Next Article in Journal
Enzymatic Synthesis of 2-Chloropurine Arabinonucleosides with Chiral Amino Acid Amides at the C6 Position and an Evaluation of Antiproliferative Activity In Vitro
Next Article in Special Issue
Role of the ABCA4 Gene Expression in the Clearance of Toxic Vitamin A Derivatives in Human Hair Follicle Stem Cells and Keratinocytes
Previous Article in Journal
Exploiting the Integrated Valorization of Eucalyptus globulus Leaves: Chemical Composition and Biological Potential of the Lipophilic Fraction before and after Hydrodistillation
Previous Article in Special Issue
ABCB1 and ABCC1 Function during TGF-β-Induced Epithelial-Mesenchymal Transition: Relationship between Multidrug Resistance and Tumor Progression
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

ABC Transporters in Bacterial Nanomachineries

Institute of Biochemistry, Heinrich Heine University, Universitätsstr. 1, 40225 Düsseldorf, Germany
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(7), 6227; https://doi.org/10.3390/ijms24076227
Submission received: 24 February 2023 / Revised: 18 March 2023 / Accepted: 23 March 2023 / Published: 25 March 2023
(This article belongs to the Special Issue ABC Transporters: Where Are We 45 Years On?)

Abstract

:
Members of the superfamily of ABC transporters are found in all domains of life. Most of these primary active transporters act as isolated entities and export or import their substrates in an ATP-dependent manner across biological membranes. However, some ABC transporters are also part of larger protein complexes, so-called nanomachineries that catalyze the vectorial transport of their substrates. Here, we will focus on four bacterial examples of such nanomachineries: the Mac system providing drug resistance, the Lpt system catalyzing vectorial LPS transport, the Mla system responsible for phospholipid transport, and the Lol system, which is required for lipoprotein transport to the outer membrane of Gram-negative bacteria. For all four systems, we tried to summarize the existing data and provide a structure-function analysis highlighting the mechanistical aspect of the coupling of ATP hydrolysis to substrate translocation.

1. The Classical View of an ABC Exporter

ATP-Binding Cassette (ABC) transporters are present in all domains of life, from prokaryotes to eukaryotes [1]. ABC exporters consist of two nucleotide-binding domains (NDBs) and two transmembrane domains (TMDs) [2,3]. The NBDs are responsible for binding and hydrolyzing ATP, and the TMDs allow substrate translocation across the membrane [4]. The conformational changes of the NBDs, due to ATP binding and hydrolysis, are translated via coupling helices to the TMDs [5,6]. While the NBDs are highly conserved among all ABC transporters, the TMDs show a huge variety that goes in line with the huge variety of ABC transporter substrates [7]. Within this review, we will focus on prokaryotic ABC exporters functioning in nanomachineries. Some of the prokaryotic ABC exporters are “half-size” transporters with one NBD and one TMD fused together on a single polypeptide chain, while most harbor NBDs and TMDs as separate polypeptide chains allowing heterooligomeric assemblies (see Figure 1) [3].
Without any structural information, Jardetzky et al. proposed already in 1966 two configurations for a membrane transporter enabling a ‘two-sided access’ to a central cavity within the transporter [8]. Today we know from multiple protein structures that most of the ABC exporters change during substrate translocation from an inward-facing (IF) state to an outward-facing (OF) state [9,10,11,12,13]. This model of a translocation cycle (see Figure 2) was extended by the outward-occluded state observed for the peptide transporter McjD [14] and for other exporters as well [15]. New findings revealed even more transition states within one translocation cycle [16]. Besides this classical understanding of changing between IF and OF states, there are further models expanding the mechanisms for substrate translocation. On the one hand, there is the alternate access model, where a substrate is not entering from the cytoplasm but from a lateral opening in the TMDs (PCAT1 [17]). On the other hand, there is the outward-only model, where the substrate cavity is open and closing without an inward-facing state (PglK [18]). Similar to this outward-only model, simulations suggested a Constant Contact Model of the NBDs without a wide open IF state [19].
The power stroke for transmission from IF to OF state, and along with this, the substrate translocation, is supposed to be the binding of ATP. ATP hydrolysis, therefore, only resets the system back into the IF state [20,21]. Moreover, it is conceivable that there is a difference in the power stroke for heterodimeric ABC exporters that contain an intrinsically impaired nucleotide binding site, as it is proposed for BmrCD. Here, turnover from IF to OF state is exclusively attributed to ATP hydrolysis [22]. In addition, novel findings on the homodimeric ABC exporter MacB award at least a part of the power stroke for substrate translocation to the ATP hydrolysis step [23].
Based on the variety of TMDs, a new nomenclature was introduced to categorize ABC transporters on the basis of the different TMD folds into seven different types [24]. Type I–III cover classical ABC importers, and type IV to VII mostly cover ABC exporters. A large number of structures and functional information are known for the type IV ABC transporters that are defined by the architecture of Sav1866. Type V also includes a few ABC transporters with importer functions. Type VI ABC transporters are defined by the fold of LptB2FG and Type VII by the fold of MacB [24]. These classes of Type VI and VII are of special interest for this review and are complemented by the MlaFEDB system (recently assigned as founding member of type VIII [25]) and the LolCDE system (type VII). Although the classical understanding of the transport cycle from an IF state to an OF state with an occluded state trapping the substrate in a binding pocket is feasible for small substrates such as ions, amino acids, or even smaller peptides, this model cannot apply to large substrates such as lipopolysaccharide (LPS) or lipoproteins. In the following sections, we will summarize the recent findings with respect to the structure and function of these selected ABC transporter systems of MacB, LptB2FGC, MlaFEDB, and LolCDE (Figure 3) as a wide range of information is available for these nanomachineries.

2. The MacAB-TolC System

The first experimental evidence of an ABC antibiotic efflux transporter in Gram-negative organisms was published in 2001 by Kobayashi et al. [26]. They investigated a system that gave resistance to erythromycin and other macrolide-type antibiotics and was therefore termed macrolide-specific ABC-type efflux carrier, short “MacAB”. Furthermore, they showed that TolC is necessary for MacAB efflux activity [26]. Today it is known that the MacAB-TolC system also exports the extracellular peptide toxin STII [27], the heme precursor protoporphyrin IX [28], cyclic peptides like bacitracin and colistin [29], and penicillin-type antibiotics as well as arsenite [30]. Furthermore, overexpression of MacAB in Klebsiella pneumoniae increased resistance against the synthetic tetracycline-class antibiotic eravacycline [31]. All substrates of the MacAB-TolC system are transported from the periplasm across the outer membrane to the extracellular space [27,29]. Interestingly macB homologs can be found in Gram-positive bacteria as well, although they lack a periplasm and a second membrane (reviewed in Greene et al. [32]). The expression of macA and macB is controlled by the PhoP/PhoQ system [33] and therefore downregulated in case of low Mg2+ levels, which was shown using real-time quantitative polymerase chain reaction (rt-qPCR) [34].
The structure of the outer membrane protein TolC was determined in 2000 by Koronakis et al. [35]. Due to DNA sequence analysis, it was already proposed in 2001 without having a high-resolution structure that MacB is a transmembrane protein with four transmembrane helices (TMH) and that MacA is a peripheral membrane protein belonging to the membrane fusion protein (MFP) family [26]. Biochemical and biophysical data by Lin et al. revealed in 2009 that MacA stabilizes the tripartite assembly of the MacAB-TolC efflux system through specific interactions with MacB as well as TolC. In addition, it was shown that the N-terminal transmembrane helix of MacA, anchoring it into the inner membrane, is not essential for the functional assembly of the system [36]. In contrast, Tikhonova et al. could show that a MacA mutant lacking the N-terminal transmembrane helix was not able to confer an increase in erythromycin resistance in vivo, although in vitro studies showed that this mutant can still interact with MacB [37]. Furthermore, MacA increases MacB’s affinity for ATP and the substrate erythromycin. Mass spectrometry (MS) and atomic force microscopy (AFM) revealed a dimeric organization of MacB [36]. Bound lipids such as phosphatidylethanolamine (PE) and cardiolipin could only be observed for the dimeric protein using detailed MS analysis [38]. In 2009, the first crystal structure of the periplasmic domain of MacB [39], as well as the crystal structure of a hexameric arrangement of the periplasmic part of MacA from Actinobacillus actinomycetemcomitans [40], became available. Further analysis of the interface between MacA and TolC supported a strongly conserved tip-to-tip interaction between those two proteins [41,42,43]. In 2017, the single-particle cryo-EM structure of the fully assembled MacAB-TolC efflux system (see Figure 4) from Escherichia coli (E. coli) was published in a nucleotide- and substrate-free conformation by Fitzpatrick et al. [44]. This structural information on the MacAB-TolC system was complemented by an ADP-bound crystal structure of MacB from Acinetobacter baumannii published by Okada et al. [45] and an ATP-bound crystal structure of MacB from Aggregatibacter actinomycetemcomitans published by Crow et al. [29] in the same year (see Figure 5). Moreover, the crystal structure of a Gram-positive MacAB-like efflux pump from Streptococcus pneumoniae was published by Yang et al. in 2018 [46].
The structure of the full MacAB-TolC efflux system was derived by Fitzpatrick et al. [44] from a hybrid electron density map combining density maps from two different stabilizing approaches of the tripartite assembly. The first approach was a fusion of the C-terminus of MacB to the N-terminus of MacA as they were expected to be in close proximity. Interestingly, the fusion construct resolved only two copies of MacB, which were forming the dimer in the inner membrane. The other four copies of MacB, each fused to one MacA, were not resolved in the electron density map but produced diffuse density in the two-dimensional classifications. The second approach was a stabilization using disulfide bonds between MacA and MacB introduced via cysteine mutations. Only two of the MacA protomers formed a disulfide bond with MacB. Both approaches stabilized MacAB-TolC sufficiently for the acquisition of high-resolution structural data without loss of functionality in vivo [44]. This confirmed the overall organization of the MacAB-TolC system comprising a dimer of MacB in the inner membrane linked to a hexamer of MacA in the periplasm, which in turn is connected to a trimer of TolC in the outer membrane. This tripartite assembly with approximately 320 Å spans the inner and outer membrane in vivo [44].
Within the fully assembled complex, MacA adopts a hexameric structure, as seen for MacA alone [40], with four domains: a cylindrical α-helical hairpin domain, a ring-forming lipoyl domain, a ring-forming β-barrel domain, and a membrane-proximal domain. A density for the N-terminal transmembrane helix was missing, just as for the already existing crystal structure [40]. The already proposed tip-to-tip interaction between the trimeric TolC and MacA [41,42] was located between the α-helical hairpin region of MacA and the intra- and inter-protomer grooves of TolC [44]. A loop in the lipoyl domain of MacA allows highly conserved glutamine residues [40] to form an inter-protomer hydrogen-bond network that acts like a gating ring. Molecular dynamics simulations suggested that this gating ring acts as a one-way valve for the outward-directed transfer of the substrate [44]. The β-barrel domain and the membrane-proximal domain from three MacA protomers interact with the periplasmic domain of one MacB protomer [44].
MacB forms a homodimer in the tripartite assembly, with each protomer consisting of a NBD, a TMD, and a periplasmic domain (PD). The NBD is connected to the TMD via a long loop and an amphipathic helix. The TMD of MacB is unlike other ABC transporters, built up by only four TMH. TMH1 and TMH2 are elongated and reach above the membrane plane into the periplasm. In between these two helices, the PD is located. This PD contains a so-called porter domain that is a structural homolog to the AcrB “porter domain” and a sabre (small alpha beta-rich extracytoplasmic) domain. The porter domain is formed by two subunits that are located before and after the sabre domain. TMH3 and TMH4 are shorter than TMH1 and TMH2 and are connected via the so-called shoulder loop [29]. The major coupling helix is located in between TMH2 and TMH3; the minor coupling helix is located C-terminal of TMH4 [44]. The major and minor coupling helix interact with the NBD from the same protomer as there is no domain swapping [29]. The role of the minor coupling helix is not fully understood as, according to Crow et al. [29], the deletion of this helix did not influence MacB activity significantly, but for Okada et al. [45], deletion of the minor coupling helix resulted in the loss of drug export.
Although substrate-bound structures are lacking, the nucleotide-free MacAB-TolC cryo-EM structure [44], together with the ATP-bound crystal structure [29] and the ADP-bound crystal structure [45], allow prediction about the function of MacAB-TolC mediated efflux.
In the nucleotide-free state (see Figure 5A, left panel), the transmembrane dimer interface adopts a V-shape form having the periplasmic parts of TMH1 and TMH2 of one protomer far away from the other protomer and thereby forming a cavity on the periplasmic site margined by the periplasmic domains of MacB. This arrangement also brings the periplasmic domains into an open conformation forming a small gap towards the periplasm (see Figure 5B, left panel). In the cryo-EM structure, an additional electron density was observed in this periplasmic cavity between the periplasmic elongations of the TMHs. The orientation of this additional density already indicates a lateral substrate entrance. Nevertheless, it was not possible to examine the identity of this molecule. The NBDs are also separated and far away from each other in the nucleotide-free structure [44]. In the ATP-bound state (see Figure 5A, middle panel), the NBDs are dimerized and form a tightly packed classical head-to-tail arrangement. TMH1 and TMH2 of each protomer align parallel to each other in a rigid dimer interface, omitting the V-shaped form but adopting a so-called “zipped stalk” conformation. The PD adopts a closed form without an opening towards the periplasmic side (see Figure 5B, middle panel). The NBDs dimerization is supposed to mediate the “zipping” of the stalk and closure of the PD via the major coupling helix and movement of TMH2, which changes the dimer interface favoring close proximity of TMH1 and TMH2 of each protomer. This transfer of structural movements from one side of the membrane to the other via the transmembrane helices is referred to as mechanotransmission. The mechanotransmission mechanism was investigated in vivo by using cysteine mutants that lock MacB TMH1 and TMH2 in the zipped conformation. Those mutants showed decreased resistance to erythromycin [29]. In the ADP-bound state (see Figure 5A, right panel), the NBDs are dimerized but not as tight as in the ATP-bound state. TMH1 and TMH2 adopt a V-shape-like open conformation as in the nucleotide-free state, but the PDs arrange differently, and the opening towards the periplasm is absent (see Figure 5B, right panel) [45].
Biochemical data of Tikhonova et al. showed that MacB exhibits basal ATPase activity in detergent, which is little to none affected by the addition of either substrates or MacA or TolC [37]. When reconstituted into proteoliposomes, the basal ATPase activity of MacB is reduced about 10-fold but is strongly stimulated in the presence of MacA. MacA mutants missing either an N- or C-terminal part of MacA could not stimulate MacB ATPase activity in proteoliposomes in vitro and were also not able to increase erythromycin resistance in vivo, although the N-terminal truncated mutant was capable of binding MacB in vitro [37]. Modali et al. revealed that MacA stabilizes the ATP-bound state of MacB, and a single mutation in the membrane-proximal domain of MacA abolishes the macrolide efflux function of MacAB-TolC [47]. Lu et al. discovered that the periplasmic domain of MacB is essential for MacA-dependent stimulation of MacB ATPase activity. Furthermore, ATP binding of MacB increases affinity towards MacA [48]. In 2021, Souabni et al. determined a transport rate of three molecules per hydrolyzed ATP molecule for the substrate roxithromycin [23]. Recently, Batista dos Santos et al. [49] showed that the presence of TolC increases MacB ATPase activity in detergent as well as in lipid nanodiscs. Furthermore, they could also show a substrate-induced increase in MacB ATPase activity for the fully assembled MacAB-TolC system in a lipid environment [49].
Based on the available structures, Crow et al. [29] postulated a mechanism for substrate efflux of the MacAB-TolC system in 2017 termed the “molecular bellows” mechanism. In the nucleotide-free state, the cavity between the periplasmic part of TMH1 and TMH2 of MacB and MacA is open to the membrane, and substrates can enter. Upon ATP binding, the dimerization of the NBDs on the cytosolic side causes closure of the TMDs of MacB and constrains the volume of the cavity. This generates pressure and pushes the substrate through MacA towards TolC. The gating ring inside MacA prevents backflow of the substrate once the pressure is balanced. After ATP hydrolysis, the system switches back to the nucleotide-free open state [29]. This mechanism was further expanded by Souabni et al. [23] in 2021, investigating the role of ATP hydrolysis using a quantum dots-based real-time analysis of substrate transport and ATP turnover. They suggested a “modified bellows mechanism” where ATP hydrolysis is additionally energizing the transport of the substrate across the MacA gating ring towards TolC. In this modified mechanism, MacA and its transmembrane helix act as a mediator for the additional energetic input. In their study, substrate translocation and ATP hydrolysis are shown to be synchronous events [23].
Although much is known due to biochemical and structural data, many open questions remain regarding the function of the MacAB-TolC system. Examples are: How is the substrate recognized? Is there a specific substrate binding site? Is there a feedback mechanism from the periplasm to the NBDs? How is lateral substrate leakage prevented? These questions require further research using intermediate or substrate-bound structures.

