.1126/science.1107679. [DOI] [PubMed] [Google Scholar]
  • 63.Kubori T, Sukhan A, Aizawa SI, Galan JE. Molecular characterization and assembly of the needle complex of the Salmonella typhimurium type III protein secretion system. Proc Natl Acad Sci U S A. 2000;97:10225–10230. doi: 10.1073/pnas.170128997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Tamano K, Katayama E, Toyotome T, Sasakawa C. Shigella Spa32 is an essential secretory protein for functional type III secretion machinery and uniformity of its needle length. J Bacteriol. 2002;184:1244–1252. doi: 10.1128/JB.184.5.1244-1252.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Journet L, Agrain C, Broz P, Cornelis GR. The needle length of bacterial injectisomes is determined by a molecular ruler. Science. 2003;302:1757–1760. doi: 10.1126/science.1091422. [DOI] [PubMed] [Google Scholar]
  • 66.Lefebre MD, Galan JE. The inner rod protein controls substrate switching and needle length in a Salmonella type III secretion system. Proc Natl Acad Sci U S A. 2014;111:817–822. doi: 10.1073/pnas.1319698111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Mueller CA, Broz P, Muller SA, Ringler P, Erne-Brand F, Sorg I, Kuhn M, Engel A, Cornelis GR. The V-antigen of Yersinia forms a distinct structure at the tip of injectisome needles. Science. 2005;310:674–676. doi: 10.1126/science.1118476. [DOI] [PubMed] [Google Scholar]
  • 68.Epler CR, Dickenson NE, Bullitt E, Picking WL. Ultrastructural analysis of IpaD at the tip of the nascent MxiH type III secretion apparatus of Shigella flexneri. J Mol Biol. 2012;420:29–39. doi: 10.1016/j.jmb.2012.03.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Goure J, Broz P, Attree O, Cornelis GR, Attree I. Protective anti-V antibodies inhibit Pseudomonas and Yersinia translocon assembly within host membranes. J Infect Dis. 2005;192:218–225. doi: 10.1086/430932. [DOI] [PubMed] [Google Scholar]
  • 70.Mueller CA, Broz P, Cornelis GR. The type III secretion system tip complex and translocon. Mol Microbiol. 2008;68:1085–1095. doi: 10.1111/j.1365-2958.2008.06237.x. [DOI] [PubMed] [Google Scholar]
  • 71.Lawton WD, Surgalla MJ. Immunization against Plague by a Specific Fraction of Pasteurella Pseudotuberculosis. J Infect Dis. 1963;113:39–42. doi: 10.1093/infdis/113.1.39. [DOI] [PubMed] [Google Scholar]
  • 72.DiGiandomenico A, Keller AE, Gao C, Rainey GJ, Warrener P, Camara MM, Bonnell J, Fleming R, Bezabeh B, Dimasi N, Sellman BR, Hilliard J, Guenther CM, Datta V, Zhao W, Gao C, Yu XQ, Suzich JA, Stover CK. A multifunctional bispecific antibody protects against Pseudomonas aeruginosa. Sci Transl Med. 2014;6:262ra155. doi: 10.1126/scitranslmed.3009655. [DOI] [PubMed] [Google Scholar]
  • 73.Derewenda U, Mateja A, Devedjiev Y, Routzahn KM, Evdokimov AG, Derewenda ZS, Waugh DS. The structure of Yersinia pestis V-antigen, an essential virulence factor and mediator of immunity against plague. Structure. 2004;12:301–306. doi: 10.1016/j.str.2004.01.010. [DOI] [PubMed] [Google Scholar]
  • 74.Johnson S, Roversi P, Espina M, Olive A, Deane JE, Birket S, Field T, Picking WD, Blocker AJ, Galyov EE, Picking WL, Lea SM. Self-chaperoning of the type III secretion system needle tip proteins IpaD and BipD. J Biol Chem. 2007;282:4035–4044. doi: 10.1074/jbc.M607945200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Lunelli M, Hurwitz R, Lambers J, Kolbe M. Crystal structure of PrgI-SipD: insight into a secretion competent state of the type three secretion system needle tip and its interaction with host ligands. PLoS Pathog. 2011;7:e1002163. doi: 10.1371/journal.ppat.1002163. doi: 10.1371/journal.ppat.1002163. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Cheung M, Shen DK, Makino F, Kato T, Roehrich AD, Martinez-Argudo I, Walker ML, Murillo I, Liu X, Pain M, Brown J, Frazer G, Mantell J, Mina P, Todd T, Sessions RB, Namba K, Blocker AJ. Three-dimensional electron microscopy reconstruction and cysteine-mediated crosslinking provide a model of the type III secretion system needle tip complex. Mol Microbiol. 2015;95:31–50. doi: 10.1111/mmi.12843. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Rathinavelan T, Lara-Tejero M, Lefebre M, Chatterjee S, McShan AC, Guo DC, Tang C, Galan JE, De Guzman RN. NMR model of PrgI-SipD interaction and its implications in the needle-tip assembly of the Salmonella type III secretion system. J Mol Biol. 2014;426:2958–2969. doi: 10.1016/j.jmb.2014.06.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Knutton S, Rosenshine I, Pallen MJ, Nisan I, Neves BC, Bain C, Wolff C, Dougan G, Frankel G. A novel EspA-associated surface organelle of enteropathogenic Escherichia coli involved in protein translocation into epithelial cells. EMBO J. 1998;17:2166–2176. doi: 10.1093/emboj/17.8.2166. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Daniell SJ, Kocsis E, Morris E, Knutton S, Booy FP, Frankel G. 3D structure of EspA filaments from enteropathogenic Escherichia coli. Mol Microbiol. 2003;49:301–308. doi: 10.1046/j.1365-2958.2003.03555.x. [DOI] [PubMed] [Google Scholar]
  • 80.Picking WL, Nishioka H, Hearn PD, Baxter MA, Harrington AT, Blocker A, Picking WD. IpaD of Shigella flexneri is independently required for regulation of Ipa protein secretion and efficient insertion of IpaB and IpaC into host membranes. Infect Immun. 2005;73:1432–1440. doi: 10.1128/IAI.73.3.1432-1440.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Barta ML, Dickenson NE, Patil M, Keightley A, Wyckoff GJ, Picking WD, Picking WL, Geisbrecht BV. The structures of coiled-coil domains from type III secretion system translocators reveal homology to pore-forming toxins. J Mol Biol. 2012;417:395–405. doi: 10.1016/j.jmb.2012.01.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Mizusaki H, Takaya A, Yamamoto T, Aizawa S. Signal pathway in salt-activated expression of the Salmonella pathogenicity island 1 type III secretion system in Salmonella enterica serovar Typhimurium. J Bacteriol. 2008;190:4624–4631. doi: 10.1128/JB.01957-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Arnold R, Brandmaier S, Kleine F, Tischler P, Heinz E, Behrens S, Niinikoski A, Mewes HW, Horn M, Rattei T. Sequence-based prediction of type III secreted proteins. PLoS Pathog. 2009;5:e1000376. doi: 10.1371/journal.ppat.1000376. doi: 10.1371/journal.ppat.1000376. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Samudrala R, Heffron F, McDermott JE. Accurate prediction of secreted substrates and identification of a conserved putative secretion signal for type III secretion systems. PLoS Pathog. 2009;5:e1000375. doi: 10.1371/journal.ppat.1000375. doi: 10.1371/journal.ppat.1000375. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.McDermott JE, Corrigan A, Peterson E, Oehmen C, Niemann G, Cambronne ED, Sharp D, Adkins JN, Samudrala R, Heffron F. Computational prediction of type III and IV secreted effectors in gram-negative bacteria. Infect Immun. 2011;79:23–32. doi: 10.1128/IAI.00537-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Lloyd SA, Norman M, Rosqvist R, Wolf-Watz H. Yersinia YopE is targeted for type III secretion by N-terminal, not mRNA, signals. Mol Microbiol. 2001;39:520–531. doi: 10.1046/j.1365-2958.2001.02271.x. [DOI] [PubMed] [Google Scholar]
  • 87.Lee SH, Galan JE. Salmonella type III secretion-associated chaperones confer secretion-pathway specificity. Mol Microbiol. 2004;51:483–495. doi: 10.1046/j.1365-2958.2003.03840.x. [DOI] [PubMed] [Google Scholar]
  • 88.Stebbins CE, Galan JE. Maintenance of an unfolded polypeptide by a cognate chaperone in bacterial type III secretion. Nature. 2001;414:77–81. doi: 10.1038/35102073. [DOI] [PubMed] [Google Scholar]
  • 89.Stebbins CE, Galan JE. Priming virulence factors for delivery into the host. Nat Rev Mol Cell Biol. 2003;4:738–743. doi: 10.1038/nrm1201. [DOI] [PubMed] [Google Scholar]
  • 90.Izore T, Job V, Dessen A. Biogenesis, regulation, and targeting of the type III secretion system. Structure. 2011;19:603–612. doi: 10.1016/j.str.2011.03.015. [DOI] [PubMed] [Google Scholar]
  • 91.DeBord KL, Lee VT, Schneewind O. Roles of LcrG and LcrV during type III targeting of effector Yops by Yersinia enterocolitica. J Bacteriol. 2001;183:4588–4598. doi: 10.1128/JB.183.15.4588-4598.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Lilic M, Vujanac M, Stebbins CE. A common structural motif in the binding of virulence factors to bacterial secretion chaperones. Mol Cell. 2006;21:653–664. doi: 10.1016/j.molcel.2006.01.026. [DOI] [PubMed] [Google Scholar]
  • 93.Janjusevic R, Quezada CM, Small J, Stebbins CE. Structure of the HopA1(21-102)-ShcA chaperone-effector complex of Pseudomonas syringae reveals conservation of a virulence factor binding motif from animal to plant pathogens. J Bacteriol. 2013;195:658–664. doi: 10.1128/JB.01621-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Costa SC, Schmitz AM, Jahufar FF, Boyd JD, Cho MY, Glicksman MA, Lesser CF. A new means to identify type 3 secreted effectors: functionally interchangeable class IB chaperones recognize a conserved sequence. MBio. 2012;3:e00243–11. doi: 10.1128/mBio.00243-11. doi: 10.1128/mBio.00243-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Michiels T, Cornelis GR. Secretion of hybrid proteins by the Yersinia Yop export system. J Bacteriol. 1991;173:1677–1685. doi: 10.1128/jb.173.5.1677-1685.1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Chen LM, Briones G, Donis RO, Galan JE. Optimization of the delivery of heterologous proteins by the Salmonella enterica serovar Typhimurium type III secretion system for vaccine development. Infect Immun. 2006;74:5826–5833. doi: 10.1128/IAI.00375-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Widmaier DM, Tullman-Ercek D, Mirsky EA, Hill R, Govindarajan S, Minshull J, Voigt CA. Engineering the Salmonella type III secretion system to export spider silk monomers. Mol Syst Biol. 2009;5:309. doi: 10.1038/msb.2009.62. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Akeda Y, Galan JE. Genetic analysis of the Salmonella enterica type III secretion-associated ATPase InvC defines discrete functional domains. J Bacteriol. 2004;186:2402–2412. doi: 10.1128/JB.186.8.2402-2412.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Zarivach R, Vuckovic M, Deng W, Finlay BB, Strynadka NC. Structural analysis of a prototypical ATPase from the type III secretion system. Nat Struct Mol Biol. 2007;14:131–137. doi: 10.1038/nsmb1196. [DOI] [PubMed] [Google Scholar]
  • 100.Claret L, Calder SR, Higgins M, Hughes C. Oligomerization and activation of the FliI ATPase central to bacterial flagellum assembly. Mol Microbiol. 2003;48:1349–1355. doi: 10.1046/j.1365-2958.2003.03506.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Ibuki T, Imada K, Minamino T, Kato T, Miyata T, Namba K. Common architecture of the flagellar type III protein export apparatus and F- and V-type ATPases. Nat Struct Mol Biol. 2011;18:277–282. doi: 10.1038/nsmb.1977. [DOI] [PubMed] [Google Scholar]
  • 102.Pozidis C, Chalkiadaki A, Gomez-Serrano A, Stahlberg H, Brown I, Tampakaki AP, Lustig A, Sianidis G, Politou AS, Engel A, Panopoulos NJ, Mansfield J, Pugsley AP, Karamanou S, Economou A. Type III protein translocase: HrcN is a peripheral ATPase that is activated by oligomerization. J Biol Chem. 2003;278:25816–25824. doi: 10.1074/jbc.M301903200. [DOI] [PubMed] [Google Scholar]
  • 103.Akeda Y, Galan JE. Chaperone release and unfolding of substrates in type III secretion. Nature. 2005;437:911–915. doi: 10.1038/nature03992. [DOI] [PubMed] [Google Scholar]
  • 104.Minamino T, Namba K. Distinct roles of the FliI ATPase and proton motive force in bacterial flagellar protein export. Nature. 2008;451:485–488. doi: 10.1038/nature06449. [DOI] [PubMed] [Google Scholar]
  • 105.Paul K, Erhardt M, Hirano T, Blair DF, Hughes KT. Energy source of flagellar type III secretion. Nature. 2008;451:489–492. doi: 10.1038/nature06497. [DOI] [PubMed] [Google Scholar]
  • 106.Erhardt M, Mertens ME, Fabiani FD, Hughes KT. ATPase-Independent Type-III Protein Secretion in Salmonella enterica. PLoS Genet. 2014;10:e1004800. doi: 10.1371/journal.pgen.1004800. doi: 10.1371/journal.pgen.1004800. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Minamino T, Morimoto YV, Kinoshita M, Aldridge PD, Namba K. The bacterial flagellar protein export apparatus processively transports flagellar proteins even with extremely infrequent ATP hydrolysis. Sci Rep. 2014;4:7579. doi: 10.1038/srep07579. doi: 10.1038/srep07579. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Chen S, Beeby M, Murphy GE, Leadbetter JR, Hendrixson DR, Briegel A, Li Z, Shi J, Tocheva EI, Muller A, Dobro MJ, Jensen GJ. Structural diversity of bacterial flagellar motors. EMBO J. 2011;30:2972–2981. doi: 10.1038/emboj.2011.186. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Kawamoto A, Morimoto YV, Miyata T, Minamino T, Hughes KT, Kato T, Namba K. Common and distinct structural features of Salmonella injectisome and flagellar basal body. Sci Rep. 2013;3:3369. doi: 10.1038/srep03369. doi: 10.1038/srep03369. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Lorenzini E, Singer A, Singh B, Lam R, Skarina T, Chirgadze NY, Savchenko A, Gupta RS. Structure and protein-protein interaction studies on Chlamydia trachomatis protein CT670 (YscO Homolog). J Bacteriol. 2010;192:2746–2756. doi: 10.1128/JB.01479-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Wilharm G, Lehmann V, Krauss K, Lehnert B, Richter S, Ruckdeschel K, Heesemann J, Trulzsch K. Yersinia enterocolitica type III secretion depends on the proton motive force but not on the flagellar motor components MotA and MotB. Infect Immun. 2004;72:4004–4009. doi: 10.1128/IAI.72.7.4004-4009.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Galan JE. Energizing type III secretion machines: what is the fuel? Nat Struct Mol Biol. 2008;15:127–128. doi: 10.1038/nsmb0208-127. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Allison SE, Tuinema BR, Everson ES, Sugiman-Marangos S, Zhang K, Junop MS, Coombes BK. Identification of the docking site between a type III secretion system ATPase and a chaperone for effector cargo. J Biol Chem. 2014;289:23734–23744. doi: 10.1074/jbc.M114.578476. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Roblin P, Dewitte F, Villeret V, Biondi EG, Bompard C. A Salmonella type three secretion effector/chaperone complex adopts a hexameric ring-like structure. J Bacteriol. 2015;197:688–98. doi: 10.1128/JB.02294-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Tsai CL, Burkinshaw BJ, Strynadka NC, Tainer JA. The Salmonella type III secretion system virulence effector forms a new hexameric chaperone assembly for export of effector/chaperone complexes. J Bacteriol. 2015;197:672–5. doi: 10.1128/JB.02524-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Chen L, Balabanidou V, Remeta DP, Minetti CA, Portaliou AG, Economou A, Kalodimos CG. Structural instability tuning as a regulatory mechanism in protein-protein interactions. Mol Cell. 2011;44:734–744. doi: 10.1016/j.molcel.2011.09.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Chen L, Ai X, Portaliou AG, Minetti CA, Remeta DP, Economou A, Kalodimos CG. Substrate-activated conformational switch on chaperones encodes a targeting signal in type III secretion. Cell Rep. 2013;3:709–715. doi: 10.1016/j.celrep.2013.02.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Ohgita T, Hayashi N, Hama S, Tsuchiya H, Gotoh N, Kogure K. A novel effector secretion mechanism based on proton-motive force-dependent type III secretion apparatus rotation. FASEB J. 2013;27:2862–2872. doi: 10.1096/fj.13-229054. [DOI] [PubMed] [Google Scholar]
  • 119.Morita-Ishihara T, Ogawa M, Sagara H, Yoshida M, Katayama E, Sasakawa C. Shigella Spa33 is an essential C-ring component of type III secretion machinery. J Biol Chem. 2006;281:599–607. doi: 10.1074/jbc.M509644200. [DOI] [PubMed] [Google Scholar]
  • 120.Hu B, Morado DR, Margolin W, Rohde JR, Arizmendi O, Picking WL, Picking WD, Liu J. Visualization of the type III secretion sorting platform of Shigella flexneri. Proc Natl Acad Sci U S A. 2015;112:1047–1052. doi: 10.1073/pnas.1411610112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Lara-Tejero M, Kato J, Wagner S, Liu X, Galan JE. A sorting platform determines the order of protein secretion in bacterial type III systems. Science. 2011;331:1188–1191. doi: 10.1126/science.1201476. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Gonzalez-Pedrajo B, Fraser GM, Minamino T, Macnab RM. Molecular dissection of Salmonella FliH, a regulator of the ATPase FliI and the type III flagellar protein export pathway. Mol Microbiol. 2002;45:967–982. doi: 10.1046/j.1365-2958.2002.03047.x. [DOI] [PubMed] [Google Scholar]
  • 123.Minamino T, MacNab RM. FliH, a soluble component of the type III flagellar export apparatus of Salmonella, forms a complex with FliI and inhibits its ATPase activity. Mol Microbiol. 2000;37:1494–1503. doi: 10.1046/j.1365-2958.2000.02106.x. [DOI] [PubMed] [Google Scholar]
  • 124.Lane MC, O'Toole PW, Moore SA. Molecular basis of the interaction between the flagellar export proteins FliI and FliH from Helicobacter pylori. J Biol Chem. 2006;281:508–517. doi: 10.1074/jbc.M507238200. [DOI] [PubMed] [Google Scholar]
  • 125.Minamino T, Yoshimura SD, Morimoto YV, Gonzalez-Pedrajo B, Kami-Ike N, Namba K. Roles of the extreme N-terminal region of FliH for efficient localization of the FliH-FliI complex to the bacterial flagellar type III export apparatus. Mol Microbiol. 2009;74:1471–1483. doi: 10.1111/j.1365-2958.2009.06946.x. [DOI] [PubMed] [Google Scholar]
  • 126.McMurry JL, Murphy JW, Gonzalez-Pedrajo B. The FliN-FliH interaction mediates localization of flagellar export ATPase FliI to the C ring complex. Biochemistry. 2006;45:11790–11798. doi: 10.1021/bi0605890. [DOI] [PubMed] [Google Scholar]
  • 127.Minamino T, Gonzalez-Pedrajo B, Kihara M, Namba K, Macnab RM. The ATPase FliI can interact with the type III flagellar protein export apparatus in the absence of its regulator, FliH. J Bacteriol. 2003;185:3983–3988. doi: 10.1128/JB.185.13.3983-3988.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Bai F, Morimoto YV, Yoshimura SD, Hara N, Kami-Ike N, Namba K, Minamino T. Assembly dynamics and the roles of FliI ATPase of the bacterial flagellar export apparatus. Sci Rep. 2014;4:6528. doi: 10.1038/srep06528. doi: 10.1038/srep06528. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Diepold A, Kudryashev M, Delalez NJ, Berry RM, Armitage JP. Composition, formation, and regulation of the cytosolic c-ring, a dynamic component of the type III secretion injectisome. PLoS Biol. 2015;13:e1002039. doi: 10.1371/journal.pbio.1002039. doi: 10.1371/journal.pbio.1002039. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Fadouloglou VE, Tampakaki AP, Glykos NM, Bastaki MN, Hadden JM, Phillips SE, Panopoulos NJ, Kokkinidis M. Structure of HrcQB-C, a conserved component of the bacterial type III secretion systems. Proc Natl Acad Sci U S A. 2004;101:70–75. doi: 10.1073/pnas.0304579101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Brown PN, Mathews MA, Joss LA, Hill CP, Blair DF. Crystal structure of the flagellar rotor protein FliN from Thermotoga maritima. J Bacteriol. 2005;187:2890–2902. doi: 10.1128/JB.187.8.2890-2902.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Bzymek KP, Hamaoka BY, Ghosh P. Two translation products of Yersinia yscQ assemble to form a complex essential to type III secretion. Biochemistry. 2012;51:1669–1677. doi: 10.1021/bi201792p. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Paul K, Blair DF. Organization of FliN subunits in the flagellar motor of Escherichia coli. J Bacteriol. 2006;188:2502–2511. doi: 10.1128/JB.188.7.2502-2511.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Paul K, Harmon JG, Blair DF. Mutational analysis of the flagellar rotor protein FliN: identification of surfaces important for flagellar assembly and switching. J Bacteriol. 2006;188:5240–5248. doi: 10.1128/JB.00110-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Fadouloglou VE, Bastaki MN, Ashcroft AE, Phillips SE, Panopoulos NJ, Glykos NM, Kokkinidis M. On the quaternary association of the type III secretion system HrcQB-C protein: experimental evidence differentiates among the various oligomerization models. J Struct Biol. 2009;166:214–225. doi: 10.1016/j.jsb.2009.01.008. [DOI] [PubMed] [Google Scholar]
  • 136.Lee SY, Cho HS, Pelton JG, Yan D, Henderson RK, King DS, Huang L, Kustu S, Berry EA, Wemmer DE. Crystal structure of an activated response regulator bound to its target. Nat Struct Biol. 2001;8:52–56. doi: 10.1038/83053. [DOI] [PubMed] [Google Scholar]
  • 137.Vartanian AS, Paz A, Fortgang EA, Abramson J, Dahlquist FW. Structure of flagellar motor proteins in complex allows for insights into motor structure and switching. J Biol Chem. 2012;287:35779–35783. doi: 10.1074/jbc.C112.378380. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Kubori T, Galan JE. Salmonella type III secretion-associated protein InvE controls translocation of effector proteins into host cells. J Bacteriol. 2002;184:4699–4708. doi: 10.1128/JB.184.17.4699-4708.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Deane JE, Abrusci P, Johnson S, Lea SM. Timing is everything: the regulation of type III secretion. Cell Mol Life Sci. 2010;67:1065–1075. doi: 10.1007/s00018-009-0230-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Magdalena J, Hachani A, Chamekh M, Jouihri N, Gounon P, Blocker A, Allaoui A. Spa32 regulates a switch in substrate specificity of the type III secreton of Shigella flexneri from needle components to Ipa proteins. J Bacteriol. 2002;184:3433–3441. doi: 10.1128/JB.184.13.3433-3441.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Agrain C, Callebaut I, Journet L, Sorg I, Paroz C, Mota LJ, Cornelis GR. Characterization of a Type III secretion substrate specificity switch (T3S4) domain in YscP from Yersinia enterocolitica. Mol Microbiol. 2005;56:54–67. doi: 10.1111/j.1365-2958.2005.04534.x. [DOI] [PubMed] [Google Scholar]
  • 142.Mizuno S, Amida H, Kobayashi N, Aizawa S, Tate S. The NMR structure of FliK, the trigger for the switch of substrate specificity in the flagellar type III secretion apparatus. J Mol Biol. 2011;409:558–573. doi: 10.1016/j.jmb.2011.04.008. [DOI] [PubMed] [Google Scholar]
  • 143.Mukerjea R, Ghosh P. Functionally essential interaction between Yersinia YscO and the T3S4 domain of YscP. J Bacteriol. 2013;195:4631–4638. doi: 10.1128/JB.00876-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Botteaux A, Sani M, Kayath CA, Boekema EJ, Allaoui A. Spa32 interaction with the inner-membrane Spa40 component of the type III secretion system of Shigella flexneri is required for the control of the needle length by a molecular tape measure mechanism. Mol Microbiol. 2008;70:1515–1528. doi: 10.1111/j.1365-2958.2008.06499.x. [DOI] [PubMed] [Google Scholar]
  • 145.Wood SE, Jin J, Lloyd SA. YscP and YscU switch the substrate specificity of the Yersinia type III secretion system by regulating export of the inner rod protein YscI. J Bacteriol. 2008;190:4252–4262. doi: 10.1128/JB.00328-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Botteaux A, Kayath CA, Page AL, Jouihri N, Sani M, Boekema E, Biskri L, Parsot C, Allaoui A. The 33 carboxyl-terminal residues of Spa40 orchestrate the multi-step assembly process of the type III secretion needle complex in Shigella flexneri. Microbiology. 2010;156:2807–2817. doi: 10.1099/mic.0.039651-0. [DOI] [PubMed] [Google Scholar]
  • 147.Cherradi Y, Schiavolin L, Moussa S, Meghraoui A, Meksem A, Biskri L, Azarkan M, Allaoui A, Botteaux A. Interplay between predicted inner-rod and gatekeeper in controlling substrate specificity of the type III secretion system. Mol Microbiol. 2013;87:1183–1199. doi: 10.1111/mmi.12158. [DOI] [PubMed] [Google Scholar]
  • 148.Sorg I, Wagner S, Amstutz M, Muller SA, Broz P, Lussi Y, Engel A, Cornelis GR. YscU recognizes translocators as export substrates of the Yersinia injectisome. EMBO J. 2007;26:3015–3024. doi: 10.1038/sj.emboj.7601731. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Botteaux A, Sory MP, Biskri L, Parsot C, Allaoui A. MxiC is secreted by and controls the substrate specificity of the Shigella flexneri type III secretion apparatus. Mol Microbiol. 2009;71:449–460. doi: 10.1111/j.1365-2958.2008.06537.x. [DOI] [PubMed] [Google Scholar]
  • 150.Cheng LW, Kay O, Schneewind O. Regulated secretion of YopN by the type III machinery of Yersinia enterocolitica. J Bacteriol. 2001;183:5293–5301. doi: 10.1128/JB.183.18.5293-5301.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Rimpilainen M, Forsberg A, Wolf-Watz H. A novel protein, LcrQ, involved in the low-calcium response of Yersinia pseudotuberculosis shows extensive homology to YopH. J Bacteriol. 1992;174:3355–3363. doi: 10.1128/jb.174.10.3355-3363.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Skryzpek E, Straley SC. LcrG, a secreted protein involved in negative regulation of the low-calcium response in Yersinia pestis. J Bacteriol. 1993;175:3520–3528. doi: 10.1128/jb.175.11.3520-3528.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Schubot FD, Jackson MW, Penrose KJ, Cherry S, Tropea JE, Plano GV, Waugh DS. Three-dimensional structure of a macromolecular assembly that regulates type III secretion in Yersinia pestis. J Mol Biol. 2005;346:1147–1161. doi: 10.1016/j.jmb.2004.12.036. [DOI] [PubMed] [Google Scholar]
  • 154.Deane JE, Roversi P, King C, Johnson S, Lea SM. Structures of the Shigella flexneri type 3 secretion system protein MxiC reveal conformational variability amongst homologues. J Mol Biol. 2008;377:985–992. doi: 10.1016/j.jmb.2008.01.072. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Weise CF, Login FH, Ho O, Grobner G, Wolf-Watz H, Wolf-Watz M. Negatively Charged Lipid Membranes Promote a Disorder-Order Transition in the Yersinia YscU Protein. Biophys J. 2014;107:1950–1961. doi: 10.1016/j.bpj.2014.09.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Stensrud KF, Adam PR, La Mar CD, Olive AJ, Lushington GH, Sudharsan R, Shelton NL, Givens RS, Picking WL, Picking WD. Deoxycholate interacts with IpaD of Shigella flexneri in inducing the recruitment of IpaB to the type III secretion apparatus needle tip. J Biol Chem. 2008;283:18646–18654. doi: 10.1074/jbc.M802799200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Prouty AM, Gunn JS. Salmonella enterica serovar typhimurium invasion is repressed in the presence of bile. Infect Immun. 2000;68:6763–6769. doi: 10.1128/iai.68.12.6763-6769.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Barta ML, Guragain M, Adam P, Dickenson NE, Patil M, Geisbrecht BV, Picking WL, Picking WD. Identification of the bile salt binding site on IpaD from Shigella flexneri and the influence of ligand binding on IpaD structure. Proteins. 2012;80:935–945. doi: 10.1002/prot.23251. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Chatterjee S, Zhong D, Nordhues BA, Battaile KP, Lovell S, De Guzman RN. The crystal structures of the Salmonella type III secretion system tip protein SipD in complex with deoxycholate and chenodeoxycholate. Protein Sci. 2011;20:75–86. doi: 10.1002/pro.537. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Menard R, Sansonetti P, Parsot C. The secretion of the Shigella flexneri Ipa invasins is activated by epithelial cells and controlled by IpaB and IpaD. EMBO J. 1994;13:5293–5302. doi: 10.1002/j.1460-2075.1994.tb06863.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Rosqvist R, Magnusson KE, Wolf-Watz H. Target cell contact triggers expression and polarized transfer of Yersinia YopE cytotoxin into mammalian cells. EMBO J. 1994;13:964–972. doi: 10.1002/j.1460-2075.1994.tb06341.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Zierler MK, Galan JE. Contact with cultured epithelial cells stimulates secretion of Salmonella typhimurium invasion protein InvJ. Infect Immun. 1995;63:4024–4028. doi: 10.1128/iai.63.10.4024-4028.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Lara-Tejero M, Galan JE. Salmonella enterica serovar typhimurium pathogenicity island 1-encoded type III secretion system translocases mediate intimate attachment to nonphagocytic cells. Infect Immun. 2009;77:2635–2642. doi: 10.1128/IAI.00077-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Epler CR, Dickenson NE, Olive AJ, Picking WL, Picking WD. Liposomes recruit IpaC to the Shigella flexneri type III secretion apparatus needle as a final step in secretion induction. Infect Immun. 2009;77:2754–2761. doi: 10.1128/IAI.00190-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165.Blocker AJ, Deane JE, Veenendaal AK, Roversi P, Hodgkinson JL, Johnson S, Lea SM. What's the point of the type III secretion system needle? Proc Natl Acad Sci U S A. 2008;105:6507–6513. doi: 10.1073/pnas.0708344105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Cordes FS, Daniell S, Kenjale R, Saurya S, Picking WL, Picking WD, Booy F, Lea SM, Blocker A. Helical packing of needles from functionally altered Shigella type III secretion systems. J Mol Biol. 2005;354:206–211. doi: 10.1016/j.jmb.2005.09.062. [DOI] [PubMed] [Google Scholar]
  • 167.Yu XJ, McGourty K, Liu M, Unsworth KE, Holden DW. pH sensing by intracellular Salmonella induces effector translocation. Science. 2010;328:1040–1043. doi: 10.1126/science.1189000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Figueira R, Holden DW. Functions of the Salmonella pathogenicity island 2 (SPI-2) type III secretion system effectors. Microbiology. 2012;158:1147–1161. doi: 10.1099/mic.0.058115-0. [DOI] [PubMed] [Google Scholar]
  • 169.Torruellas J, Jackson MW, Pennock JW, Plano GV. The Yersinia pestis type III secretion needle plays a role in the regulation of Yop secretion. Mol Microbiol. 2005;57:1719–1733. doi: 10.1111/j.1365-2958.2005.04790.x. [DOI] [PubMed] [Google Scholar]
  • 170.Birtalan SC, Phillips RM, Ghosh P. Three-dimensional secretion signals in chaperone-effector complexes of bacterial pathogens. Mol Cell. 2002;9:971–980. doi: 10.1016/s1097-2765(02)00529-4. [DOI] [PubMed] [Google Scholar]
  • 171.Phan J, Tropea JE, Waugh DS. Structure of the Yersinia pestis type III secretion chaperone SycH in complex with a stable fragment of YscM2. Acta Crystallogr D Biol Crystallogr. 2004;60:1591–1599. doi: 10.1107/S0907444904017597. [DOI] [PubMed] [Google Scholar]
  • 172.Schreiner M, Niemann HH. Crystal structure of the Yersinia enterocolitica type III secretion chaperone SycD in complex with a peptide of the minor translocator YopD. BMC Struct Biol. 2012;12:13. doi: 10.1186/1472-6807-12-13. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES