Skip to main content
Microbiology logoLink to Microbiology
. 2023 Jun 13;169(6):001328. doi: 10.1099/mic.0.001328

Virulence-associated type III secretion systems in Gram-negative bacteria

Sara Vilela Pais 1,, Eunjin Kim 1,, Samuel Wagner 1,2,3,*
PMCID: PMC10333785  PMID: 37310005

Abstract

Virulence-associated bacterial type III secretion systems are multiprotein molecular machines that promote the pathogenicity of bacteria towards eukaryotic host cells. These machines form needle-like structures, named injectisomes, that span both bacterial and host membranes, forming a direct conduit for the delivery of bacterial proteins into host cells. Once within the host, these bacterial effector proteins are capable of manipulating a multitude of host cell functions. In recent years, the knowledge of assembly, structure and function of these machines has grown substantially and is presented and discussed in this review.

Keywords: bacterial pathogenicity, type III secretion systems

Introduction

Virulence-associated-type III secretion systems (T3SS) utilize a supramolecular protein complex that is also termed injectisome for secretion of bacterial effector proteins into eukaryotic host cells. The injectisome evolved from the flagellar T3SS that serves bacterial motility, and a high level of structural similarity exists between the proteins that constitute the injectisome and those that constitute the flagella [1]. The injectisome has a molecular mass of around 7 MDa and is composed of more than 200 subunits of up to 20 different proteins [2]. It spans the inner (IM) and outer (OM) membranes of Gram-negative bacteria as well as a membrane of the host. It is built up of a base, an export apparatus, cytosolic components including an ATPase and a needle filament including a needle adapter and a needle tip (Fig. 1). T3SS adapted to the needs of different pathogens and as a result, several families of T3SS can be found [3]. While the overall structure of the injectisome and the export mechanism are conserved, the components of T3SS that directly contact the host, like the needle tip and translocon, are very different between bacterial species [4]. Also, the regulatory mechanisms of gene expression, assembly and secretion as well as the secreted effectors vary among bacteria and even in the same bacteria (i.e. Salmonella ’s T3SS-1 and T3SS-2) depending on pathogen needs.

Fig. 1.

Fig. 1.

Structure and components of a type III secretion system (T3SS) injectisome. Left: Subtomogram of T3SS-1 injectisome of Salmonella enterica serovar Typhimurium, (accession code EMD-8544) [35]. Right: Overlay of sliced-tomogram with a model of T3SS depicting all the individual components of the injectisome. Below: Table showing conserved proteins between virulence T3SS (vT3SS) and flagellar T3SS (fT3SS). Protein names follow the unified nomenclature according to Wagner and Diepold [155] and are depicted as letters (e.g. SctO: O). The different structural units are colour-coded. Adapted from Wagner [102] . IM: inner membrane, PG: peptidoglycan, OM: outer membrane.

Much new structural and functional information on the injectisome has been produced in recent years, which has led to a substantial elucidation of the organization of this machinery, of the understanding of the dynamic mechanisms that drive injectisome assembly, and of the targeting and passage of substrates through the injectisome channel. In this review, we summarize and discuss these recent findings of the field, starting with the structural components.

Export apparatus

The export apparatus is composed of five highly conserved proteins, SctR, SctS, SctT, SctU and SctV, with a stoichiometry of 5 : 4 : 1 : 1 : 9 [2, 5–7] (where proteins are equivalent, we also refer to publications on the bacterial flagellum). The minor export apparatus proteins SctR, SctS and SctT contain four, two, and six predicted transmembrane segments (TMS), respectively, and together form a unique helical structure called the core export apparatus [5, 8]. One SctR and one SctS are structurally equivalent to SctT, so that the SctR5S4T1 complex is of pseudo-hexameric symmetry [5, 8]. Despite its substantial hydrophobicity, the SctR5S4T1 complex is not integrated in the inner bacterial membrane in the final structure but it is located on the periplasmic side at the centre of the base with hydrophobic contacts to a nonameric ring of the transmembrane domains of SctV [5, 9, 10].

As the core export apparatus forms a pore in the IM, it is crucial that the pore is gated and opens only during substrate translocation. The export apparatus has several constriction points that undergo conformational changes upon substrate engagement [11, 12]. There are hydrophilic Q1 and Q2 belts that sandwich a hydrophobic M-gate [11, 12]. The substrate is first engaged by the Q1 belt, which is formed by glutamic acid residues from the four SctS subunits. Once stably engaged by the Q1 belt, the substrate is engaged by the M-gate, which is formed by conserved methionines in the SctR proteins. Subsequent movement of the flexible SctT plug opens the upper atrium of the export apparatus where is located the Q2 belt, formed by SctR subunits. The Q2 belt further engages the substrate during translocation. These constriction points are narrow and substrates must be unfolded to go through the passage; the narrowest is the M-gate, which measures~5.5 Å in diameter [6, 12]. Several conserved inter- and intramolecular salt bridges are present in the export apparatus, which have been shown to be important for assembly and stabilization of the complex [13], and some are also involved in gating of the export apparatus [5, 11–13].

The switch protein SctU contains four predicted TMS and a cytoplasmic domain. These TMS form two α-helical hairpins that resemble the structure of SctS, making SctU the closing component of the SctR5S4T1 complex [6]. SctU contains a highly conserved loop between the two hairpins that wraps around the entrance of the export gate [6]. Although the specific role of the loop is unclear, the interaction of the loop with each SctS subunit has been shown to be important for secretion function and the loop likely plays a role in the gating mechanism [6]. The cytoplasmic domain of SctU contains a highly conserved NPTH motif, which undergoes autocleavage [14–16]. It has been proposed that the autocleavage is the trigger for the substrate switching [17, 18] but it is possible that the autocleavage causes conformational changes in SctU that are required for its function in substrate specificity switching [19].

The major export apparatus protein SctV forms a sea-horse like structure as shown by cryo-electron tomography (cryo-ET) and contains a transmembrane domain (TMD) with eight predicted TMS, a cytoplasmic domain, and a linker that connects the two domains [9]. The atomic structure of the TMD of SctV has not been determined yet but the cryo-ET structure revealed that the TMDs of SctV wrap around the lower part of the export apparatus through hydrophobic interactions [9]. The cytoplasmic domain forms a nonameric ring with a pore size of 50 Å [9, 20–22]. It is postulated to serve as a docking platform for substrates and chaperones and to be involved in the specificity switching of secretion [9, 23, 24].

Base

The base is composed of the IM and OM rings. SctD and SctJ are components of the IM ring, where a SctD 24-mer ring (~250 Å wide) encompasses a smaller SctJ 24-mer ring (~180 Å wide) [25, 26]. SctD is anchored in the IM with a TMS located between three periplasmic domains and a cytoplasmic domain [25]. The cytoplasmic domain of SctD functions as a platform for interaction with the cytoplasmic components of the injectisome. The three periplasmic domains each contain topologically similar ring building motifs (RBMs) [27]. The RBM is composed of two α-helices folded against a β-sheet. It is present in all proteins that build the base structure (SctCDJ) and aids in the formation of the ring [27].

The smaller IM ring protein SctJ is anchored to the IM by an N-terminal lipid anchor and in addition by a C-terminal TMS in some T3SS [25, 28]. The two periplasmic domains of SctJ contain RBMs that constitute the concentric ring architecture of SctJ and provide inter-subunit interactions [27, 28]. The SctJ ring clamps the core export apparatus and supposedly brings it into its juxtamembrane position during assembly [5].

Above the IM rings, SctC comprises the OM ring. SctC contains three N-terminal domains (N0, N1 and N3), a secretin domain, and an S domain. The N-terminal domains contain the modular RBM [27], which aid in the formation of the SctC ring [26]. N0 and N1 domains constitute the ring below the periplasmic constriction and the N0 domain connects the OM ring to the IM ring of SctD. The N3 domain, together with the secretin and S domains, forms a pore structure above the periplasmic gate [29]. The secretin domain displays a double-walled β-barrel structure. In the lumen of the β-barrel, the inner wall acts as a transient periplasmic gate (~15 Å wide) to block access from the extracellular space, before needle filament polymerization opens up the gate [10, 29, 30]. The S domain is located outside of the barrel and functions as a binding site for pilotin SctG, a chaperone that is involved in SctC targeting and assembly in the OM [31, 32]. The outer wall of the secretin contains an amphipathic loop and hairpin at the distal end, which anchor the SctC ring to the OM [29]. SctC has repeatedly been shown to form a 15-mer concentric ring, which seemed unconventional to assemble with the 24 subunits of the SctD IM ring (ratio of 5 : 8) [25, 26, 29]. It has been revealed recently that the SctC ring possesses a sixteenth SctC molecule that only contains the N0 and N1 domains. This accessory domain resolves the symmetry mismatch between SctC and SctD in such a way that always two adjacent C-terminal domains of SctD interact with two N-terminal domains of SctC. The C-terminal domain of each third SctD does not participate in this interaction and folds somewhat to the outside of the base [10, 33, 34]. Interestingly, this partial SctC molecule only exists in the presence of the export apparatus. The absence of the latter results in a 15 : 23 stoichiometry between the OM and IM rings [34], showing how critical a precisely orchestrated assembly is for a functional injectisome. However, it is unclear at this point when the sixteenth SctC molecule is recruited and processed to its final form.

Cytoplasmic components

The cytoplasmic components of the injectisome comprise the proteins SctQ, SctL, SctO, SctN and SctK (Fig. 2b). The structural analysis of these components has been hampered by the difficulty of isolating this structure, possibly due to its dynamic nature. Cryo-ET analysis revealed that the cytoplasmic components form a cage of six pod-like structures comprising SctK and SctQ. These emerge from the needle complex towards the cytoplasmic side, forming a scaffold for a centrally positioned ATPase (SctN) that connects to the pods through six spoke-like structures of SctL. The central ATPase is connected to SctV by one molecule of SctO [35, 36].

Fig. 2.

Fig. 2.

Fig. 2.

Assembly of the type III secretion system. (a) Assembly of T3SS begins with formation of the export apparatus and the membrane rings, which follows either ‘inside-out’ or ‘outside-in’ models. In both models, the export apparatus self-sufficiently assembles and the smaller IM ring (SctJ) forms around the export apparatus. In the inside-out model, it is followed by the formations of the bigger IM ring (SctD) and OM secretin ring (SctC), in that order. SctC is recruited and oligomerized in the OM often with help of pilotin. In the outside-in model, the SctC ring formation is required for the SctD ring assembly. The export apparatus with the SctJ ring is then incorporated into the SctCD rings. (b) The sorting platform assembly and recruitment to the export apparatus and base is a dynamic and cyclic process. The sorting platform is capable of interacting with T3SS chaperone-substrate. (c) The assembly of the needle involves the secretion of early substrates SctI, the needle adaptor, followed by secretion of SctF, the needle filament. Needle length is regulated by SctP, which further interaction with SctU leads to switch of secretion to the intermediate substrates. (d-e) Translocators (SctBE) are then secreted and inserted into the host membrane where they form a pore. The needle tip guides translocon insertion. SctW regulates intermediate substrate secretion, by binding to SctV. Once SctW is degraded or secreted, switch to secretion of late substrates occurs and effectors are deployed into the host cells. Protein names follow the unified nomenclature according to Wagner and Diepold [155] and are depicted as letters (e.g. SctO: O). The different structural units are colour coded. Adapted from Wagner [102]. IM: inner membrane, PG: peptidoglycan, OM: outer membrane.

The pods connect to the cytoplasmic domain of the base inner ring protein SctD through SctK [35, 37]. SctK is composed of nine α-helices and folds into a kidney-like structure [38]. In addition to its base-associated form, SctK was also shown to form soluble complexes with the other cytosolic components [39–41]. The other pod component, SctQ, contains an internal translation start site, resulting in the expression of both full-length (SctQFL) and C-terminal fragment (SctQC) [42–44]. SctQFL and SctQC interact to form heterotrimers in a stoichiometry 1 SctQFL : 2 SctQC [45], where SctQC binds to the N-terminal domain of SctQFL. This interaction probably stabilizes SctQFL which is unstable without SctQC [42, 45–47]. SctQFL and SctQC are essential for function but the exact role of SctQC is still controversially discussed [45, 47].

SctL connects the pods in a spoke-like fashion to the ATPases SctN, forming a hub. The N-terminus of SctL interacts with SctQFL [35, 45, 48], while its C-terminus interacts with the N-terminus of SctN [45]. In vitro, SctL is insoluble by itself but forms soluble, stable complexes of SctQFL/SctQC/SctL, SctQFL/SctQC/SctL/SctN and SctL/SctN [45].

The ATPase SctN shares structural homology with the F1-ATPases and forms a hexameric ring containing six catalytic sites in different states [49], consistent with the rotary mechanism adopted by the F1-ATPases. Each SctN monomer comprises three subdomains. First, a C-terminal domain containing five α-helices forms the surface of the SctN pore. Then, a central ATPase, essential for catalytic activity, contains mixed A/B Rossman fold and Walker A and B motifs, typical of ATPases. Lastly, a N-terminal domain is required for oligomerization. The ATPase ring complex is similar to the cytoplasmic part of the FoF1 ATP synthases, where a rotary motor couples proton flow through Fo with ATP synthesis by F1. SctN was shown to dissociate the chaperone-substrate complexes by hydrolysis of ATP and to unfold substrates [50]. However, it is difficult to integrate this observation into a model where substrate-chaperone complexes also bind to the cytoplasmic ring of SctV.

SctO forms helical coiled-coil hairpins like its flagella counterpart FliJ [36, 49, 51–53]. The latter is structurally similar to the central stalk proteins of rotary ATPases. SctO links SctN and SctV. On one side, the N- and C-termini of SctO locate to the SctN C-terminal pore, where a SctN negatively charged region adjoins a positively charged region of SctO [49]. On the other side, SctO binds to a cleft between two different SctV protomers [51]. This interaction occurs in a region of SctV also described to bind chaperone-substrate complexes [54].

Needle structure

The needle filament is a helical conduit emerging from the export apparatus. It is anchored to the export apparatus by six molecules of the needle adapter protein SctI (also known as inner rod) [2, 7, 10, 12, 55]. SctI forms an alpha helical hairpin with the termini facing towards the core export apparatus and it is able to adapt to different structures, depending on the position. SctI1 binds extensively to the interface between SctR and SctT, while being partially ordered. The SctI2-6 molecules display more ordered helices and interact at the SctR-R and SctR-T interfaces, as well as with the surrounding SctC structure [10, 12]. The ability to adopt different structural conformations and establish different interactions allows SctI to function as a needle adaptor providing a homogenous platform for the following needle filament assembly [10, 12]. The needle filament is a stiff rod composed of SctF protomers that are structurally similar to SctI [56]. The C-terminal half of SctF faces the lumen of the needle filament. It measures about 15 Å in diameter, exhibits a right-handed spiral shape and is lined with conserved charged amino acid residues that are critical for the substrate secretion process [30, 57]. The N-terminus of SctF extends to the outside of the filament and lacks secondary structure [30]. The needle filament measures around 50 nm in length and its extent varies between bacterial species [58–60].

Needle tip

At the distal end of the needle filament, SctA assembles a needle tip complex, which is mostly found to be pentameric [61–63]. A recent structural study on the Salmonella needle tip complex has revealed that the complex forms a continuous conduit from the needle filament (~80 Å wide and ~800 Å long) but with a narrower opening at the distal end (~10 Å) [62]. The SctA structure is composed of a central coiled-coil domain in between the N- and C-terminal α/β domains [61, 64–67]. The coiled-coil domain is a conserved structural component across the T3SS needle tip homologs. It is involved in the interaction with the needle filament and oligomerization, and makes up the lumen of the tip complex [61, 64, 68]. Interestingly, the surface electrostatic properties of the lumen of the tip complex and of the needle complex are distinct, as the lumen of the tip complex is predominantly negatively charged [62]. However, the physiological relevance of this difference remains unclear. On the other hand, the N-terminal domain seems to differ in the structure depending on the bacterial species and the presence of a cognate chaperone. In Salmonella , Shigella , and Burkholderia , the N-terminal domain exhibits an α-helical hairpin structure and has been shown to act as a self-chaperone to inhibit premature oligomerization in the bacterial cytosol by binding to the C-terminal coiled-coil domain [64]. In Yersinia and Pseudomonas , the N-terminus of SctA has a globular domain and a separate cognate chaperone is required to prevent premature oligomerization and for successful secretion of SctA [63, 68–71]. In EPEC and EHEC, SctA does not form a pentameric complex but extends as a unique filamentous structure on top of the needle [72–74]. The lumen of the SctA filament has a similar diameter as the one of the needle filament and also similar electrostatic properties featuring a spiral hydrophobic groove lined with positively charged residues [75, 76]. The SctA homolog of EPEC and EHEC mainly consists of a conserved coiled-coil domain with a small, kinked N-terminal helix [28, 75, 76]. In addition, SctA in EPEC and EHEC also requires a cognate chaperone to inhibit premature polymerization [28, 77].

Translocators

The hydrophobic translocator proteins SctB and SctE form the translocon pore in the host cell membrane. SctB is the major and SctE is the minor translocator and both share a dedicated chaperone (see Targeting and transport across the injectisome). The major and minor translocators are predicted to contain two and one transmembrane helices, respectively. Additionally, these proteins contain coiled-coils and intrinsically disordered regions. Due to their nature, only the structures of translocators interacting with their respective T3SS chaperones have been solved, but not those of full-length translocators assembled in a membrane-inserted translocon structure [78, 79]. Cryo-ET analysis revealed that the formation of the translocon leads to membrane bending and takes place in regions where the needle contacts the host cell, suggesting that a close association of bacteria and host is required for an efficient translocation [58, 80]. Moreover, the needle tip was shown to function as a guide for translocon formation [81]. The translocon diameter varies between approx. 13.5–60 nm, depending on the conditions in which was determined (i.e. in vivo/in vitro) [82].

Type III secretion-independent assembly

For a successful infection, in competition with the host’s immune defence and the commensal microbiota, a timely and effective assembly of the intricate injectisome complex is essential.

Assembly of the injectisome can be divided into two parts: T3SS-independent and -dependent assembly. The first one utilizes the bacterial housekeeping protein secretion and membrane protein integration machinery (e.g. Sec, Bam, Lol) for building up the minimal secretion complex (export apparatus, base rings), which is complemented by the cytoplasmic components and then finished in a T3SS-dependent way.

Once the bacteria sense the right environmental signals, which vary among bacterial species, expression of the injectisome components is triggered. For example, in Yersinia , the host temperature of 37 °C activates the T3SS expression, whereas osmolarity, pH, and oxygen availability in the host intestinal environment serve as a trigger in Salmonella [83, 84]. Expression of the components leads to assembly of the injectisome complex to occur in about 30 min [19, 85]. Injectisome assembly begins with the formation of the core export apparatus in the IM [86] and with the formation of the secretin complex in the OM [87].

The core export apparatus assembles from a pseudo-hexameric SctR5T1 subcomplex through recruitment of four SctS followed by one SctU (Fig. 2a) [5–8]. During the assembly steps, inter- and intramolecular salt-bridges formed by highly conserved charged residues in the hydrophobic domains of the subunits play a crucial role in stabilization of the unique helical structure of the core export apparatus [13]. The SctV nonamer assembles around the core export apparatus which may support the displacement of the SctR5S4T1U1 complex from the IM [5, 9]. Surprisingly, SctV did not prove to be required for assembly of the base in S. Typhimurium T3SS-1 even though it is essential for secretion [86]. Instead, the core export apparatus was shown to nucleate assembly of the base, possibly by scaffolding SctJ oligomerization. In the absence of the core export apparatus, the base assembles inefficiently into a structure of wrong SctC (15 instead of 15.5) and SctDJ (23 instead of 24) stoichiometry [9, 34], highlighting the critical importance of the core export apparatus for base assembly. For SctD assembly, two models were proposed: in Yersinia , SctD assembles onto the pre-formed secretin [87], after which the export apparatus-SctJ complex and the secretin-SctD complex are thought to integrate. This model was termed the outside-in assembly model. Data from Salmonella T3SS-1 on the other hand suggest that SctD assembles onto or together with SctJ, after which the entire IM complex docks onto the OM secretin (inside-out model) [25, 88]. Both models have their challenges. For the outside-in model to work it is necessary that the SctD ring does not close completely as the export apparatus-SctJ complex must be able to integrate in the two-dimensional confinement of the IM. In the inside-out model it is in principle conceivable that the cytoplasmic components are recruited in the absence of the secretin, leading to the (premature) onset of secretion. Thus, checks and balances must exist to ensure the proper timing of the assembly and secretion processes that await elucidation.

A small lipoprotein called pilotin is important for assembly of the SctC secretin (Fig. 2a). The pilotin is anchored to the inner leaflet of the OM by a lipid anchor. It interacts with SctC monomers at the S domain, and targets them to the OM [31]. SctC then oligomerizes into a double-walled β-barrel [29]. In some pathogens including Yersinia , the pilotin is also implicated in oligomerization of SctC [32]. While in most studies, the secretin was shown to assemble independently of other T3SS components, except the pilotin, the formation of an SDS-resistant secretin oligomer depended on SctD in EPEC [89].

The cytoplasmic components were shown to exhibit a dynamic nature with cycles of assembly and disassembly [39, 44] (Fig. 2b). Since the cytoplasmic components are soluble, their assembly is not confined in a two-dimensional space and only limited by the affinity of the individual components to each other. Pod substructures may pre-form in the cytoplasm and play a role in targeting of substrates in a hierarchical fashion, for which reason the term ‘sorting platform’ was coined. The major integration point of the base and the cytoplasmic components is the SctK-SctD interaction [40]. It is unlikely, though, that SctK is recruited independently to SctD as the SctD ring possesses 24 potential binding sites while the cytoplasmic components feature a sixfold symmetry. This notion is supported by the observation that SctK shows a strongly diminished binding to SctD in the absence of SctQ when observed by fluorescence microscopy in Yersinia or cryo-ET in Salmonella [35, 39]. Interestingly, in vivo photocrosslinking showed an interaction of SctK with SctD in the absence of SctQ in Salmonella , even though its extent was reduced [37]. It is conceivable that in this case the binding of SctK to one of the 24 sites on SctD is possible albeit with an incorrect overall symmetry. In Salmonella , SctD organizes in six patches of four SctD N-terminal domains each, suggesting conformational changes to accommodate the hexameric symmetry of the cytoplasmic components [35]. However, this was not seen in Shigella ’s T3SS [90], so the interaction between SctD and SctK may vary from system to system. Overall, it is conceivable that the hexameric ATPase enforces the symmetry of the pods. Pre-formed complexes of the pods (SctQK), the spoke (SctL) and one subunit of the ATPase (SctN) dock onto SctD and are forced by SctN-SctN-interactions into the hexameric symmetry. Alternatively, the cytoplasmic components pre-form the complete hexameric complex before docking onto SctD altogether. In vitro data point at the first model [45], which would also accommodate the observation of the dynamic exchange of substructures of the cytoplasmic components more easily. The last component to assemble is possibly SctO, which integrates into the pore of the hexameric SctN and protrudes towards and connects to SctV.

After assembly of the cytoplasmic components, the T3SS is thought to be secretion competent.

Type III secretion-dependent assembly

Assembly of the injectisome is finalized by T3S-dependent processes that build up the needle filament and its associated structures. These processes require the targeting and secretion of proteins, i.e. substrates, through the otherwise assembled T3S apparatus (see Targeting and transport across the injectisome). Type III secretion-dependent assembly begins with the formation of the needle of the injectisome, followed by the insertion of translocators at the host membrane, which culminates in the secretion of effectors (Fig. 2c, d and e). Depending on the time of secretion, substrates of the T3SS can be grouped as: i) early substrates (components required for assembly of the needle filament); ii) intermediate substrates (needle tip/translocators); and iii) late substrates (effectors). The sequential secretion of the different substrates requires precise orchestration and distinct substrate specificity switches.

There are three early substrates common to all T3SS, the needle adapter SctI, the needle filament protein SctF and the needle length regulator SctP [88, 91]. In the Salmonella T3SS-1, a fourth early substrate, OrgC, has been reported to aid in polymerization of the needle filament at the distal end [92]. Whether there is a temporal regulation of secretion between the early substrates is not clear. It is clear though, that six subunits of the needle adapter SctI need to assemble onto the core export apparatus before assembly of the needle filament can begin, even though it was observed that SctI mutants with defects in needle assembly could be suppressed by overexpression of SctF, although inefficiently [93]. This suggests that SctI is not essential but facilitates the efficient anchoring of the needle filament to the core export apparatus, functioning as a flexible adapter protein. Interestingly, assembly of the SctI does not require prior assembly of SctC [55] and needle polymerization is still possible without SctC [25]. It has been postulated that secretion of SctI prior to secretin assembly activates a peptidoglycan-lytic enzyme in the periplasm of Salmonella in order to create space for subsequent assembly of the secretin [94]. However, it is unclear in this model how the secretin could assemble around a growing needle or how secretin and needle filament assembly are otherwise coordinated.

SctF polymerizes at the distal end of the needle [95] after travelling through the already assembled needle conduit (Fig. 2c). It is thought that the growing needle opens the secretin gate and then resumes in the extracellular space [29]. A variety of models of needle length control have been proposed over time [96]. It is now believed that the termination of needle elongation is regulated by the needle length regulator SctP and recent cryo-EM structure of the needle [30] suggests that it is independent of SctI. Deletion of SctP leads to extremely long needles and the oversecretion of SctI and SctF [18, 97–101]. How exactly SctP controls needle length has been debated for a long time [96, 102]. The most widely accepted model based on injectisome and flagella data states that SctP regulates needle length as an intermittently secreted infrequent ruler triggering the stop of needle elongation once the needle length exceeds the maximum extension of the unfolded SctP inside the needle conduit [59, 103]. SctP’s folded C-terminal domain is thought to trigger the substrate specificity switching to the secretion of intermediate substrates by interaction with the autocleaved C-terminal domain of SctU [99]. This interaction may induce a conformational change in SctU that disallows further secretion of early substrates and facilitates the binding of the gatekeeper SctW onto the cytoplasmic ring of SctV, which is postulated to be a prerequisite for secretion of intermediate substrates [17–19].

Upon substrate switching SctA (tip) is secreted and forms a pentameric ring whose subunits interact with SctF subunits in the same way as SctF subunits interact with each other [62]. This is in contrast with the previous model whereby the needle filament was thought to be topped by a tetrameric SctA together with an SctB (translocator) monomer [104]. Moreover, the insertion of SctA subunits at the top of the needle filament was shown to induce conformational changes in SctF [62]. Upon host cell contact, the needle tip may guide translocon formation [81]. The understanding of the mechanism of translocon assembly in the host membrane falls still short, due to a lack of high-resolution structures, but it seems to vary between different pathogens. After translocon insertion, assembly of the complete T3SS is finished and injection of effector proteins into the host cell can take place.

Targeting and transport across the injectisome

For injection into host cells, T3SS substrates must be targeted to the injectisome, unfolded, and transported across it.

An essential requirement for T3S is the presence of a non-cleavable N-terminal signal within the first 20 amino acid residues [105–107]. This signal is necessary and sufficient for T3S and is universal, meaning the T3SSs of different bacteria can recognize it [108–111], however, there is no clear consensus sequence. The signal region is enriched in serine, threonine, isoleucine and proline, and is unstructured. Bioinformatic tools allow more accurate predictions of the signal [112, 113]. Exceptions, where T3S signals can be located in the middle of [100] and at both N- and C-terminus of the substrates have also been reported [114–116]. Also a potential relevance of mRNA signals has been discussed [107, 117].

In addition to the presence of the N-terminal signal sequence, many substrates also contain a chaperone binding domain (CBD) located downstream of the T3S signal that serves as a binding site for cognate T3SS chaperones [118]. T3SS chaperones have been shown to play different roles such as maintaining substrates unfolded [119], avoiding protein degradation [120], preventing mistargeting of transmembrane substrates [121], prevent self-aggregation [122] and facilitating the hierarchy of secretion [40, 123]. However, their functions seem specific for each substrate as not each chaperone performs all the functions. Moreover, not each T3SS substrate requires cognate chaperones for secretion [124].

Chaperones are classified depending on their substrate specificity. Class I chaperones bind to T3SS effectors. These are acidic proteins that share a similar 3D structure with αβα sandwich fold. They form homodimers. CDBs of effectors wrap around the chaperone dimers and prevent complete folding of the effector [125]. Class IA chaperones exclusively bind one effector whereas class IB chaperones bind multiple effector proteins. Class II chaperones are encoded in close proximity to and bind to the CBD and TMD of intermediate substrates such as translocators. They are also structurally similar and are composed of α-helices and tetratricopeptide repeat (TPR) domains [125]. A concave face of the chaperones binds the CDB, which contains a specific motif, and the convex back face binds the TMD [78]. They have been suggested to form dimers in solution, but the biological relevance of dimerization is unclear. Class III chaperones are structurally similar α-helical chaperones that bind flagellin and needle subunits (SctF), to the C-termini of their substrates [126–129]. Class IV chaperones are a group structurally different from the other classes, comprising the chaperone of the EPEC filament [130]. Class V chaperones also bind needle subunits and mediate contact of SctF with the T3SS apparatus [131] and regulate the levels of class III chaperone to translocate early substrates.

The presence of a T3S signal sequence and of a CBD may not be sufficient for the successful secretion of a protein through T3SS. On the contrary, several studies have shown that so far not all proteins are compatible with T3S. They probably cannot be unfolded sufficiently and thus obstruct the secretion channel [132–134], meaning that the mechanical stability of a protein can inform its compatibility with T3S [135].

The chain of events that occur from the expression of a substrate protein to its secretion by the injectisome are still elusive on the molecular level. Effector proteins seem to be immobilized in the cytosol, whether the needle complex and cytoplasmic components are present or not [136]. Moreover, the cytosolic components SctKQL appear to bind chaperones with different affinities, thus influencing effector export [40]. Dynamic assembly and disassembly of these components and their cycling between the cytoplasm and the injectisome were proposed to increase the probability of encountering a chaperone-substrate pair. However, cycles of assembly and disassembly were reported to have a minimum half-time of 1 min [44], while the rate of substrate secretion was shown to be in the range of 7–60 effectors per second [137]. It is difficult to match these different figures, questioning the relevance of this targeting hypothesis. Moreover, it was observed that chaperone-substrate pairs were capable of binding SctN and SctV [24, 50, 54, 138, 139]. It is not understood whether these interactions required a previous interaction with the SctKQL complex (multi-step targeting) or occur directly (one-step targeting).

Secretion of intermediate substrates depends also on the so-called gatekeeper SctW (Fig. 2d) that was shown to bind to the cytoplasmic ring of SctV [140]. SctW may act as a specific adapter for class II chaperone-substrate complexes to ensure timely secretion of intermediate substrates [138–142]. After host cell sensing and T3S activation, SctW dissociates from SctV and is secreted or degraded, so that the system switches to secretion of late substrates (Fig. 2d). Accordingly, in the absence of SctW late substrates are oversecreted [140, 143, 144]. Moreover, SctW expression was shown to be regulated by temperature through an upstream RNA thermometer [145]. Additionally, interaction of SctW with the ATPase was suggested to play a role in different binding affinities of early and intermediate substrates [146]. The hierarchy of secretion of different effectors may be regulated, e.g. by the regulation of the timing and strength of substrate expression and by the binding affinities between chaperones, substrates and injectisome components (Fig. 2e).

The secretion process itself is energized by the utilization of the proton motive force by the transmembrane domains of SctV, as well as by the hydrolysis of ATP by SctN [147–149]. Since secretion was also observed independently of ATPase activity, it seems that the main driver for secretion is the proton motive force, possibly leveraged by the ATPase [150, 151]. The underlying molecular mechanism of secretion is, however, not known to date. The actual crossing of the secretion conduit may also be driven by diffusion as it was proposed for bacterial flagella [152] and by folding of the substrate proteins when they emerge from the secretion conduit.

Outlook

In recent years, the architecture of T3SS has been unveiled using different structural biology techniques such as cryo-ET, cryo-EM, and X-ray crystallography. The integrated picture obtained from all this structural information has led to a much improved understanding of the structure of the injectisome and to the generation of solid hypotheses on its function. However, some pieces are still missing, particularly transmembrane segments/regions and disordered or dynamic regions. The machine learning-based structure prediction tool Alphafold2 will certainly help to fill these gaps [153, 154]. Despite its only short time of availability, it has already been used to predict structures of needle tips and needle tip chimaeras [81], supported the solution of crystal structures by molecular replacement using Alphafold2 prediction [139] and helped to generate an integrated model of the cytoplasmic components of the Salmonella T3SS-1, which was then validated by in vivo photocrosslinking [37]. Beyond AlphaFold, molecular dynamics simulations may become increasingly useful to predict the movements of these large machines and thus infer function more accurately.

Funding information

The work in the laboratory of S.W. related to this review was supported by the Deutsche Forschungsgemeinschaft (DFG) grant WA3299/5-1, by the cluster of excellence EXC2124 Controlling Microbes to Fight Infections (CMFI) and by the German Center for Infection research (DZIF) grant TTU 06 819.

Acknowledgements

We thank Libera Lo Presti for critical reading of the manuscript.

Conflicts of interest

The authors declare that there are no conflicts of interest.

Footnotes

Abbreviations: CBD, chaperone binding domain; cryo-EM, cryo-electron microscopy; cryo-ET, cryo-electron tomography; EHEC, enterohemorrhagic Escherichia coli; EPEC, enteropathogenic Escherichia coli; fT3SS, flagellar type III secretion system; IM, inner membrane; OM, outer membrane; PG, peptidoglycan; RBM, ring building motif; TMD, transmembrane domain; TMS, transmembrane segment; TPR, tetratricopeptide repeat; T3SS, type III secretion system; vT3SS, virulence type III secretion system.

References

  • 1.Abby SS, Rocha EPC. The non-flagellar type III secretion system evolved from the bacterial flagellum and diversified into host-cell adapted systems. PLoS Genet. 2012;8:e1002983. doi: 10.1371/journal.pgen.1002983. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Zilkenat S, Franz-Wachtel M, Stierhof Y-D, Galán JE, Macek B, et al. Determination of the stoichiometry of the complete bacterial type III secretion needle complex using a combined quantitative proteomic approach. Mol Cell Proteomics. 2016;15:1598–1609. doi: 10.1074/mcp.M115.056598. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Troisfontaines P, Cornelis GR. Type III secretion: more systems than you think. Physiology. 2005;20:326–339. doi: 10.1152/physiol.00011.2005. [DOI] [PubMed] [Google Scholar]
  • 4.Heuck AP, Brovedan MA. Evolutionary conservation, variability, and adaptation of Type III secretion systems. J Membr Biol. 2022;255:599–612. doi: 10.1007/s00232-022-00247-9. [DOI] [PubMed] [Google Scholar]
  • 5.Kuhlen L, Abrusci P, Johnson S, Gault J, Deme J, et al. Author correction: structure of the core of the type III secretion system export apparatus. Nat Struct Mol Biol. 2018;25:743. doi: 10.1038/s41594-018-0095-8. [DOI] [PubMed] [Google Scholar]
  • 6.Kuhlen L, Johnson S, Zeitler A, Bäurle S, Deme JC, et al. The substrate specificity switch FlhB assembles onto the export gate to regulate type three secretion. Nat Commun. 2020;11:1296. doi: 10.1038/s41467-020-15071-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Dietsche T, Tesfazgi Mebrhatu M, Brunner MJ, Abrusci P, Yan J, et al. Structural and functional characterization of the bacterial Type III secretion export apparatus. PLoS Pathog. 2016;12:e1006071. doi: 10.1371/journal.ppat.1006071. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Johnson S, Kuhlen L, Deme JC, Abrusci P, Lea SM. The structure of an injectisome export gate demonstrates conservation of architecture in the core export gate between flagellar and virulence type III secretion systems. mBio. 2019;10:e00818-19. doi: 10.1128/mBio.00818-19. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Butan C, Lara-Tejero M, Li W, Liu J, Galán JE. High-resolution view of the type III secretion export apparatus in situ reveals membrane remodeling and a secretion pathway. Proc Natl Acad Sci. 2019;116:24786–24795. doi: 10.1073/pnas.1916331116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Hu J, Worrall LJ, Vuckovic M, Hong C, Deng W, et al. T3S injectisome needle complex structures in four distinct states reveal the basis of membrane coupling and assembly. Nat Microbiol. 2019;4:2010–2019. doi: 10.1038/s41564-019-0545-z. [DOI] [PubMed] [Google Scholar]
  • 11.Hüsing S, Halte M, van Look U, Guse A, Gálvez EJC, et al. Control of membrane barrier during bacterial type-III protein secretion. Nat Commun. 2021;12:3999. doi: 10.1038/s41467-021-24226-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Miletic S, Fahrenkamp D, Goessweiner-Mohr N, Wald J, Pantel M, et al. Substrate-engaged type III secretion system structures reveal gating mechanism for unfolded protein translocation. Nat Commun. 2021;12:1546. doi: 10.1038/s41467-021-21143-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Singh N, Kronenberger T, Eipper A, Weichel F, Franz-Wachtel M, et al. Conserved salt bridges facilitate assembly of the helical core export apparatus of a Salmonella enterica Type III secretion system. J Mol Biol. 2021;433:167175. doi: 10.1016/j.jmb.2021.167175. [DOI] [PubMed] [Google Scholar]
  • 14.Ferris HU, Furukawa Y, Minamino T, Kroetz MB, Kihara M, et al. FlhB regulates ordered export of flagellar components via autocleavage mechanism. J Biol Chem. 2005;280:41236–41242. doi: 10.1074/jbc.M509438200. [DOI] [PubMed] [Google Scholar]
  • 15.Minamino T, Macnab RM. Domain structure of Salmonella FlhB, a flagellar export component responsible for substrate specificity switching. J Bacteriol. 2000;182:4906–4914. doi: 10.1128/JB.182.17.4906-4914.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Zarivach R, Deng W, Vuckovic M, Felise HB, Nguyen HV, et al. 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]
  • 17.Björnfot A-C, Lavander M, Forsberg A, Wolf-Watz H. Autoproteolysis of YscU of Yersinia pseudotuberculosis is important for regulation of expression and secretion of Yop proteins. J Bacteriol. 2009;191:4259–4267. doi: 10.1128/JB.01730-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Edqvist PJ, Olsson J, Lavander M, Sundberg L, Forsberg A, et al. YscP and YscU regulate substrate specificity of the Yersinia type III secretion system. J Bacteriol. 2003;185:2259–2266. doi: 10.1128/JB.185.7.2259-2266.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Monjarás Feria JV, Lefebre MD, Stierhof Y-D, Galán JE, Wagner S. Role of autocleavage in the function of a type III secretion specificity switch protein in Salmonella enterica serovar Typhimurium. mBio. 2015;6:e01459–15. doi: 10.1128/mBio.01459-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Abrusci P, Vergara-Irigaray M, Johnson S, Beeby MD, Hendrixson DR, et al. 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]
  • 21.Inoue Y, Kinoshita M, Namba K, Minamino T. Mutational analysis of the C-terminal cytoplasmic domain of FlhB, a transmembrane component of the flagellar type III protein export apparatus in Salmonella . Genes Cells. 2019;24:408–421. doi: 10.1111/gtc.12684. [DOI] [PubMed] [Google Scholar]
  • 22.Worrall LJ, Vuckovic M, Strynadka NCJ. 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]
  • 23.Minamino T, Kinoshita M, Inoue Y, Kitao A, Namba K. Conserved GYXLI Motif of FlhA Is Involved in dynamic domain motions of FlhA required for flagellar protein export. Microbiol Spectr. 2022;10:e0111022. doi: 10.1128/spectrum.01110-22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Yuan B, Portaliou AG, Parakra R, Smit JH, Wald J, et al. Structural dynamics of the functional nonameric Type III translocase export gate. J Mol Biol. 2021;433:167188. doi: 10.1016/j.jmb.2021.167188. [DOI] [PubMed] [Google Scholar]
  • 25.Schraidt O, Lefebre MD, Brunner MJ, Schmied WH, Schmidt A, et al. 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] [PMC free article] [PubMed] [Google Scholar]
  • 26.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]
  • 27.Spreter T, Yip CK, Sanowar S, André I, Kimbrough TG, et al. 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]
  • 28.Yip CK, Kimbrough TG, Felise HB, Vuckovic M, Thomas NA, et al. 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]
  • 29.Worrall LJ, Hong C, Vuckovic M, Deng W, Bergeron JRC, et al. Near-atomic-resolution cryo-EM analysis of the Salmonella T3S injectisome basal body. Nature. 2016;540:597–601. doi: 10.1038/nature20576. [DOI] [PubMed] [Google Scholar]
  • 30.Hu J, Worrall LJ, Hong C, Vuckovic M, Atkinson CE, et al. Cryo-EM analysis of the T3S injectisome reveals the structure of the needle and open secretin. Nat Commun. 2018;9:3840. doi: 10.1038/s41467-018-06298-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Majewski DD, Okon M, Heinkel F, Robb CS, Vuckovic M, et al. Characterization of the pilotin-secretin complex from the Salmonella enterica Type III secretion system using hybrid structural methods. Structure. 2021;29:125–138. doi: 10.1016/j.str.2020.08.006. [DOI] [PubMed] [Google Scholar]
  • 32.Burghout P, Beckers F, de Wit E, van Boxtel R, Cornelis GR, et al. 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]
  • 33.Lunelli M, Kamprad A, Bürger J, Mielke T, Spahn CMT, et al. Cryo-EM structure of the Shigella type III needle complex. PLoS Pathog. 2020;16:e1008263. doi: 10.1371/journal.ppat.1008263. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Goessweiner-Mohr N, Kotov V, Brunner MJ, Mayr J, Wald J, et al. Structural control for the coordinated assembly into functional pathogenic type-3 secretion systems. Mol Biol. 2019:714097. doi: 10.1101/714097. [DOI] [Google Scholar]
  • 35.Hu B, Lara-Tejero M, Kong Q, Galán JE, Liu J. In situ molecular architecture of the Salmonella Type III secretion machine. Cell. 2017;168:1065–1074. doi: 10.1016/j.cell.2017.02.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Ibuki T, Imada K, Minamino T, Kato T, Miyata T, et al. 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]
  • 37.Soto JE, Galán JE, Lara-Tejero M. Assembly and architecture of the type III secretion sorting platform. Proc Natl Acad Sci U S A. 2022;119:e2218010119. doi: 10.1073/pnas.2218010119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Muthuramalingam M, Whittier SK, Lovell S, Battaile KP, Tachiyama S, et al. The structures of SctK and SctD from Pseudomonas aeruginosa reveal the interface of the type III secretion system basal body and sorting platform. J Mol Biol. 2020;432:166693. doi: 10.1016/j.jmb.2020.10.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Diepold A, Sezgin E, Huseyin M, Mortimer T, Eggeling C, et al. A dynamic and adaptive network of cytosolic interactions governs protein export by the T3SS injectisome. Nat Commun. 2017;8:15940. doi: 10.1038/ncomms15940. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Lara-Tejero M, Kato J, Wagner S, Liu X, Galán 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]
  • 41.Johnson S, Blocker A. Characterization of soluble complexes of the Shigella flexneri type III secretion system ATPase. FEMS Microbiol Lett. 2008;286:274–278. doi: 10.1111/j.1574-6968.2008.01284.x. [DOI] [PubMed] [Google Scholar]
  • 42.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]
  • 43.McDowell MA, Marcoux J, McVicker G, Johnson S, Fong YH, et al. Characterisation of Shigella Spa33 and Thermotoga FliM/N reveals a new model for C-ring assembly in T3SS. Mol Microbiol. 2016;99:749–766. doi: 10.1111/mmi.13267. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.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] [PMC free article] [PubMed] [Google Scholar]
  • 45.Bernal I, Börnicke J, Heidemann J, Svergun D, Horstmann JA, et al. Molecular organization of soluble type III secretion system sorting platform complexes. J Mol Biol. 2019;431:3787–3803. doi: 10.1016/j.jmb.2019.07.004. [DOI] [PubMed] [Google Scholar]
  • 46.Yu X-J, Liu M, Matthews S, Holden DW. Tandem translation generates a chaperone for the Salmonella type III secretion system protein SsaQ. J Biol Chem. 2011;286:36098–36107. doi: 10.1074/jbc.M111.278663. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Lara-Tejero M, Qin Z, Hu B, Butan C, Liu J, et al. Role of SpaO in the assembly of the sorting platform of a Salmonella type III secretion system. PLoS Pathog. 2019;15:e1007565. doi: 10.1371/journal.ppat.1007565. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Notti RQ, Bhattacharya S, Lilic M, Stebbins CE. A common assembly module in injectisome and flagellar type III secretion sorting platforms. Nat Commun. 2015;6:7125. doi: 10.1038/ncomms8125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Majewski DD, Worrall LJ, Hong C, Atkinson CE, Vuckovic M, et al. Cryo-EM structure of the homohexameric T3SS ATPase-central stalk complex reveals rotary ATPase-like asymmetry. Nat Commun. 2019;10:626. doi: 10.1038/s41467-019-08477-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Akeda Y, Galán JE. Chaperone release and unfolding of substrates in type III secretion. Nature. 2005;437:911–915. doi: 10.1038/nature03992. [DOI] [PubMed] [Google Scholar]
  • 51.Jensen JL, Yamini S, Rietsch A, Spiller BW. The structure of the Type III secretion system export gate with CdsO, an ATPase lever arm. PLoS Pathog. 2020;16:e1008923. doi: 10.1371/journal.ppat.1008923. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Romo-Castillo M, Andrade A, Espinosa N, Monjarás Feria J, Soto E, et al. EscO, a functional and structural analog of the flagellar FliJ protein, is a positive regulator of EscN ATPase activity of the enteropathogenic Escherichia coli injectisome. J Bacteriol. 2014;196:2227–2241. doi: 10.1128/JB.01551-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Lorenzini E, Singer A, Singh B, Lam R, Skarina T, et al. 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]
  • 54.Xing Q, Shi K, Portaliou A, Rossi P, Economou A, et al. Structures of chaperone-substrate complexes docked onto the export gate in a type III secretion system. Nat Commun. 2018;9:1773. doi: 10.1038/s41467-018-04137-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Torres-Vargas CE, Kronenberger T, Roos N, Dietsche T, Poso A, et al. The inner rod of virulence-associated type III secretion systems constitutes a needle adapter of one helical turn that is deeply integrated into the system’s export apparatus. Mol Microbiol. 2019;112:918–931. doi: 10.1111/mmi.14327. [DOI] [PubMed] [Google Scholar]
  • 56.Loquet A, Sgourakis NG, Gupta R, Giller K, Riedel D, et al. 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]
  • 57.Guo EZ, Desrosiers DC, Zalesak J, Tolchard J, Berbon M, et al. A polymorphic helix of A Salmonella needle protein relays signals defining distinct steps in type III secretion. PLoS Biol. 2019;17:e3000351. doi: 10.1371/journal.pbio.3000351. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Park D, Lara-Tejero M, Waxham MN, Li W, Hu B, et al. Visualization of the type III secretion mediated Salmonella-host cell interface using cryo-electron tomography. Elife. 2018;7:e39514. doi: 10.7554/eLife.39514. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.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]
  • 60.Tamano K, Aizawa S, Katayama E, Nonaka T, Imajoh-Ohmi S, et al. Supramolecular structure of the Shigella type III secretion machinery: the needle part is changeable in length and essential for delivery of effectors. EMBO J. 2000;19:3876–3887. doi: 10.1093/emboj/19.15.3876. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Rathinavelan T, Lara-Tejero M, Lefebre M, Chatterjee S, McShan AC, et al. 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]
  • 62.Guo EZ, Galán JE. Cryo-EM structure of the needle filament tip complex of the Salmonella type III secretion injectisome. Proc Natl Acad Sci. 2021;118:e2114552118. doi: 10.1073/pnas.2114552118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Broz P, Mueller CA, Müller SA, Philippsen A, Sorg I, et al. Function and molecular architecture of the Yersinia injectisome tip complex. Mol Microbiol. 2007;65:1311–1320. doi: 10.1111/j.1365-2958.2007.05871.x. [DOI] [PubMed] [Google Scholar]
  • 64.Johnson S, Roversi P, Espina M, Olive A, Deane JE, et al. 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]
  • 65.Chatterjee S, Zhong D, Nordhues BA, Battaile KP, Lovell S, et al. 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]
  • 66.Erskine PT, Knight MJ, Ruaux A, Mikolajek H, Wong Fat Sang N, et al. High resolution structure of BipD: an invasion protein associated with the type III secretion system of Burkholderia pseudomallei . J Mol Biol. 2006;363:125–136. doi: 10.1016/j.jmb.2006.07.069. [DOI] [PubMed] [Google Scholar]
  • 67.Chaudhury S, Battaile KP, Lovell S, Plano GV, De Guzman RN. Structure of the Yersinia pestis tip protein LcrV refined to 1.65 Å resolution. Acta Crystallogr Sect F Struct Biol Cryst Commun. 2013:477–481. doi: 10.1107/S1744309113008579. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Gébus C, Faudry E, Bohn Y-ST, Elsen S, Attree I. Oligomerization of PcrV and LcrV, protective antigens of Pseudomonas aeruginosa and Yersinia pestis . J Biol Chem. 2008;283:23940–23949. doi: 10.1074/jbc.M803146200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Lee P-C, Stopford CM, Svenson AG, Rietsch A. Control of effector export by the Pseudomonas aeruginosa type III secretion proteins PcrG and PcrV. Mol Microbiol. 2010;75:924–941. doi: 10.1111/j.1365-2958.2009.07027.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Matson JS, Nilles ML. LcrG-LcrV interaction is required for control of Yops secretion in Yersinia pestis . J Bacteriol. 2001;183:5082–5091. doi: 10.1128/JB.183.17.5082-5091.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Matson JS, Nilles ML. Interaction of the Yersinia pestis type III regulatory proteins LcrG and LcrV occurs at a hydrophobic interface. BMC Microbiol. 2002;2:16. doi: 10.1186/1471-2180-2-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Sekiya K, Ohishi M, Ogino T, Tamano K, Sasakawa C, et al. Supermolecular structure of the enteropathogenic Escherichia coli type III secretion system and its direct interaction with the EspA-sheath-like structure. Proc Natl Acad Sci U S A. 2001;98:11638–11643. doi: 10.1073/pnas.191378598. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Wang YA, Yu X, Yip C, Strynadka NC, Egelman EH. Structural polymorphism in bacterial EspA filaments revealed by cryo-EM and an improved approach to helical reconstruction. Structure. 2006;14:1189–1196. doi: 10.1016/j.str.2006.05.018. [DOI] [PubMed] [Google Scholar]
  • 74.Daniell SJ, Kocsis E, Morris E, Knutton S, Booy FP, et al. 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]
  • 75.Lyons BJE, Atkinson CE, Deng W, Serapio-Palacios A, Finlay BB, et al. Cryo-EM structure of the EspA filament from enteropathogenic Escherichia coli: revealing the mechanism of effector translocation in the T3SS. Structure. 2021;29:479–487. doi: 10.1016/j.str.2020.12.009. [DOI] [PubMed] [Google Scholar]
  • 76.Zheng W, Peña A, Ilangovan A, Baghshomali YN, Frankel G, et al. Cryoelectron-microscopy structure of the enteropathogenic Escherichia coli type III secretion system EspA filament. Proc Natl Acad Sci U S A. 2021;118:e2022826118. doi: 10.1073/pnas.2022826118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Creasey EA, Friedberg D, Shaw RK, Umanski T, Knutton S, et al. CesAB is an enteropathogenic Escherichia coli chaperone for the type-III translocator proteins EspA and EspB. Microbiology. 2003;149:3639–3647. doi: 10.1099/mic.0.26735-0. [DOI] [PubMed] [Google Scholar]
  • 78.Nguyen VS, Jobichen C, Tan KW, Tan YW, Chan SL, et al. Structure of AcrH-AopB chaperone-translocator complex reveals a role for membrane hairpins in Type III secretion system translocon assembly. Structure. 2015;23:2022–2031. doi: 10.1016/j.str.2015.08.014. [DOI] [PubMed] [Google Scholar]
  • 79.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]
  • 80.Lara-Tejero M, Galán 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]
  • 81.Kundracik E, Trichka J, Díaz Aponte J, Roistacher A, Rietsch A. PopB-PcrV interactions are essential for pore formation in the Pseudomonas aeruginosa Type III secretion system translocon. mBio. 2022;13:e0238122. doi: 10.1128/mbio.02381-22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Ide T, Laarmann S, Greune L, Schillers H, Oberleithner H, et al. Characterization of translocation pores inserted into plasma membranes by type III-secreted Esp proteins of enteropathogenic Escherichia coli . Cell Microbiol. 2001;3:669–679. doi: 10.1046/j.1462-5822.2001.00146.x. [DOI] [PubMed] [Google Scholar]
  • 83.Srikanth CV, Mercado-Lubo R, Hallstrom K, McCormick BA. Salmonella effector proteins and host-cell responses. Cell Mol Life Sci. 2011;68:3687–3697. doi: 10.1007/s00018-011-0841-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Volk M, Vollmer I, Heroven AK, Dersch P. Transcriptional and post-transcriptional regulatory mechanisms controlling type III secretion. Curr Top Microbiol Immunol. 2020;427:11–33. doi: 10.1007/82_2019_168. [DOI] [PubMed] [Google Scholar]
  • 85.Westerhausen S, Nowak M, Torres-Vargas CE, Bilitewski U, Bohn E, et al. A NanoLuc luciferase-based assay enabling the real-time analysis of protein secretion and injection by bacterial type III secretion systems. Mol Microbiol. 2020;113:1240–1254. doi: 10.1111/mmi.14490. [DOI] [PubMed] [Google Scholar]
  • 86.Wagner S, Königsmaier L, Lara-Tejero M, Lefebre M, Marlovits TC, et al. Organization and coordinated assembly of the type III secretion export apparatus. Proc Natl Acad Sci. 2010;107:17745–17750. doi: 10.1073/pnas.1008053107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Diepold A, Amstutz M, Abel S, Sorg I, Jenal U, et al. Deciphering the assembly of the Yersinia type III secretion injectisome. EMBO J. 2010;29:1928–1940. doi: 10.1038/emboj.2010.84. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Sukhan A, Kubori T, Wilson J, Galán JE. Genetic analysis of assembly of the Salmonella enterica serovar Typhimurium type III secretion-associated needle complex. J Bacteriol. 2001;183:1159–1167. doi: 10.1128/JB.183.4.1159-1167.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Tseytin I, Dagan A, Oren S, Sal-Man N. The role of EscD in supporting EscC polymerization in the type III secretion system of enteropathogenic Escherichia coli . Biochim Biophys Acta Biomembr. 2018;1860:384–395. doi: 10.1016/j.bbamem.2017.10.001. [DOI] [PubMed] [Google Scholar]
  • 90.Tachiyama S, Chang Y, Muthuramalingam M, Hu B, Barta ML, et al. The cytoplasmic domain of MxiG interacts with MxiK and directs assembly of the sorting platform in the Shigella type III secretion system. J Biol Chem. 2019;294:19184–19196. doi: 10.1074/jbc.RA119.009125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Diepold A, Wiesand U, Amstutz M, Cornelis GR. Assembly of the Yersinia injectisome: the missing pieces. Mol Microbiol. 2012;85:878–892. doi: 10.1111/j.1365-2958.2012.08146.x. [DOI] [PubMed] [Google Scholar]
  • 92.Kato J, Dey S, Soto JE, Butan C, Wilkinson MC, et al. A protein secreted by the Salmonella type III secretion system controls needle filament assembly. Elife. 2018;7:e35886. doi: 10.7554/eLife.35886. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Monlezun L, Liebl D, Fenel D, Grandjean T, Berry A, et al. PscI is a type III secretion needle anchoring protein with in vitro polymerization capacities. Mol Microbiol. 2015;96:419–436. doi: 10.1111/mmi.12947. [DOI] [PubMed] [Google Scholar]
  • 94.Burkinshaw BJ, Deng W, Lameignère E, Wasney GA, Zhu H, et al. Structural analysis of a specialized type III secretion system peptidoglycan-cleaving enzyme. J Biol Chem. 2015;290:10406–10417. doi: 10.1074/jbc.M115.639013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Poyraz O, Schmidt H, Seidel K, Delissen F, Ader C, et al. 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]
  • 96.Galán 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]
  • 97.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]
  • 98.Sukhan A, Kubori T, Galán JE. Synthesis and localization of the Salmonella SPI-1 type III secretion needle complex proteins PrgI and PrgJ. J Bacteriol. 2003;185:3480–3483. doi: 10.1128/JB.185.11.3480-3483.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Monjarás Feria J, García-Gómez E, Espinosa N, Minamino T, Namba K, et al. Role of EscP (Orf16) in injectisome biogenesis and regulation of type III protein secretion in enteropathogenic Escherichia coli . J Bacteriol. 2012;194:6029–6045. doi: 10.1128/JB.01215-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Agrain C, Sorg I, Paroz C, Cornelis GR. Secretion of YscP from Yersinia enterocolitica is essential to control the length of the injectisome needle but not to change the type III secretion substrate specificity. Mol Microbiol. 2005;57:1415–1427. doi: 10.1111/j.1365-2958.2005.04758.x. [DOI] [PubMed] [Google Scholar]
  • 101.Bergeron JRC, Fernández L, Wasney GA, Vuckovic M, Reffuveille F, et al. The structure of a type 3 secretion system (T3SS) ruler protein suggests a molecular mechanism for needle length sensing. J Biol Chem. 2016;291:1676–1691. doi: 10.1074/jbc.M115.684423. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Wagner S, Grin I, Malmsheimer S, Singh N, Torres-Vargas CE, et al. Bacterial type III secretion systems: a complex device for the delivery of bacterial effector proteins into eukaryotic host cells. FEMS Microbiol Lett. 2018;365:fny201. doi: 10.1093/femsle/fny201. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Erhardt M, Singer HM, Wee DH, Keener JP, Hughes KT. An infrequent molecular ruler controls flagellar hook length in Salmonella enterica . EMBO J. 2011;30:2948–2961. doi: 10.1038/emboj.2011.185. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Cheung M, Shen D-K, Makino F, Kato T, Roehrich AD, et al. 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]
  • 105.Sory MP, Boland A, Lambermont I, Cornelis GR. Identification of the YopE and YopH domains required for secretion and internalization into the cytosol of macrophages, using the cyaA gene fusion approach. Proc Natl Acad Sci. 1995;92:11998–12002. doi: 10.1073/pnas.92.26.11998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.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]
  • 107.Rüssmann H, Kubori T, Sauer J, Galán JE. Molecular and functional analysis of the type III secretion signal of the Salmonella enterica InvJ protein. Mol Microbiol. 2002;46:769–779. doi: 10.1046/j.1365-2958.2002.03196.x. [DOI] [PubMed] [Google Scholar]
  • 108.Rosqvist R, Håkansson S, Forsberg A, Wolf-Watz H. Functional conservation of the secretion and translocation machinery for virulence proteins of Yersiniae, Salmonellae and Shigellae . EMBO J. 1995;14:4187–4195. doi: 10.1002/j.1460-2075.1995.tb00092.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Frithz-Lindsten E, Holmström A, Jacobsson L, Soltani M, Olsson J, et al. Functional conservation of the effector protein translocators PopB/YopB and PopD/YopD of Pseudomonas aeruginosa and Yersinia pseudotuberculosis . Mol Microbiol. 1998;29:1155–1165. doi: 10.1046/j.1365-2958.1998.00994.x. [DOI] [PubMed] [Google Scholar]
  • 110.Anderson DM, Fouts DE, Collmer A, Schneewind O. Reciprocal secretion of proteins by the bacterial type III machines of plant and animal pathogens suggests universal recognition of mRNA targeting signals. Proc Natl Acad Sci. 1999;96:12839–12843. doi: 10.1073/pnas.96.22.12839. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Subtil A, Parsot C, Dautry-Varsat A. Secretion of predicted Inc proteins of Chlamydia pneumoniae by a heterologous type III machinery. Mol Microbiol. 2001;39:792–800. doi: 10.1046/j.1365-2958.2001.02272.x. [DOI] [PubMed] [Google Scholar]
  • 112.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] [PMC free article] [PubMed] [Google Scholar]
  • 113.Wang Y, Zhang Q, Sun M-A, Guo D. High-accuracy prediction of bacterial type III secreted effectors based on position-specific amino acid composition profiles. Bioinformatics. 2011;27:777–784. doi: 10.1093/bioinformatics/btr021. [DOI] [PubMed] [Google Scholar]
  • 114.Allen-Vercoe E, Toh MCW, Waddell B, Ho H, DeVinney R. A carboxy-terminal domain of Tir from enterohemorrhagic Escherichia coli O157:H7 (EHEC O157:H7) required for efficient type III secretion. FEMS Microbiol Lett. 2005;243:355–364. doi: 10.1016/j.femsle.2004.12.027. [DOI] [PubMed] [Google Scholar]
  • 115.Kim BH, Kim HG, Kim JS, Jang JI, Park YK. Analysis of functional domains present in the N-terminus of the SipB protein. Microbiology. 2007;153:2998–3008. doi: 10.1099/mic.0.2007/007872-0. [DOI] [PubMed] [Google Scholar]
  • 116.Tomalka AG, Stopford CM, Lee PC, Rietsch A. A translocator-specific export signal establishes the translocator-effector secretion hierarchy that is important for type III secretion system function. Mol Microbiol. 2012;86:1464–1481. doi: 10.1111/mmi.12069. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Ramamurthi KS, Schneewind O. Yersinia yopQ mRNA encodes a bipartite type III secretion signal in the first 15 codons. Mol Microbiol. 2003;50:1189–1198. doi: 10.1046/j.1365-2958.2003.03772.x. [DOI] [PubMed] [Google Scholar]
  • 118.Cheng LW, Anderson DM, Schneewind O. Two independent type III secretion mechanisms for YopE in Yersinia enterocolitica . Mol Microbiol. 1997;24:757–765. doi: 10.1046/j.1365-2958.1997.3831750.x. [DOI] [PubMed] [Google Scholar]
  • 119.Stebbins CE, Galán 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]
  • 120.Tucker SC, Galán JE. Complex function for SicA, a Salmonella enterica serovar typhimurium type III secretion-associated chaperone. J Bacteriol. 2000;182:2262–2268. doi: 10.1128/JB.182.8.2262-2268.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Krampen L, Malmsheimer S, Grin I, Trunk T, Lührmann A, et al. Revealing the mechanisms of membrane protein export by virulence-associated bacterial secretion systems. Nat Commun. 2018;9:3467. doi: 10.1038/s41467-018-05969-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Letzelter M, Sorg I, Mota LJ, Meyer S, Stalder J, et al. The discovery of SycO highlights a new function for type III secretion effector chaperones. EMBO J. 2006;25:3223–3233. doi: 10.1038/sj.emboj.7601202. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Portaliou AG, Tsolis KC, Loos MS, Balabanidou V, Rayo J, et al. Hierarchical protein targeting and secretion is controlled by an affinity switch in the type III secretion system of enteropathogenic Escherichia coli . EMBO J. 2017;36:3517–3531. doi: 10.15252/embj.201797515. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Ernst NH, Reeves AZ, Ramseyer JE, Lesser CF. High-throughput screening of Type III secretion determinants reveals a major chaperone-independent pathway. mBio. 2018;9:e01050-18. doi: 10.1128/mBio.01050-18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Parsot C, Hamiaux C, Page A-L. The various and varying roles of specific chaperones in type III secretion systems. Curr Opin Microbiol. 2003;6:7–14. doi: 10.1016/s1369-5274(02)00002-4. [DOI] [PubMed] [Google Scholar]
  • 126.Sun P, Tropea JE, Austin BP, Cherry S, Waugh DS. Structural characterization of the Yersinia pestis type III secretion system needle protein YscF in complex with its heterodimeric chaperone YscE/YscG. J Mol Biol. 2008;377:819–830. doi: 10.1016/j.jmb.2007.12.067. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Quinaud M, Chabert J, Faudry E, Neumann E, Lemaire D, et al. The PscE-PscF-PscG complex controls type III secretion needle biogenesis in Pseudomonas aeruginosa . J Biol Chem. 2005;280:36293–36300. doi: 10.1074/jbc.M508089200. [DOI] [PubMed] [Google Scholar]
  • 128.Evdokimov AG, Phan J, Tropea JE, Routzahn KM, Peters HK, et al. Similar modes of polypeptide recognition by export chaperones in flagellar biosynthesis and type III secretion. Nat Struct Biol. 2003;10:789–793. doi: 10.1038/nsb982. [DOI] [PubMed] [Google Scholar]
  • 129.Muskotál A, Király R, Sebestyén A, Gugolya Z, Végh BM, et al. Interaction of FliS flagellar chaperone with flagellin. FEBS Lett. 2006;580:3916–3920. doi: 10.1016/j.febslet.2006.06.024. [DOI] [PubMed] [Google Scholar]
  • 130.Yip CK, Finlay BB, Strynadka NCJ. Structural characterization of a type III secretion system filament protein in complex with its chaperone. Nat Struct Mol Biol. 2005;12:75–81. doi: 10.1038/nsmb879. [DOI] [PubMed] [Google Scholar]
  • 131.Souza C de A, Richards KL, Park Y, Schwartz M, Torruellas Garcia J, et al. The YscE/YscG chaperone and YscF N-terminal sequences target YscF to the Yersinia pestis type III secretion apparatus. Microbiology. 2018;164:338–348. doi: 10.1099/mic.0.000610. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Lee VT, Schneewind O. Yop fusions to tightly folded protein domains and their effects on Yersinia enterocolitica type III secretion. J Bacteriol. 2002;184:3740–3745. doi: 10.1128/JB.184.13.3740-3745.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Feldman MF, Müller S, Wüest E, Cornelis GR. SycE allows secretion of YopE-DHFR hybrids by the Yersinia enterocolitica type III Ysc system. Mol Microbiol. 2002;46:1183–1197. doi: 10.1046/j.1365-2958.2002.03241.x. [DOI] [PubMed] [Google Scholar]
  • 134.Radics J, Königsmaier 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]
  • 135.LeBlanc MA, Fink MR, Perkins TT, Sousa MC. Type III secretion system effector proteins are mechanically labile. Proc Natl Acad Sci. 2021;118:12. doi: 10.1073/pnas.2019566118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Zhang Y, Lara-Tejero M, Bewersdorf J, Galán JE. Visualization and characterization of individual type III protein secretion machines in live bacteria. Proc Natl Acad Sci U S A. 2017;114:6098–6103. doi: 10.1073/pnas.1705823114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Schlumberger MC, Müller AJ, Ehrbar K, Winnen B, Duss I, et al. 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]
  • 138.Portaliou AG, Tsolis KC, Loos MS, Balabanidou V, Rayo J, et al. Hierarchical protein targeting and secretion is controlled by an affinity switch in the type III secretion system of enteropathogenic Escherichia coli . EMBO J. 2017;36:3517–3531. doi: 10.15252/embj.201797515. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Gilzer D, Schreiner M, Niemann HH. Direct interaction of a chaperone-bound type III secretion substrate with the export gate. Nat Commun. 2022;13:2858. doi: 10.1038/s41467-022-30487-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Yu X-J, Grabe GJ, Liu M, Mota LJ, Holden DW. SsaV interacts with SsaL to control the translocon-to-effector switch in the Salmonella SPI-2 Type three secretion system. mBio. 2018;9:e01149-18. doi: 10.1128/mBio.01149-18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Shen D-K, Blocker AJ. MxiA, MxiC and IpaD regulate substrate selection and secretion mode in the T3SS of Shigella flexneri . PLoS One. 2016;11:e0155141. doi: 10.1371/journal.pone.0155141. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Gaytán MO, Monjarás Feria J, Soto E, Espinosa N, Benítez JM, et al. Novel insights into the mechanism of SepL-mediated control of effector secretion in enteropathogenic Escherichia coli . MicrobiologyOpen. 2018;7:e00571. doi: 10.1002/mbo3.571. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Kubori T, Galán 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]
  • 144.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]
  • 145.Pienkoß S, Javadi S, Chaoprasid P, Nolte T, Twittenhoff C, et al. The gatekeeper of Yersinia type III secretion is under RNA thermometer control. PLoS Pathog. 2021;17:e1009650. doi: 10.1371/journal.ppat.1009650. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Ngo T-D, Perdu C, Jneid B, Ragno M, Novion Ducassou J, et al. The PopN gate-keeper complex acts on the ATPase PscN to regulate the T3SS secretion switch from early to middle substrates in Pseudomonas aeruginosa . J Mol Biol. 2020;432:166690. doi: 10.1016/j.jmb.2020.10.024. [DOI] [PubMed] [Google Scholar]
  • 147.Erhardt M, Wheatley P, Kim EA, Hirano T, Zhang Y, et al. Mechanism of type-III protein secretion: Regulation of FlhA conformation by a functionally critical charged-residue cluster. Mol Microbiol. 2017;104:234–249. doi: 10.1111/mmi.13623. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Minamino T, Morimoto YV, Hara N, Namba K. An energy transduction mechanism used in bacterial flagellar type III protein export. Nat Commun. 2011;2:475. doi: 10.1038/ncomms1488. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.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]
  • 150.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] [PMC free article] [PubMed] [Google Scholar]
  • 151.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]
  • 152.Renault TT, Abraham AO, Bergmiller T, Paradis G, Rainville S, et al. Bacterial flagella grow through an injection-diffusion mechanism. Elife. 2017;6:e23136. doi: 10.7554/eLife.23136. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Jumper J, Evans R, Pritzel A, Green T, Figurnov M, et al. Highly accurate protein structure prediction with AlphaFold. Nature. 2021;596:583–589. doi: 10.1038/s41586-021-03819-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Bryant P, Pozzati G, Elofsson A. Improved prediction of protein-protein interactions using AlphaFold2. Nat Commun. 2022;13:1265. doi: 10.1038/s41467-022-28865-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Wagner S, Diepold A. A unified nomenclature for injectisome-type type III secretion systems. Curr Top Microbiol Immunol. 2020;427:1–10. doi: 10.1007/82_2020_210. [DOI] [PubMed] [Google Scholar]

Articles from Microbiology are provided here courtesy of Microbiology Society

RESOURCES