3. The Lpt System

The cell envelope of Gram-negative bacteria exhibits a complex architecture. It not only consists of an inner membrane (IM), made up of phospholipids in the inner and outer leaflet, and an additional outer membrane (OM) but also a periplasm in between with a cell wall formed by crosslinked peptidoglycan. The OM protects the cell from the environment and serves as a potent barrier for hydrophobic molecules. Its composition is highly asymmetrical, with glycerophospholipids in the inner leaflet and mainly LPS molecules in the outer leaflet [50,51]. The interested reader is referred to comprehensive reviews focusing on the structure and synthesis of LPS [52,53,54], as it will be only briefly described here. The LPS structure can vary among different bacteria; typically, it can be divided into three parts: LPS is anchored to the membrane via its Lipid A moiety, a β-1′-6-linked glucosamine disaccharide, every sugar being acylated with fatty acid chains. Additionally, in E. coli, the glucosamine is phosphorylated at positions 1 and 4′ and can be further modified, e.g., upon polymyxin exposure with ethanolamine or 4-amino-4-deoxy-L-arabinose to decrease the negative net charge [55,56]. The core oligosaccharide moiety is connected to Lipid A via 3-deoxy-D-manno-oct-2-ulosonic acid (Kdo), which itself is linked with several heptose and hexose molecules. On top of the core oligosaccharide is the O antigen, also called O-antigenic polysaccharide (O-PS); it represents the terminal part of LPS and, as the name suggests, is a polymer made up of different kinds of oligosaccharides. The composition of the O antigen varies greatly, not only between different but also within the same species. E. coli, as one example, displays over one hundred different serotypes [52]. LPS is found in most, but not all, Gram-negative bacteria, with exceptions like Sphingomonas paucimobilis and Treponema pallidum [57,58]. In E. coli and Salmonella, the presence of LPS is essential, although this is not the case for all LPS-containing bacteria. Certain strains of Neisseria, Moraxella, and Acinetobacter can live without genes necessary for the synthesis or transport of LPS to the OM [59].
The LPS homeostasis displays a delicate challenge. Even though it is exclusively present in the outer leaflet of the OM, the biogenesis starts on the complete opposite side of the cell envelope: The inner leaflet of the IM (and the cytoplasm). Therefore, it does not only need to be extracted from the IM—an energy-consuming process—but the amphipathic LPS molecule also needs to traverse the aqueous periplasm on its route to the OM. Ultimately, the LPS needs to be incorporated into the outer leaflet of the OM. This process must occur repetitively during the life cycle of the cell in a highly ordered manner, as processes in the periplasm and outer membrane cannot be energized directly through ATP hydrolysis.
This part of the review will deal with the LPS transport (Lpt) machinery (see Figure 6), which mediates the extraction of LPS from the IM, its transport across the periplasm, and insertion into the OM. A special focus lies on the ABC transporter, including a step-by-step examination of its functionality and recent discoveries.
Intensive research on the LPS transport over more than five decades, starting with genetic approaches along with biochemical and structural studies, gradually revealed the Lpt system to be a multiprotein assembly with an ABC transporter LptB2FGC in the inner membrane [60] and a translocon LptDE in the outer membrane [61] which are connected by a periplasmic bridge built up by (most likely) oligomeric LptA [62] (see Figure 6). In this complex, LptB2FGC provides the energy for the extraction and transport of LPS by ATP binding and hydrolysis. Although its TMD fold is reminiscent to type V ABC transporters, LptB2GFC is considered to be the first member of the new type VI ABC transporter family, as there are several aspects that led researchers to classify this transporter into its distinct group [24,63].
In the domain organization of the LPS extractor, the two homologous proteins, LptF and LptG, form the TMD [64,65]. Both proteins have an additional periplasmic domain, the so-called β-jellyroll domain. This protein fold is a hallmark of the Lpt system, as it occurs in five of the seven Lpt proteins (LptA, LptC, LptD, LptF, and LptG; see Figure 6). The homo-dimeric LptB shows the fold of a canonical NBD and serves as the motor domain of the extractor. Its structure as an isolated protein was resolved by two independent groups in the same year, Sherman et al. and Wang et al. [66,67]. It is present as a dimer in the cytoplasm, where it binds and hydrolyzes ATP to energize the LPS translocation and forms the functional transporter with LptFG [60]. What makes the LPS transporter complex unique is the presence of a mysterious protein within the inner membrane complex: LptC. It is a small protein, only consisting of an N-terminal TMH and a periplasmic β-jellyroll domain [68]. Like all other Lpt proteins, it is essential for LPS transport, as deletion strains show phenotypes attributed to defective LPS transport, such as increased sensitivity to hydrophobic compounds [69].
LptD and LptE build the translocon in the outer membrane [61,70,71,72,73]. LptD is a large, 26-stranded β-barrel protein in which the lipoprotein LptE resides, forming a barrel-and-plug complex. The N-terminal part of LptD features a β-jellyroll domain as well and is connected to the jellyrolls of LptB2FGC via the periplasmic LptA, in which this protein fold was first discovered [74]. The N-terminal part of LptAs β-jellyroll domain interacts with LptC, while its C-terminal part showed crosslinks to the β-jellyroll domain of LptD. Studies also showed that LptA is capable of forming oligomeric structures in vivo and in vitro, although the exact number of LptA molecules necessary for establishing the periplasmic bridge between the complexes in the IM and OM remains unknown [74,75,76].
Early models based on initial structures of LptB2FG by Dong et al. and Luo et al., who only had the transporter in a single functional state available, proposed a sequence in which ATP binding would open the cavity, allowing LPS to enter and subsequent ATP hydrolysis and release of ADP would push LPS out of the transporter [65,77]. Today, additional structures of LptB2FG together with LptC, bound LPS, and/or bound nucleotide by the groups of Li, Luo, Tang, and Owens suggest the following model [78,79,80,81]: In the initial state of the transport cycle, the TMH of LptC (LptCTM) resides between LptF and LptG (see Figure 7, state i). LPS enters the cavity first (see Figure 7, state ii) and triggers in a not completely understood manner the release of LptCTM. Here, a strict NBD-TMD coupling is apparently involved (see Figure 7, state iii). By this, the cavity narrows, and its residues bind tightly to the LPS molecule and elevate it inside the cavity. The NBDs of LptB2 bind tightly to each other, which results in a complete collapse of the cavity and expulsion of LPS (see Figure 7, state iv). ATP, which can bind before ejection of LptCTM, as new data suggest, is hydrolyzed to ADP and released from the NBDs. This reopens the cavity, and LptCTM can bind again between TMH5 of LptF (THM5F) and TMH1 of LptG (TMH1G) to allow the next transport cycle (see Figure 7, state v) [82].
Together, LptF and LptG are adopting a V-shaped fold with an opening to the periplasm and only a few contacts between their interfaces, namely TMH1 and TMH5 (see Figure 7B and C). The limited interaction sites suggested early on that one, or both interfaces may open further to allow the lateral entry of LPS into the cavity [65,77]. The formed cavity is covered with residues of hydrophobic amino acids, which was confirmed by the structures of LptB2FG(C) with bound LPS. Additional charged residues stabilize bound LPS via salt bridges to the phosphate groups and glucosamine disaccharides of the lipid A moiety as well as the core oligosaccharide [78,79]. Even before the structure of LptB2FGC with LPS was revealed, Hamad et al. and Bertano et al. used mutational studies and bacterial strains, which constitutively modify lipid A phosphates to show that a cluster of charged residues in the TMH1 of LptG is important for LPS transport, most likely by binding LPS through establishing contact sites with the phosphate moieties of lipid A [83,84]. The first step in the LPS transport is the entry of the LPS molecule into this cavity for extraction, and this process already raises two questions: (i) From which side of the transporter does LPS enter? The structure of LptB2FGC allows LPS entry in principle from both LptFG interfaces, TMH1F:TMH5G or TMH5F:TMH1G. (ii) How does the transporter differentiate between LPS and other phospholipids, which are also present in the outer leaflet of the inner membrane?
The first question was answered by crosslinking studies using unnatural amino acids as photo-crosslinkers, together with structures of LptB2FGC, in which the periplasmic domains of LptF and LptG were resolved. Owens et al. detected crosslinks of LPS to residues in TMH1 of LptG and TMH5 of LptF as well as LptC, but not with the possible cavity opening formed by TMH1 of LptF and TMH5 of LptG [81]. Moreover, structures of the LPS ABC transporter with LptC revealed that the LptCTM is positioned between the interface of TMH5F:TMH1G, which is in line with the aforementioned crosslinking studies. So far, no data supported the proposal that LPS might enter from the TMH5G:TMH1F interface or LptC to reside in that position. Additionally, the structure of LptB2FGC showed that the β-jellyroll domains of the transporter are placed above the interface formed by TMH1F and TMH5G, possibly blocking the entry of LPS from this side due to the bulky core oligosaccharide and O-antigen moieties [81].
The second question, why phospholipids are not transported by the Lpt system, is not easily answered. The aforementioned study by Owens et al. showed that even in the absence of ATP, LPS is able to enter the cavity of the transporter [81]. One possibility is that phospholipids can enter the cavity through the TMD interface as well but are not recognized by the transporter as a substrate due to the lack of interactions mentioned above. Since the simplest LPS structure, enabling cell viability, contains lipid A and Kdo, it is likely, that the presence of these key components is necessary for substrate recognition by the Lpt transporter [52,85].
We know now that the TMH5F:TMH1G interface—in which LptCTM, at least at some point of the transport cycle, resides—is the entry for LPS. However, how does LPS enter the cavity of LptFG with a TMH of another protein in its way? The LptB2FG(C) structures of Li et al. revealed that LptCTM pushes away TMH1-3 of LptG, preventing the interaction of bound LPS with the charged residues mentioned above (see Figure 7B and C). The tighter binding of LPS might therefore push LptCTM out of the cavity [79]. The constriction of the cavity ultimately results in its full collapse and the transfer of LPS from the cavity to the β-jellyroll domains. This movement, as well as the reopening of the cavity, is performed by the NBDs LptB2, which bind and hydrolyze ATP. Closure and opening of the NBD interface are thought to be tied to the collapse and opening of the TMD cavity by a rigid body mechanism, as suggested for other transporters as well [5,86,87]. The LptB structure features a groove region in which the coupling helices of LptF and LptG are embedded [66]. These coupling helices are conserved in the TMD of ABC transporters and are crucial for the TMD-NBD interaction and can be found between TMH2 and TMH3 in LptFG [7,65,77]. They were identified using photo-crosslinking and mutational studies, in which defects induced by substituting a conserved glutamate in both LptF (E84) and LptG (E88) could be suppressed by altering an amino acid in the groove region of LptB (R91) [88]. Interestingly, the same study showed that LptF and LptG do not act symmetrically, as identical changes at equivalent positions led to different defects. Following the signature motif is the signature helix, in which Simpson et al. identified an arginine residue (R144) to be important for forming contacts with the Q-Loop and Walker B motif. Altering this residue led to reduced LPS extraction due to a lower ATP binding affinity, favoring the open conformation of the transporter. Strikingly, an alteration in the C-terminal domain (CTD) of LptB (F239), a domain unique to this transporter, is usually lethal but could complement the change in the signature helix. The data for this mutation showed a decreased ATP hydrolysis, indicating that this mutant is favoring the closed conformation, and the combination of these two defects complement each other, resulting in a functional transporter with decreased ATPase activity. Since full restoration of ATP hydrolysis in this double mutant is not necessary for LPS transport, the researchers proposed that the binding of ATP leads to the collapse of the LptFG cavity and hydrolysis to its reopening [89]. This correlation is supported by structures of the transporter in its closed form with bound β-γ-imidoadenosine 5′-triphosphate (AMP-PNP), a nonhydrolyzable ATP analog, and with bound ADP-vanadate [78,79]. ATP binding leads to the closure of the LptB2 dimer interface and the concomitant anti-clockwise rotation of the LptFG TMH1-5, ultimately closing its cavity [78].
Since this review focuses on the transporter of the Lpt system, subsequent transport steps will be only briefly described here. At the end of each transport cycle, a new LPS molecule is placed into the β-jellyroll domains of the transporter. The functionality of the transport system is often compared with a PEZ candy dispenser, where the LptB2FGC transporter acts like a spring, with each cycle loading a new LPS “candy” onto the bridge, pushing the former one further ahead on the LptA bridge towards LptDE and the OM [90]. Although both LptF and LptG feature β-jellyroll domains, structural, mutational, and crosslinking studies so far only showed that LPS travels from the cavity to the β-jellyroll domain of LptF, but not LptG, and further to the one of LptC [81]. The backflow into the cavity must be prevented if the entry of new LPS is not much faster since LPS does not diffuse away from the transporter. Owens et al. proposed on the basis of their LptB2FGC structure that the β-jellyroll domain of LptF can adopt a closed conformation, preventing a backward flow much like a valve [81]. Another possibility for the unidirectional flow might be different binding affinities of LPS for the different proteins, as this was at least shown for the transfer of LPS from LptC to LptA [68]. Even though LptA is known to form oligomers in vivo and in vitro, the exact number forming the bridge is unknown [62,74]. Sherman et al. succeeded with the in vitro reconstitution of the complete system, proving that LptA physically connects LptB2FGC with LptDE [91]. Once LPS reaches LptDE in the outer membrane, it is first placed into the β-jellyroll domain of LptD. The lipid A moiety of LPS is proposed to enter the membrane directly through a cavity between its β-jellyroll domain and the 26-strand β-barrel. The hydrophilic part of LPS would then first enter the β-barrel from the periplasmic side before exiting it through a lateral gate between helices 1 and 26, which showed to have only a few interacting residues [73,92,93]. LptE, the barrel’s plug, is not only important for the biogenesis and proper folding of LptD but was also shown to bind LPS and extract it from aggregates, suggesting that its role is to accept LPS coming from the periplasm, weaken neighboring LPS-LPS interactions and assisting its insertion into the OM [61,93,94,95,96,97]. Once the hydrophilic part of LPS has passed the lateral gate, it becomes part of the already existing LPS network. Intriguingly, there is evidence that the activity of the LPS transporter LptB2FGC in the inner membrane is influenced by the translocon LptDE. For an in vitro setup, inhibition of LPS transport and ATPase activity of LptB2FGC in liposomes was observed when adding LPS-preloaded LptDE-containing liposomes [98]. Lately, LptB2FGC has been shown to exhibit an adenylate kinase activity in addition to the ATPase activity, as it was reported for other ABC transporters like MsbA and CFTR [99,100,101].
Due to recent advances in structural biology, many answers regarding LPS transport could be answered. However, still some key questions remain, especially regarding the enigmatic LptC: What role does LptC play in the transport process since no other ABC transporter features a protein alike? Even though LptC is found in structures with LptB2FG, its TMH is neither required for LPS to enter the cavity nor for transport in general. Studies showed that LptC variants, in which the TMH was deleted, are still able to form a complex with the other Lpt proteins to transport LPS [102]. Experimental data suggest that the LptCTM plays a regulatory role in the ATPase activity of the transporter. In vitro, LptB2FG displays an increased ATP hydrolysis when no LptC or only variants without TM helix (LptC∆TM) are present. Therefore, the helix might reduce futile ATP hydrolysis by coupling it efficiently to substrate extraction [78,79,81]. Lethal LptC deletions can be suppressed when the arginine residue R212 in the β-jellyroll domain of LptF is substituted with glycine, restoring wildtype LPS transport and ATPase activity of the transporter, even though affinity to LptA is reduced [103,104]. Interestingly, this point mutation is still able to form a complex with LptC when it is present. This indicates that the periplasmic β-jellyroll domain of LptC is responsible for the interaction with the other Lpt components and enhances the stability of the transporter to the periplasmic bridge. The availability of LPS-bound LptB2FG structures with and without LptC allows a comparison of the cavity and interactions with the LPS molecule. With the LptCTM missing, the transporter cavity is significantly smaller (see Figure 7C). Simultaneously, the LPS structure in those structures is better resolved, indicating an improved binding of LPS once the LptCTM is removed from the transporter. When, how, and why the LptCTM is removed during the transport cycle is still not completely clear.
Recently, Wilson et al. approached these questions by combining LptC∆TM with mutations in other Lpt transporter components and searched for synergistic or suppressive effects on the phenotypes. They aimed to elucidate the exact role of LptCTM in the different steps of the extraction cycle, as deletion of the TM helix should strengthen or weaken the phenotype when a mutant is affected in the same step, while no change to the phenotype was expected when LptC∆TM was combined with mutants, in which the TM helix plays no role. Two observations were made: (i) The presence of the LptCTM increases the stability of the protein, possibly by benefiting from a complex formation with LptB2FG, and ii) the phenotype of LptB and LptF/G mutants with defects in ATP binding and NBD-TMD coupling were affected, indicating that the LptCTM plays a role in these steps. This new data hints that i) binding of ATP to LptB can occur with the LptCTM associated with LptFG (although it is not obligatory) and (ii) the coupling helices in LptFG, as well as the corresponding groove region in LptB, take part in LptCTM displacement from the transporters cavity, meaning that the sole entry of LPS into the cavity is not sufficient for this process [82]. The proposal that ATP can bind to LptB2FG with LptCTM still present is contrary to former models, in which LPS binding and formation of tighter contacts to LptFG is displacing LptCTM with ATP binding afterwards, particularly as structures of LptB2FG with bound LPS and displaced LptCTM were lacking nucleotides so far [79]. It is surprising that the essential process of LPS extraction, which takes place countless times during the lifetime of a cell, features an apparently useless step. One possible explanation is that the intercalation of LptCTM slows down ATP hydrolysis [78,79,81,87,99] to synchronize the hydrolysis of ATP with the binding of LPS to the cavity of the extractor, preventing futile ATP hydrolysis. It also remains enigmatic if the placement of LptCTM is a step that is not performed during active LPS transport but rather to slow it down before turning it off.

4. The Mla System

As mentioned in the previous section, the OM is an asymmetric bilayer, and all the components for building the OM envelope are synthesized either in the cytoplasm or the IM before being transported across the periplasm to be inserted into the OM [105]. As mentioned in the previous section, LPS is trafficked across the cell envelope via the Lpt system. The maintenance of lipid asymmetry (Mla) pathway is involved in maintaining the asymmetry of the OM by trafficking phospholipids (PLs) between the IM and OM [106,107,108]. The Mla pathway is a multi-component system and uses a ferry-like mechanism to shuttle phospholipids across the periplasm. Its mutation leads to the accumulation of PLs in the outer leaflet of the OM, increased OM permeability, and increased susceptibility to antibiotics [106,107,108,109]. Although translocation of phospholipids between outer and inner membrane was initially discovered in Salmonella typhimurium as early as 1977, it was not until 2009 that the components of the phospholipid transport system in Gram-negative bacteria were discovered via homology of conserved transporters in Actinobacteria and chloroplasts [106,110,111,112]. The orthologous TGD pathway of plants transports phosphatidic acid from the OM to the IM of chloroplasts [110], while the Mce4 pathway in Actinobacteria is paralogous and imports exogenous cholesterol [111,112].
In E. coli, the genes representing the Mla system are located on the mlaFEDCB operon, which is conserved among Gram-negative bacteria. Although mlaA and ompC/G are part of the Mla system, these genes are located outside the mla locus [106].
The Mla system is a six-component system with components present in each compartment of the cell envelope. It includes an OM lipoprotein termed MlaA; the OM major porins OmpF/OmpC, which act as a scaffold for MlaA; a soluble periplasmic component known as MlaC, which acts as a carrier of PLs, shuttling them between the membranes; a mammalian cell entry (MCE) domain protein called MlaD which is anchored in the plasma membrane; the transmembrane domain of the ABC transporter MlaE; a sulphate transporter and anti-sigma factor antagonist (STAS) domain protein called MlaB and the ATPase MlaF. These components create the following three main parts of the Mla system [113,114]: (a) the trimeric porin OmpC, which forms a complex with lipoprotein MlaA at the OM; (b) a soluble lipid-binding protein, MlaC, located in the periplasm; and (c) MlaFEDB, an ABC transporter localized in the plasma membrane (see Figure 8).
MlaA is a lipid transport protein and assembles into a ring-shaped α-helical structure that contains a central pore [115]. It forms a complex with both the OM porin proteins, OmpC and OmpF [115,116]. However, the complex of OmpC-MlaA is the active species, as MlaA copurifies with OmpC. MlaA binds in the groove between the two OmpF/C monomers, and the interaction between MlaA and OmpF/C is mainly mediated via van der Waals forces [115]. Since MlaA might be unstable on its own in the lipid bilayer, the porins might function as a scaffold to ensure the proper functioning of MlaA [115].
MlaC is a periplasmic lipid-binding protein. The crystal structure of MlaC was resolved and comprised four ß-sheets and seven helices with a large hydrophobic pocket in the core of the protein. Since MlaC can bind to both IM and OM complexes, it probably exhibits a central role in the transport of PLs between the membranes [117]. MlaC has a high affinity towards phospholipids and can bind three different PLs: phosphatidylglycerol, phosphatidylethanolamine, and cardiolipin [118,119].
MlaD is anchored to the IM through a single N-terminal TMH with its MCE domain residing in the periplasm. The MCE domains have been involved in lipid uptake in Gram-negative bacteria and retrograde transport of PLs in chloroplasts [110,117,120]. MlaD forms a ring-shaped homo-hexamer with a central hydrophobic pore that allows movement of PLs [114,117,118,119].
MlaE represents the transmembrane domain of the ABC transporter, which forms a homodimer. In contrast to the previously described TMDs of MacB and LptF/G (Section 3 and Section 4), MlaE has five TMHs. Recently, it has been assigned as a founding member of the type VIII group of ABC transporters [24,117]. Each subunit contains one elbow helix (EH), five transmembrane helices (TMH1–5), one coupling helix (CH), and one periplasmic helix (PH). TMH1, TMH2 and TMH5 of the two subunits of MlaE form a central hydrophobic cavity. EH runs parallel to the inner membrane plane while the CH connects TMH2 and TMH3 and is involved in interaction with MlaF. The PH is placed between TMH3 and TMH4 and interacts with MlaD [121].
MlaF represents the nucleotide-binding domain of the ABC transporter with a conserved structure representative of the ABC superfamily and is present as a dimer [121]. MlaB contains a STAS domain and represents the accessory protein, which is involved in cross-talk with the NBDs of ABC transporters. The two MlaB subunits are located on the opposite side of the MlaF dimer [122,123] and are involved in stabilizing the complex and ATP hydrolysis [109,124].
The inner membrane complex (IMC) of the Mla pathway is composed of MlaFEDB, which represents the ABC transporter. Overall, the complex comprises of four different proteins: MlaF, MlaE, MlaD, and MlaB in a stoichiometry of 2:2:6:2 [109,117,121,125]. The homodimers of MlaE and MlaF function as the TMDs and the NBDs, respectively. Associated with the homodimers are the unique auxiliary proteins: MlaB and MlaD. While there are two copies of MlaB, MlaD is present as a hexamer in the complex [109]. The two MlaB proteins in the cytoplasm do not contact each other, and each MlaB molecule interacts with one MlaF [121]. The MlaD hexamer rests on top of the periplasmic side of the MlaE dimer [117]. The six α-helices form a hollow hydrophobic channel to allow the transport of lipids. A total of six MlaD TMHs in the MlaFEDB complex are inserted into the membrane [121]. Each three MlaD subunits incorporate one MlaE where the interaction between the TMH of MlaD and the EH of MlaE is critical for the phospholipid transport function of MlaFEDB [121].
Recent cryo-EM structural work on the Mla system has provided further insights into this intriguing complex. The high-resolution structures from E. coli [114,121,126], Pseudomonas aeruginosa [127], and Acinetobacter baumannii [128,129] clearly indicate that the overall architecture of the whole complex is conserved throughout these species.
These structures show that in the absence of ATP, MlaE embraces a V-shaped open conformation to frame a cavity with the wider side confronting the hydrophobic channel of MlaD. In the case of E. coli, the lipid binding site has been shown to be the outward-open pocket of MlaE [114,126], while for A. baumannii, lipid binding is described to happen at the pore of MlaD and in between the pore loops of the MCE domain of the MlaD hexamer [128]. A few of the structures contain residual electron density in the cavity and in different locations and could represent PLs or bound detergent [114,121,126,127,128,129] (see Figure 9A). The direction of lipid transport by the Mla system is discussed as controversial. A retrograde transport mechanism was originally described for the Mla system, where it maintained the lipid asymmetry by removing mislocated phospholipids from the outer leaflet of the OM and importing them back to the plasma membrane [106,107,108,109,113,115,116,130]. However, there is also data that supports the export of phospholipids (anterograde transport) from the IM to the OM [119,125,126,131] or even a bi-directional transport between both membranes [114].
For simplicity, we only describe here the proposed retrograde transport mechanism in the context of phospholipid import (see Figure 9B). However, reversing the outlined steps could result in anterograde transport of phospholipids, as also described [119,125].
As part of the OmpF/C complex, MlaA is embedded inside the OM, and together they form a channel across the membrane [115]. Phospholipids are extracted from the outer leaflet of the OM into the channel via a lateral pathway. Interaction of MlaC to the MlaA-OmpC/OmpF complex results in phospholipid transfer to the hydrophobic pocket of MlaC [114,117,118]. Then, MlaC diffuses across the periplasm to deliver the lipids to the MlaFEDB complex in the IM. A hexameric ring formed by MlaD subunits creates a central hydrophobic tunnel for the transport of lipids [117]. MlaC acts as a chaperone and directly binds to MlaD [117], transferring the lipid into a continuous channel from MlaD to an outward-facing MlaE [114,121,127,128,129]. ATP binding induces a conformational change in MlaE, resulting in the collapse of the lipid-binding pocket, thereby facilitating the incorporation of the lipids into the IM [121]. The auxiliary protein MlaB is known to regulate the transport [109,124].
Although the recent advances in research have added a plethora of information about the Mla system, a major unresolved question is still the directionality of lipid transport. MCEs are usually involved in the retrograde transport of misplaced phospholipids; however, there are still crucial details missing regarding the transfer of PLs from MlaC to MlaD.
Another open question is the reason behind the formation of stable complexes of MlaA with both OmpF and OmpC. Interestingly only one of these assemblies is functional, although functionality towards the transport of PLs lies within MlaA. These questions need to be addressed in the future to better understand the system.

5. The Lol System

Lipoproteins are crucial elements in bacteria. They are either located in the outer leaflet of the cytoplasmic membrane or in the leaflets of the OM. The latter case is most often true for Gram-negative bacteria, which is also the focus of this section. Lipoproteins are a compelling object of study for many reasons. One of them is their strong involvement in building and maintaining the OM of Gram-negative bacteria. As the OM is the first line of defense against xenobiotics, lipoproteins, their synthesis, and their transport pathway are attractive targets for novel antibiotics [132,133]. This section focuses specifically on the ABC transporter, which is involved in the transport of lipoproteins, and shows a transport mechanism differing from the classical ABC transporter mechanism. For a better overview regarding the synthesis, sorting, and function of lipoproteins, the interested reader is referred to other reviews [132,133], as these aspects will only be briefly mentioned here.
Lipoproteins are synthesized in the cytoplasm together with a cleavable signal peptide, which contains a consensus sequence that is highly conserved among lipoproteins, called the lipobox. The consensus sequence is L-A/S-G/A-C and was confirmed by various sequence analyses [134,135]. After translocation to the IM via the Sec [132,136] or, in some cases, via the Tat machinery [132,137], lipoproteins undergo further maturation steps in the outer leaflet of the IM. In the first step, the enzyme Lgt attaches a diacyl moiety to the cysteine in the aforementioned consensus sequence [138]. Subsequently, the enzyme Lsp [139,140] cleaves the signal sequence from the cysteine, and therefore this residue is named Cys+1 and becomes the new N-Terminus of the lipoprotein. In the last maturation step, another acyl group is attached to the free amino group of the cysteine by Lnt [141,142]. The sorting of lipoproteins either to the IM or the OM is determined in E. coli by the amino acid next to the N-terminal cysteine of mature lipoproteins (the +2 position). If this amino acid is aspartate, the lipoprotein is retained in the IM [143]. In the case of most other naturally occurring residues, the lipoprotein is localized to the OM [144].
The trafficking of lipoproteins to the OM of Gram-negative bacteria is mediated by the localization of lipoproteins (Lol) system [133]. The discovery of the Lol system started with the discovery of the periplasmic chaperone LolA (called p20 at that time), which was found to form a soluble complex with a lipoprotein [145]. Soon, the interaction partner of LolA, namely LolB, was also discovered [146]. LolB was localized to the OM and was identified as an essential protein for E. coli, as depletion of LolB is lethal. It was further shown that the incubation of a lipoprotein-LolA complex with a soluble LolB derivative led to the transfer of the lipoprotein to the OM. Through these findings, an initial mechanism for lipoprotein localization was already emerging. Soon after that, the crystal structures of LolA and LolB were solved [147]. Remarkably, both proteins showed a similar overall structure, despite their different amino acid sequences. Both proteins show a beta-barrel structure with a hydrophobic inside and alpha-helical lid. The hydrophobic cavities were identified as the possible binding sites for the acyl chains of the transported lipoproteins. This was further supported by finding polyethylene glycol 2000 monomethyl ether (PEGMME2000), which was used for the crystallization, in the hydrophobic cavity in one of the LolB structures [147].
It was later discovered that the detachment of lipoproteins from the membrane is ATP-dependent, and the corresponding protein is an ABC transporter [148]. Yakushi et al. solved the stoichiometry of this transporter termed LolCDE. It is a tetrameric and asymmetric ABC transporter composed of the proteins LolC and LolE, which make up the TMDs, and two copies of LolD, which form the NBDs. It was predicted that both proteins, LolC and LolE, each possess four transmembrane helices and a large periplasmic domain [148]. Due to the topology of LolCDE, the transporter can be assigned to the group of type VII ABC transporters [24]. Taking together all main findings, a general transport cycle, as shown in Figure 10, was derived.
One major unknown factor was certainly the ABC transporter LolCDE. Its mechanism has to surely differ from the mechanism of classical ABC transporter, as the substrate is located in the outer leaflet of the inner membrane and not in the cytoplasm. However, for a relatively long time after the discovery of the Lol pathway, the ABC transporter LolCDE lacked structural analysis, and thus, a detailed understanding of its molecular mechanisms was missing. The first published structure related to LolCDE comprised a crystal structure of a soluble periplasmic domain of LolC solved by the Koronakis lab [29]. This structural analysis was done to confirm the structural similarity of the periplasmic domain of LolC to that of the homologous ABC transporter MacB, which is involved inter alia in the efflux of antibiotic macrolides (described in Section 2). The confirmation of the similarity of the two periplasmic domains of LolC and MacB led to the conclusion that LolC could also follow the mechanotransmission mechanism. This characteristic periplasmic domain is divided into two subdomains named sabre and porter. The superposition of the sabre domains of LolC and MacB showed a prominent loop in LolC which is not present in MacB. Accordingly, this loop was investigated further by the Koronakis lab, and they solved the crystal structure of the periplasmic domain of LolC in complex with the periplasmic chaperone LolA [149]. This structure sheds light on the molecular details of the interactions within the complex and highlights the importance of this loop, which was termed “hook”. Furthermore, an additional important interacting domain was found in LolC, which was termed “pad”. The hook and the pad were determined to be essential for the recruitment of LolA. These findings further underpin the assignment of LolCDE into the group of type VII ABC transporters, as MacB is the founding member of this group [24]. In a more recent study, the crystal structure of LolA bound to a ligand was also solved by the Koronakis lab [150]. The structure shows the precise interaction of the acyl chains of the lipoprotein with the hydrophobic cavity in LolA. In addition, an overlap between the acyl binding sites and the LolC binding sites of LolA were found by comparison with other LolC-LolA structures. This indicates that the substrate binding to LolA is inducing the detachment from LolC. These findings further complete our understanding of the Lol system.
Since 2021, major breakthroughs have been made regarding the structural characterization of LolCDE. Tang et al. started this by solving six structures of the transporter representing different states of the transport via cryo-EM [151]. The structures represent LolCDE solubilized in the detergent LMNG. In the apo state, the NBDs show a relaxed conformation, whereas the rest of the transporter shows a rather compact conformation. More precisely, the transmembrane (TM) segments of LolC twist around the TM segments of LolE, and the PD of LolC rotates to the front of the PD of LolE. In the substrate-bound conformation, the TMDs of LolCDE are outward-opened, leading to a V-shaped central channel. In addition, two lateral gates are observable, which are formed by TMH1 and TMH2 of each of the proteins, LolC and LolE. The triacyl chains and a small N-terminal segment of the lipoprotein were resolved in a vertical arrangement in the upper part of the V-shaped channel. This suggests that the lipoprotein was extracted laterally from the periplasmic side of the IM, as this is also the arrangement of the lipoprotein in the IM. Furthermore, two AMP-PNP-bound LolCDE structures were solved by Tang et al. [151]. Only one of these structures contains a bound lipoprotein. Since AMP-PNP is a non-hydrolyzable ATP analog, these structures represent LolCDE after ATP binding. In the lipoprotein-bound structure, the transporter maintains its V-shaped cavity, and the NBDs are open. In the structure without the lipoprotein, the NBDs are closed. The closure of the NBDs leads to conformational changes in the TMDs, which leads to the closure of the central channel. In particular, TMH1 and TMH2 of LolE are shifted towards LolC. Additionally, the TMH2 of LolE clashes with the triacylcysteine of the lipoprotein of the ligand-bound state, which leads to the extrusion of the lipoprotein.
Sharma and colleagues published cryo-EM structures of LolCDE in a nucleotide-free and a nucleotide-bound state reconstituted in nanodiscs, which represent a native-like lipid environment [152]. In summary, the two determined structures show a very high similarity to the structural counterparts determined by Tang et al. [151]. The nucleotide-bound state was determined using an ADP-vanadate complex trapped in the ATP binding site. In addition, in the vanadate-trapped structure, the two LolD proteins come into close contact, which leads to the uplifting of the TMH2. Thus, ATP-binding leads to the extrusion of the lipoproteins from the TMDs to LolA. Sharma et al. further points out that in this conformation, TMH2 closes the lateral opening between LolC and LolE [152]. This could serve as a further mechanism for preventing substrate entry before the completion of a full transport cycle. In comparison to the structures of Tang et al. [151], the nucleotide-free structure of Sharma et al. [152] shows a significant difference in the binding of the lipoproteins. In the structure of Sharma et al. [152], the N-terminally linked acyl chain of the lipoprotein adopts a rather horizontal conformation in the cavity of LolCDE compared to the rather vertical arrangement of this acyl chain in the structure of Tang et al. [151].
In a relatively recent study, Bei et al. presented cryo-EM structures of nanodisc-reconstituted LolCDE in the apo, lipoprotein-bound, and AMP-PNP-bound states [153]. In general, the structures are highly similar to the corresponding structures published by Tang et al. [151] and Sharma et al. [152]. However, Bei et al. [153] pointed out a remarkable difference in the distances of the periplasmatic domains of LolC and LolE in the apo state compared to the apo state determined by Tang et al. [151]. Nevertheless, the structures of Bei et al. [153] also suggested that the binding of ATP leads to the closure of the central cavity and to the extrusion of the lipoprotein. The three structures determined by Bei et al. [153] are depicted exemplarily for the LolCDE transport mechanism in Figure 11. Panel B shows how the movement of TMH2 of LolE leads to the blockage of the central cavity and, thus, to the extrusion of the lipoprotein.
Putting it all together, the published LolCDE structures in different transport states show a high similarity and correspond well with each other, especially regarding the connection of ATP-binding and release of the substrate to the periplasmic chaperone LolA [152,153]. All studies suggested a similar transport mechanism. In the apo state, the lipoprotein is extracted from the outer leaflet of the IM in an energy-independent manner into the V-shaped cavity inside LolCDE. Through ATP-binding, the NBDs dimerize, which leads to movements in the TMDs, resulting in the shuffling of the lipoprotein to LolA. After the dissociation of ADP and the lipoprotein-LolA complex, the central cavity opens again, and the transporter is primed for the next transport cycle. Despite the vast similarities, slight differences exist between specific structures. There are also differences between the ATPase activities of the different structures. For instance, LolCDE prepared in nanodiscs by Sharma and colleagues shows a several-fold higher activity than LolCDE purified in LMNG by Tang et al. [151,152]. The differences could be attributed to the different detergent and lipid environments in which the structures were solved. This further emphasizes the significance and involvement of lipids in the function of membrane transporters and membrane proteins in general. The ABC transporter LolCDE falls in line with a group of emerging ABC transporters, which show a non-classical transport mechanism. The initial suggestion that LolCDE also follows a mechanotransmission mechanism due to the resemblance to MacB [29] was confirmed by the various discussed LolCDE structures. The study of such ABC transporters with non-conventional transport mechanisms enhances our understanding of living systems and could pave the way for novel drug targets and biotechnological applications.

6. Conclusions

Here, we have reviewed four ABC transporters that form the energizing component of four bacterial nanomachineries. We also tried to highlight how the use of ATP deviates from the classic “two-side access” [8] and how these machineries adapted to the particular needs of transporting quite different substrates. Although we have witnessed a tremendous increase in knowledge about the function and structure of the nanomachineries, we have also summarized the obvious questions.
Despite the availability of structural information on the whole MacAB-TolC nanomachinery, detailed insights about substrate binding and recognition are still lacking, including intermediate or substrate-bound structures. Especially the role of ATP hydrolysis in substrate translocation and possible feedback from the putative binding site to the NBDs require further investigations.
Similarly, there are still some intriguing features of the Lpt system, such as the role of LptC during the inactive state or the influence of LptC on the binding of LptB2FG to ATP and substrate, that need to be studied in detail towards an in-depth understanding of the Lpt system. Additionally, the oligomeric organization of LptA forming the periplasmic bridge is still unknown.
For the Mla system, the fundamental question regarding the directionality of the transport remains open. The influence of the cell wall on the movement of MlaC is also unknown. Furthermore, the role of a non-functional OMP-MlaA assembly remains elusive.
In terms of the Lol system, the variety of structures is impressive though a detailed functional understanding of the shuttling process between IM and OM remains unclear. This emphasizes the necessity of a combination of all kinds of structural, biochemical, and biophysical data to understand these complex systems in detail.
These questions need to be addressed in the future to fully understand and maybe even exploit the beautiful variety of these different nanomachineries in prokaryotes.

Author Contributions

All authors wrote, drafted, and revised the review and searched for literature. L.S. was responsible for the ideation. All authors have read and agreed to the published version of the manuscript.

Funding

Work on ABC transporters is funded by the DFG (CRC 1208 project A01 and Schm1279/17-1 to L.S.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

We thank all current and former members of the Institute of Biochemistry for their support and many stimulating discussions. Apologies to all our colleagues whose work was not referenced. We thank the Center for Structural Sciences (CSS) at Heinrich Heine University for excellent support and infrastructure, which enables structural studies of our ABC systems.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Higgins, C.F. ABC transporters: Physiology, structure and mechanism--an overview. Res. Microbiol. 2001, 152, 205–210. [Google Scholar] [CrossRef] [PubMed]
  2. Davidson, A.L.; Dassa, E.; Orelle, C.; Chen, J. Structure, function, and evolution of bacterial ATP-binding cassette systems. Microbiol. Mol. Biol. Rev. 2008, 72, 317–364. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Biemans-Oldehinkel, E.; Doeven, M.K.; Poolman, B. ABC transporter architecture and regulatory roles of accessory domains. FEBS Lett. 2006, 580, 1023–1035. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Thomas, C.; Tampe, R. Structural and Mechanistic Principles of ABC Transporters. Annu. Rev. Biochem. 2020, 89, 605–636. [Google Scholar] [CrossRef] [PubMed]
  5. Oldham, M.L.; Davidson, A.L.; Chen, J. Structural insights into ABC transporter mechanism. Curr. Opin. Struct. Biol. 2008, 18, 726–733. [Google Scholar] [CrossRef] [Green Version]
  6. Higgins, C.F.; Linton, K.J. The ATP switch model for ABC transporters. Nat. Struct. Mol. Biol. 2004, 11, 918–926. [Google Scholar] [CrossRef]
  7. Hollenstein, K.; Dawson, R.J.; Locher, K.P. Structure and mechanism of ABC transporter proteins. Curr. Opin. Struct. Biol. 2007, 17, 412–418. [Google Scholar] [CrossRef]
  8. Jardetzky, O. Simple allosteric model for membrane pumps. Nature 1966, 211, 969–970. [Google Scholar] [CrossRef]
  9. Dawson, R.J.; Locher, K.P. Structure of a bacterial multidrug ABC transporter. Nature 2006, 443, 180–185. [Google Scholar] [CrossRef]
  10. Bonifer, C.; Glaubitz, C. MsbA: An ABC transporter paradigm. Biochem. Soc. Trans. 2021, 49, 2917–2927. [Google Scholar] [CrossRef]
  11. Liu, F.; Zhang, Z.; Csanády, L.; Gadsby, D.C.; Chen, J. Molecular Structure of the Human CFTR Ion Channel. Cell 2017, 169, 85–95.e8. [Google Scholar] [CrossRef] [Green Version]
  12. Chen, S.; Oldham, M.L.; Davidson, A.L.; Chen, J. Carbon catabolite repression of the maltose transporter revealed by X-ray crystallography. Nature 2013, 499, 364–368. [Google Scholar] [CrossRef] [Green Version]
  13. Aller, S.G.; Yu, J.; Ward, A.; Weng, Y.; Chittaboina, S.; Zhuo, R.; Harrell, P.M.; Trinh, Y.T.; Zhang, Q.; Urbatsch, I.L.; et al. Structure of P-glycoprotein reveals a molecular basis for poly-specific drug binding. Science 2009, 323, 1718–1722. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Choudhury, H.G.; Tong, Z.; Mathavan, I.; Li, Y.; Iwata, S.; Zirah, S.; Rebuffat, S.; Van Veen, H.W.; Beis, K. Structure of an antibacterial peptide ATP-binding cassette transporter in a novel outward occluded state. Proc. Natl. Acad. Sci. USA 2014, 111, 9145–9150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Kopcho, N.; Chang Gand Komives, E.A. Dynamics of ABC Transporter P-glycoprotein in Three Conformational States. Sci. Rep. 2019, 9, 15092. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Hofmann, S.; Januliene, D.; Mehdipour, A.R.; Thomas, C.; Stefan, E.; Brüchert, S.; Kuhn, B.T.; Geertsma, E.R.; Hummer, G.; Tampé, R.; et al. Conformation space of a heterodimeric ABC exporter under turnover conditions. Nature 2019, 571, 580–583. [Google Scholar] [CrossRef]
  17. Lin, D.Y.; Huang, S.; Chen, J. Crystal structures of a polypeptide processing and secretion transporter. Nature 2015, 523, 425–430. [Google Scholar] [CrossRef]
  18. Perez, C.; Mehdipour, A.R.; Hummer, G.; Locher, K.P. Structure of Outward-Facing PglK and Molecular Dynamics of Lipid-Linked Oligosaccharide Recognition and Translocation. Structure 2019, 27, 669–678.e5. [Google Scholar] [CrossRef] [Green Version]
  19. George, A.M.; Jones, P.M. An asymmetric post-hydrolysis state of the ABC transporter ATPase dimer. PLoS ONE 2013, 8, e59854. [Google Scholar] [CrossRef] [Green Version]
  20. Prieß, M.; Göddeke, H.; Groenhof, G.; Schäfer, L.V. Molecular Mechanism of ATP Hydrolysis in an ABC Transporter. ACS Cent. Sci. 2018, 4, 1334–1343. [Google Scholar] [CrossRef] [Green Version]
  21. Stefan, E.; Hofmann, S.; Tampé, R. A single power stroke by ATP binding drives substrate translocation in a heterodimeric ABC Transporter. Elife 2020, 9, e55943. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Mishra, S.; Verhalen, B.; Stein, R.A.; Wen, P.C.; Tajkhorshid, E.; Mchaourab, H.S. Conformational dynamics of the nucleotide binding domains and the power stroke of a heterodimeric ABC transporter. Elife 2014, 3, e02740. [Google Scholar] [CrossRef] [PubMed]
  23. Souabni, H.; Batista dos Santos, W.; Cece, Q.; Catoire, L.J.; Puvanendran, D.; Bavro, V.N.; Picard, M. Quantitative real-time analysis of the efflux by the MacAB-TolC tripartite efflux pump clarifies the role of ATP hydrolysis within mechanotransmission mechanism. Commun. Biol. 2021, 4, 493. [Google Scholar] [CrossRef] [PubMed]
  24. Thomas, C.; Aller, S.G.; Beis, K.; Carpenter, E.P.; Chang, G.; Chen, L.; Dassa, E.; Dean, M.; Duong Van Hoa, F.; Ekiert, D.; et al. Structural and functional diversity calls for a new classification of ABC transporters. FEBS Lett. 2020, 594, 3767–3775. [Google Scholar] [CrossRef]
  25. Ekiert, D.C.; Coudray, N.; Bhabha, G. Structure and mechanism of the bacterial lipid ABC transporter, MlaFEDB. Curr. Opin. Struct. Biol. 2022, 76, 102429. [Google Scholar] [CrossRef] [PubMed]
  26. Kobayashi, N.; Nishino, K.; Yamaguchi, A. Novel macrolide-specific ABC-type efflux transporter in Escherichia coli. J. Bacteriol. 2001, 183, 5639–5644. [Google Scholar] [CrossRef] [Green Version]
  27. Yamanaka, H.; Kobayashi, H.; Takahashi, E.; Okamoto, K. MacAB is involved in the secretion of Escherichia coli heat-stable enterotoxin II. J. Bacteriol. 2008, 190, 7693–7698. [Google Scholar] [CrossRef] [Green Version]
  28. Turlin, E.; Heuck, G.; Simões Brandão, M.I.; Szili, N.; Mellin, J.R.; Lange, N.; Wandersman, C. Protoporphyrin (PPIX) efflux by the MacAB-TolC pump in Escherichia coli. Microbiologyopen 2014, 3, 849–859. [Google Scholar] [CrossRef]
  29. Crow, A.; Greene, N.P.; Kaplan, E.; Koronakis, V. Structure and mechanotransmission mechanism of the MacB ABC transporter superfamily. Proc. Natl. Acad. Sci. USA 2017, 114, 12572–12577. [Google Scholar] [CrossRef] [Green Version]
  30. Shi, K.; Cao, M.; Li, C.; Huang, J.; Zheng, S.; Wang, G. Efflux proteins MacAB confer resistance to arsenite and penicillin/macrolide-type antibiotics in Agrobacterium tumefaciens 5A. World J. Microbiol. Biotechnol. 2019, 35, 115. [Google Scholar] [CrossRef]
  31. Zheng, J.X.; Lin, Z.W.; Sun, X.; Lin, W.H.; Chen, Z.; Wu, Y.; Qi, G.B.; Deng, Q.W.; Qu, D.; Yu, Z.J. Overexpression of OqxAB and MacAB efflux pumps contributes to eravacycline resistance and heteroresistance in clinical isolates of Klebsiella pneumoniae. Emerg. Microbes. Infect. 2018, 7, 139. [Google Scholar] [CrossRef] [Green Version]
  32. Greene, N.P.; Kaplan, E.; Crow, A.; Koronakis, V. Antibiotic Resistance Mediated by the MacB ABC Transporter Family: A Structural and Functional Perspective. Front. Microbiol. 2018, 9, 950. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Groisman, E.A. The pleiotropic two-component regulatory system PhoP-PhoQ. J. Bacteriol. 2001, 183, 1835–1842. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Nishino, K.; Latifi, T.; Groisman, E.A. Virulence and drug resistance roles of multidrug efflux systems of Salmonella enterica serovar Typhimurium. Mol. Microbiol. 2006, 59, 126–141. [Google Scholar] [CrossRef] [PubMed]
  35. Koronakis, V.; Sharff, A.; Koronakis, E.; Luisi, B.; Hughes, C. Crystal structure of the bacterial membrane protein TolC central to multidrug efflux and protein export. Nature 2000, 405, 914–919. [Google Scholar] [CrossRef]
  36. Lin, H.T.; Bavro, V.N.; Barrera, N.P.; Frankish, H.M.; Velamakanni, S.; van Veen, H.W.; Robinson, C.V.; Borges-Walmsley, M.I.; Walmsley, A.R. MacB ABC transporter is a dimer whose ATPase activity and macrolide-binding capacity are regulated by the membrane fusion protein MacA. J. Biol. Chem. 2009, 284, 1145–1154. [Google Scholar] [CrossRef] [Green Version]
  37. Tikhonova, E.B.; Devroy, V.K.; Lau, S.Y.; Zgurskaya, H.I. Reconstitution of the Escherichia coli macrolide transporter: The periplasmic membrane fusion protein MacA stimulates the ATPase activity of MacB. Mol. Microbiol. 2007, 63, 895–910. [Google Scholar] [CrossRef]
  38. Barrera, N.P.; Isaacson, S.C.; Zhou, M.; Bavro, V.N.; Welch, A.; Schaedler, T.A.; Seeger, M.A.; Miguel, R.N.; Korkhov, V.M.; Van Veen, H.W.; et al. Mass spectrometry of membrane transporters reveals subunit stoichiometry and interactions. Nat. Methods 2009, 6, 585–587. [Google Scholar] [CrossRef] [Green Version]
  39. Xu, Y.; Sim, S.H.; Nam, K.H.; Jin, X.L.; Kim, H.M.; Hwang, K.Y.; Lee, K.; Ha, N.C. Crystal structure of the periplasmic region of MacB, a noncanonic ABC transporter. Biochemistry 2009, 48, 5218–5225. [Google Scholar] [CrossRef]
  40. Yum, S.; Xu, Y.; Piao, S.; Sim, S.H.; Kim, H.M.; Jo, W.S.; Kim, K.J.; Kweon, H.S.; Jeong, M.H.; Jeon, H.; et al. Crystal structure of the periplasmic component of a tripartite macrolide-specific efflux pump. J. Mol. Biol. 2009, 387, 1286–1297. [Google Scholar] [CrossRef]
  41. Xu, Y.; Sim, S.H.; Song, S.; Piao, S.; Kim, H.M.; Jin, X.L.; Lee, K.; Ha, N.C. The tip region of the MacA alpha-hairpin is important for the binding to TolC to the Escherichia coli MacAB-TolC pump. Biochem. Biophys Res. Commun. 2010, 394, 962–965. [Google Scholar] [CrossRef] [PubMed]
  42. Xu, Y.; Song, S.; Moeller, A.; Kim, N.; Piao, S.; Sim, S.H.; Kang, M.; Yu, W.; Cho, H.S.; Chang, I.; et al. Functional implications of an intermeshing cogwheel-like interaction between TolC and MacA in the action of macrolide-specific efflux pump MacAB-TolC. J. Biol. Chem. 2011, 286, 13541–13549. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Lee, M.; Kim, H.L.; Song, S.; Joo, M.; Lee, S.; Kim, D.; Hahn, Y.; Ha, N.C.; Lee, K. The alpha-barrel tip region of Escherichia coli TolC homologs of Vibrio vulnificus interacts with the MacA protein to form the functional macrolide-specific efflux pump MacAB-TolC. J. Microbiol. 2013, 51, 154–159. [Google Scholar] [CrossRef] [PubMed]
  44. Fitzpatrick, A.W.; Llabrés, S.; Neuberger, A.; Blaza, J.N.; Bai, X.C.; Okada, U.; Murakami, S.; Van Veen, H.W.; Zachariae, U.; Scheres, S.H.; et al. Structure of the MacAB-TolC ABC-type tripartite multidrug efflux pump. Nat. Microbiol. 2017, 2, 17070. [Google Scholar] [CrossRef] [Green Version]
  45. Okada, U.; Yamashita, E.; Neuberger, A.; Morimoto, M.; van Veen, H.W.; Murakami, S. Crystal structure of tripartite-type ABC transporter MacB from Acinetobacter baumannii. Nat. Commun. 2017, 8, 1336. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Yang, H.B.; Hou, W.T.; Cheng, M.T.; Jiang, Y.L.; Chen, Y.; Zhou, C.Z. Structure of a MacAB-like efflux pump from Streptococcus pneumoniae. Nat. Commun. 2018, 9, 196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Modali, S.D.; Zgurskaya, H.I. The periplasmic membrane proximal domain of MacA acts as a switch in stimulation of ATP hydrolysis by MacB transporter. Mol. Microbiol. 2011, 81, 937–951. [Google Scholar] [CrossRef] [Green Version]
  48. Lu, S.; Zgurskaya, H.I. Role of ATP binding and hydrolysis in assembly of MacAB-TolC macrolide transporter. Mol. Microbiol. 2012, 86, 1132–1143. [Google Scholar] [CrossRef] [Green Version]
  49. Batista Dos Santos, W.; Souabni, H.; Picard, M. Corseting a tripartite ABC transporter to make it fit for transport. Biochimie 2022, 205, 117–123. [Google Scholar] [CrossRef]
  50. Mühlradt, P.F.; Golecki, J.R. Asymmetrical distribution and artifactual reorientation of lipopolysaccharide in the outer membrane bilayer of Salmonella typhimurium. Eur. J. Biochem. 1975, 51, 343–352. [Google Scholar] [CrossRef]
  51. Beveridge, T.J. Structures of gram-negative cell walls and their derived membrane vesicles. J. Bacteriol. 1999, 181, 4725–4733. [Google Scholar] [CrossRef] [Green Version]
  52. Raetz, C.R.; Whitfield, C. Lipopolysaccharide endotoxins. Annu. Rev. Biochem. 2002, 71, 635. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Whitfield, C.; Trent, M.S. Biosynthesis and export of bacterial lipopolysaccharides. Annu. Rev. Biochem. 2014, 83, 99–128. [Google Scholar] [CrossRef] [PubMed]
  54. Bertani, B.; Ruiz, N. Function and biogenesis of lipopolysaccharides. EcoSal Plus 2018, 8. [Google Scholar] [CrossRef] [PubMed]
  55. Trent, M.S.; Ribeiro, A.A.; Lin, S.; Cotter, R.J.; Raetz, C.R. An Inner Membrane Enzyme in Salmonella and Escherichia coli That Transfers 4-Amino-4-deoxy-L-arabinose to Lipid A: Induction in Polymyxin-Resistant Mutants and Role of a Novel Lipid-Linked Donor. J. Biol. Chem. 2001, 276, 43122–43131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Nummila, K.; Kilpeläinen, I.; Zähringer, U.; Vaara, M.; Helander, I.M. Lipopolysaccharides of polymyxin B-resistant mutants of Escherichia coii are extensively substituted by 2-aminoethyl pyrophosphate and contain aminoarabinose in lipid A. Mol. Microbiol. 1995, 16, 271–278. [Google Scholar] [CrossRef] [PubMed]
  57. Kawasaki, S.; Moriguchi, R.; Sekiya, K.; Nakai, T.; Ono, E.; Kume, K.; Kawahara, K. The cell envelope structure of the lipopolysaccharide-lacking gram-negative bacterium Sphingomonas paucimobilis. J. Bacteriol. 1994, 176, 284–290. [Google Scholar] [CrossRef] [Green Version]
  58. Salazar, J.C.; Hazlett, K.R.; Radolf, J.D. The immune response to infection with Treponema pallidum, the stealth pathogen. Microbes Infect. 2002, 4, 1133–1140. [Google Scholar] [CrossRef]
  59. Zhang, G.; Meredith, T.C.; Kahne, D. On the essentiality of lipopolysaccharide to Gram-negative bacteria. Curr. Opin. Microbiol. 2013, 16, 779–785. [Google Scholar] [CrossRef] [Green Version]
  60. Narita S-i Tokuda, H. Biochemical characterization of an ABC transporter LptBFGC complex required for the outer membrane sorting of lipopolysaccharides. FEBS Lett. 2009, 583, 2160–2164. [Google Scholar] [CrossRef] [Green Version]
  61. Freinkman, E.; Chng, S.-S.; Kahne, D. The complex that inserts lipopolysaccharide into the bacterial outer membrane forms a two-protein plug-and-barrel. Proc. Natl. Acad. Sci. USA 2011, 108, 2486–2491. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Freinkman, E.; Okuda, S.; Ruiz, N.; Kahne, D. Regulated Assembly of the Transenvelope Protein Complex Required for Lipopolysaccharide Export. Biochemistry 2012, 51, 4800–4806. [Google Scholar] [CrossRef] [PubMed]
  63. Thomas, C.; Tampé, R. Multifaceted structures and mechanisms of ABC transport systems in health and disease. Curr. Opin. Struct. Biol. 2018, 51, 116–128. [Google Scholar] [CrossRef] [PubMed]
  64. Ruiz, N.; Gronenberg, L.S.; Kahne, D.; Silhavy, T.J. Identification of two inner-membrane proteins required for the transport of lipopolysaccharide to the outer membrane of Escherichia coli. Proc. Natl. Acad. Sci. USA 2008, 105, 5537–5542. [Google Scholar] [CrossRef] [Green Version]
  65. Luo, Q.; Yang, X.; Yu, S.; Shi, H.; Wang, K.; Xiao, L.; Zhu, G.; Sun, C.; Li, T.; Li, D.; et al. Structural basis for lipopolysaccharide extraction by ABC transporter LptB2FG. Nat. Struct. Mol. Biol. 2017, 24, 469–474. [Google Scholar] [CrossRef]
  66. Sherman, D.J.; Lazarus, M.B.; Murphy, L.; Liu, C.; Walker, S.; Ruiz, N.; Kahne, D. Decoupling catalytic activity from biological function of the ATPase that powers lipopolysaccharide transport. Proc. Natl. Acad. Sci. USA 2014, 111, 4982–4987. [Google Scholar] [CrossRef] [Green Version]
  67. Wang, Z.; Xiang, Q.; Zhu, X.; Dong, H.; He, C.; Wang, H.; Zhang, Y.; Wang, W.; Dong, C. Structural and functional studies of conserved nucleotide-binding protein LptB in lipopolysaccharide transport. Biochem. Biophys. Res. Commun. 2014, 452, 443–449. [Google Scholar] [CrossRef]
  68. Tran, A.X.; Dong, C.; Whitfield, C. Structure and functional analysis of LptC, a conserved membrane protein involved in the lipopolysaccharide export pathway in Escherichia coli. J. Biol. Chem. 2010, 285, 33529–33539. [Google Scholar] [CrossRef] [Green Version]
  69. Sperandeo, P.; Lau, F.K.; Carpentieri, A.; De Castro, C.; Molinaro, A.; Deho, G.; Silhavy, T.J.; Polissi, A. Functional analysis of the protein machinery required for transport of lipopolysaccharide to the outer membrane of Escherichia coli. J. Bacteriol. 2008, 190, 4460–4469. [Google Scholar] [CrossRef] [Green Version]
  70. Bos, M.P.; Tefsen, B.; Geurtsen, J.; Tommassen, J. Identification of an outer membrane protein required for the transport of lipopolysaccharide to the bacterial cell surface. Proc. Natl. Acad. Sci. USA 2004, 101, 9417–9422. [Google Scholar] [CrossRef] [Green Version]
  71. Wu, T.; McCandlish, A.C.; Gronenberg, L.S.; Chng, S.S.; Silhavy, T.J.; Kahne, D. Identification of a protein complex that assembles lipopolysaccharide in the outer membrane of Escherichia coli. Proc. Natl. Acad. Sci. USA 2006, 103, 11754–11759. [Google Scholar] [CrossRef] [Green Version]
  72. Dong, H.; Xiang, Q.; Gu, Y.; Wang, Z.; Paterson, N.G.; Stansfeld, P.J.; He, C.; Zhang, Y.; Wang, W.; Dong, C. Structural basis for outer membrane lipopolysaccharide insertion. Nature 2014, 511, 52–56. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Qiao, S.; Luo, Q.; Zhao, Y.; Zhang, X.C.; Huang, Y. Structural basis for lipopolysaccharide insertion in the bacterial outer membrane. Nature 2014, 511, 108–111. [Google Scholar] [CrossRef]
  74. Suits, M.D.; Sperandeo, P.; Dehò, G.; Polissi, A.; Jia, Z. Novel structure of the conserved gram-negative lipopolysaccharide transport protein A and mutagenesis analysis. J. Mol. Biol. 2008, 380, 476–488. [Google Scholar] [CrossRef]
  75. Bowyer, A.; Baardsnes, J.; Ajamian, E.; Zhang, L.; Cygler, M. Characterization of interactions between LPS transport proteins of the Lpt system. Biochem. Biophys. Res. Commun. 2011, 404, 1093–1098. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Sperandeo, P.; Villa, R.; Martorana, A.M.; Šamalikova, M.; Grandori, R.; Deho, G.; Polissi, A. New insights into the Lpt machinery for lipopolysaccharide transport to the cell surface: LptA-LptC interaction and LptA stability as sensors of a properly assembled transenvelope complex. J. Bacteriol. 2011, 193, 1042–1053. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Dong, H.; Zhang, Z.; Tang, X.; Paterson, N.G.; Dong, C. Structural and functional insights into the lipopolysaccharide ABC transporter LptB2FG. Nat. Commun. 2017, 8, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Tang, X.; Chang, S.; Luo, Q.; Zhang, Z.; Qiao, W.; Xu, C.; Zhang, C.; Niu, Y.; Yang, W.; Wang, T.; et al. Cryo-EM structures of lipopolysaccharide transporter LptB(2)FGC in lipopolysaccharide or AMP-PNP-bound states reveal its transport mechanism. Nat. Commun. 2019, 10, 4175. [Google Scholar] [CrossRef] [Green Version]
  79. Li, Y.; Orlando, B.J.; Liao, M. Structural basis of lipopolysaccharide extraction by the LptB2FGC complex. Nature 2019, 567, 486–490. [Google Scholar] [CrossRef]
  80. Luo, Q.; Shi, H.; Xu, X. Cryo-EM structures of LptB2FG and LptB2FGC from Klebsiella pneumoniae in complex with lipopolysaccharide. Biochem. Biophys. Res. Commun. 2021, 571, 20–25. [Google Scholar] [CrossRef]
  81. Owens, T.W.; Taylor, R.J.; Pahil, K.S.; Bertani, B.R.; Ruiz, N.; Kruse, A.C.; Kahne, D. Structural basis of unidirectional export of lipopolysaccharide to the cell surface. Nature 2019, 567, 550–553. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Wilson, A.; Ruiz, N. The transmembrane alpha-helix of LptC participates in LPS extraction by the LptB2 FGC transporter. Mol. Microbiol. 2022, 118, 61–76. [Google Scholar] [CrossRef] [PubMed]
  83. Bertani, B.R.; Taylor, R.J.; Nagy, E.; Kahne, D.; Ruiz, N. A cluster of residues in the lipopolysaccharide exporter that selects substrate variants for transport to the outer membrane. Mol. Microbiol. 2018, 109, 541–554. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Hamad, M.A.; Di Lorenzo, F.; Molinaro, A.; Valvano, M.A. Aminoarabinose is essential for lipopolysaccharide export and intrinsic antimicrobial peptide resistance in Burkholderia cenocepacia. Mol. Microbiol. 2012, 85, 962–974. [Google Scholar] [CrossRef] [PubMed]
  85. Gronow, S.; Brade, H. Invited review: Lipopolysaccharide biosynthesis: Which steps do bacteria need to survive? J. Endotoxin Res. 2001, 7, 3–23. [Google Scholar] [CrossRef] [Green Version]
  86. Khare, D.; Oldham, M.L.; Orelle, C.; Davidson, A.L.; Chen, J. Alternating access in maltose transporter mediated by rigid-body rotations. Mol. Cell 2009, 33, 528–536. [Google Scholar] [CrossRef] [Green Version]
  87. Lundstedt, E.A.; Simpson, B.W.; Ruiz, N. LptB-LptF coupling mediates the closure of the substrate-binding cavity in the LptB2FGC transporter through a rigid-body mechanism to extract LPS. Mol. Microbiol. 2020, 114, 200–213. [Google Scholar] [CrossRef]
  88. Simpson, B.W.; Owens, T.W.; Orabella, M.J.; Davis, R.M.; May, J.M.; Trauger, S.A.; Kahne, D.; Ruiz, N. Identification of residues in the lipopolysaccharide ABC transporter that coordinate ATPase activity with extractor function. MBio 2016, 7, e01729-16. [Google Scholar] [CrossRef] [Green Version]
  89. Simpson, B.W.; Pahil, K.S.; Owens, T.W.; Lundstedt, E.A.; Davis, R.M.; Kahne, D.; Ruiz, N. Combining mutations that inhibit two distinct steps of the ATP hydrolysis cycle restores wild-type function in the lipopolysaccharide transporter and shows that ATP binding triggers transport. MBio 2019, 10, e01931-19. [Google Scholar] [CrossRef] [Green Version]
  90. Okuda, S.; Sherman, D.J.; Silhavy, T.J.; Ruiz, N.; Kahne, D. Lipopolysaccharide transport and assembly at the outer membrane: The PEZ model. Nat. Rev. Microbiol. 2016, 14, 337–345. [Google Scholar] [CrossRef] [Green Version]
  91. Sherman, D.J.; Xie, R.; Taylor, R.J.; George, A.H.; Okuda, S.; Foster, P.J.; Needleman, D.J.; Kahne, D. Lipopolysaccharide is transported to the cell surface by a membrane-to-membrane protein bridge. Science 2018, 359, 798–801. [Google Scholar] [CrossRef] [Green Version]
  92. Botte, M.; Ni, D.; Schenck, S.; Zimmermann, I.; Chami, M.; Bocquet, N.; Egloff, P.; Bucher, D.; Trabuco, M.; Cheng, R.K.; et al. Cryo-EM structures of a LptDE transporter in complex with Pro-macrobodies offer insight into lipopolysaccharide translocation. Nat. Commun. 2022, 13, 1–10. [Google Scholar] [CrossRef] [PubMed]
  93. Gu, Y.; Stansfeld, P.J.; Zeng, Y.; Dong, H.; Wang, W.; Dong, C. Lipopolysaccharide is inserted into the outer membrane through an intramembrane hole, a lumen gate, and the lateral opening of LptD. Structure 2015, 23, 496–504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Malojčić, G.; Andres, D.; Grabowicz, M.; George, A.H.; Ruiz, N.; Silhavy, T.J.; Kahne, D. LptE binds to and alters the physical state of LPS to catalyze its assembly at the cell surface. Proc. Natl. Acad. Sci. USA 2014, 111, 9467–9472. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Chng, S.S.; Ruiz, N.; Chimalakonda, G.; Silhavy, T.J.; Kahne, D. Characterization of the two-protein complex in Escherichia coli responsible for lipopolysaccharide assembly at the outer membrane. Proc. Natl. Acad. Sci. USA 2010, 107, 5363–5368. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Ruiz, N.; Chng, S.S.; Hiniker, A.; Kahne, D.; Silhavy, T.J. Nonconsecutive disulfide bond formation in an essential integral outer membrane protein. Proc. Natl. Acad. Sci. USA 2010, 107, 12245–12250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Chimalakonda, G.; Ruiz, N.; Chng, S.S.; Garner, R.A.; Kahne, D.; Silhavy, T.J. Lipoprotein LptE is required for the assembly of LptD by the β-barrel assembly machine in the outer membrane of Escherichia coli. Proc. Natl. Acad. Sci. USA 2011, 108, 2492–2497. [Google Scholar] [CrossRef] [Green Version]
  98. Xie, R.; Taylor, R.J.; Kahne, D. Outer Membrane Translocon Communicates with Inner Membrane ATPase To Stop Lipopolysaccharide Transport. J. Am. Chem. Soc. 2018, 140, 12691–12694. [Google Scholar] [CrossRef]
  99. Baeta, T.; Giandoreggio-Barranco, K.; Ayala, I.; Moura, E.C.; Sperandeo, P.; Polissi, A.; Simorre, J.P.; Laguri, C. The lipopolysaccharide-transporter complex LptB2FG also displays adenylate kinase activity in vitro dependent on the binding partners LptC/LptA. J. Biol. Chem. 2021, 297, 101313. [Google Scholar] [CrossRef]
  100. Randak, C.; Welsh, M.J. An intrinsic adenylate kinase activity regulates gating of the ABC transporter CFTR. Cell 2003, 115, 837–850. [Google Scholar] [CrossRef] [Green Version]
  101. Kaur, H.; Lakatos-Karoly, A.; Vogel, R.; Nöll, A.; Tampé, R.; Glaubitz, C. Coupled ATPase-adenylate kinase activity in ABC transporters. Nat. Commun. 2016, 7, 13864. [Google Scholar] [CrossRef] [PubMed]
  102. Villa, R.; Martorana, A.M.; Okuda, S.; Gourlay, L.J.; Nardini, M.; Sperandeo, P.; Dehò, G.; Bolognesi, M.; Kahne, D.; Polissi, A. The Escherichia coli Lpt transenvelope protein complex for lipopolysaccharide export is assembled via conserved structurally homologous domains. J. Bacteriol. 2013, 195, 1100–1108. [Google Scholar] [CrossRef] [PubMed]
  103. Falchi, F.A.; Taylor, R.J.; Rowe, S.J.; Moura, E.C.; Baeta, T.; Laguri, C.; Simorre, J.P.; Kahne, D.E.; Polissi, A.; Sperandeo, P. Suppressor Mutations in LptF Bypass Essentiality of LptC by Forming a Six-Protein Transenvelope Bridge That Efficiently Transports Lipopolysaccharide. mBio 2022, 0, e02202-22. [Google Scholar] [CrossRef] [PubMed]
  104. Benedet, M.; Falchi, F.A.; Puccio, S.; Di Benedetto, C.; Peano, C.; Polissi, A.; Deho, G. The lack of the essential LptC protein in the trans-envelope lipopolysaccharide transport machine is circumvented by suppressor mutations in LptF, an inner membrane component of the Escherichia coli transporter. PLoS ONE 2016, 11, e0161354. [Google Scholar] [CrossRef] [Green Version]
  105. Silhavy, T.J.; Kahne, D.; Walker, S. The bacterial cell envelope. Cold Spring Harb Perspect. Biol. 2010, 2, a000414. [Google Scholar] [CrossRef]
  106. Malinverni, J.C.; Silhavy, T.J. An ABC transport system that maintains lipid asymmetry in the gram-negative outer membrane. Proc. Natl. Acad. Sci. USA 2009, 106, 8009–8014. [Google Scholar] [CrossRef] [Green Version]
  107. Powers, M.J.; Trent, M.S. Phospholipid retention in the absence of asymmetry strengthens the outer membrane permeability barrier to last-resort antibiotics. Proc. Natl. Acad. Sci. USA 2018, 115, E8518–E8527. [Google Scholar] [CrossRef] [Green Version]
  108. Sutterlin, H.A.; Shi, H.; May, K.L.; Miguel, A.; Khare, S.; Huang, K.C.; Silhavy, T.J. Disruption of lipid homeostasis in the Gram-negative cell envelope activates a novel cell death pathway. Proc. Natl. Acad. Sci. USA 2016, 113, E1565–E1574. [Google Scholar] [CrossRef] [Green Version]
  109. Thong, S.; Ercan, B.; Torta, F.; Fong, Z.Y.; Wong, H.Y.A.; Wenk, M.R.; Chng, S.S. Defining key roles for auxiliary proteins in an ABC transporter that maintains bacterial outer membrane lipid asymmetry. Elife 2016, 5, e19042. [Google Scholar] [CrossRef]
  110. Benning, C. A role for lipid trafficking in chloroplast biogenesis. Prog. Lipid Res. 2008, 47, 381–389. [Google Scholar] [CrossRef]
  111. Mohn, W.W.; Van Der Geize, R.; Stewart, G.R.; Okamoto, S.; Liu, J.; Dijkhuizen, L.; Eltis, L.D. The actinobacterial mce4 locus encodes a steroid transporter. J. Biol. Chem. 2008, 283, 35368–35374. [Google Scholar] [CrossRef] [Green Version]
  112. Pandey, A.K.; Sassetti, C.M. Mycobacterial persistence requires the utilization of host cholesterol. Proc. Natl. Acad. Sci. USA 2008, 105, 4376–4380. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Low, W.-Y.; Thong, S.; Chng, S.-S. ATP disrupts lipid-binding equilibrium to drive retrograde transport critical for bacterial outer membrane asymmetry. Proc. Natl. Acad. Sci. USA 2021, 118, e2110055118. [Google Scholar] [CrossRef] [PubMed]
  114. Tang, X.; Chang, S.; Qiao, W.; Luo, Q.; Chen, Y.; Jia, Z.; Coleman, J.; Zhang, K.; Wang, T.; Zhang, Z.; et al. Structural insights into outer membrane asymmetry maintenance in Gram-negative bacteria by MlaFEDB. Nat. Struct. Mol. Biol. 2021, 28, 81–91. [Google Scholar] [CrossRef] [PubMed]
  115. Abellón-Ruiz, J.; Kaptan, S.S.; Baslé, A.; Claudi, B.; Bumann, D.; Kleinekathöfer, U.; van den Berg, B. Structural basis for maintenance of bacterial outer membrane lipid asymmetry. Nat. Microbiol. 2017, 2, 1616–1623. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Chong, Z.S.; Woo, W.F.; Chng, S.S. Osmoporin OmpC forms a complex with MlaA to maintain outer membrane lipid asymmetry in E scherichia coli. Mol. Microbiol. 2015, 98, 1133–1146. [Google Scholar] [CrossRef] [Green Version]
  117. Ekiert, D.C.; Bhabha, G.; Isom, G.L.; Greenan, G.; Ovchinnikov, S.; Henderson, I.R.; Cox, J.S.; Vale, R.D. Architectures of lipid transport systems for the bacterial outer membrane. Cell 2017, 169, 273–285.e17. [Google Scholar] [CrossRef] [Green Version]
  118. Ercan, B.; Low, W.Y.; Liu, X.; Chng, S.S. Characterization of interactions and phospholipid transfer between substrate binding proteins of the OmpC-Mla system. Biochemistry 2018, 58, 114–119. [Google Scholar] [CrossRef]
  119. Hughes, G.W.; Hall, S.C.; Laxton, C.S.; Sridhar, P.; Mahadi, A.H.; Hatton, C.; Piggot, T.J.; Wotherspoon, P.J.; Leney, A.C.; Ward, D.G.; et al. Evidence for phospholipid export from the bacterial inner membrane by the Mla ABC transport system. Nat. Microbiol. 2019, 4, 1692–1705. [Google Scholar] [CrossRef]
  120. Isom, G.L.; Davies, N.J.; Chong, Z.S.; Bryant, J.A.; Jamshad, M.; Sharif, M.; Cunningham, A.F.; Knowles, T.J.; Chng, S.S.; Cole, J.A.; et al. MCE domain proteins: Conserved inner membrane lipid-binding proteins required for outer membrane homeostasis. Sci. Rep. 2017, 7, 8608. [Google Scholar] [CrossRef]
  121. Chi, X.; Fan, Q.; Zhang, Y.; Liang, K.; Wan, L.; Zhou, Q.; Li, Y. Structural mechanism of phospholipids translocation by MlaFEDB complex. Cell Res. 2020, 30, 1127–1135. [Google Scholar] [CrossRef] [PubMed]
  122. Oldham, M.L.; Chen, J. Crystal structure of the maltose transporter in a pretranslocation intermediate state. Science 2011, 332, 1202–1205. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Shibagaki, N.; Grossman, A.R. The role of the STAS domain in the function and biogenesis of a sulfate transporter as probed by random mutagenesis. J. Biol. Chem. 2006, 281, 22964–22973. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Kolich, L.R.; Chang, Y.T.; Coudray, N.; Giacometti, S.I.; MacRae, M.R.; Isom, G.L.; Teran, E.M.; Bhabha, G.; Ekiert, D.C. Structure of MlaFB uncovers novel mechanisms of ABC transporter regulation. Elife 2020, 9, e60030. [Google Scholar] [CrossRef] [PubMed]
  125. Kamischke, C.; Fan, J.; Bergeron, J.; Kulasekara, H.D.; Dalebroux, Z.D.; Burrell, A.; Kollman, J.M.; Miller, S.I. The Acinetobacter baumannii Mla system and glycerophospholipid transport to the outer membrane. Elife 2019, 8, e40171. [Google Scholar] [CrossRef]
  126. Coudray, N.; Isom, G.L.; MacRae, M.R.; Saiduddin, M.N.; Bhabha, G.; Ekiert, D.C. Structure of bacterial phospholipid transporter MlaFEDB with substrate bound. Elife 2020, 9, e62518. [Google Scholar] [CrossRef]
  127. Zhou, C.; Shi, H.; Zhang, M.; Zhou, L.; Xiao, L.; Feng, S.; Im, W.; Zhou, M.; Zhang, X.; Huang, Y. Structural insight into phospholipid transport by the MlaFEBD complex from P. aeruginosa. J. Mol. Biol. 2021, 433, 166986. [Google Scholar] [CrossRef]
  128. Mann, D.; Fan, J.; Somboon, K.; Farrell, D.P.; Muenks, A.; Tzokov, S.B.; DiMaio, F.; Khalid, S.; Miller, S.I.; Bergeron, J.R. Structure and lipid dynamics in the maintenance of lipid asymmetry inner membrane complex of A. baumannii. Commun. Biol. 2021, 4, 1–9. [Google Scholar] [CrossRef]
  129. Zhang, Y.; Fan, Q.; Chi, X.; Zhou, Q.; Li, Y. Cryo-EM structures of Acinetobacter baumannii glycerophospholipid transporter. Cell Discov. 2020, 6, 86. [Google Scholar] [CrossRef]
  130. Yeow, J.; Tan, K.W.; Holdbrook, D.A.; Chong, Z.S.; Marzinek, J.K.; Bond, P.J.; Chng, S.S. The architecture of the OmpC–MlaA complex sheds light on the maintenance of outer membrane lipid asymmetry in Escherichia coli. J. Biol. Chem. 2018, 293, 11325–11340. [Google Scholar] [CrossRef] [Green Version]
  131. Hughes, G.W.; Sridhar, P.; Nestorow, S.A.; Wotherspoon, P.J.; Cooper, B.F.; Knowles, T.J. MlaFEDB displays flippase activity to promote phospholipid transport towards the outer membrane of Gram-negative bacteria. bioRxiv 2020. [Google Scholar] [CrossRef]
  132. Narita, S.I.; Tokuda, H. Bacterial lipoproteins; biogenesis, sorting and quality control. Biochim. Biophys. Acta. Mol. Cell Biol. Lipids. 2017, 1862, 1414–1423. [Google Scholar] [CrossRef]
  133. Grabowicz, M. Lipoproteins and Their Trafficking to the Outer Membrane. EcoSal Plus 2019, 8. [Google Scholar] [CrossRef] [PubMed]
  134. Hayashi, S.; Wu, H.C. Lipoproteins in bacteria. J. Bioenerg. Biomembr. 1990, 22, 451–471. [Google Scholar] [CrossRef] [PubMed]
  135. Babu, M.M.; Priya, M.L.; Selvan, A.T.; Madera, M.; Gough, J.; Aravind, L.; Sankaran, K. A database of bacterial lipoproteins (DOLOP) with functional assignments to predicted lipoproteins. J. Bacteriol. 2006, 188, 2761–2773. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Pugsley, A.P. The complete general secretory pathway in gram-negative bacteria. Microbiol. Rev. 1993, 57, 50–108. [Google Scholar] [CrossRef]
  137. Giménez, M.I.; Dilks, K.; Pohlschröder, M. Haloferax volcanii twin-arginine translocation substates include secreted soluble, C-terminally anchored and lipoproteins. Mol. Microbiol. 2007, 66, 1597–1606. [Google Scholar] [CrossRef]
  138. Sankaran, K.; Wu, H.C. Lipid modification of bacterial prolipoprotein. Transfer of diacylglyceryl moiety from phosphatidylglycerol. J. Biol. Chem. 1994, 269, 19701–19706. [Google Scholar] [CrossRef]
  139. Inouye, S.; Franceschini, T.; Sato, M.; Itakura, K.; Inouye, M. Prolipoprotein signal peptidase of Escherichia coli requires a cysteine residue at the cleavage site. EMBO J. 1983, 2, 87–91. [Google Scholar] [CrossRef]
  140. Vogeley, L.; El Arnaout, T.; Bailey, J.; Stansfeld, P.J.; Boland, C.; Caffrey, M. Structural basis of lipoprotein signal peptidase II action and inhibition by the antibiotic globomycin. Science 2016, 351, 876–880. [Google Scholar] [CrossRef] [Green Version]
  141. Jackowski, S.; Rock, C.O. Transfer of fatty acids from the 1-position of phosphatidylethanolamine to the major outer membrane lipoprotein of Escherichia coli. J. Biol. Chem. 1986, 261, 11328–11333. [Google Scholar] [CrossRef] [PubMed]
  142. Gupta, S.D.; Wu, H.C. Identification and subcellular localization of apolipoprotein N-acyltransferase in Escherichia coli. FEMS Microbiol. Lett. 1991, 62, 37–41. [Google Scholar] [CrossRef] [PubMed]
  143. Yamaguchi, K.; Yu, F.; Inouye, M. A single amino acid determinant of the membrane localization of lipoproteins in E. coli. Cell 1988, 53, 423–432. [Google Scholar] [CrossRef] [PubMed]
  144. Gennity, J.M.; Inouye, M. The protein sequence responsible for lipoprotein membrane localization in Escherichia coli exhibits remarkable specificity. J. Biol. Chem. 1991, 266, 16458–16464. [Google Scholar] [CrossRef] [PubMed]
  145. Matsuyama, S.; Tajima, T.; Tokuda, H. A novel periplasmic carrier protein involved in the sorting and transport of Escherichia coli lipoproteins destined for the outer membrane. EMBO J. 1995, 14, 3365–3372. [Google Scholar] [CrossRef]
  146. Matsuyama, S.; Yokota, N.; Tokuda, H. A novel outer membrane lipoprotein, LolB (HemM), involved in the LolA (p20)-dependent localization of lipoproteins to the outer membrane of Escherichia coli. EMBO J. 1997, 16, 6947–6955. [Google Scholar] [CrossRef]
  147. Takeda, K.; Miyatake, H.; Yokota, N.; Matsuyama, S.I.; Tokuda, H.; Miki, K. Crystal structures of bacterial lipoprotein localization factors, LolA and LolB. EMBO J. 2003, 22, 3199–3209. [Google Scholar] [CrossRef] [Green Version]
  148. Yakushi, T.; Masuda, K.; Narita, S.I.; Matsuyama, S.I.; Tokuda, H. A new ABC transporter mediating the detachment of lipid-modified proteins from membranes. Nat. Cell Biol. 2000, 2, 212–218. [Google Scholar] [CrossRef]
  149. Kaplan, E.; Greene, N.P.; Crow, A.; Koronakis, V. Insights into bacterial lipoprotein trafficking from a structure of LolA bound to the LolC periplasmic domain. Proc. Natl. Acad. Sci. USA 2018, 115, E7389–E7397. [Google Scholar] [CrossRef] [Green Version]
  150. Kaplan, E.; Greene, N.P.; Jepson, A.E.; Koronakis, V. Structural basis of lipoprotein recognition by the bacterial Lol trafficking chaperone LolA. Proc. Natl. Acad. Sci. USA 2022, 119, e2208662119. [Google Scholar] [CrossRef]
  151. Tang, X.; Chang, S.; Zhang, K.; Luo, Q.; Zhang, Z.; Wang, T.; Qiao, W.; Wang, C.; Shen, C.; Zhang, Z.; et al. Structural basis for bacterial lipoprotein relocation by the transporter LolCDE. Nat. Struct. Mol. Biol. 2021, 28, 347–355. [Google Scholar] [CrossRef] [PubMed]
  152. Sharma, S.; Zhou, R.; Wan, L.; Feng, S.; Song, K.; Xu, C.; Li, Y.; Liao, M. Mechanism of LolCDE as a molecular extruder of bacterial triacylated lipoproteins. Nat. Commun. 2021, 12, 4687. [Google Scholar] [CrossRef] [PubMed]
  153. Bei, W.; Luo, Q.; Shi, H.; Zhou, H.; Zhou, M.; Zhang, X.; Huang, Y. Cryo-EM structures of LolCDE reveal the molecular mechanism of bacterial lipoprotein sorting in Escherichia coli. PLoS Biol. 2022, 20, e3001823. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Domain organization of prokaryotic ABC exporters. (A): Half-size homodimer. (B): Single domain heterodimer. TMD = transmembrane domain. NBD = nucleotide-binding domain. The TMD and NBD are either fused on one polypeptide chain (A) or exist as separate polypeptides (B).
Figure 1. Domain organization of prokaryotic ABC exporters. (A): Half-size homodimer. (B): Single domain heterodimer. TMD = transmembrane domain. NBD = nucleotide-binding domain. The TMD and NBD are either fused on one polypeptide chain (A) or exist as separate polypeptides (B).
Ijms 24 06227 g001
Figure 2. Schematic of a transport cycle of a type IV ABC exporter. (A): In the inward-facing (IF) conformation, the transporter binds the substrate. (B): During substrate translocation across the membrane, the occluded state is formed with the substrate pocket closed to the inside and outside. (C): The transporter adopts the outward-facing (OF) conformation, and the substrate leaves the transporter to the outside.
Figure 2. Schematic of a transport cycle of a type IV ABC exporter. (A): In the inward-facing (IF) conformation, the transporter binds the substrate. (B): During substrate translocation across the membrane, the occluded state is formed with the substrate pocket closed to the inside and outside. (C): The transporter adopts the outward-facing (OF) conformation, and the substrate leaves the transporter to the outside.
Ijms 24 06227 g002
Figure 3. Schematic representation of the Mac, Lpt, Mla, and Lol system. For the Mac system, the ABC transporter MacB is shown in purple. The membrane fusion protein MacA and the outer membrane protein TolC are shown in grey. For the Lpt system, the ABC transporter LptB2FGC is shown in green. The periplasmic protein LptA and the translocon in the outer membrane LptDE are shown in grey. For the Mla system, the ABC transporter MlaFEDB is shown in red. The periplasmic protein MlaC and the outer membrane protein MlaA, together with OmpF, are shown in grey. For the Lol system, the ABC transporter LolCDE is shown in yellow. The periplasmic protein LolA and the outer membrane protein LolB are shown in grey. Inner membrane (IM) and outer membrane (OM) are depicted as grey bars. The peptidoglycan layer is omitted for reasons of clarity.
Figure 3. Schematic representation of the Mac, Lpt, Mla, and Lol system. For the Mac system, the ABC transporter MacB is shown in purple. The membrane fusion protein MacA and the outer membrane protein TolC are shown in grey. For the Lpt system, the ABC transporter LptB2FGC is shown in green. The periplasmic protein LptA and the translocon in the outer membrane LptDE are shown in grey. For the Mla system, the ABC transporter MlaFEDB is shown in red. The periplasmic protein MlaC and the outer membrane protein MlaA, together with OmpF, are shown in grey. For the Lol system, the ABC transporter LolCDE is shown in yellow. The periplasmic protein LolA and the outer membrane protein LolB are shown in grey. Inner membrane (IM) and outer membrane (OM) are depicted as grey bars. The peptidoglycan layer is omitted for reasons of clarity.
Ijms 24 06227 g003
Figure 4. Single particle cryo-EM structure of the assembled MacAB-TolC efflux system (PDB entry 5NIL). Trimeric TolC and hexameric MacA are shown in grey and dimeric MacB in color. The nucleotide-binding domains of MacB are shown in deep teal and marine, and the transmembrane domains, plus the periplasmic part, are shown in salmon and light magenta. The outer membrane (OM) and inner membrane (IM) are displayed as grey boxes. The peptidoglycan layer is omitted for reasons of clarity.
Figure 4. Single particle cryo-EM structure of the assembled MacAB-TolC efflux system (PDB entry 5NIL). Trimeric TolC and hexameric MacA are shown in grey and dimeric MacB in color. The nucleotide-binding domains of MacB are shown in deep teal and marine, and the transmembrane domains, plus the periplasmic part, are shown in salmon and light magenta. The outer membrane (OM) and inner membrane (IM) are displayed as grey boxes. The peptidoglycan layer is omitted for reasons of clarity.
Ijms 24 06227 g004
Figure 5. Structures of MacB in a nucleotide-free (PDB entry 5NIL), ATP-bound (PDB entry 5LIL), and ADP-bound (PDB entry 5WS4) state. NBDs are shown in deep teal and marine, and the TMDs with the periplasmic part are shown in salmon and light magenta. (A): Side view of MacB in the different nucleotide-free/-bound states. (B): Top view of the periplasmic part of MacB in the different nucleotide-free/-bound states.
Figure 5. Structures of MacB in a nucleotide-free (PDB entry 5NIL), ATP-bound (PDB entry 5LIL), and ADP-bound (PDB entry 5WS4) state. NBDs are shown in deep teal and marine, and the TMDs with the periplasmic part are shown in salmon and light magenta. (A): Side view of MacB in the different nucleotide-free/-bound states. (B): Top view of the periplasmic part of MacB in the different nucleotide-free/-bound states.
Ijms 24 06227 g005
Figure 6. Overview of the Lpt machinery spanning both membranes and the periplasm. Note that the number of LptA oligomers forming the periplasmic bridge is unknown. The figure was created using PyMOL and the PDB entries of LptDE (5IV9, grey in the outer membrane), LptA (2R19, grey in the periplasm), and LptB2FGC (6MJP, orange, salmon, light magenta, deep teal and marine in the inner membrane). The outer membrane (OM) and inner membrane (IM) are displayed as grey boxes. The peptidoglycan layer is omitted for reasons of clarity.
Figure 6. Overview of the Lpt machinery spanning both membranes and the periplasm. Note that the number of LptA oligomers forming the periplasmic bridge is unknown. The figure was created using PyMOL and the PDB entries of LptDE (5IV9, grey in the outer membrane), LptA (2R19, grey in the periplasm), and LptB2FGC (6MJP, orange, salmon, light magenta, deep teal and marine in the inner membrane). The outer membrane (OM) and inner membrane (IM) are displayed as grey boxes. The peptidoglycan layer is omitted for reasons of clarity.
Ijms 24 06227 g006
Figure 7. Model and structures of the Lpt systems transport cycle. LptC is shown in orange, while LptF and LptG are shown in light magenta and salmon, respectively. Both LptB protomers are shown in marine and deep teal. LPS is shown in red. Lipids and detergent molecules in the structure are shown as grey sticks, while nucleotides are shown as spheres. (A): Schematic model of LPS extraction by the extractor LptB2FGC. The different states of the transporter are labeled i-v with the PDB entry for the respective structure. i,ii: LPS enters the LptB2FGC cavity from the reader’s side. ii,iii: The LptC TMH leaves the LptFG interface, the cavity tightens and elevates LPS, forming tighter contacts. iii,iv: The LptB2 dimer closes, causing the cavity to collapse and pushes LPS upwards to the β-jellyroll domain of LptF. iv,v: ATP (yellow dots) is hydrolyzed to ADP (grey circles) and Pi, thereby opening LptB2 and the cavity for a new extraction cycle. (B): Crystal structures and single particle cryo-EM structures according to the different states (C): View on the cavity from the periplasmic side (β-jellyroll domains are omitted for clarity). Note that only structures of states ii and iii show LPS. Even though only the structure of state iv shows bound nucleotides, latest data suggest that ATP can bind already during earlier states. Structures of states ii–iv did not resolve the β-jellyroll domain of LptFG. The structure of state ii did not resolve the β-jellyroll domain of LptC, while the structures of states iii–v were lacking LptC completely.
Figure 7. Model and structures of the Lpt systems transport cycle. LptC is shown in orange, while LptF and LptG are shown in light magenta and salmon, respectively. Both LptB protomers are shown in marine and deep teal. LPS is shown in red. Lipids and detergent molecules in the structure are shown as grey sticks, while nucleotides are shown as spheres. (A): Schematic model of LPS extraction by the extractor LptB2FGC. The different states of the transporter are labeled i-v with the PDB entry for the respective structure. i,ii: LPS enters the LptB2FGC cavity from the reader’s side. ii,iii: The LptC TMH leaves the LptFG interface, the cavity tightens and elevates LPS, forming tighter contacts. iii,iv: The LptB2 dimer closes, causing the cavity to collapse and pushes LPS upwards to the β-jellyroll domain of LptF. iv,v: ATP (yellow dots) is hydrolyzed to ADP (grey circles) and Pi, thereby opening LptB2 and the cavity for a new extraction cycle. (B): Crystal structures and single particle cryo-EM structures according to the different states (C): View on the cavity from the periplasmic side (β-jellyroll domains are omitted for clarity). Note that only structures of states ii and iii show LPS. Even though only the structure of state iv shows bound nucleotides, latest data suggest that ATP can bind already during earlier states. Structures of states ii–iv did not resolve the β-jellyroll domain of LptFG. The structure of state ii did not resolve the β-jellyroll domain of LptC, while the structures of states iii–v were lacking LptC completely.
Ijms 24 06227 g007
Figure 8. Structural components of the Mla system in a cartoon representation. The homo-trimeric complex of OmpF and MlaA (PDB: 5NUO) situated in the outer membrane is shown in grey color. MlaC (PDB: 6GKI) is shown in the periplasm in grey color. Note: the exact number of MlaC molecules in the periplasm is unknown. The inner membrane complex of MlaFEDB (PDB: 6ZY2) is shown in color: hexameric MlaD, and both MlaB molecules are shown in orange. The two MlaE molecules are shown in salmon and light magenta. The two molecules of MlaF are shown in marine and deep teal. Outer membrane (OM) and inner membrane (IM) are displayed as grey boxes. The peptidoglycan layer is omitted for reasons of clarity.
Figure 8. Structural components of the Mla system in a cartoon representation. The homo-trimeric complex of OmpF and MlaA (PDB: 5NUO) situated in the outer membrane is shown in grey color. MlaC (PDB: 6GKI) is shown in the periplasm in grey color. Note: the exact number of MlaC molecules in the periplasm is unknown. The inner membrane complex of MlaFEDB (PDB: 6ZY2) is shown in color: hexameric MlaD, and both MlaB molecules are shown in orange. The two MlaE molecules are shown in salmon and light magenta. The two molecules of MlaF are shown in marine and deep teal. Outer membrane (OM) and inner membrane (IM) are displayed as grey boxes. The peptidoglycan layer is omitted for reasons of clarity.
Ijms 24 06227 g008
Figure 9. Structures and model representing the transport cycle of the Mla pathway. MlaC is shown in grey. The hexameric MlaD and both MlaB are shown in orange. MlaE is shown in salmon and light magenta. MlaF is shown in marine and deep teal. Lipid molecules are shown in red, and nucleotides are shown as spheres. The IM is displayed as a grey box. The peptidoglycan layer is omitted for reasons of clarity. (A): Different states of the ABC transporter during the transport cycle: nucleotide-free (PDB: 6ZY2), AMP-PNP-bound (PDB: 6ZY9), ADP-bound(PDB: 6ZY4), and substrate-bound (PDB: 6ZY3). (B): A schematic model of the retrograde transport of the lipid via MlaFEDB. Lipid-loaded MlaC binds to MlaD in the resting state of MlaFEDB. Binding of ATP (yellow dots) prompts the exit of the lipid molecule present in the cavity of MlaE from the last transport cycle. ATP hydrolysis to ADP (grey circles) and Pi prompts dimerization of MlaF and conformational changes in MlaE, which ultimately lead to the extraction of lipid from MlaD-MlaC into the cavity of MlaE. Upon release of hydrolyzed products, the conformation gets back to the resting state.
Figure 9. Structures and model representing the transport cycle of the Mla pathway. MlaC is shown in grey. The hexameric MlaD and both MlaB are shown in orange. MlaE is shown in salmon and light magenta. MlaF is shown in marine and deep teal. Lipid molecules are shown in red, and nucleotides are shown as spheres. The IM is displayed as a grey box. The peptidoglycan layer is omitted for reasons of clarity. (A): Different states of the ABC transporter during the transport cycle: nucleotide-free (PDB: 6ZY2), AMP-PNP-bound (PDB: 6ZY9), ADP-bound(PDB: 6ZY4), and substrate-bound (PDB: 6ZY3). (B): A schematic model of the retrograde transport of the lipid via MlaFEDB. Lipid-loaded MlaC binds to MlaD in the resting state of MlaFEDB. Binding of ATP (yellow dots) prompts the exit of the lipid molecule present in the cavity of MlaE from the last transport cycle. ATP hydrolysis to ADP (grey circles) and Pi prompts dimerization of MlaF and conformational changes in MlaE, which ultimately lead to the extraction of lipid from MlaD-MlaC into the cavity of MlaE. Upon release of hydrolyzed products, the conformation gets back to the resting state.
Ijms 24 06227 g009
Figure 10. Overview of the Lol pathway. After insertion of lipoproteins (red, labeled one time with LP) via the Sec or the Tat pathway into the IM, they are diacylated (black) by Lgt in a first modification step. This is followed by cleavage of the N-terminal signal sequence (orange) by Lsp. Subsequently, Lnt acylates the newly N-terminally located cysteine, which makes the now mature lipoprotein ready for transport via the Lol machinery. This starts with the extraction of the lipoprotein from the cytoplasmic membrane via the ABC transporter LolCDE (salmon, light magenta, deep teal, and marine), which leads to the delivery of the lipoprotein to the periplasmic chaperone LolA (grey). LolA shuffles the lipoprotein to the last checkpoint LolB (grey), which finally inserts the lipoprotein into the outer membrane.
Figure 10. Overview of the Lol pathway. After insertion of lipoproteins (red, labeled one time with LP) via the Sec or the Tat pathway into the IM, they are diacylated (black) by Lgt in a first modification step. This is followed by cleavage of the N-terminal signal sequence (orange) by Lsp. Subsequently, Lnt acylates the newly N-terminally located cysteine, which makes the now mature lipoprotein ready for transport via the Lol machinery. This starts with the extraction of the lipoprotein from the cytoplasmic membrane via the ABC transporter LolCDE (salmon, light magenta, deep teal, and marine), which leads to the delivery of the lipoprotein to the periplasmic chaperone LolA (grey). LolA shuffles the lipoprotein to the last checkpoint LolB (grey), which finally inserts the lipoprotein into the outer membrane.
Ijms 24 06227 g010
Figure 11. Structures of LolCDE in the Apo (PDB: 7V8M), lipoprotein-bound (PDB: 7V8L), and AMP-PNP-bound (PDB: 7V8I) form. LolC is shown in salmon, LolE is shown in light magenta, one LolD monomer is shown in deep teal, and the other LolD monomer is shown in marine. The lipoprotein is shown in red spheres. (A): Side-view of LolCDE. In the Apo state, LolCDE exhibits a V-shaped cavity that the lipoprotein enters from the membrane. Upon ATP binding (AMP-PNP-bound state), the central cavity is closed, and the substrate is shuffled out of LolCDE to LolA. (B): Cross-sectional view of the TMDs of LolCDE in the different substrate- or nucleotide-bound states. Movement of TMH2 of LolE upon ATP-binding into the central cavity extrudes the substrate out of the transporter.
Figure 11. Structures of LolCDE in the Apo (PDB: 7V8M), lipoprotein-bound (PDB: 7V8L), and AMP-PNP-bound (PDB: 7V8I) form. LolC is shown in salmon, LolE is shown in light magenta, one LolD monomer is shown in deep teal, and the other LolD monomer is shown in marine. The lipoprotein is shown in red spheres. (A): Side-view of LolCDE. In the Apo state, LolCDE exhibits a V-shaped cavity that the lipoprotein enters from the membrane. Upon ATP binding (AMP-PNP-bound state), the central cavity is closed, and the substrate is shuffled out of LolCDE to LolA. (B): Cross-sectional view of the TMDs of LolCDE in the different substrate- or nucleotide-bound states. Movement of TMH2 of LolE upon ATP-binding into the central cavity extrudes the substrate out of the transporter.
Ijms 24 06227 g011
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Bilsing, F.L.; Anlauf, M.T.; Hachani, E.; Khosa, S.; Schmitt, L. ABC Transporters in Bacterial Nanomachineries. Int. J. Mol. Sci. 2023, 24, 6227. https://doi.org/10.3390/ijms24076227

AMA Style

Bilsing FL, Anlauf MT, Hachani E, Khosa S, Schmitt L. ABC Transporters in Bacterial Nanomachineries. International Journal of Molecular Sciences. 2023; 24(7):6227. https://doi.org/10.3390/ijms24076227

Chicago/Turabian Style

Bilsing, Florestan L., Manuel T. Anlauf, Eymen Hachani, Sakshi Khosa, and Lutz Schmitt. 2023. "ABC Transporters in Bacterial Nanomachineries" International Journal of Molecular Sciences 24, no. 7: 6227. https://doi.org/10.3390/ijms24076227

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop