Skip to main content
The Journal of Biological Chemistry logoLink to The Journal of Biological Chemistry
. 2020 Jun 4;295(30):10340–10367. doi: 10.1074/jbc.REV120.011473

Role of the lipid bilayer in outer membrane protein folding in Gram-negative bacteria

Jim E Horne 1,2, David J Brockwell 1, Sheena E Radford 1,*
PMCID: PMC7383365  PMID: 32499369

Abstract

β-Barrel outer membrane proteins (OMPs) represent the major proteinaceous component of the outer membrane (OM) of Gram-negative bacteria. These proteins perform key roles in cell structure and morphology, nutrient acquisition, colonization and invasion, and protection against external toxic threats such as antibiotics. To become functional, OMPs must fold and insert into a crowded and asymmetric OM that lacks much freely accessible lipid. This feat is accomplished in the absence of an external energy source and is thought to be driven by the high thermodynamic stability of folded OMPs in the OM. With such a stable fold, the challenge that bacteria face in assembling OMPs into the OM is how to overcome the initial energy barrier of membrane insertion. In this review, we highlight the roles of the lipid environment and the OM in modulating the OMP-folding landscape and discuss the factors that guide folding in vitro and in vivo. We particularly focus on the composition, architecture, and physical properties of the OM and how an understanding of the folding properties of OMPs in vitro can help explain the challenges they encounter during folding in vivo. Current models of OMP biogenesis in the cellular environment are still in flux, but the stakes for improving the accuracy of these models are high. OMP folding is an essential process in all Gram-negative bacteria, and considering the looming crisis of widespread microbial drug resistance it is an attractive target. To bring down this vital OMP-supported barrier to antibiotics, we must first understand how bacterial cells build it.

Keywords: β-barrel assembly machinery (BAM) complex, outer membrane, membrane protein folding, protein-lipid interactions, Gram-negative bacteria, lipid membrane, folding kinetics, OMPome, antibiotic resistance, disorderase, protein folding, membrane bilayer, membrane protein, lipid, BAM complex


Proteins that span lipid bilayers come in two types, either α-helical or β-barrels. Whereas the cytosolic inner membranes (IMs) of bacteria and the plasma membrane of eukaryotes are comprised only of α-helical membrane proteins, β-barrel outer membrane proteins (OMPs) are found exclusively in the outer membranes (OMs) of diderm bacteria as well as in bacterially derived eukaryotic organelles, such as mitochondria and chloroplasts. The “OMPome” (the complement of OMPs encoded for by a genome) of Escherichia coli consists of a large number of proteins ranging in barrel size from 8 to 26 β-strands and includes monomers, small assemblies (dimers, trimers etc.), and oligomeric structures that can form up to 60-stranded pores (Fig. 1). Some OMPs comprise only the integral membrane β-barrel structure, whereas others have soluble domains in the periplasm or on the extracellular surface of the OM. Some OMPs have low copy number or can be absent in the OM under “standard” growth conditions (e.g. the E. coli porin OmpN) (14), and others are present in large copy number (e.g. OmpA is estimated to have >100,000 copies in the OM of E. coli, whereas OmpX, OmpC, and OmpF are estimated to have >20,000 copies each) (37). The functions of OMPs are also very diverse, including passive pores and ion channels (811), antibiotic efflux channels (1215), nutrient uptake systems (1618), maintenance of structural integrity (1921), biogenesis and upkeep of the OM (2226), host cell adhesion and invasion (2729), biofilm formation (3033), and cell defense (34, 35). Despite the enormous diversity of OMPs in E. coli, it is perhaps surprising that only two are essential: the 16-stranded BamA and 26-stranded LptD (36) (Fig. 1). This is perhaps even more remarkable considering that LptD itself relies on BamA for its assembly (37). LptD's biological role is to insert the lipid component of the outer leaflet of the OM (22, 38). BamA (part of the β-barrel assembly machinery, BAM) is required to fold and insert most (but not all) OMPs into the OM in vivo (39) (Table 1). The importance of BAM for the biogenesis of the OM is illustrated by the observation that despite the evolutionary distance between bacteria and eukaryotes, a homologue of BamA, Sam50, is retained in all mitochondria (70). Although only BamA and LptD are essential in E. coli under laboratory conditions, it is likely that many more OMPs will be necessary for bacteria to survive, invade new niches, and thrive in diverse environments. Understanding how OMPs fold has been the goal of researchers for approximately the last 3 decades, since the first observations were made that OMPs are capable of folding spontaneously into reconstituted lipid bilayers (71). Initially, the study of the structure and folding mechanisms of OMPs lagged behind those of their α-helical membrane protein counterparts, because the latter are more abundant in eukaryotes and were considered, initially at least, to be more important from the perspective of human health, as half of all approved drugs target α-helical membrane proteins (72, 73). However, in the last 15 years, it has become clear that OMPs are ubiquitous, and some are essential in bacteria (i.e. BamA and LptD) or in mitochondria (i.e. Sam50 and Tom40) (22, 23, 7477). Furthermore, the growth in antibiotic-resistant pathogens has highlighted the importance of the OM as a formidable barrier to the entry of antibiotics into bacteria as well as a site of efflux out (78) and as a shield against recognition of surface epitopes by natural or designed antibodies (7982). Hence, insights gained from studies of OMP folding and biogenesis are also vital for our understanding of human physiology (83) and will be key in guiding our choice of targets for the generation of new antibiotics and vaccines against Gram-negative bacteria (84). Consequently, a number of academic groups and drug companies have ongoing research projects targeting the essential OMPs BamA (the central β-barrel-containing subunit of BAM) and LptD (80, 82, 8592), with at least six reports of inhibitors of their function in 2018-2019 alone (9398).

Figure 1.

Figure 1.

Structures of transmembrane proteins found in the OM of E. coli K-12 MG1655. A list of all known and predicted transmembrane proteins in the OM of E. coli K-12 strain MG1655 was manually curated, creating the “OMP-ome.” The Protein Data Bank was then searched for solved structures of these proteins or close homologues. Where no high-resolution solved 3D structures were available, homology models were generated using the I-TASSER server (RRID:SCR_014627) (396). For two proteins, NfrA (the N4 bacteriophage receptor), and FlgH (the flagellar L-ring protein), no homology models could be generated. Predictions for YaiO, YcgI, YdbH, and YhjY generated deformed or broken barrels (possibly due to a lack of homology to existing structures), but their predictions are displayed to indicate their approximate structure. Extracellular domains of autotransporters have only been shown where accurate models could be built or crystal structures were available. OMPs are grouped here by the number of β-strands and then by protein family. The non-OMP subunits of the BAM complex are labeled below the central BamA subunit. Protein names are in red if they represent pseudogenes (inactivated by mutation in this strain) and blue if they are encoded on the F plasmid. The color of the box surrounding the protein names represents the number of β-strands in the β-barrel. Light Orange, 8; red, 10; light blue, 12; violet, 14; pink, 16; purple, 18; light green, 22; dark green, 24; brown, 26; black, oligomeric split β-barrel; gray, α-helical transmembrane region. Structures were aligned with each other by their β-barrel domains and rendered individually in PyMOL 2.X (Schrödinger, LLC). A list of the proteins with their associated family and PDB code can be found in Table S1.

Table 1.

Summary of BAM-dependent and BAM-independent OMPs in the OM of different bacteria

Listed are studies that present evidence a link is present (BAM catalysis–involved) or absent (BAM-independent folding) between the biogenesis of a particular OMP and the presence of the BAM complex. This list includes in vivo studies and in vitro folding studies performed with polar lipid extract from E. coli.

OMP(s) Family No. of β-strands Organism Reference
BAM catalysis–involved
    OmpA, OmpX, OmpT, OmPLA, OmpG Varied small barrels 8–14 E. colia Refs. 4046
    Various Autotransporters 12 E. coli Refs. 4652
    OprD Outer membrane porin 18 P. aeruginosa Ref. 53
    LamB Sugar porin 18 E. coli Refs. 54, 55
    Various TonB-dependent receptors 22 Caulobacter crescentus Refs. 56, 57
    TolC Outer membrane factor 3 × 4 (12) E. coli Ref. 58
    FimDb Fimbrial usher 24 E. coli Ref. 59
    LptD LPS assembly 26 E. coli Refs. 37, 54, 6062
    PilQb Type IV pilus secretin 14 × 4 (56) Neisseria meningitidis Refs. 23, 63
BAM-independent folding
    PulDb, XcpQb, GspDb T2SS secretin 15 × 4 (60) Klebsiella oxytoca, P. aeruginosa, E. coli Refs. 6467
    pIVb Phage secretin ? Phage f1 Refs. 68, 69
    CsgGb,c Curli secretion 9 × 4 (36) E. coli Ref. 67

a These studies were all performed in vitro.

b These proteins are often assembled as part of larger protein machineries or export/import pathways and may also include their own targeting and assembly factors.

c Also contains an N-terminal lipid anchor.

This review aims to provide a holistic view of our current understanding of the process of OMP biogenesis, including 1) the composition and physical and chemical properties of the OM in vivo; 2) current knowledge of the determinants of OMP folding through in vitro studies; and 3) how OMP folding depends on parameters such as the lipid composition, physical environment, and the presence/absence of BAM. Although information is drawn from different organisms, we focus on OMPs and the OM of E. coli because of the position of this bacterium as the de facto model organism for studying these processes.

Another brick in the wall: Building the OM

To understand OMP folding and biogenesis, it is first important to review our current understanding of the composition and architecture of the environment in which this process takes place: the complex and crowded bacterial OM.

Lipid types found in the OM

The OM of Gram-negative bacteria is unusual in that it is a highly asymmetric lipid bilayer, comprising an inner leaflet enriched in phospholipids and an outer leaflet containing lipopolysaccharide (LPS) (Fig. 2) (99). This is in contrast to the IM in Gram-negative and Gram-positive bacteria, which mostly contain phospholipids mixed between both leaflets. In Gram-negative bacteria, phospholipids in the OM generally have the canonical structure expected of a phospholipid, containing two hydrophobic acyl chains, with different length and degree of saturation. These are connected via an ester linkage to a headgroup that can be zwitterionic, or positively or negatively charged (Fig. 2). Cardiolipin (CL) is also found in the OM and has the appearance of a phospholipid dimer (Fig. 2). LPS is a much bulkier molecule, made up of a variable number of acyl chains (between 4 and 8, depending on the species) (100). The acyl chain can vary in length both within each molecule and between species (e.g. C10:0–C14:0 in Bordetella pertussis, C12:0–C14:0 in E. coli, C14:0–C21:0 in Chlamydia trachomatis, and C14:0–C28:0 in Agrobacterium tumefaciens) (101104). Furthermore, the acyl chains of LPS are usually shorter than those of the average phospholipid and are almost always saturated (105) (Fig. 2). The acyl chains in LPS are connected to a disaccharide diphosphate headgroup, which in turn is connected to a conserved “core” region made up of chained sugar groups, and then finally a highly variable sugar-containing O-antigen region (Fig. 2) (106, 107). Together, these sugar regions convey a large net negative charge to the outer surface of bacteria (107).

Figure 2.

Figure 2.

Common lipid types found in bacterial outer membranes and/or used in in vitro studies of OMP folding. Top, schematic of the generic structure of phospholipids and LPS. Bacterial lipids can be conceptualized as having two “domains”: a polar headgroup and a hydrophobic acyl tail region. In phospholipids, the acyl tails are connected by an ester linkage to a phosphate group and a variable headgroup region. PC and PE are zwitterionic, whereas PG carries a net negative charge. Note that PC lipids are not commonly found in bacterial membranes but are often used for OMP folding-studies in vitro due to their net neutral charge and propensity to form bilayers. Cardiolipin comprises two acyl tail regions connected by phosphate groups via a glycerol linkage and carries a net double negative charge. LPS is found exclusively in the OM of Gram-negative bacteria and varies considerably between species in both the number and length of acyl tails in the lipid A region and the sugar composition in the polysaccharide region (shown below). Here the most common structure of lipid A-Kdo2 for E. coli K-12 LPS is shown in full. Bottom, the architecture of a generic LPS is shown. The lipid A and core region are consistent with LPS found in E. coli K-12; however, this strain does not naturally produce an O-antigen, whereas many environmental and clinical strains do. Strains lacking the O-antigen region are said to contain “rough” LPS, and this can further be divided into subtypes dependent on truncations in the core region. The most extreme of these that is still viable at 37 °C under laboratory growth conditions is “deep rough” LPS, containing only lipid A-Kdo2. The O-antigen region is highly variable within species and can contain as many as 40 glycan repeats. Kdo, keto-deoxyoctulosonate; LDmanHep, l-glycero-d-manno-heptose; Glc, glucose; Gal, galactose; P, phosphate group; PEtN, phosphorylethanolamine.

Essentiality of specific lipids

E. coli is remarkably tolerant of modifications in its lipid biosynthesis pathways, with viable strains including bacteria in which synthesis of phosphatidylethanolamine (PE) (108, 109), phosphatidylglycerol (PG) and CL (110, 111), or CL alone (109, 112) is eliminated (see Fig. 2 for the structure of common lipid types); phosphatidylcholine (PC) synthesis is induced synthetically (E. coli lacks PC in its IM or OM, although this phospholipid is present in some bacterial membranes) (113, 114); gluco- or galactolipids are utilized (115); or even archaeal lipids are incorporated into the membrane (116). Although these strains are able to survive under laboratory conditions, their growth and virulence are affected (in some cases severely), stress responses are up-regulated, and defects of varying acuteness are seen in the structure and permeability of the cell envelope (109, 117). The effect of such changes in lipid composition in the OM on OMP biogenesis has not been investigated in detail for all of these strains. However, in PE-deficient strains, OmpF folding is impaired in a titratable manner, with complete lack of PE reducing folding yields from ∼100% in WT to <15% (109). Lack of CL causes less severe defects but still reduces OmpF folding yields to ∼25% (109) and has also been shown to cause mislocalization of the OMP IcsA, which normally resides at the cell pole in Shigella flexneri (117). Interestingly, in E. coli, lack of CL causes severe distention/detachment of the OM from the IM at the cell poles, and CL and PG have been observed to accumulate at cell poles and division sites (118), suggesting a role for CL in maintaining cell shape and integrity at sites of negative curvature (119). PG null only mutants have not been described, as CL utilizes PG for its biosynthesis. However, the creation of viable strains absent in PG synthesis also requires mutations in the major E. coli OM lipoprotein Lpp (Braun's lipoprotein), suggesting that lack of PG causes lethality primarily though lethal accumulation of Lpp at the IM (111, 120). The first step in lipoprotein maturation after translocation into the periplasm involves the transfer of a diacyl moiety from PG (121), and its absence presumably stalls maturation at this point. However, the OM lipoproteins LptE and BamD are essential in E. coli, so the viability of these lpp mutant strains suggests that alternate maturation pathways or sources of diacylglycerol must exist (122, 123). As BamA and LptD are essential OMPs, the fact that bacteria can still grow and divide in these strains (albeit poorly) suggests that other lipids can moonlight for the loss of PE, PG, or CL or that there is no absolute need for a particular phospholipid type as a minimum requirement for OMP biogenesis in E. coli. Nonetheless, the severe defects observed in these strains show that outside the laboratory, all of these components are needed for bacterial viability. This highlights that whereas a stable bilayer is the minimum requirement to fold an OMP, to understand how OMP biogenesis occurs in biologically relevant environments, consideration of the complexity of the OM environment is crucial.

Organization of lipid types within the OM

The E. coli OM contains PE, PG, CL, and LPS (Fig. 2). These lipid types are divided asymmetrically between the inner and outer leaflets of the OM, with the outer leaflet containing almost 100% LPS and the inner leaflet containing ∼80% PE, ∼15% PG, and ∼5% CL (Figs. 2 and 3). By contrast, the IM also contains ∼5% CL with a lower ratio of PE/PG of ∼70%/25% (124). Although it is physically possible for phospholipids to flip from the inner leaflet to the outer leaflet, this process is likely to be intrinsically slow (occurring on the order of hours or longer in vesicles in vitro) (125, 126). However, this process may be accelerated under conditions of OM/bilayer stress, such as exposure to antimicrobial peptides or detergents, in strains with truncated LPS (see Fig. 2), or after loss of OMPs (78, 127129). This process is therefore associated with increased permeability of the OM. Regardless of such events, the asymmetry of the OM is actively maintained in E. coli by the maintenance of lipid asymmetry (Mla) system, which removes errant phospholipids specifically from the outer leaflet of the OM to maintain its barrier function (130132).

Figure 3.

Figure 3.

Model depicting the structural organization of the E. coli OM. A schematic displays the degree of crowding in the OM. A, view of an imagined OM showing the dense packing of different size OMPs in monomers, dimers, and trimers interspersed with LPS in the outer leaflet (top) and phospholipids in the inner leaflet (bottom). Phospholipids are represented as dark gray circles with a diameter proportional to the headgroup size of PE/PG and LPS as light gray circles with a diameter proportional to the size of lipid A. Different OMPs are represented as idealized circles with diameters proportional to their number of strands. Blue outline, abundant 8-stranded; orange outline, rare 8-stranded; black, 16-stranded porin trimers; red, other 16-stranded; yellow, 22-stranded; green, 26-stranded. The overall LPR in this schematic is ∼9:1, with ∼2 LPS and ∼7 phospholipids per OMP, consistent with estimates for the LPR of the E. coli OM. B, left, high-resolution AFM image of OM extracts from Roseobacter dentrificans imaged from the periplasmic side showing a dense lattice of porin trimers. Right, atomic model of the packing of porin trimers derived from the AFM data. Reproduced with permission from Jarosławski et al. (156). This research was originally published in Molecular Microbiology. Jarosławski, S., Duquesne, K., Sturgis, J. N., and Scheuring, S. High-resolution architecture of the outer membrane of the Gram-negative bacteria Roseobacter denitrificans. Molecular Microbiology 2009; 74:1211–1222. © Wiley–Blackwell. C, view of an imagined OM with the same LPR as in A but assuming a more extreme clustering of most OMPs. Only the inner leaflet is shown. Despite having the same LPR values, the buried surface area of the clustered OMPs frees up more lipid to form larger bulk lipid domains.

The lipid acyl chain composition is diverse

The acyl chain composition of lipids in the OM of E. coli is more variable than that of their headgroups. The acyl chain composition of the OM depends on the growth conditions, with acyl chains varying in length from C12 to C18, as well as in the degree of saturation or the presence of cyclopropyl modifications (Fig. 2) (133138). Lipidomics has provided insights into the range of acyl groups found in E. coli membranes. Early experiments in E. coli K1062 reported the presence of C12:0, C14:0, C16:0, C16:1, C18:1, and cyclo-C19:0 acyl chains in phospholipids from the IM and OM (139). Examining total lipid content in E. coli K-12 strain LM3118 grown at 37 °C and harvested in stationary phase showed that the acyl chains of PE and PG lipids were comprised primarily of C12:0, C14:0, C16:0, C16:1, C18:0, and C18:1 (with C16:0 being about 3 times more abundant than the other acyl chains), with a lesser contribution from C15:0, cyclo-C17:0, and cyclo-C19:0 (140, 141). These acyl chains are combined to form a large variety of phospholipid types, most containing at least one unsaturated acyl bond or cyclo-propyl group, although diC16:0PE, diC16:0PG, C16:0C14:0PG, C16:0C12:0PE, diC14:0PE, and diC12:0PE lipids were also observed. Despite cyclo-propyl acyl chain–containing lipids being relatively understudied, the most common lipid species detected under these conditions was C16:0/cyclo-C17:0 (Fig. 2). Cyclo-propyl–containing lipids are produced in large quantities in stationary phase cultures from the conversion of double bonds in unsaturated chains to cyclo-propyl groups (142, 143). Although the physiological functions of these modifications are unclear, they appear to be related to protection of the bacteria against a variety of adverse environmental conditions (143), including acid shock (144, 145), osmotic shock (146), high alcohol concentrations (147), or high temperature (148). Furthermore, OMP folding in E. coli occurs primarily during exponential growth (56, 149), while expression is down-regulated (150, 151) and OMPs are lost from the OM (152) when bacteria enter stationary phase. Thus, the relevance of this lipid modification for OMP biogenesis may be minor under favorable growth conditions.

Few studies have examined whether significant differences or biases exist between the acyl chain composition of phospholipids in the IM and those of the OM of Gram-negative bacteria. Some have reported an enrichment of shorter acyl chains (C12:0 and C14:0) (139), saturated fatty acids (153), lyso-PE lipids (154), and C16:0 acyl chains in the OM and a depletion of polyunsaturated acyl chains (155). These biases, however, generally vary with growth conditions, and it is unclear to what extent this simply reflects the presence of the lipid A component of the outer leaflet. Regulatory systems that alter the acyl chain composition of lipids specific to the OM are known for LPS, including enzymes that alter the acyl chains attached to lipid A to modulate the endotoxicity of this lipid type during growth in a host (99). Modulation of the acyl chain content of lipid A has also been shown to occur in E. coli when under selective pressure from an external insult by the addition of a bactericidal BamA-specific antibody. This suggests a direct link between modulation of lipid content and a selective pressure to efficiently fold OMPs (93). However, it is not clear whether this change in lipid A reflects a need to aid the function of an essential BAM client (e.g. LptD), is related to conformational changes in BamA, or is simply a response to a defective permeability barrier.

This diversity of acyl chain types gives E. coli a wide range of lipids with which it can tailor the biophysical properties of its membranes both globally and locally. This variety may allow it to deal with local minor deformations of the membrane, due to either random thermal fluctuations or the presence of membrane-bound or embedded proteins, as acyl chains can diffuse laterally and occupy the most energetically favorable position, dependent on the match between their own physicochemical properties (length and saturation) and those of the membrane environment.

A crowded environment

Although the familiar fluid-mosaic model of membranes found in most textbooks depicts a biological membrane with just a few proteins floating in a “sea” of lipid, the OM is markedly different, containing instead a much higher fraction of protein by weight, with lipid/protein ratios (LPRs) (w/w) estimated to be between 0.14 and 0.36 (139, 156), corresponding to only 2–4 LPS and 4–10 phospholipid molecules per OMP (157). Estimates based on biochemical studies suggest that as much as 50% of the surface area of the OM may be occupied by OMPs (7), whereas AFM studies (156) (Fig. 3B), extrapolation from the copy numbers of OMPs measured by proteomics (3, 4), and the above measurements of the LPR suggest that this value may be even higher. For example, a copy number of 100,000 for OmpA would imply that ∼6–20% of the surface area of E. coli (dependent on the size of the bacterium) would be occupied by this protein alone. Hence, the OM could be considered more like a protein-rich layer solubilized in a relatively small amount of lipid (Fig. 3, A and B). Despite the low LPR of the OM, the diffusion rates of OMPs in the OM of E. coli are similar to those of inner membrane proteins (IMPs) but are, on average, slower (diffusion coefficients of 0.006–0.15 μm2/s for OMPs versus 0.001–0.4 μm2/s for IMPs) (56, 158170) (Fig. 4A). For comparison, the length elongation rate of E. coli alone is ∼0.006 μm/s (171), whereas the diffusion coefficients of LPS in the OM of Salmonella typhimurium are ∼0.00005 and 0.02 μm2 s−1 (for O-antigen–containing and truncated “deep rough” LPS, respectively (Fig. 2)) (172, 173), lipid probes in the IM of E. coli ∼0.8-1.5 μm2/s (162, 174, 175), and the periplasm, cytoplasm, and buffer ∼3, 0.4-9, and ∼87 μm2/s, respectively (159, 168, 176180) (Fig. 4B). What particularly distinguishes OMPs from IMPs is their restricted diffusion areas, with diffusion being confined within clusters in the OM, compared with free diffusion of most IMPs in the IM (56, 166, 181, 182). These observations can be explained by the propensities of OMPs to form clusters (56, 149, 156, 181188) and/or to interact strongly with LPS or components of the cell envelope (189, 190). OMPs that have a lower tendency to cluster and/or interact with cell envelope components less strongly may exhibit higher diffusion coefficients but will ultimately become “corralled” within OMP-LPS domains. On this point, an abundance of clustered OMPs with low mobility may make the OM locally rigid. Indeed, molecular dynamics (MD) simulations have shown that membranes containing 8–12-stranded OMPs are much stiffer than membranes containing only DMPC (diC14:0PC) (157). However, simulations investigating larger length scales have shown that crowding a bilayer with some OMPs (i.e. BtuB), but not others (i.e. OmpF), can reduce the global bending rigidity of a POPE (C16:0C18:1PE)/POPC (C16:0C18:1PC) membrane (191)—an effect that would explain how a cell with a protein-rich OM could still maintain its shape. An interesting alternative possibility to explain the potentially incompatible concepts of OMP folding and a protein-dense, lipid-poor OM would be to consider the OM as an inhomogeneous mixture of protein and lipid (Fig. 3C). The view of OMP monomers or trimers, well-solubilized with lipid, would leave little bulk lipid available for nascent OMPs to fold. However, by clustering OMPs together into regions resembling two-dimensional crystals (i.e. forming regions highly enriched with OMPs and little to no lipid—a local LPR closer to 1:1 (mol/mol)), sufficient lipid-rich regions would be created to enable OMP folding and insertion (Fig. 3, B and C). Regardless of which model is correct, this locally stiff, crowded, and confined environment, with a relative paucity of free lipid, poses a challenging environment into which OMPs must fold.

Figure 4.

Figure 4.

Comparison of physical properties of bacterial membranes. A, box plots showing the range of diffusion coefficients reported for OMPs and IMPs (see “A crowded environment”). Boxes show interquartile range calculated by the Tukey method with the median indicated as a boldface horizontal line. Whiskers show the minimum and maximum values. B, comparison of the diffusion coefficients of membrane proteins with other components of bacteria. Whiskers are only shown for components that have three or more values reported in the literature. All values are reported from in vivo studies. LPS, diffusion of LPS molecules in S. typhimurium. Lipid, diffusion rate of a fluorescent lipid reporter probe in E. coli membranes. Peri, diffusion of soluble protein in the E. coli periplasm. Cyto, diffusion of soluble protein in the E. coli cytoplasm. C, viscosities of different membrane environments as measured by the use of fluorescent BODIPY C10 lipid reporter probes. E. coli data are from Mika et al. (175), and synthetic phospholipid data are from Wu et al. (225). BODIPY C10 specifically incorporates into the IM of E. coli, and removal of the OM minimally affects the measured viscosity. Synthetic phospholipid 200 nm LUVs were comprised of DLPC, DMPC, POPC, or DOPC.

More than a mix of lipid types

The physical properties of lipid membranes

Lipid bilayers can be characterized by a number of physical, mechanical, and chemical parameters, including stored curvature elastic stress (lateral pressure), melting temperature, the bulk lipid phase, the presence of lipid rafts, membrane viscosity, and headgroup charge (192194) (Fig. 5). Many of these properties are interrelated and can be modulated by altering the acyl chain composition and/or the phospholipid headgroup and by altering the relative amounts of phospholipid, CL, and LPS (195199). Stored curvature elastic stress makes membranes more rigid and less elastic in terms of their ability to deform or bend. This property can be introduced by the presence of nonbilayer-forming lipids in otherwise bilayer-forming membranes (Fig. 5). For example, PE lipids create negative curvature, whereas PG and PC lipids have zero or low spontaneous curvature, which allows them to readily form bilayers (200). Doping bilayers containing PC or PG with PE or CL generates a tension in packing of the different lipid types, creating a crowding, or pressure, near the center of the bilayer where each leaflet meets (Fig. 5), which is further altered by the length of the acyl chains (with shorter chains reducing this elastic stress) (200).

Figure 5.

Figure 5.

Physical and mechanical properties of a lipid bilayer. A, top, the phase of a lipid bilayer depends on the temperature, with the lipids being in an ordered (gel) phase below the Tm and in a (liquid) disordered phase above the Tm. At the transition temperature, frustration between packing of regions of gel and liquid phase causes defects to occur at these boundaries. Bottom, a typical differential scanning calorimetry curve illustrating the thermal response of a DMPC (diC14:0PC) bilayer with the regions of each phase colored as above. B, the hydrophobic thickness of a membrane depends on the lipid acyl chain length. However, when an OMP becomes embedded in a lipid bilayer, the membrane responds by trying to “match” the hydrophobic thickness of the bilayer to that of the protein to minimize the energetic penalty of exposing polar lipid headgroups to a hydrophobic OMP surface or hydrophobic acyl tails to polar OMP loops. C, mixtures of lipids can separate, forming “rafts” or domains dependent on the physical conditions and lipid type. CL has a high propensity for negative curvature and has been shown to be enriched at cell poles and division sites where the membrane constricts. CL has also been observed to bind to membrane protein complexes such as the BAM complex (397) and cluster under patches of LPS in MD simulations (209, 210), suggesting that it may help stabilize bilayer packing defects (which might be induced by LPS) and stabilize regions of large hydrophobic mismatch (e.g. around embedded proteins such as BamA). D, schematic describing stored curvature elastic stress and how this depends on lipid type. Adapted from Booth and Curnow (398). This research was originally published in Current Opinion in Structural Biology. Booth, P. J., and Curnow, P. Folding scene investigation: membrane proteins. Current Opinion in Structural Biology 2009; 19:8–13. © Elsevier. Attractive and repulsive interactions driven by the packing of lipids create a pressure differential along the normal of the membrane that must be overcome to deform or alter lipid packing. Incorporation of lipids that have a tendency toward negative curvature (due to the relative size of the headgroup versus the acyl chain (e.g. PE lipids)) into a bilayer formed from lipids with a neutral or low curvature tendency (e.g. PC lipids) generates a stress force within the bilayer due to the opposing tendencies for bilayer formation of these lipids.

The lipid phase of a membrane is also dependent on its lipid composition and on the temperature (Fig. 5). Bilayers exist primarily in one of two major states, a solid “gel” phase in which the acyl chains are tightly packed and the mobility of lipid molecules is low, and a “liquid” phase, where lipid mobility is higher (201203). Furthermore, analogous to the familiar phase change of ice to water, lipids in a gel-phase membrane can “melt” to become the liquid phase at a temperature characteristic of the particular lipid type—called the transition temperature, Tm. Lipid mixtures can also adopt a “liquid-ordered” phase (with the classical pure liquid phase described as “liquid-disordered”). This liquid-ordered phase contains lipids that are highly mobile but have well-ordered acyl chains and is often associated with the formation of lipid rafts in cholesterol-containing membranes in eukaryotes (204206). Although sterol lipids are rare in bacteria, CL may play a similar role in increasing lipid order, and there is evidence that CL can participate in the formation of rafts/domains in membranes in vitro and in vivo (Fig. 5) (although it is likely this mechanism is distinct from that of cholesterol) (207, 208). In silico, CL has been observed to form clusters under patches of LPS in simulations of bacterial OMs in a role that may compensate for packing defects in the outer leaflet of the OM (209, 210).

The physical properties of native lipid extracts

The OM differs from the IM in its highly asymmetric structure, large fraction of proteins by weight, permeability to small molecules (<600 Da), and lack of energization across it (11, 99, 139, 156, 211). However, relatively little is known about how these unique features of the OM affect its mechanical properties and how this differs from the IM. E. coli is known to alter the lipid content of its membranes, particularly the length and degree of saturation of acyl chains, in response to changes in growth temperature. This process, termed “homeoviscous adaptation” (212), suggests that bacteria actively maintain their membranes at a constant level of “fluidity,” or in a particular phase, that enables them to adjust to different environments. For example, whereas total lipid extract from E. coli K-12 W3110 showed approximately the same headgroup content (with a minor monotonic increase in CL and decrease in PG) when grown at 30, 37, 42, or 45 °C (213), the ratio of saturated over unsaturated acyl chains increased with temperature. This suggests a change to a more “rigid” mixture of phospholipids at higher growth temperatures to balance the increased fluidity caused by the input of thermal energy.

Membrane lipid properties can be probed using fluorescent reporter dyes that partition into membranes (either globally or into specific lipid phases) and alter their excitation or emission properties according to the local lipid environment. Hence, these dyes can be used as reporters of membrane viscosity, degree of hydration, phase, and mobility (214). The fluorophore laurdan partitions into membranes through its acyl tail, whereas its naphthalene-based headgroup resides in the interfacial region of bilayers, where its fluorescence excitation and emission spectra are sensitive to the degree of hydration of the bilayer, allowing it to report on the phase and order of lipids in a bilayer (215). Using laurdan, a Tm of E. coli total lipid extract of <14 °C was determined for bacteria grown at 30 °C or 37 °C, whereas the Tm was higher (Tm ∼20-22 °C) for bacteria grown at 42 °C and elevated again (Tm ∼27 °C) for bacteria grown at 45 °C (213). This shows that the lipids in E. coli are natively in the liquid phase, and their Tm is as much as 20 °C below the growth temperature. However, when the OM and IM are considered separately, it becomes clear that “global” lipid properties do not accurately capture the differences between these two different membranes. Deuterium NMR studies, used to measure the order of acyl chains, found that lipids in an OM preparation of E. coli L51 were less fluid than those from the IM and that the Tm was ∼7 °C higher than the IM (216, 217). Electron spin resonance experiments on IMs and OMs of E. coli W1485 doped with a spin-labeled stearic acid probe found the Tm of the OM (26 °C) to be ∼13 °C higher than the IM (13 °C) when the bacteria were grown at 37 °C (196). Fluorescence polarization studies using IM and OM extracts from E. coli B doped with parinaric acid also found that the phase transition of the OM was ∼15 °C higher than the IM, initiating its phase transition at 40 °C (218).

E. coli membranes in situ

Few studies on the properties of bacterial membrane lipid order (or phase) in vivo have been reported to date, but current data suggest that the organization of the OM is more complicated than that derived using lipid extracts, as described above. Differential scanning calorimetry studies on whole cells of E. coli W945 grown at 20 or 37 °C observed two reversible transitions, one well below the growth temperature, which was suggested to correspond to the IM, and another slightly above the growth temperature, which was assigned to the OM (219). Experiments on fixed cells using laurdan and the dye 1,3-diphenyl-1,3,5-hexatriene as a probe of local viscosity (214) have shown that E. coli membranes are predominantly in the liquid phase. These experiments also revealed that these membranes are heterogeneous and contain at least two distinct phases, one more liquid and one less so, possibly indicating the presence of distinct lipid domains (220). However, an alternative explanation of these data is that they reflect a difference between the IM and the OM, because it is not clear to which membrane these probes localize. Indeed, whereas early studies suggested that these dyes (and others, including FM 4-64) localize primarily to the IM (221), it is now known that FM 4-64 partitions specifically into the OM immediately after labeling (19), and 1,3-diphenyl-1,3,5-hexatriene may equilibrate between both membranes (222, 223) or be trapped mainly in the first hydrophobic surface encountered (224). Measurements of membrane viscosity using “molecular rotor” dyes such as BODIPY C10 (viscosity alters the fluorescence lifetime of the probe) showed that the IM of E. coli is more viscous than previously believed, with an average viscosity of 980 cP for intact E. coli, 950 cP for spheroplasted cells at 37 °C, and 200 cP for large unilamellar vesicles (LUVs) of E. coli total lipid extract at 37 °C (175). By contrast, a lower viscosity (∼60 cP) was consistently observed for 200-nm LUVs formed from DLPC (diC12:0PC), DMPC (diC14:0PC), POPC (C16:0C18:1PC), or DOPC (diC18:1PC) in their liquid phase at 37 °C (225) (Fig. 4C). Although not yet measured directly, the lower diffusion coefficients of OMPs compared with IMPs (Fig. 4, A and B) suggest that the viscosity of the OM may be even higher still, and this is a characteristic that model lipid membranes commonly used for OMP-folding studies (see below) clearly cannot capture.

The influence of LPS on the mechanical properties of the OM

On the basis of the physicochemical properties of LPS extracts (which generally show Tm values at or above the growth temperature) (226230) and those of the OM (discussed above), some authors have argued that the OM is more likely to exist in the gel phase at physiological temperatures (78). However, the exact thermotropic response of LPS in the outer leaflet may vary, depending on the composition (particularly the presence of hydrogen bond donors and acceptors) and size of the polysaccharide region of the LPS molecules incorporated (78, 229, 231, 232). In vitro models of the OM using an asymmetric bilayer deposited on silicon or lipid-coated gold surfaces showed that the membrane has unusual mixed characteristics with elements of both liquid- and gel-phase lipids (233, 234). The OM in these studies comprised “rough” LPS (containing the conserved polysaccharide core but lacking the O-antigen) (Fig. 2) in the outer leaflet and phospholipid (DPPC (diC16:0PC)) in the inner leaflet, with lipid order parameters measured using neutron reflectometry and attenuated total reflection FTIR spectroscopy (ATR-FTIR) (233). Two transition midpoints (Tm) were observed, one just below (∼36 °C) the physiological growth temperature of E. coli (37 °C) for the outer leaflet of LPS and the other above the Tm (∼39 °C) for the inner leaflet comprising DPPC (diC16:0PC) (233). Even though the composition of the inner leaflet lipids and LPS differ in vivo from those used in this study, the results suggest that the LPS component of the asymmetric OM may confer greater rigidity to the OM. This agrees with deuterium NMR studies of preparations of E. coli OM and IM, which found that at a given temperature, phospholipids were more ordered and a larger fraction were in the gel phase in the OM than the IM (217, 235). However, it should be noted that extraction of the OM likely causes mixing of the inner and outer leaflets, reducing the asymmetry (236).

The fluidity of the OM of E. coli may also be controlled by temperature-dependent modification of LPS in the outer leaflet. The lpxT gene in E. coli transfers a phosphate group onto lipid A, which may alter the rigidity of the OM, and expression of this gene is regulated by an mRNA thermostat (237). LpxT is an IM protein that covalently modifies LPS before it reaches the OM. The protein was shown to be stable between 28 and 42 °C; however, its mRNA levels fall dramatically across this same temperature range (237). Another LPS biosynthesis pathway gene, lpxP, replaces the C12:0 chain in E. coli lipid A (which is normally installed by lpxL) with a C16:1 chain, but expression of this protein is only induced at low temperature (12 °C) and may also be regulated by an mRNA thermostat similar to lpxT (238). Francisella novicida synthesizes lipid A with shorter acyl chains at low growth temperatures (25 °C compared with 37 °C), an effect that has been linked to both the differential expression and enzyme activity of the lpxD2 and lpxD1 genes (which add different length acyl chains to lipid A) at 25 and 37 °C, respectively (239). In Yersinia pestis, the lpxR gene, which is responsible for the removal of acyl chains from lipid A, is inactive at 21 °C (resulting in a hexa-acylated LPS) but functional at 37 °C, resulting in a tetra-acylated LPS in the OM of bacteria grown at this temperature (240). The deacylase enzyme and OMP, PagL, alters lipid A through the removal of acyl chains. The activity of this enzyme in Pseudomonas aeruginosa is also affected by growth temperature, with activity falling at low growth temperature (≤ 21 °C) (241). The extent to which these temperature-dependent modifications are a mechanical response to maintain or alter membrane rigidity/fluidity or an immune-modulating response triggered by the detection of a host environment, or both, remains unclear (242). Despite this, it is likely that the structure and mechanical properties of LPS are important for OMP folding. For example, genetic alterations to the biosynthetic pathway of LPS, which cause changes in its degree of acylation or its sugar headgroup have been shown to cause severe defects in the biogenesis of OMPs (243245).

The effect of asymmetry

The asymmetric architecture of the OM, with an inner leaflet containing canonical phospholipids and an outer leaflet containing LPS with its highly acylated lipid A attached to large sugar groups that protrude into solution, impacts the physical properties of the OM and creates a lipid environment that is very different from that of the IM in vivo and vesicles of lipids commonly used in in vitro experiments. Representative models of the OM have been built in silico and studied using coarse-grained molecular dynamics (CG-MD) and atomistic molecular dynamics (A-MD). A model was built of the P. aeruginosa PAO1 OM, with an LPS outer leaflet and a DPPE (diC16:0PE) inner leaflet and studied by A-MD (246). At 37 °C, molecules of DPPE in the inner leaflet showed diffusive movement consistent with a liquid phase, whereas LPS in the outer leaflet showed an order of magnitude lower mean-square displacement. Despite this low lateral mobility, calculation of the lipid order parameters of the acyl tails of LPS indicated that they are fluid and not ordered as would be expected for lipids in the gel phase (246). Similar results have been observed in other A-MD simulations of E. coli and P. aeruginosa OMs with an outer leaflet containing LPS with a short O-antigen region (247, 248) or a “rough” LPS (248, 249) (Fig. 2) and an inner leaflet comprised of a 75:20:5 (mol/mol/mol) mix of PPPE (C16:0C16:1PE)/PVPG (C16:0C18:1PG)/CL. CG-MD simulations of the E. coli OM with a DPPE (diC16:0PE) inner leaflet and an outer leaflet containing a different ratio of LPS/DPPE (diC16:0PE) from 10:90 to 100:0 (mol/mol) showed that increasing the fraction of LPS lowered the simulated Tm from 73 to 15 °C (250) (for reference, a symmetric bilayer composed solely of DPPE has an experimental Tm of ∼64 °C, showing interleaflet coupling can have hard-to-predict consequences on phase behavior) (251). A-MD simulations of a model of a P. aeruginosa OM containing only lipid A in the outer leaflet and a mix of DPPE (diC16:0PE), DPPG (diC16:0PG), DOPE (diC18:1PE), and DOPG (diC18:1PG) in the inner leaflet showed that the acyl chains of lipid A were also consistent with a liquid phase but were overall less disordered than observed with larger sugar regions—again emphasizing the importance of the polysaccharide region of LPS in modulating its packing behavior (252). Other CG-MD experiments showed that E. coli asymmetric membranes with a lipid A outer leaflet and a DPPE (diC16:0PE) inner leaflet had a lower Tm than a “rough” LPS outer leaflet and DPPE (diC16:0PE) inner leaflet (∼41-46 °C versus ∼55 °C) (253), although it is not clear why the Tm values are much higher in these simulations than observed previously from the same group (250). Preparations of the OM and IM from E. coli J5 doped with spin label probes showed that membrane order is higher, and lipid mobility lower, in the OM, and the magnitude of this difference decreases when a large fraction of LPS is removed (254). Furthermore, in the absence of galactose, this strain is unable to synthesize the full polysaccharide region and produces a short “rough” LPS (Fig. 2) (255). The presence of the O-antigen was shown to confer even greater rigidity to the OM extracts than when they contained only truncated “rough” LPS (254). Deuterium solid-state NMR experiments on multilamellar vesicles containing a mix of partially deuterated DPPC (diC16:0PC) and “rough” LPS from E. coli J5 or E. coli EH100 also found an “ordering” effect on the acyl chains of DPPC conferred by LPS (256). Despite this ordering of acyl chains, experiments with SUVs composed of E. coli B LPS and phospholipid extract doped with spin-labeled PE and PG lipids, and S. typhimurium OM lipid preparations doped with a spin-labeled stearic acid probe found that phospholipids remained freely diffusive and segregated away from LPS (226, 257, 258). The slower diffusion of phospholipids in the inner leaflet of the OM observed in the previous study (which retained its complement of OMPs) (254) may therefore be due to the transient clustering and reduced diffusion of lipids around embedded OMPs (186, 259261). These in vitro studies, although not on fully asymmetric membranes (236), broadly validate the observations of the above in silico studies.

Measurement of water permeation in P. aeruginosa PAO1 LPS (outer)–DPPE (diC16:0PE) (inner), P. aeruginosa LPS (outer)–PPPE (C16:0C16:1PE)/PVPG (C16:0C18:1PG)/CL (inner), and E. coli LPS (outer)–PPPE/PVPG/CL (inner) asymmetric bilayers by A-MD showed that the outer leaflet is relatively permeable to water when compared with the inner leaflet (with water reaching the terminal methyl groups of lipid A acyl chains) with both “rough” LPS and LPS containing an O-antigen region (246, 248,249, 262). This polarity gradient is also apparent through interactions between loop regions of OMPs and charged and polar groups on LPS that have been observed by A-MD and CG-MD (79, 210, 247249, 253, 263274) and specific LPS-binding sites that have been validated experimentally for trimeric porins such as OmpF (6, 189, 275278). One area that remains understudied but warrants further investigation is the effect of the acyl tails of lipoproteins in the inner leaflet of the asymmetric OM on membrane properties. The copy number of lipoproteins at the IM is small enough to, presumably, have a negligible effect on the membrane's bulk physicochemical properties. However, at the OM there may be over 1 million lipoproteins anchored to the inner leaflet, composed primarily of Lpp (Braun's lipoprotein)—one of the most abundant proteins in E. coli—and Pal (3, 4, 279, 280). Each lipoprotein is triacylated at an N-terminal cysteine (121), and each acyl chain can be approximated to occupy ∼0.28 nm2 in a bilayer (100, 262,263, 281285). Assuming a lower bound on the surface area of an average E. coli as 3.7 μm2 and an upper bound of 13 μm2 (286288), and the total area occupied by lipoprotein tails similarly bounded between 0.25 μm2 (3 × 105 proteins) and 1.1 μm2 (1.3 × 106 proteins) (3, 4), we come to an estimate of 2–30% of the inner leaflet of the OM occupied by these lipid anchors. Lpp and Pal both bind peptidoglycan through their protein domains. This interaction restricts their mobility (and therefore the lateral mobility of their lipid anchors) and stiffens the OM, and the protein itself physically occludes the headgroups of the OM inner leaflet lipids (171, 289, 290). The impact that this has on the bilayer properties of the OM requires further study, but the presence of these lipoproteins would act to further reduce the “accessible” surface for OMPs to bind and initiate folding.

Just as early in vitro experiments on synthetic lipid vesicles allowed hypotheses to be generated about the properties of biological membranes, we can look to in vitro experimental models of asymmetric phospholipid membranes to infer information about how asymmetry may affect the OM. These systems remain experimentally challenging due to problems with mixing/scrambling of the inner and outer leaflets in both surface-deposited (291, 292) and liposome-based asymmetric membranes (293), as well as difficulties in accurately controlling the composition of the outer leaflet in asymmetric liposome-based studies (294, 295). Nonetheless, such studies have suggested that asymmetry results in coupling between leaflets, which alters the physical properties of the bilayer distinctly from those of the same lipid types mixed symmetrically. These include changes to the membrane potential difference, lateral pressure differential, and the packing of lipids (296300).

Synthesizing the data described above from in vitro and in silico studies allows insights into the view of the membrane encountered by a nascent OMP as it approaches and is inserted into the OM. As the OMP approaches and moves through the membrane, it begins folding on a 'typical' liquid-disordered bilayer leaflet where phospholipids are free to diffuse before entering a region of low lateral mobility and increased hydration in the outer leaflet. These gradients of lateral mobility, lipid packing, hydration, and lipid headgroup polarity as an OMP inserts across the membrane normal could help to stabilize the tertiary structure of OMPs, particularly the hydrophilic loop regions (which can be >20 residues in length) (301), and drive the folding of the β-barrel domain to completion. Overall, therefore, the above studies have shown that the physical properties of the OM are highly complex and can vary dependent on the underlying lipid phase and the elastic stress, as well as the presence of OMPs and lipoproteins, and the LPR (175, 214, 302). These parameters, in turn, can be tuned by the lipids incorporated into each leaflet, especially by modifications to the LPS in the outer leaflet of the OM. Hence, bacteria have to be adaptable so that they can form and maintain their OM whatever the nature of their environment and the environmental stresses that they encounter.

Membrane-spanning OMPs need to insert through both leaflets of the OM to adopt their native, functional folds, yet how this unique asymmetry and changing lipid content of the OM affects the folding and function of OMPs is currently unknown. We now need to take a deeper look into what we know about OMP folding, starting with the basics and then building more complexity into experimental and theoretical models to allow us to understand how the unusual properties of the OM might influence the process of OMP biogenesis in bacteria.

How do OMPs fold?

30. years of experiments on OMP folding in vitro

Likely due their high thermodynamic stability (ΔG°F = −10 to −140 kJ mol−1) (303312), relatively low hydrophobicity (by all-residue average on the Kyte–Doolittle scale most OMPs are hydrophilic due their alternating hydrophobic membrane facing/generally hydrophilic lumen facing patterning), and ease of recombinant expression and purification, OMPs are unusually tractable models for in vitro studies of membrane protein folding. The first published study on the successful folding of OMPs in vitro was carried out in 1978, when it was shown that SDS-boiled and denatured OmpA (Fig. 6) could be refolded with high yield into LPS (but not into solutions containing total E. coli phospholipids, DMPC (diC14:0PC), or the sugar moiety of LPS) (313). This refolded OmpA was shown to be natively structured because it had regained function (activity in phage receptor binding assays), was protected from protease digestion, and migrated at an anomalous molecular weight in SDS-polyacrylamide gels when loaded without boiling (cold SDS-PAGE) (313). Hence, the stage was set for detailed studies of the mechanisms of OMP folding, at least for this and other small (8-stranded) OMPs.

Figure 6.

Figure 6.

The structure and architecture of OmpA. The 8-stranded β-barrel OmpA has been used for many studies of OMP folding in vitro. It comprises a transmembrane β-barrel domain and a soluble periplasmic peptidoglycan-binding domain. Neighboring β-strands are connected on their extracellular side by a long disordered “loop” and on the periplasmic side by a short “turn.” OMP β-strands are usually numbered from the N terminus (NT) to the C terminus (CT), and the C-terminal β-strand often contains a conserved motif of Gly-X-X-Ar-X-Ar (where Ar represents any aromatic residue), indicated in purple on OmpA, thought to be important for recognition by BAM (386, 399). Many OMPs contain an enrichment in aromatic Trp and Tyr residues in their β-barrel domain, particularly at the interfacial region between the lipid headgroups (approximate position indicated by the gray line) and acyl tails (approximate position indicated by the gray box) known as an “aromatic girdle.” Trp (red) and Tyr (blue) residues found in the β-barrel domain of OmpA are indicated above. This model of OmpA was created in PyMol 2.X (Schrödinger, LLC) by fusing the NMR structures of the E. coli OmpA β-barrel (PDB code 1G90) (400) and its periplasmic domain (PDB code 2MQE) (401).

The next major breakthroughs in our understanding of OMP folding were made in the 1990s, when it was shown that OmpA and an unnamed porin could be spontaneously refolded in vitro into detergent (octyl glucoside), small unilamellar vesicles (SUVs) of DMPC (diC14:0PC), or mixed lipid-detergent micelles (octyl-poly-oxyethyleneoxide (C8POE) and soybean lecithin) in the presence of SDS without the addition of an external energy source or other proteins (314316). Soon after, several groups began to experiment with the refolding conditions to identify the determinants that allow OMPs to attain their native structure. These studies showed that detergents are required to be above their critical micelle concentration for successful OMP folding (317, 318). This highlighted the importance of a surface to initiate folding and showed that simple binding of hydrophobic molecules around an unfolded OMP chain is insufficient to enable folding. Instead, some degree of preorganization is required. These studies also showed that folding of OMPs in vitro is a remarkably slow process (most OMPs taking on the order of minutes to hours to fold), too slow to be physiologically relevant (the doubling time of E. coli is ∼20 min at 37 °C). They also showed that refolding yields are improved by the addition of urea, implying that the OMP-folding landscape contains kinetic traps and/or off-pathway intermediates or aggregates that can be suppressed by the addition of chaotrope (317, 319). These experiments also showed that the β-barrel transmembrane fold is extremely stable once formed, with native OMPs being resilient to denaturation by SDS (unless heated to high temperatures), enabling kinetic measurement of folding using cold SDS-PAGE (71, 313, 320). This is due to the high kinetic barrier to unfolding in SDS (on the order of years at 30 °C for OmpA) (314, 321).

Due to their high kinetic and thermodynamic stability, high concentrations of denaturant often fail to unfold OMPs that have been refolded into detergent micelles or lipid bilayers. For example, OmPLA remains enzymatically active in 8 m urea and in 6 m guanidine-HCl (317). Experiments measuring the activity of OmPLA following refolding into LPS and a range of detergents showed not only that the protein can reacquire a stable fold, but also that it regained its phospholipase activity, confirming reversible folding to a functional state (317). Thus, OMPs obey Anfinsen's dogma that all of the information for a protein to reach its thermodynamically stable native structure is contained in its amino acid sequence (322).

First glimpses of an OMP-folding mechanism

The next era of work on OMP folding focused on attempts to determine the mechanisms of folding, including why the process is so slow, whether partially folded intermediates are formed, the nature of folding transition states, and the role of the protein sequence, lipid composition, and folding factors (including BAM and molecular chaperones) in the folding process. In some of the earliest studies of this kind, Surrey et al. (316) measured the kinetics of OmpA folding into SUVs formed from 95% DMPC (diC14:0PC), 5% DMPG (diC14:0PG) (mol/mol) at a temperature above the lipid Tm by rapid dilution from urea in vitro. The techniques employed and optimized in these early studies of OMP folding form the toolset used for such studies to this day (323). They also reinforced the idea garnered from studies on the folding of water-soluble proteins (324327) that insights into the mechanisms of folding can be best learned by taking a kinetic approach to dissect each step in the folding pathway. These methods include (i) monitoring the change in tryptophan (and/or tyrosine) fluorescence as an OMP folds from aqueous solution into a nonpolar membrane (Fig. 6 shows the location of aromatic residues on OmpA); (ii) measuring the formation of secondary structure using far-UV CD, and (iii) following the formation of SDS-resistant molecules (presumably containing a correctly folded β-barrel domain) using cold SDS-PAGE (316, 328). This work showed that OmpA folds into lipid bilayers in vitro via a multiphasic mechanism involving rapid hydrophobic collapse that occurs in the experimental dead time (∼1 s), followed by two slower phases occurring in minutes (phase 1) and tens of minutes (phase 2) corresponding to structural rearrangement and concurrent formation of secondary and tertiary structure (Fig. 7A) (316). It was also shown that pH affects the yield and rate of OMP folding, with pH values close to the pI of OmpA (pI ∼5.7) increasing the folding rate but decreasing the folding yield. The average pI of the OMPs shown in Fig. 1 is ∼5.4, and the environment in which E. coli has evolved (the large colon) is mildly acidic (pH 5.5-7.5) (329), suggesting that E. coli has adapted to favor rapid folding and has systems to handle or circumvent the lower yields. Higher LPRs were also shown to increase the folding rate and yield, consistent with models suggesting that the OMP first binds to the surface of the lipid bilayer before folding and insertion are completed (316, 330332). Indeed, refolding studies of OmpA into lipids above their Tm confirmed this membrane-binding step with elegant experiments using membranes composed of lipids that had been brominated at different positions along the acyl chain (bromination quenches fluorescence from tryptophan only when in close proximity, thereby allowing depth-dependent changes in fluorescence) (333, 334). Earlier data from ATR-FTIR studies showed that this membrane-adsorbed (but not stably integrated) folding intermediate had significant β-sheet content, with β-strands that have still to adopt the orientation found in the native β-barrel (335). Finally, the finding that tryptophan residues located in β-strands, near extracellular loops, or in intracellular turns show similar rates of membrane insertion (judged by tryptophan fluorescence quenching) suggested that folding occurs via a concerted mechanism in which all four hairpins of this 8-stranded OMP move synchronously across the bilayer (330), increasing in tilt angle as the protein becomes fully membrane embedded (335). More recent experiments employing single-Cys single-Trp variants of OmpA labeled at the Cys with Trp-quenching nitroxyl spin labels have shown that despite being concerted, OmpA's β-strands form in a particular order (336), with β-strands initially associating with each other at their “extracellular” sides like a tipi (i.e. before the “periplasmic” side forms) (Fig. 7A, step 4) and showing that the N- and C-terminal strands are already in close proximity in the membrane-adsorbed state (Fig. 7A, step 3), providing evidence against models invoking β-barrel closure as the last step (336). These experiments also failed to detect native β-strand association in aqueous solution, ruling out models in which the OmpA β-barrel “pre-folds” in solution before inserting as a single unit into the membrane. This complex, multistep process where formation of secondary and tertiary structure is coordinated contrasts markedly with the two-stage mechanism proposed for α-helical membrane proteins in which the formation of α-helices precedes formation of the native tertiary structure (337, 338). The overall general model for spontaneous OMP folding in vitro is summarized in Fig. 7A.

Figure 7.

Figure 7.

Mechanisms of OMP folding in vitro and in vivo. A, Mechanism of spontaneous OMP folding in vitro as described for OmpA. In vitro studies have shed light on the folding pathway of model OMPs, particularly OmpA, after dilution out of high concentrations of denaturant in the presence of a lipid bilayer. Adapted from Danoff and Fleming (351). This research was originally published in Biochemistry. Danoff, E. J., and Fleming, K. G. Membrane defects accelerate outer membrane β-barrel protein folding. Biochemistry 2015; 54:97–99. © American Chemical Society. 1, immediately after dilution out of denaturant, the chain undergoes hydrophobic collapse; 2, the polypeptide chain then binds to the surface of a membrane; 3, the nascent OMP then begins to form secondary structure as it brings together neighboring β-strands to form β-hairpins while still mostly exposed to the aqueous environment; 4, these β-hairpins associate and begin to insert into the acyl tail region of the membrane; 5, the tertiary structure of the barrel is complete with the final step likely being a slower equilibration of side chains and extrusion of hydrophilic loops from the barrel lumen. B, proposed mechanisms of BAM-catalyzed folding of OMPs in vivo. The nascent OMP is shown in red or purple, BamA is shown in green, and BamD is shown in yellow (other subunits have been omitted for clarity). 1, BamA-assisted. Substrate OMPs are delivered to BAM or directly to the membrane by periplasmic chaperones. These nascent OMPs then fold spontaneously into a region of destabilized membrane in front of the lateral gate of BamA, essentially following the same pathway as described in A. 2, BamA-budding. Binding/recognition of the OMP occurs on BAM before β-strands are added a β-hairpin at a time between β1 and β16 of BamA, forming a semisymmetric hybrid-barrel intermediate. Once all β-hairpins are added, this folded OMP then buds off from BamA. 3, BamA-swing/elongation. Binding/recognition of the OMP occurs on BAM and folding starts with templating of the C-terminal β-strand of a nascent OMP against β1 of BamA. Folding proceeds in the periplasm through the stepwise addition of more β-strands. Once all β-strands have been added, a conformational change in BamA “swings” the folded β-barrel into the membrane. 4, BamA lumen–catalyzed. This model begins as described in 3 with templating against BamA β1. However, formation of further β-strands is catalyzed against the lumen wall of BamA with a conserved motif in loop 6 of BamA (not shown) possibly stabilizing this interaction. In all of these models, BamD (yellow) may play an important role in substrate recognition and/or the conformational cycle.

The role of the membrane in OMP folding

Studies varying the lipid headgroup, acyl chain length, lipid phase, membrane curvature, and elastic curvature stress (the change in lateral pressure across the membrane) (Fig. 5) have shown that the properties of the membrane can have dramatic effects on the folding of OMPs. For example, whereas OMPs such as OmpA and PagP (both 8-stranded; Fig. 1) are able to fold into highly curved SUVs comprised of DOPC (diC18:1PC) or DMPC (diC14:0PC) in minutes, folding is slower or effectively prevented (beyond the timescale employed in the study) when these proteins are folding into LUVs formed from DOPC (diC18:1PC) under the same experimental conditions (40). Similar studies have recapitulated these results and shown that the elastic curvature stress, hydrophobic mismatch (hydrophobic thickness of the OMP versus the bilayer), and membrane curvature (SUV versus LUV) affect OMP folding, emphasizing the importance of the physical properties of the lipid environment in determining the folding process (339, 340). The steep decrease in refolding rates observed as the acyl chain length of PC lipids increases (330) and the absence of reports of spontaneous folding into LUVs formed from pure PC lipids with saturated acyl chain lengths >14 carbon units (i.e. >DMPC (diC14:0PC)) can be rationalized by the elastic free energy of the membrane, as this parameter is expected to increase to the fourth power of membrane thickness (341). Membrane thickness is also thought to relate to the incidence of packing defects and thermal fluctuations in the bilayer, with thinner membranes having more defects. Such packing defects may be responsible for the more rapid folding of OMPs into thinner bilayers (341). OMPs have also been shown to be unable to fold into some lipids, such as DMPC (diC14:0PC) and DPPC (diC16:0PC), in the gel phase, but they fold readily when the same lipids are in the liquid-disordered phase, suggesting a significant activation energy barrier for the protein to insert across the bilayer (333, 334). Accordingly, the folding rate and yield of tOmpA (the β-barrel domain of OmpA) and OmpX (8-stranded; Fig. 1) into micelles (or mixed micelles of different lipid/detergent types) can be increased 10−100-fold by applying a transient “heat shock” (∼70 °C) during folding (342). This heat shock presumably confers enough thermal energy to rapidly take the unfolded ensemble over the activation energy barrier. It has also been shown that whereas gel-phase lipids (below their Tm) prevent or retard folding, and more rapid folding is observed into lipids in their liquid-disordered phase, folding is most rapid at the interface between these phases (i.e. at the Tm) (343). At this temperature, regions of gel-phase (ordered acyl chains) and liquid-phase (disordered acyl chains) lipids coexist (344, 345), and frustration between the packing requirements of these phases is believed to generate packing defects (as shown by an increase in solute permeability) (346350), which may be responsible for the acceleration of OMP folding. Kinetic modeling experiments have suggested that this phenomenon involves an acceleration in the formation of an early membrane-inserted folding intermediate (351).

Together, these data have established the importance of the chemical and physical properties of the membrane environment (lipid headgroup, acyl chain length, size, and curvature of the membrane vesicle) and the intimate connection between the protein sequence and membrane into which the OMP must fold.

Commonalities in the folding mechanisms of OMPs in vitro

OMP-folding intermediates formed by initial adsorption to the lipid bilayer surface have been observed for several small OMPs, including the 8-stranded OmpA (see above) and PagP (332). An extensive mutational study of the folding kinetics and equilibrium stability of single point mutants of PagP revealed the first evidence for the nature of an OMP-folding transition state (in DLPC (diC12:0PC) LUVs), revealing that this species, which is formed subsequent to initial folding on the membrane surface, contains a partially folded β-barrel, in which the C-terminal strands have formed native-like contacts, but the N-terminal strands remain largely unstructured (352). The results from this study also suggested that PagP is tilted in the membrane in the transition state ensemble, reminiscent of CG-MD studies observing the insertion of preformed OmpA into a DPPC (diC16:0PC) bilayer (353). Kinetic analysis of PagP folding also showed that this OMP folds via parallel pathways, with the route taken depending on the nature of the lipid employed (332). Similar results showing parallel folding pathways were also obtained for FomA (331), yet again highlighting the importance of the membrane surface and the physical state of the lipid bilayer for OMPs to fold (331, 332).

The view of OMP folding in vitro that emerges from these studies shows a common first step involving folding on the lipid surface followed by preorganization of the approximate structure of the OMP, with translocating aromatic residues anchoring the nascent β-barrel to the membrane, and association of neighboring β-strands causing a hydrophobic surface to be displayed toward the membrane. This would, in turn, drive the energetically favorable partitioning of the hydrophobic protein surface deeper into the acyl chains (354). At the same time, hydrogen bonding between β-strands during insertion may be driven by the energetic penalty associated with displaying the unbonded polar peptide backbone to the nonpolar membrane environment. Finally, the association of β-strands in the correct order may be a rate-limiting step for insertion into the bilayer.

BAM: Nature's answer to the challenges of OMP folding

Whereas many OMPs can be folded in vitro into SUVs and LUVs composed of short chain lipids, attempts to fold these proteins into liposomes comprised of E. coli polar lipid extract result in moderate folding yields for some OMPs (OmpA (8 strands), OmpT (10 strands), and BamA (16 strands)) and poor or no folding for others (OmpX (8 strands), PagP (8 strands), OmpW (8 strands), OmPLA (12 strands), and FadL (14 strands)) (Fig. 1) (40, 41). Spontaneous folding of OMPs of many sizes has been observed into DDPC (diC10:0PC) LUVs (40) but has been shown to be suppressed upon incorporation of lipids with headgroups native to the E. coli OM that confer negative spontaneous curvature (PE) or net negative charge (PG) (41). BamA is a 16-stranded essential OMP (Fig. 1) and is the most conserved subunit of the multiprotein BAM complex, which is involved in the biogenesis of other OMPs. BamA or a BamA variant lacking most (4 of 5) of its periplasmic polypeptide transport-associated (POTRA) domains can partially rescue the poor folding efficiency of OMPs into these liposomes (41). These results highlight the vital role of BAM for OMP folding into lipids commensurate with those found in vivo, suggesting that spontaneous folding of OMPs into a native lipid bilayer in vivo is kinetically unfavorable unless BAM is present.

BAM as a modulator of the physical properties of membranes

Until recently, how BAM folds OMPs remained mysterious. The solution of several structures of BamA and BAM, as well as kinetic experiments, are now starting to provide glimpses of how this amazing machinery functions and how it might involve remodeling of the lipid bilayer. MD simulations of BamA and the full BAM complex have shown that the presence of the 16-stranded BamA β-barrel causes thinning and disordering of the membrane in the vicinity of its β1–β16 seam and that BamA can switch between conformations at this interface (355360) (Fig. 8). Other OMPs have been shown to generate variable (i.e. anisotropic) membrane thickness around their circumference experimentally (e.g. BtuB, 22-stranded) (361), and more isotropic bilayer thinning is seen in A-MD and CG-MD simulations in the vicinity of many OMPs (OprH (8-stranded from P. aeruginosa), OmpA, LpxR (12-stranded from S. typhimurium), Hia (3 × 4 strands from Hemophilus influenzae), OmPLA, NanC, OmpF, LamB, and FhuA) (186, 209, 248, 249, 267, 274, 359, 362). Although the dynamics of lipids surrounding BamA have not yet been studied in vitro or in vivo (e.g. by using spin-labeled lipids), experiments examining the lipids directly surrounding other OMPs have found them to be motionally restricted in the vicinity of the β-barrel, as seen for FomA (8-stranded from Fusobacterium nucleatum), OmpA, OmpG, and FhuA (363365). This has also been observed in CG-MD simulations of OmpA, NanC, OmpF, LamB, and FhuA in a bilayer formed from 75% POPE (C16:0C18:1PE), 25% POPG (C16:0C18:1PG) (186). It should be noted, however, that it is not certain that the two parameters of lipid order and motional restriction are always correlated—consider a situation where disordered lipids are corralled by ordered lipids. With current data, it is unclear whether the lipid disordering observed around the β1–β16 seam of BamA is purely a consequence of membrane thinning (i.e. a large hydrophobic mismatch) or is accentuated by some other mechanism in this OMP. Evidence for such a reaction cycle that could support a membrane-remodeling mechanism for BamA was suggested from crystal structures of BamA in isolation and from crystal and cryo-EM structures of BamA in the BAM complex. These studies showed that the BamA β-barrel is shorter in the β1–β16 region than in the rest of the barrel and that it can explore at least three distinct and potentially membrane-influencing conformations: a closed fully zipped (i.e. hydrogen-bonded along its whole seam) barrel (Fig. 8, closed no kink and zipped), a closed partially zipped barrel (Fig. 8, closed kink), and an open barrel (Fig. 8, open kinked) (42, 91, 355, 357, 366371). This open conformation was surprising, as it would intuitively seem highly energetically unfavorable to break hydrogen bonds in a hydrophobic environment. However, WT BamA is only capable of forming at most 6-8 backbone hydrogen bonds between its first and last β-strands (Fig. 8, closed no kink). MD simulations showed that this fully hydrogen-bonded conformation is unstable and eventually forms the “partially zipped/closed kink” conformation seen in structures of the full BAM complex (Fig. 8, closed kink) where the terminal strand bends back into the barrel lumen, leaving just 2–3 hydrogen bonds (360). This partially zipped state may lower the energetic cost of fully opening the BamA barrel (360), leaving at most one hydrogen bond between β1 and the periplasmic turn between β15 and β16. This periplasmic turn is likely to be important in stabilizing this open state, as its mutation or truncation is lethal in vivo (370). This still requires the breaking of 5–7 hydrogen bonds, but some of this may be compensated by the kinked residues hydrogen-bonding to water within the lumen of the BamA barrel. Furthermore, the cost for OMPs to break hydrogen bonds in the membrane may be lower than previously assumed (372), and the outer leaflet of the OM may be more hydrated than a symmetric phospholipid bilayer. These effects could help stabilize the structure of BamA while effecting local changes in packing or stability of the membrane. The importance of this BamA barrel opening was highlighted by the lethality of disulfide bonds engineered to lock the barrel closed (356) or open (357). Interestingly, folding of OmpT and OmpX via BamA or the full BAM complex in vitro is still catalyzed in the locked closed state (42, 373). This implies that the structure of BamA alone, with its reduced hydrophobic thickness around β1–β16, may intrinsically accelerate OMP folding by distorting the membrane—even in the absence of the open state. In vitro, BamA has been shown to have a greater catalytic effect on tOmpA folding (higher catalytic fold rate enhancement) as the hydrophobic thickness of a bilayer is increased from ∼19.5 Å in DLPC (diC12:0PC) LUVs to ∼23.0 Å in DMPC (diC14:0PC) LUVs, showing that hydrophobic mismatch and/or lipid disordering plays an important role in the mechanism of BamA-assisted folding (359). These data illustrate how both structural and biochemical approaches will be required to fully understand the function of the BAM complex.

Figure 8.

Figure 8.

Conformations of the BamA lipid-facing lateral gate. Shown are example structures of E. coli BamA adopting different conformations around the location of the β1–β16 seam/gate. BamA has been observed in both gate open and closed states with the open state observed in the presence of other BAM subunits but not in structures of BamA in isolation. Furthermore, in all structures of the full BAM complex, β16 of BamA adopts a kinked conformation at a highly conserved glycine (Gly807) in both the open (PDB code 5EKQ; BamACDE) (368) and closed (PDB code 5D0O; BamABCDE) (357) states of the gate. Residues Asn427–Gly433 (β1) are highlighted in light blue, residues Phe802–Trp810 (β16) are indicated in orange, and the kink is further highlighted with spheres (Ile806–Trp810). This kink is also observed in structures of BamA from Salmonella enterica (PDB code 5OR1) (370) and Neisseria gonorrhoeae (PDB code 4K3B) (355) (not shown) and in the BamA homologue, TamA, which also plays a role in OMP assembly (PDB codes 4N74 and 4C00) (387) (not shown). BamA with a closed gate and no kink has been observed in isolation (PDB code 4N75; BamAΔ1–427) (366) and in a hybrid BamA containing a C-terminal 9-residue extension (colored yellow) comprised of part of turn 3 and β7 from OmpX, which may represent a mimic of an OMP-BamA folding intermediate (PDB code 6FSU) (371). Structures are represented in an asymmetric bilayer with a mixture of phospholipids with 14-18 carbon acyl chains (shown in violet) in the inner leaflet and E. coli rough LPS in the outer leaflet (acyl chains in white). Note the different hydrophobic thickness between each leaflet. Asymmetric bilayer was built using the GNOMM server (402).

Recent studies have also shown the importance of BAM and membrane fluidity in folding OMPs in vivo. These studies exploited a mAb that was found to be bactericidal, binding to extracellular loop 6 of BamA (93). Interestingly, bacteria showing spontaneous resistance to this BAM-mediated toxicity were found to have mutations in the lpxM gene. This protein transfers a C14 (myristoyl) chain to penta-acylated LPS, creating hexa-acylated LPS (93). Antibody sensitivity was restored (i.e. bactericidal effects of the antibody were reinstated) when lpxM was expressed from a plasmid. Assays of membrane fluidity using a pyrene-based probe showed that membrane fluidity decreased in the resistant strains, and this effect was recapitulated in other conditions that decrease membrane fluidity (high salt, longer LPS sugar region, lower temperature) (80, 93). The levels of OMPs were not reduced in ΔlpxM strains in the absence of the antibody. This suggests that there is a mechanistic link between BamA/BAM and membrane fluidity, as BAM is most sensitive to inhibition when the membrane is more fluid. Hence BamA activity (as part of the BAM complex) may be lower when the OM is excessively fluid, confirming a link between OMP-folding efficiency, BAM activity, and membrane fluidity in vivo. More generally, it is interesting to consider the role of the lipids that directly surround BAM and how they might interact with it. Most MD studies on BamA have utilized simple lipid mixtures (355360). However, it is known that membrane proteins can enrich lipids or other proteins around them that match their hydrophobic thickness (374376). In the context of BAM, this may have two consequences. First, shorter-chain lipids may be enriched around the β1–β16 seam to reduce the energetic penalty of membrane thinning, with C12:0 and C14:0 acyl tails found in E. coli and also reported to be more abundant in the OM than the IM (139). Second, the hydrophobic thickness of substrate OMPs is generally greater than that of BamA. This means immediately after insertion into the region surrounding the BamA β1–β16 seam, the newly folded or folding OMPs will encounter a region of positive hydrophobic mismatch (Fig. 5), and it may be energetically favorable for them to diffuse away to regions of greater hydrophobic thickness with a more optimal match to their size (376). This could provide a mechanism for release and local clearance of newly folded OMPs from BAM. Furthermore, although substrate binding to BAM subunits or the POTRA domains of BamA is likely to be important in conformational cycling (60, 377382), the lateral pressure or fluidity (Fig. 5) within the OM might also play a role in controlling opening and closing of the BamA barrel. These dynamics are thought to be essential for its role in catalyzing OMP folding in vivo (356, 383), so modulation of the BamA barrel dynamics by the lipid environment may also provide a secondary mechanism of controlling the function of BAM.

A multifaceted mechanism for the BAM complex

Growing evidence points to a mechanism of BAM function in catalyzing OMP folding and membrane insertion that involves the templating of C-terminal β-strands of the nascent substrate OMPs onto the β1 strand of BamA (Figs. 68) (47, 61, 384). The C-terminal strands of OMPs in bacteria and mitochondria contain a conserved aromatic-rich motif termed the β-signal (Fig. 6), which may be important in the recognition of OMP clients by the BAM complex (385, 386). The initial recognition of the OMP's C-terminal strand is thought to then trigger nucleation of further β-hairpins to the growing β-barrel, perhaps favored by the reduced entropic penalty of folding (as the degrees of freedom of the unfolded chain are reduced). However, the steps following this initial binding event, particularly how the β-barrel of the newly forming OMP folds and inserts into the membrane, remain unclear. A role for the lateral gate of BamA in this process was proposed after the first structures of this family of proteins were solved (355, 387, 388). Further experimental evidence suggests four possible models for how this is achieved (Fig. 7B). 1) BamA plays a passive role in OMP folding/insertion and merely targets nascent OMPs to a destabilized bilayer in front of its β1–β16 seam. Substrate binding possibly initiates a conformational change in BamA, increasing its “lipid disorderase” activity, but folding otherwise proceeds as in the in vitro pathway (Fig. 7A) (the assisted model) (388). 2) The β-barrel grows laterally into the membrane after templating onto the BamA gate (the budding model) (384, 388). 3) Prefolding/elongation of the OMP occurs in the periplasm after binding to β1 of BamA. A conformational change in BAM then inserts the already-folded β-barrel (the swing/elongation models) (47, 389). 4) Substrates fold against the interior wall of BamA while keeping their N and C termini in close proximity ready for β-barrel closure and release into the membrane (the lumen-catalyzed model) (61). The last two models are particularly intriguing as they suggest that the hydrophobic surface of the folding OMP is partially exposed to an aqueous or polar environment. How this step could be energetically favorable is not yet clear, but some authors have proposed that the cradle created by the BamA lumen and POTRA domains may aid folding by acting like an entropic cage, analogous to chaperonins such as GroEL/ES (61, 389). This may drive folding by reducing the conformational entropy of the unfolded state but may also restrict the mobility of water, perhaps offsetting some of the cost of exposing hydrophobic residues.

Through its mechanism of β-strand capture, the BAM complex may also act to suppress the reversible off-pathway intermediates in OMP folding that cause kinetic retardation of folding in vitro and which may be related to both misfolded monomeric intermediates and aberrant transient intermolecular interactions (351). Furthermore, these aberrant multimers (so called “elusive” states as they are not easily observed by SDS-PAGE) are not observed in the absence of a lipid bilayer (351), implicating the OMP-membrane interaction as an important control point where these states could form. The nature of the terminal stages of folding, including how the OMP is able to partition rapidly into the OM and how it overcomes the activation energy barrier associated with membrane insertion, remain unresolved. Finally, the enrichment of shorter-chain phospholipids and depletion of OMPs around BamA proposed above would essentially “clear some space” for folding to proceed (e.g. see Fig. 3C), providing a mechanism that could overcome the remarkably low LPR of the OM and an inner leaflet crowded with OMPs and lipoproteins. The proposed formation of “supercomplexes” that span the periplasm linking the IM SecYEG translocon and the BAM complex at the OM could also provide a direct conduit for OMP biogenesis (390392). If only limited patches of free lipid exist, it would be important to direct OMPs to these regions of the OM before they misfold or aggregate.

Comparison of folding in vitro and in vivo

Our current knowledge about the physical constraints of OMP folding into lipid bilayers, the properties of the OM, the folding of OMPs in vitro, and the mechanism of BAM action all point to the membrane as an important interface for preorganization of OMPs into an insertion-competent state. Despite clear evidence for the importance of this early folding step, the unusually low LPR of the OM of E. coli (which may be as low as 6–14:1) (139, 156, 157) calls into question the direct interaction of the OMP and the OM during folding in vivo. Given this dearth of lipid surface and the observation that the rate of OMP folding falls dramatically with decreasing LPR (316, 331, 332), how do OMPs insert into the OM on a biologically meaningful timescale? Large lipid-rich patches could exist in the OM, but in vivo imaging studies have shown that BAM and newly inserted OMPs appear together in protein-rich clusters (56, 149, 181). This means that there are either very few free lipids (Fig. 3A) or only small, local, lipid-enriched domains (Fig. 3C). Hence, BAM may have evolved to provide an initial nucleation site for OMP folding in a generally lipid-poor environment and to accelerate this process, particularly for rapidly growing organisms such as E. coli. This would also be necessary to pack the OM to the high protein density that is observed in vivo. Although the recognition of nascent OMPs and the nucleation of β-barrel formation through templating/recognition of the newly folding OMP's β-strands on BamA would provide a rate enhancement in folding, the dramatic differences of folding rates into detergent micelles (fast) versus LUVs (slow) suggest that the ability of an OMP to navigate its folding pathway to find the native fold is not necessarily a major rate-limiting factor. Instead, it appears that membrane insertion and disruption of lipid packing could play a more significant role in controlling the rate of folding. In vitro experiments have shown that the greater the lateral pressure, lipid order, and packing of lipids within a bilayer, the greater the activation energy barrier for folding (41, 339, 340). However, local or transient defects in the packing of lipids in a membrane can allow OMPs to bypass a slower folding pathway and instead fold more rapidly (as shown by studies on the propensity of thinner bilayers and lipids at their Tm to accelerate OMP folding) (40, 343). The physical structure of the OM remains poorly defined, but what we know is this: already-folded OMPs are resistant to deformation in the OM, LPS can rigidify the outer leaflet, and in vivo measurements suggest that at least some proportion of lipids in the OM are relatively ordered and less mobile. BAM may introduce local defects into this environment (by thinning of the membrane and causing local disorder of the lipids) to lower the energy barrier to insertion and thus accelerate OMP folding into this otherwise impenetrable membrane barrier.

Concluding remarks and open questions

The OMP-folding problem in vivo can be simplified as the interplay between three factors: enzyme (BAM), substrate (OMP), and solvent (lipid). To fully understand OMP folding in vivo, it is necessary to characterize the relationships between these three dominating factors to understand their holistic function. BAM acts to catalyze the folding of OMPs into a lipid bilayer, and it may do this by nucleating the folding of the unfolded OMP substrate (BAM-OMP interaction) and/or by locally disrupting the packing of lipids (BAM-lipid interaction) to lower the activation energy of folding (OMP-lipid interaction). The evidence for the action of BAM on lipids is currently restricted to in silico experiments, although we can imply its importance from in vitro studies of the effect of the physical and chemical properties of the bilayer on OMP folding. More studies are now needed to determine the relative importance of this “disorderase” activity of BAM in ensuring that OMPs can gain access to the highly crowded and lipid-poor OM.

Our current lack of understanding of the physical properties of the OM prevents the generation of accurate models for the mechanism(s) of BAM in the OM, the effect of BAM on the structure and dynamics of the OM lipids (and vice versa), and how these changes impact OMP folding. Although levels of OMPs fall upon depletion of BAM in vivo, and OMP insertion kinetics are slowed in the absence of BAM in vitro, which OMPs require BAM for folding into the OM in vivo is still not known unequivocally. Indeed, some OMPs have been shown to be capable of folding without the aid of BAM, and others are aided by other proteins such as those of the localization of lipoproteins (Lol) machinery (Table 1). The asymmetry of the OM, combined with the mixture of lipid types, high fraction of proteins, and tethering to the peptidoglycan layer, makes the journey taken by a nascent OMP to the OM and its insertion in vivo an immensely challenging task to replicate in vitro. Future in vitro work should focus on understanding the OM and OMP folding in a context closer to that found in vivo and also on better understanding the effects of suppressor variants on the catalytic activity of BAM, so as to narrow the gap between in vivo and in vitro insights. At the same time, modern biophysical and biochemical tools are also needed to make these same measurements directly on OMPs in their native context within bacteria. En route to this, the use of outer membrane vesicles directly derived from bacteria may provide an excellent stepping stone (393395).

In summary, despite enormous progress in our understanding of how OMPs fold in vitro and in dissecting the interactions and mechanisms of BAM, we still do not fully understand how BAM folds OMPs of different size and sequence; nor do we understand fully the role of the asymmetric OM lipid bilayer in defining this process. OMP folding has moved in the last few years from a question of fundamental importance and interest to one having direct implications for the development of new antibiotics targeting OM biogenesis. The challenge has thus been set to define how OMPs fold in vivo and to solve the remaining mysteries, so that BAM can be targeted to break the OM barrier and render pathogenic bacteria susceptible to attack by new antimicrobial agents.

Acknowledgments

We thank members of the S. E. R. and D. J. B. laboratories and our collaborators for many helpful discussions, particularly the OMP team, including Bob Schiffrin, Anna Higgins, Paul White, Anton Calabrese, James Whitehouse, Sam Haysom, Matt Iadanza, Matt Watson, Roman Tuma, and Neil Ranson.

Note added in proof

After this manuscript was accepted for publication, a cryo-EM structure of a substrate-engaged BAM complex purified after stalling in vivo was reported supported by in vivo cross-linking. This structure shows a late-stage assembly intermediate and, along with the cross-linking data, appears to support model 3 (BamA-swing/elongation) or 4 (BamA lumen-catalyzed) of BAM-catalyzed OMP folding shown in Fig. 7 (403).

This article contains supporting information.

Funding and additional information—J. E. H. was supported by Biological and Biotechnology Research Council (BBSRC) Grant BB/M01151/1. Work on BAM and OMPs by S. E. R. and D. J. B. was supported by BBSRC Grants BB/K000659/1, BB/T000635/1, and BB/N007603/1 and Medical Research Council Grant MR/P018491/1.

Conflict of interestThe authors declare that they have no conflicts of interest with the contents of this article.

Abbreviations—The abbreviations used are:
IM
inner membrane
A-MD
atomistic molecular dynamics
BAM
β-barrel assembly infection with five pathogenic microbes whichmachinery
CG-MD
coarse-grained molecular dynamics
CL
cardiolipin
DDPC
1,2-didecanoyl-sn-glycero-3-phosphocholine
DLPC
1,2-dilauroyl-sn-glycero-3-phosphocholine
DMPC
1,2-dimyristoyl-sn-of C. elegansglycero-3-phosphocholine
DMPG
1,2-dimyristoyl-sn-glycero-3-phosphoglycerol
DOPC
1,2-dioleoyl-sn-glycero-3-phosphocholine
DOPE
1,2-dioleoyl-sn-glycero-3-phosphoethanolamine
DOPG
1,2-dioleoyl-sn-glycero-3-phosphoglycerol
DPPC
1,2-dipalmitoyl-sn-glycero-3-phosphocholine
DPPE
1,2-dipalmitoyl-sn-glycero-3-phosphoethanolamine
DPPG
1,2-dipalmitoyl-sn-glycero-3-phosphoglycerol
IMP
inner membrane protein
LPR
lipid-to-protein ratio
LPS
lipopolysaccharide
LUV
large unilamellar vesicle
MD
molecular dynamics
OM
outer membrane
OMP
outer membrane protein
PC
phosphatidylcholine
PE
phosphatidylethanolamine
PG
phosphatidylglycerol
POPC
1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine
POPE
1-palmitoyl-2-oleoyl-sn-glycero-3-phosphoethanolamine
POPG
1-palmitoyl-2-oleoyl-sn-glycero-3-phosphoglycerol
POTRA domain
polypeptide transport–associated domain
PPPE
1-palmitoyl-2-palmitoleoyl-sn-glycero-3-phosphoethanolamine
PVPG
1-palmitoyl-2-vacenoyl-sn-glycero-3-phosphoglycerol
SUV
small unilamellar vesicle
Tm
transition temperature
tOmpA
the β-barrel domain of OmpA
cP
centipoise
AFM
atomic force microscopy
ATR
attenuated total reflection
PDB
Protein Data Bank.

References

  • 1. Prilipov A., Phale P. S., Koebnik R., Widmer C., and Rosenbusch J. P. (1998) Identification and characterization of two quiescent porin genes, nmpC and ompN, in Escherichia coli BE. J. Bacteriol. 180, 3388–3392 10.1128/JB.180.13.3388-3392.1998 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2. Fàbrega A., Rosner J. L., Martin R. G., Solé M., and Vila J. (2012) SoxS-dependent coregulation of ompN and ydbK in a multidrug-resistant Escherichia coli strain. FEMS Microbiol. Lett. 332, 61–67 10.1111/j.1574-6968.2012.02577.x [DOI] [PubMed] [Google Scholar]
  • 3. Li G.-W., Burkhardt D., Gross C., and Weissman J. S. (2014) Quantifying absolute protein synthesis rates reveals principles underlying allocation of cellular resources. Cell 157, 624–635 10.1016/j.cell.2014.02.033 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4. Soufi B., Krug K., Harst A., and Macek B. (2015) Characterization of the E. coli proteome and its modifications during growth and ethanol stress. Front. Microbiol. 6, 103 10.3389/fmicb.2015.00103 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5. Henning U., Höhn B., and Sonntag I. (1973) Cell envelope and shape of Escherichia coli K12: the ghost membrane. Eur. J. Biochem. 39, 27–36 10.1111/j.1432-1033.1973.tb03099.x [DOI] [PubMed] [Google Scholar]
  • 6. Rosenbusch J. P. (1974) Characterization of the major envelope protein from Escherichia coli. Regular arrangement on the peptidoglycan and unusual dodecyl sulfate binding. J. Biol. Chem. 249, 8019–8029 [PubMed] [Google Scholar]
  • 7. Lugtenberg B., and Van Alphen L. (1983) Molecular architecture and functioning of the outer membrane of Escherichia coli and other Gram-negative bacteria. Biochim. Biophys. Acta 737, 51–115 10.1016/0304-4157(83)90014-X [DOI] [PubMed] [Google Scholar]
  • 8. Ye J., and van den Berg B. (2004) Crystal structure of the bacterial nucleoside transporter Tsx. EMBO J. 23, 3187–3195 10.1038/sj.emboj.7600330 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9. Subbarao G. V., and van den Berg B. (2006) Crystal structure of the monomeric porin OmpG. J. Mol. Biol. 360, 750–759 10.1016/j.jmb.2006.05.045 [DOI] [PubMed] [Google Scholar]
  • 10. Wirth C., Condemine G., Boiteux C., Bernèche S., Schirmer T., and Peneff C. M. (2009) NanC crystal structure, a model for outer-membrane channels of the acidic sugar-specific KdgM porin family. J. Mol. Biol. 394, 718–731 10.1016/j.jmb.2009.09.054 [DOI] [PubMed] [Google Scholar]
  • 11. Vergalli J., Bodrenko I. V., Masi M., Moynié L., Acosta-Gutiérrez S., Naismith J. H., Davin-Regli A., Ceccarelli M., van den Berg B., Winterhalter M., and Pagès J.-M. (2019) Porins and small-molecule translocation across the outer membrane of Gram-negative bacteria. Nat. Rev. Microbiol. 18, 164–176 10.1038/s41579-019-0294-2 [DOI] [PubMed] [Google Scholar]
  • 12. Xu C., Lin X., Ren H., Zhang Y., Wang S., and Peng X. (2006) Analysis of outer membrane proteome of Escherichia coli related to resistance to ampicillin and tetracycline. Proteomics 6, 462–473 10.1002/pmic.200500219 [DOI] [PubMed] [Google Scholar]
  • 13. Beketskaia M. S., Bay D. C., and Turner R. J. (2014) Outer membrane protein OmpW participates with small multidrug resistance protein member EmrE in quaternary cationic compound efflux. J. Bacteriol. 196, 1908–1914 10.1128/JB.01483-14 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14. Wang Z., Fan G., Hryc C. F., Blaza J. N., Serysheva I. I., Schmid M. F., Chiu W., Luisi B. F., and Du D. (2017) An allosteric transport mechanism for the AcrAB-TolC multidrug efflux pump. Elife 6, e24905 10.7554/eLife.24905 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15. Fitzpatrick A. W. P., Llabrés S., Neuberger A., Blaza J. N., Bai X.-C., Okada U., Murakami S., van Veen H. W., Zachariae U., Scheres S. H. W., Luisi B. F., and Du D. (2017) Structure of the MacAB-TolC ABC-type tripartite multidrug efflux pump. Nat. Microbiol. 2, 17070 10.1038/nmicrobiol.2017.70 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16. Noinaj N., Guillier M., Barnard T. J., and Buchanan S. K. (2010) TonB-dependent transporters: regulation, structure, and function. Annu. Rev. Microbiol. 64, 43–60 10.1146/annurev.micro.112408.134247 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17. Lepore B. W., Indic M., Pham H., Hearn E. M., Patel D. R., and van den Berg B. (2011) Ligand-gated diffusion across the bacterial outer membrane. Proc. Natl. Acad. Sci. U. S. A. 108, 10121–10126 10.1073/pnas.1018532108 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18. Aunkham A., Zahn M., Kesireddy A., Pothula K. R., Schulte A., Baslé A., Kleinekathöfer U., Suginta W., and van den Berg B. (2018) Structural basis for chitin acquisition by marine Vibrio species. Nat. Commun. 9, 220 10.1038/s41467-017-02523-y [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19. Rojas E. R., Billings G., Odermatt P. D., Auer G. K., Zhu L., Miguel A., Chang F., Weibel D. B., Theriot J. A., and Huang K. C. (2018) The outer membrane is an essential load-bearing element in Gram-negative bacteria. Nature 559, 617–621 10.1038/s41586-018-0344-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20. Samsudin F., Ortiz-Suarez M. L., Piggot T. J., Bond P. J., and Khalid S. (2016) OmpA: a flexible clamp for bacterial cell wall attachment. Structure 24, 2227–2235 10.1016/j.str.2016.10.009 [DOI] [PubMed] [Google Scholar]
  • 21. Vollmer W., Von Rechenberg M., and Höltje J. V. (1999) Demonstration of molecular interactions between the murein polymerase PBP1B, the lytic transglycosylase MltA, and the scaffolding protein MipA of Escherichia coli. J. Biol. Chem. 274, 6726–6734 10.1074/jbc.274.10.6726 [DOI] [PubMed] [Google Scholar]
  • 22. Braun M., and Silhavy T. J. (2002) Imp/OstA is required for cell envelope biogenesis in Escherichia coli. Mol. Microbiol. 45, 1289–1302 10.1046/j.1365-2958.2002.03091.x [DOI] [PubMed] [Google Scholar]
  • 23. Voulhoux R., Bos M. P., Geurtsen J., Mols M., and Tommassen J. (2003) Role of a highly conserved bacterial protein in outer membrane protein assembly. Science 299, 262–265 10.1126/science.1078973 [DOI] [PubMed] [Google Scholar]
  • 24. Bishop R. E., Gibbons H. S., Guina T., Trent M. S., Miller S. I., and Raetz C. R. H. (2000) Transfer of palmitate from phospholipids to lipid A in outer membranes of Gram-negative bacteria. EMBO J. 19, 5071–5080 10.1093/emboj/19.19.5071 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25. May K. L., and Silhavy T. J. (2018) The Escherichia coli phospholipase PldA regulates outer membrane homeostasis via lipid signaling. MBio 9, 11325–11340 10.1128/mBio.00379-18 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26. Stubenrauch C. J., and Lithgow T. (2019) The TAM: a translocation and assembly module of the β-barrel assembly machinery in bacterial outer membranes. EcoSal Plus 8, 10.1128/ecosalplus.ESP-0036-2018 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27. Mecsas J., Welch R., Erickson J. W., and Gross C. A. (1995) Identification and characterization of an outer membrane protein, OmpX, in Escherichia coli that is homologous to a family of outer membrane proteins including Ail of Yersinia enterocolitica. J. Bacteriol. 177, 799–804 10.1128/jb.177.3.799-804.1995 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28. Mulvey M. A., Lopez-Boado Y. S., Wilson C. L., Roth R., Parks W. C., Heuser J., and Hultgren S. J. (1998) Induction and evasion of host defenses by type 1-piliated uropathogenic Escherichia coli. Science 282, 1494–1497 10.1126/science.282.5393.1494 [DOI] [PubMed] [Google Scholar]
  • 29. Du M., Yuan Z., Yu H., Henderson N., Sarowar S., Zhao G., Werneburg G. T., Thanassi D. G., and Li H. (2018) Handover mechanism of the growing pilus by the bacterial outer-membrane usher FimD. Nature 562, 444–447 10.1038/s41586-018-0587-z [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30. Henderson I. R., and Owen P. (1999) The major phase-variable outer membrane protein of Escherichia coli structurally resembles the immunoglobulin A1 protease class of exported protein and is regulated by a novel mechanism involving Dam and oxyR. J. Bacteriol. 181, 2132–2141 10.1128/JB.181.7.2132-2141.1999 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31. de Luna M., das G., Scott-Tucker A., Desvaux M., Ferguson P., Morin N. P., Dudley E. G., Turner S., Nataro J. P., Owen P., and Henderson I. R. (2008) The Escherichia coli biofilm-promoting protein Antigen 43 does not contribute to intestinal colonization. FEMS Microbiol. Lett. 284, 237–246 10.1111/j.1574-6968.2008.01207.x [DOI] [PubMed] [Google Scholar]
  • 32. Wang Y., Andole Pannuri A., Ni D., Zhou H., Cao X., Lu X., Romeo T., and Huang Y. (2016) Structural basis for translocation of a biofilm-supporting exopolysaccharide across the bacterial outer membrane. J. Biol. Chem. 291, 10046–10057 10.1074/jbc.M115.711762 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33. Acheson J. F., Derewenda Z. S., and Zimmer J. (2019) Architecture of the cellulose synthase outer membrane channel and its association with the periplasmic TPR domain. Structure 27, 1855–1861.e3 10.1016/j.str.2019.09.008 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34. Stumpe S., Schmid R., Stephens D. L., Georgiou G., and Bakker E. P. (1998) Identification of OmpT as the protease that hydrolyzes the antimicrobial peptide protamine before it enters growing cells of Escherichia coli. J. Bacteriol. 180, 4002–4006 10.1128/JB.180.15.4002-4006.1998 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35. Hwang B.-Y., Varadarajan N., Li H., Rodriguez S., Iverson B. L., and Georgiou G. (2007) Substrate specificity of the Escherichia coli outer membrane protease OmpP. J. Bacteriol. 189, 522–530 10.1128/JB.01493-06 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36. Ruiz N., Kahne D., and Silhavy T. J. (2006) Advances in understanding bacterial outer-membrane biogenesis. Nat. Rev. Microbiol. 4, 57–66 10.1038/nrmicro1322 [DOI] [PubMed] [Google Scholar]
  • 37. Lee J., Xue M., Wzorek J. S., Wu T., Grabowicz M., Gronenberg L. S., Sutterlin H. A., Davis R. M., Ruiz N., Silhavy T. J., and Kahne D. E. (2016) Characterization of a stalled complex on the β-barrel assembly machine. Proc. Natl. Acad. Sci. U. S. A. 113, 8717–8722 10.1073/pnas.1604100113 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38. Gu Y., Stansfeld P. J., Zeng Y., Dong H., Wang W., and Dong C. (2015) Lipopolysaccharide is inserted into the outer membrane through an intramembrane hole, a lumen gate, and the lateral opening of LptD. Structure 23, 496–504 10.1016/j.str.2015.01.001 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39. Ricci D. P., and Silhavy T. J. (2019) Outer membrane protein insertion by the β-barrel assembly machine. EcoSal Plus 8, 10.1128/ecosalplus.ESP-0035-2018 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40. Burgess N. K., Dao T. P., Stanley A. M., and Fleming K. G. (2008) β-Barrel proteins that reside in the Escherichia coli outer membrane in vivo demonstrate varied folding behavior in vitro. J. Biol. Chem. 283, 26748–26758 10.1074/jbc.M802754200 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41. Gessmann D., Chung Y. H., Danoff E. J., Plummer A. M., Sandlin C. W., Zaccai N. R., and Fleming K. G. (2014) Outer membrane β-barrel protein folding is physically controlled by periplasmic lipid head groups and BamA. Proc. Natl. Acad. Sci. U. S. A. 111, 5878–5883 10.1073/pnas.1322473111 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42. Iadanza M. G., Higgins A. J., Schiffrin B., Calabrese A. N., Brockwell D. J., Ashcroft A. E., Radford S. E., and Ranson N. A. (2016) Lateral opening in the intact β-barrel assembly machinery captured by cryo-EM. Nat. Commun. 7, 12865 10.1038/ncomms12865 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43. Hagan C. L., Kim S., and Kahne D. (2010) Reconstitution of outer membrane protein assembly from purified components. Science 328, 890–892 10.1126/science.1188919 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44. Hagan C. L., Westwood D. B., and Kahne D. (2013) Bam lipoproteins assemble BamA in vitro. Biochemistry 52, 6108–6113 10.1021/bi400865z [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45. Patel G. J., and Kleinschmidt J. H. (2013) The lipid bilayer-inserted membrane protein BamA of Escherichia coli facilitates insertion and folding of outer membrane protein A from its complex with Skp. Biochemistry 52, 3974–3986 10.1021/bi400103t [DOI] [PubMed] [Google Scholar]
  • 46. Hussain S., and Bernstein H. D. (2018) The Bam complex catalyzes efficient insertion of bacterial outer membrane proteins into membrane vesicles of variable lipid composition. J. Biol. Chem. 293, 2959–2973 10.1074/jbc.RA117.000349 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47. Doyle M. T., and Bernstein H. D. (2019) Bacterial outer membrane proteins assemble via asymmetric interactions with the BamA β-barrel. Nat. Commun. 10, 3358 10.1038/s41467-019-11230-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48. Ieva R., and Bernstein H. D. (2009) Interaction of an autotransporter passenger domain with BamA during its translocation across the bacterial outer membrane. Proc. Natl. Acad. Sci. U. S. A. 106, 19120–19125 10.1073/pnas.0907912106 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49. Sauri A., Soprova Z., Wickström D., de Gier J.-W., Van der Schors R. C., Smit A. B., Jong W. S. P., and Luirink J. (2009) The Bam (Omp85) complex is involved in secretion of the autotransporter haemoglobin protease. Microbiology 155, 3982–3991 10.1099/mic.0.034991-0 [DOI] [PubMed] [Google Scholar]
  • 50. Bodelón G., Marín E., and Fernández L. A. (2009) Role of periplasmic chaperones and BamA (YaeT/Omp85) in folding and secretion of intimin from enteropathogenic Escherichia coli strains. J. Bacteriol. 191, 5169–5179 10.1128/JB.00458-09 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51. Norell D., Heuck A., Tran-Thi T.-A., Götzke H., Jacob-Dubuisson F., Clausen T., Daley D. O., Braun V., Müller M., and Fan E. (2014) Versatile in vitro system to study translocation and functional integration of bacterial outer membrane proteins. Nat. Commun. 5, 5396 10.1038/ncomms6396 [DOI] [PubMed] [Google Scholar]
  • 52. Roman-Hernandez G., Peterson J. H., and Bernstein H. D. (2014) Reconstitution of bacterial autotransporter assembly using purified components. Elife 3, e04234 10.7554/eLife.04234 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53. Klein K., Sonnabend M. S., Frank L., Leibiger K., Franz-Wachtel M., Macek B., Trunk T., Leo J. C., Autenrieth I. B., Schütz M., and Bohn E. (2019) Deprivation of the periplasmic chaperone SurA reduces virulence and restores antibiotic susceptibility of multidrug-resistant Pseudomonas aeruginosa. Front. Microbiol. 10, 100 10.3389/fmicb.2019.00100 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54. Ruiz N., Falcone B., Kahne D., and Silhavy T. J. (2005) Chemical conditionality: a genetic strategy to probe organelle assembly. Cell 121, 307–317 10.1016/j.cell.2005.02.014 [DOI] [PubMed] [Google Scholar]
  • 55. Wu T., Malinverni J., Ruiz N., Kim S., Silhavy T. J., and Kahne D. (2005) Identification of a multicomponent complex required for outer membrane biogenesis in Escherichia coli. Cell 121, 235–245 10.1016/j.cell.2005.02.015 [DOI] [PubMed] [Google Scholar]
  • 56. Rassam P., Copeland N. A., Birkholz O., Tóth C., Chavent M., Duncan A. L., Cross S. J., Housden N. G., Kaminska R., Seger U., Quinn D. M., Garrod T. J., Sansom M. S. P., Piehler J., Baumann C. G., et al. (2015) Supramolecular assemblies underpin turnover of outer membrane proteins in bacteria. Nature 523, 333–336 10.1038/nature14461 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57. Ryan K. R., Taylor J. A., and Bowers L. M. (2010) The BAM complex subunit BamE (SmpA) is required for membrane integrity, stalk growth and normal levels of outer membrane β-barrel proteins in Caulobacter crescentus. Microbiology 156, 742–756 10.1099/mic.0.035055-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58. Werner J., and Misra R. (2005) YaeT (Omp85) affects the assembly of lipid-dependent and lipid-independent outer membrane proteins of Escherichia coli. Mol. Microbiol. 57, 1450–1459 10.1111/j.1365-2958.2005.04775.x [DOI] [PubMed] [Google Scholar]
  • 59. Palomino C., Marín E., and Fernández L. Á. (2011) The fimbrial usher FimD follows the SurA-BamB pathway for its assembly in the outer membrane of Escherichia coli. J. Bacteriol. 193, 5222–5230 10.1128/JB.05585-11 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60. Lee J., Sutterlin H. A., Wzorek J. S., Mandler M. D., Hagan C. L., Grabowicz M., Tomasek D., May M. D., Hart E. M., Silhavy T. J., and Kahne D. (2018) Substrate binding to BamD triggers a conformational change in BamA to control membrane insertion. Proc. Natl. Acad. Sci. U. S. A. 115, 2359–2364 10.1073/pnas.1711727115 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61. Lee J., Tomasek D., Santos T. M., May M. D., Meuskens I., and Kahne D. (2019) Formation of a β-barrel membrane protein is catalyzed by the interior surface of the assembly machine protein BamA. Elife e49787 10.7554/elife.49787 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62. Chimalakonda G., Ruiz N., Chng S.-S., Garner R. A., Kahne D., and Silhavy T. J. (2011) 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. U. S. A. 108, 2492–2497 10.1073/pnas.1019089108 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63. Volokhina E. B., Beckers F., Tommassen J., and Bos M. P. (2009) The β-barrel outer membrane protein assembly complex of Neisseria meningitidis. J. Bacteriol. 191, 7074–7085 10.1128/JB.00737-09 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64. Collin S., Guilvout I., Chami M., and Pugsley A. P. (2007) YaeT-independent multimerization and outer membrane association of secretin PulD. Mol. Microbiol. 64, 1350–1357 10.1111/j.1365-2958.2007.05743.x [DOI] [PubMed] [Google Scholar]
  • 65. Hoang H. H., Nickerson N. N., Lee V. T., Kazimirova A., Chami M., Pugsley A. P., and Lory S. (2011) Outer membrane targeting of Pseudomonas aeruginosa proteins shows variable dependence on the components of Bam and Lol machineries. MBio. 2, e00246–11 10.1128/mBio.00246-11 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66. Huysmans G. H. M., Guilvout I., Chami M., Nickerson N. N., and Pugsley A. P. (2015) Lipids assist the membrane insertion of a BAM-independent outer membrane protein. Sci. Rep. 5, 1–15 10.1038/srep15068 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67. Dunstan R. A., Hay I. D., Wilksch J. J., Schittenhelm R. B., Purcell A. W., Clark J., Costin A., Ramm G., Strugnell R. A., and Lithgow T. (2015) Assembly of the secretion pores GspD, Wza and CsgG into bacterial outer membranes does not require the Omp85 proteins BamA or TamA. Mol. Microbiol. 97, 616–629 10.1111/mmi.13055 [DOI] [PubMed] [Google Scholar]
  • 68. Kazmierczak B. I., Mielke D. L., Russel M., and Model P. (1994) pIV, a filamentous phage protein that mediates phage export across the bacterial cell envelope, forms a multimer. J. Mol. Biol. 238, 187–198 10.1006/jmbi.1994.1280 [DOI] [PubMed] [Google Scholar]
  • 69. Nickerson N. N., Abby S. S., Rocha E. P. C., Chami M., and Pugsley A. P. (2012) A single amino acid substitution changes the self-assembly status of a Type IV piliation secretin. J. Bacteriol. 194, 4951–4958 10.1128/JB.00798-12 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70. Paschen S. A., Neupert W., and Rapaport D. (2005) Biogenesis of β-barrel membrane proteins of mitochondria. Trends Biochem. Sci. 30, 575–582 10.1016/j.tibs.2005.08.009 [DOI] [PubMed] [Google Scholar]
  • 71. Surrey T., and Jähnig F. (1992) Refolding and oriented insertion of a membrane protein into a lipid bilayer. Proc. Natl. Acad. Sci. U. S. A. 89, 7457–7461 10.1073/pnas.89.16.7457 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72. Yildirim M. A., Goh K.-I., Cusick M. E., Barabási A.-L., and Vidal M. (2007) Drug-target network. Nat. Biotechnol. 25, 1119–1126 10.1038/nbt1338 [DOI] [PubMed] [Google Scholar]
  • 73. Yin H., and Flynn A. D. (2016) Drugging membrane protein interactions. Annu. Rev. Biomed. Eng. 18, 51–76 10.1146/annurev-bioeng-092115-025322 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74. Genevrois S., Steeghs L., Roholl P., Letesson J.-J., and van der Ley P. (2003) The Omp85 protein of Neisseria meningitidis is required for lipid export to the outer membrane. EMBO J. 22, 1780–1789 10.1093/emboj/cdg174 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75. Wiedemann N., Kozjak V., Chacinska A., Schönfisch B., Rospert S., Ryan M. T., Pfanner N., and Meisinger C. (2003) Machinery for protein sorting and assembly in the mitochondrial outer membrane. Nature 424, 565–571 10.1038/nature01753 [DOI] [PubMed] [Google Scholar]
  • 76. Paschen S. A., Waizenegger T., Stan T., Preuss M., Cyrklaff M., Hell K., Rapaport D., and Neupert W. (2003) Evolutionary conservation of biogenesis of β-barrel membrane proteins. Nature 426, 862–866 10.1038/nature02208 [DOI] [PubMed] [Google Scholar]
  • 77. Kozjak V., Wiedemann N., Milenkovic D., Lohaus C., Meyer H. E., Guiard B., Meisinger C., and Pfanner N. (2003) An essential role of Sam50 in the protein sorting and assembly machinery of the mitochondrial outer membrane. J. Biol. Chem. 278, 48520–48523 10.1074/jbc.C300442200 [DOI] [PubMed] [Google Scholar]
  • 78. Nikaido H. (2003) Molecular basis of bacterial outer membrane permeability revisited. Microbiol. Mol. Biol. Rev. 67, 593–656 10.1128/mmbr.67.4.593-656.2003 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79. Domínguez-Medina C. C., Pérez-Toledo M., Schager A. E., Marshall J. L., Cook C. N., Bobat S., Hwang H., Chun B. J., Logan E., Bryant J. A., Channell W. M., Morris F. C., Jossi S. E., Alshayea A., Rossiter A. E., et al. (2020) Outer membrane protein size and LPS O-antigen define protective antibody targeting to the Salmonella surface. Nat. Commun. 11, 851 10.1038/s41467-020-14655-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80. Storek K. M., Vij R., Sun D., Smith P. A., Koerber J. T., and Rutherford S. T. (2018) The Escherichia coli β-barrel assembly machinery is sensitized to perturbations under high membrane fluidity. J. Bacteriol. 201, 1–15 10.1128/jb.00517-18 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81. Bentley A. T., and Klebba P. E. (1988) Effect of lipopolysaccharide structure on reactivity of antiporin monoclonal antibodies with the bacterial cell surface. J. Bacteriol. 170, 1063–1068 10.1128/jb.170.3.1063-1068.1988 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82. Storek K. M., Chan J., Vij R., Chiang N., Lin Z., Bevers J., Koth C. M., Vernes J.-M., Meng Y. G., Yin J., Wallweber H., Dalmas O., Shriver S., Tam C., Schneider K., et al. (2019) Massive antibody discovery used to probe structure-function relationships of the essential outer membrane protein LptD. Elife 8, e46258 10.7554/eLife.46258 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83. Ellenrieder L., Mårtensson C. U., and Becker T. (2015) Biogenesis of mitochondrial outer membrane proteins, problems and diseases. Biol. Chem. 396, 1199–1213 10.1515/hsz-2015-0170 [DOI] [PubMed] [Google Scholar]
  • 84. Tacconelli E., Carrara E., Savoldi A., Harbarth S., Mendelson M., Monnet D. L., Pulcini C., Kahlmeter G., Kluytmans J., Carmeli Y., Ouellette M., Outterson K., Patel J., Cavaleri M., Cox E. M., et al. (2018) Discovery, research, and development of new antibiotics: the WHO priority list of antibiotic-resistant bacteria and tuberculosis. Lancet Infect. Dis. 18, 318–327 10.1016/S1473-3099(17)30753-3 [DOI] [PubMed] [Google Scholar]
  • 85. Srinivas N., Jetter P., Ueberbacher B. J., Werneburg M., Zerbe K., Steinmann J., Van der Meijden B., Bernardini F., Lederer A., Dias R. L. A., Misson P. E., Henze H., Zumbrunn J., Gombert F. O., Obrecht D., et al. (2010) Peptidomimetic antibiotics target outer-membrane biogenesis in Pseudomonas aeruginosa. Science 327, 1010–1013 10.1126/science.1182749 [DOI] [PubMed] [Google Scholar]
  • 86. Wedege E., Lie K., Bolstad K., Weynants V. E., Halstensen A., Herstad T. K., Kreutzberger J., Nome L., Naess L. M., and Aase A. (2013) Meningococcal omp85 in detergent-extracted outer membrane vesicle vaccines induces high levels of non-functional antibodies in mice. Scand. J. Immunol. 77, 452–459 10.1111/sji.12051 [DOI] [PubMed] [Google Scholar]
  • 87. Vetterli S. U., Moehle K., and Robinson J. A. (2016) Synthesis and antimicrobial activity against Pseudomonas aeruginosa of macrocyclic β-hairpin peptidomimetic antibiotics containing N-methylated amino acids. Bioorg. Med. Chem. 24, 6332–6339 10.1016/j.bmc.2016.05.027 [DOI] [PubMed] [Google Scholar]
  • 88. Machutta C. A., Kollmann C. S., Lind K. E., Bai X., Chan P. F., Huang J., Ballell L., Belyanskaya S., Besra G. S., Barros-Aguirre D., Bates R. H., Centrella P. A., Chang S. S., Chai J., Choudhry A. E., et al. (2017) Prioritizing multiple therapeutic targets in parallel using automated DNA-encoded library screening. Nat. Commun. 8, 16081 10.1038/ncomms16081 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89. Vij R., Lin Z., Chiang N., Vernes J.-M., Storek K. M., Park S., Chan J., Meng Y. G., Comps-Agrar L., Luan P., Lee S., Schneider K., Bevers J., Zilberleyb I., Tam C., et al. (2018) A targeted boost-and-sort immunization strategy using Escherichia coli: BamA identifies rare growth inhibitory antibodies. Sci. Rep. 8, 7136 10.1038/s41598-018-25609-z [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90. Ghequire M. G. K., Swings T., Michiels J., Buchanan S. K., and De Mot R. (2018) Hitting with a BAM: selective killing by lectin-like bacteriocins. MBio 9, e02138–17 10.1128/mBio.02138-17 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91. Kaur H., Hartmann J.-B., Jakob R. P., Zahn M., Zimmermann I., Maier T., Seeger M. A., and Hiller S. (2019) Identification of conformation-selective nanobodies against the membrane protein insertase BamA by an integrated structural biology approach. J. Biomol. NMR 73, 375–384 10.1007/s10858-019-00250-8 [DOI] [PubMed] [Google Scholar]
  • 92. Ghequire M. G. K., and De Mot R. (2019) LlpB represents a second subclass of lectin-like bacteriocins. Microb. Biotechnol. 12, 567–573 10.1111/1751-7915.13373 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93. Storek K. M., Auerbach M. R., Shi H., Garcia N. K., Sun D., Nickerson N. N., Vij R., Lin Z., Chiang N., Schneider K., Wecksler A. T., Skippington E., Nakamura G., Seshasayee D., Koerber J. T., et al. (2018) Monoclonal antibody targeting the β-barrel assembly machine of Escherichia coli is bactericidal. Proc. Natl. Acad. Sci. U. S. A. 115, 3692–3697 10.1073/pnas.1800043115 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94. Psonis J. J., Chahales P., Henderson N. S., Rigel N. W., Hoffman P. S., and Thanassi D. G. (2019) The small molecule nitazoxanide selectively disrupts BAM-mediated folding of the outer membrane usher protein. J. Biol. Chem. 294, 14357–14369 10.1074/jbc.RA119.009616 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95. Steenhuis M., Abdallah A. M., de Munnik S. M., Kuhne S., Sterk G.-J., van den Berg van Saparoea B., Westerhausen S., Wagner S., van der Wel N. N., Wijtmans M., van Ulsen P., Jong W. S. P., and Luirink J. (2019) Inhibition of autotransporter biogenesis by small molecules. Mol. Microbiol. 112, 81–98 10.1111/mmi.14255 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96. Luther A., Urfer M., Zahn M., Müller M., Wang S., Mondal M., Vitale A., Hartmann J., Sharpe T., Monte F. L., Kocherla H., Cline E., Pessi G., Rath P., Modaresi S. M., et al. (2019) Chimeric peptidomimetic antibiotics against Gram-negative bacteria. Nature 576, 452–458 10.1038/s41586-019-1665-6 [DOI] [PubMed] [Google Scholar]
  • 97. Imai Y., Meyer K. J., Iinishi A., Favre-Godal Q., Green R., Manuse S., Caboni M., Mori M., Niles S., Ghiglieri M., Honrao C., Ma X., Guo J. J., Makriyannis A., Linares-Otoya L., et al. (2019) A new antibiotic selectively kills Gram-negative pathogens. Nature 576, 459–464 10.1038/s41586-019-1791-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98. Hart E. M., Mitchell A. M., Konovalova A., Grabowicz M., Sheng J., Han X., Rodriguez-Rivera F. P., Schwaid A. G., Malinverni J. C., Balibar C. J., Bodea S., Si Q., Wang H., Homsher M. F., Painter R. E., et al. (2019) A small-molecule inhibitor of BamA impervious to efflux and the outer membrane permeability barrier. Proc. Natl. Acad. Sci. U. S. A. 116, 21748–21757 10.1073/pnas.1912345116 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99. Henderson J. C., Zimmerman S. M., Crofts A. A., Boll J. M., Kuhns L. G., Herrera C. M., and Trent M. S. (2016) The power of asymmetry: architecture and assembly of the Gram-negative outer membrane lipid bilayer. Annu. Rev. Microbiol. 70, 255–278 10.1146/annurev-micro-102215-095308 [DOI] [PubMed] [Google Scholar]
  • 100. Kim S., Patel D. S., Park S., Slusky J., Klauda J. B., Widmalm G., and Im W. (2016) Bilayer properties of lipid A from various Gram-negative bacteria. Biophys. J. 111, 1750–1760 10.1016/j.bpj.2016.09.001 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101. Kosma P. (1999) Chlamydial lipopolysaccharide. Biochim. Biophys. Acta 1455, 387–402 10.1016/S0925-4439(99)00061-7 [DOI] [PubMed] [Google Scholar]
  • 102. Raetz C. R. H., and Whitfield C. (2002) Lipopolysaccharide endotoxins. Annu. Rev. Biochem. 71, 635–700 10.1146/annurev.biochem.71.110601.135414 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103. Silipo A., De Castro C., Lanzetta R., Molinaro A., and Parrilli M. (2004) Full structural characterization of the lipid A components from the Agrobacterium tumefaciens strain C58 lipopolysaccharide fraction. Glycobiology 14, 805–815 10.1093/glycob/cwh096 [DOI] [PubMed] [Google Scholar]
  • 104. Albitar-Nehme S., Basheer S. M., Njamkepo E., Brisson J.-R., Guiso N., and Caroff M. (2013) Comparison of lipopolysaccharide structures of Bordetella pertussis clinical isolates from pre- and post-vaccine era. Carbohydr. Res. 378, 56–62 10.1016/j.carres.2013.05.002 [DOI] [PubMed] [Google Scholar]
  • 105. Zähringer U., Lindner B., and Rietschel E. T. (1994) Molecular structure of lipid A, the endotoxic center of bacterial lipopolysaccharides. Adv. Carbohydr. Chem. Biochem. 50, 211–276 [PubMed] [Google Scholar]
  • 106. Miller S. I., Ernst R. K., and Bader M. W. (2005) LPS, TLR4 and infectious disease diversity. Nat. Rev. Microbiol. 3, 36–46 10.1038/nrmicro1068 [DOI] [PubMed] [Google Scholar]
  • 107. Steimle A., Autenrieth I. B., and Frick J.-S. (2016) Structure and function: lipid A modifications in commensals and pathogens. Int. J. Med. Microbiol. 306, 290–301 10.1016/j.ijmm.2016.03.001 [DOI] [PubMed] [Google Scholar]
  • 108. DeChavigny A., Heacock P. N., and Dowhan W. (1991) Sequence and inactivation of the pss gene of Escherichia coli: phosphatidylethanolamine may not be essential for cell viability. J. Biol. Chem. 266, 5323–5332 [PubMed] [Google Scholar]
  • 109. Rowlett V. W., Mallampalli V. K. P. S., Karlstaedt A., Dowhan W., Taegtmeyer H., Margolin W., and Vitrac H. (2017) Impact of membrane phospholipid alterations in Escherichia coli on cellular function and bacterial stress adaptation. J. Bacteriol. 199, 1–22 10.1128/jb.00849-16 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110. Kikuchi S., Shibuya I., and Matsumoto K. (2000) Viability of an Escherichia coli pgsA null mutant lacking detectable phosphatidylglycerol and cardiolipin. J. Bacteriol. 182, 371–376 10.1128/jb.182.2.371-376.2000 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111. Matsumoto K. (2001) Dispensable nature of phosphatidylglycerol in Escherichia coli: dual roles of anionic phospholipids. Mol. Microbiol. 39, 1427–1433 10.1046/j.1365-2958.2001.02320.x [DOI] [PubMed] [Google Scholar]
  • 112. Tan B. K., Bogdanov M., Zhao J., Dowhan W., Raetz C. R. H., and Guan Z. (2012) Discovery of a cardiolipin synthase utilizing phosphatidylethanolamine and phosphatidylglycerol as substrates. Proc. Natl. Acad. Sci. U. S. A. 109, 16504–16509 10.1073/pnas.1212797109 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113. Chen F., Zhao Q., Cai X., Lv L., Lin W., Yu X., Li C., Li Y., Xiong M., and Wang X.-G. (2009) Phosphatidylcholine in membrane of Escherichia coli changes bacterial antigenicity. Can. J. Microbiol. 55, 1328–1334 10.1139/w09-082 [DOI] [PubMed] [Google Scholar]
  • 114. Geiger O., López-Lara I. M., and Sohlenkamp C. (2013) Phosphatidylcholine biosynthesis and function in bacteria. Biochim. Biophys. Acta 1831, 503–513 10.1016/j.bbalip.2012.08.009 [DOI] [PubMed] [Google Scholar]
  • 115. Wikström M., Kelly A. A., Georgiev A., Eriksson H. M., Klement M. R., Bogdanov M., Dowhan W., and Wieslander A. (2009) Lipid-engineered Escherichia coli membranes reveal critical lipid headgroup size for protein function. J. Biol. Chem. 284, 954–965 10.1074/jbc.M804482200 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116. Caforio A., Siliakus M. F., Exterkate M., Jain S., Jumde V. R., Andringa R. L. H., Kengen S. W. M., Minnaard A. J., Driessen A. J. M., and van der Oost J. (2018) Converting Escherichia coli into an archaebacterium with a hybrid heterochiral membrane. Proc. Natl. Acad. Sci. U. S. A. 115, 3704–3709 10.1073/pnas.1721604115 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117. Rossi R. M., Yum L., Agaisse H., and Payne S. M. (2017) Cardiolipin synthesis and outer membrane localization are required for Shigella flexneri virulence. MBio 8, e01199–17 10.1128/mBio.01199-17 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118. Oliver P. M., Crooks J. A., Leidl M., Yoon E. J., Saghatelian A., and Weibel D. B. (2014) Localization of anionic phospholipids in Escherichia coli cells. J. Bacteriol. 196, 3386–3398 10.1128/JB.01877-14 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119. Koppelman C. M., Den Blaauwen T., Duursma M. C., Heeren R. M., and Nanninga N. (2001) Escherichia coli minicell membranes are enriched in cardiolipin. J. Bacteriol. 183, 6144–6147 10.1128/JB.183.20.6144-6147.2001 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120. Suzuki M., Hara H., and Matsumoto K. (2002) Envelope disorder of Escherichia coli cells lacking phosphatidylglycerol. J. Bacteriol. 184, 5418–5425 10.1128/jb.184.19.5418-5425.2002 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121. Nakayama H., Kurokawa K., and Lee B. L. (2012) Lipoproteins in bacteria: structures and biosynthetic pathways. FEBS J. 279, 4247–4268 10.1111/febs.12041 [DOI] [PubMed] [Google Scholar]
  • 122. Grabowicz M., and Silhavy T. J. (2017) Redefining the essential trafficking pathway for outer membrane lipoproteins. Proc. Natl. Acad. Sci. U. S. A. 114, 4769–4774 10.1073/pnas.1702248114 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123. Grabowicz M. (2018) Lipoprotein transport: greasing the machines of outer membrane biogenesis: re-examining lipoprotein transport mechanisms among diverse Gram-negative bacteria while exploring new discoveries and questions. Bioessays 40, e1700187 10.1002/bies.201700187 [DOI] [PubMed] [Google Scholar]
  • 124. McMorran L. M., Brockwell D. J., and Radford S. E. (2014) Mechanistic studies of the biogenesis and folding of outer membrane proteins in vitro and in vivo: what have we learned to date? Arch. Biochem. Biophys. 564, 265–280 10.1016/j.abb.2014.02.011 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125. Bai J., and Pagano R. E. (1997) Measurement of spontaneous transfer and transbilayer movement of BODIPY-labeled lipids in lipid vesicles. Biochemistry 36, 8840–8848 10.1021/bi970145r [DOI] [PubMed] [Google Scholar]
  • 126. Nakano M., Fukuda M., Kudo T., Matsuzaki N., Azuma T., Sekine K., Endo H., and Handa T. (2009) Flip-flop of phospholipids in vesicles: kinetic analysis with time-resolved small-angle neutron scattering. J. Phys. Chem. B 113, 6745–6748 10.1021/jp900913w [DOI] [PubMed] [Google Scholar]
  • 127. Vaara M., and Vaara T. (1983) Polycations as outer membrane-disorganizing agents. Antimicrob. Agents Chemother. 24, 114–122 10.1128/aac.24.1.114 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128. Paul S., Chaudhuri K., Chatterjee A. N., and Das J. (1992) Presence of exposed phospholipids in the outer membrane of Vibrio cholerae. J. Gen. Microbiol. 138, 755–761 10.1099/00221287-138-4-755 [DOI] [PubMed] [Google Scholar]
  • 129. Contreras F.-X., Sánchez-Magraner L., Alonso A., and Goñi F. M. (2010) Transbilayer (flip-flop) lipid motion and lipid scrambling in membranes. FEBS Lett. 584, 1779–1786 10.1016/j.febslet.2009.12.049 [DOI] [PubMed] [Google Scholar]
  • 130. Malinverni J. C., and Silhavy T. J. (2009) An ABC transport system that maintains lipid asymmetry in the Gram-negative outer membrane. Proc. Natl. Acad. Sci. U. S. A. 106, 8009–8014 10.1073/pnas.0903229106 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131. Abellón-Ruiz J., Kaptan S. S., Baslé A., Claudi B., Bumann D., Kleinekathöfer U., and van den Berg B. (2017) Structural basis for maintenance of bacterial outer membrane lipid asymmetry. Nat. Microbiol. 2, 1616–1623 10.1038/s41564-017-0046-x [DOI] [PubMed] [Google Scholar]
  • 132. Powers M. J., and Trent M. S. (2019) Intermembrane transport: glycerophospholipid homeostasis of the Gram-negative cell envelope. Proc. Natl. Acad. Sci. U. S. A. 116, 17147–17155 10.1073/pnas.1902026116 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133. Ingram L. O. (1977) Changes in lipid composition of Escherichia coli resulting from growth with organic solvents and with food additives. Appl. Environ. Microbiol. 33, 1233–1236 10.1128/AEM.33.5.1233-1236.1977 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134. McGarrity J. T., and Armstrong J. B. (1981) The effect of temperature and other growth conditions on the fatty acid composition of Escherichia coli. Can. J. Microbiol. 27, 835–840 10.1139/m81-128 [DOI] [PubMed] [Google Scholar]
  • 135. Ingram L. O. (1982) Regulation of fatty acid composition in Escherichia coli: a proposed common mechanism for changes induced by ethanol, chaotropic agents, and a reduction of growth temperature. J. Bacteriol. 149, 166–172 10.1128/JB.149.1.166-172.1982 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136. Arneborg N., Salskov-Iversen A., and Mathiasen T. (1993) The effect of growth rate and other growth conditions on the lipid composition of Escherichia coli. Appl. Microbiol. Biotechnol. 39, 353–357 10.1007/BF00192091 [DOI] [Google Scholar]
  • 137. Gidden J., Denson J., Liyanage R., Ivey D. M., and Lay J. O. (2009) Lipid compositions in Escherichia coli and Bacillus subtilis during growth as determined by MALDI-TOF and TOF/TOF mass spectrometry. Int. J. Mass Spectrom. 283, 178–184 10.1016/j.ijms.2009.03.005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138. Jeucken A., Molenaar M. R., van de Lest C. H. A., Jansen J. W. A., Helms J. B., and Brouwers J. F. (2019) A comprehensive functional characterization of Escherichia coli lipid genes. Cell Rep. 27, 1597–1606.e2 10.1016/j.celrep.2019.04.018 [DOI] [PubMed] [Google Scholar]
  • 139. Overath P., Brenner M., Gulik-Krzywicki T., Shechter E., and Letellier L. (1975) Lipid phase transitions in cytoplasmic and outer membranes of Escherichia coli. Biochim. Biophys. Acta 389, 358–369 10.1016/0005-2736(75)90328-4 [DOI] [PubMed] [Google Scholar]
  • 140. Oursel D., Loutelier-Bourhis C., Orange N., Chevalier S., Norris V., and Lange C. M. (2007) Identification and relative quantification of fatty acids in Escherichia coli membranes by gas chromatography/mass spectrometry. Rapid Commun. Mass Spectrom. 21, 3229–3233 10.1002/rcm.3177 [DOI] [PubMed] [Google Scholar]
  • 141. Oursel D., Loutelier-Bourhis C., Orange N., Chevalier S., Norris V., and Lange C. M. (2007) Lipid composition of membranes of Escherichia coli by liquid chromatography/tandem mass spectrometry using negative electrospray ionization. Rapid Commun. Mass Spectrom. 21, 1721–1728 10.1002/rcm.3013 [DOI] [PubMed] [Google Scholar]
  • 142. Wang A. ‐Y., and Cronan J. E. (1994) The growth phase-dependent synthesis of cyclopropane fatty acids in Escherichia coli is the result of an RpoS(KatF)-dependent promoter plus enzyme instability. Mol. Microbiol. 11, 1009–1017 10.1111/j.1365-2958.1994.tb00379.x [DOI] [PubMed] [Google Scholar]
  • 143. Grogan D. W., and Cronan J. E. (1997) Cyclopropane ring formation in membrane lipids of bacteria. Microbiol. Mol. Biol. Rev. 61, 429–441 10.1128/.61.4.429-441.1997 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144. Chang Y. Y., and Cronan J. E. (1999) Membrane cyclopropane fatty acid content is a major factor in acid resistance of Escherichia coli. Mol. Microbiol. 33, 249–259 10.1046/j.1365-2958.1999.01456.x [DOI] [PubMed] [Google Scholar]
  • 145. Kim B. H., Kim S., Kim H. G., Lee J., Lee I. S., and Park Y. K. (2005) The formation of cyclopropane fatty acids in Salmonella enterica serovar Typhimurium. Microbiology 151, 209–218 10.1099/mic.0.27265-0 [DOI] [PubMed] [Google Scholar]
  • 146. Asakura H., Ekawa T., Sugimoto N., Momose Y., Kawamoto K., Makino S.-I., Igimi S., and Yamamoto S. (2012) Membrane topology of Salmonella invasion protein SipB confers osmotolerance. Biochem. Biophys. Res. Commun. 426, 654–658 10.1016/j.bbrc.2012.09.012 [DOI] [PubMed] [Google Scholar]
  • 147. Grandvalet C., Assad-García J. S., Chu-Ky S., Tollot M., Guzzo J., Gresti J., and Tourdot-Maréchal R. (2008) Changes in membrane lipid composition in ethanol- and acid-adapted Oenococcus oeni cells: characterization of the cfa gene by heterologous complementation. Microbiology 154, 2611–2619 10.1099/mic.0.2007/016238-0 [DOI] [PubMed] [Google Scholar]
  • 148. Annous B. A., Kozempel M. F., and Kurantz M. J. (1999) Changes in membrane fatty acid composition of Pediococcus sp. strain NRRL B-2354 in response to growth conditions and its effect on thermal resistance. Appl. Environ. Microbiol. 65, 2857–2862 10.1128/AEM.65.7.2857-2862.1999 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149. Ursell T. S., Trepagnier E. H., Huang K. C., and Theriot J. A. (2012) Analysis of surface protein expression reveals the growth pattern of the Gram-negative outer membrane. PLoS Comput. Biol. 8, e1002680 10.1371/journal.pcbi.1002680 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150. Johansen J., Rasmussen A. A., Overgaard M., and Valentin-Hansen P. (2006) Conserved small non-coding RNAs that belong to the sigmaE regulon: role in down-regulation of outer membrane proteins. J. Mol. Biol. 364, 1–8 10.1016/j.jmb.2006.09.004 [DOI] [PubMed] [Google Scholar]
  • 151. Guillier M., Gottesman S., and Storz G. (2006) Modulating the outer membrane with small RNAs. Genes Dev. 20, 2338–2348 10.1101/gad.1457506 [DOI] [PubMed] [Google Scholar]
  • 152. Allen R. J., and Scott G. K. (1979) Biosynthesis and turnover of outer-membrane proteins in Escherichia coli ML308-225. Biochem. J. 182, 407–412 10.1042/bj1820407 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153. Lugtenberg E. J. J., and Peters R. (1976) Distribution of lipids in cytoplasmic and outer membranes of Escherichia coli K12. Biochim. Biophys. Acta 441, 38–47 10.1016/0005-2760(76)90279-4 [DOI] [PubMed] [Google Scholar]
  • 154. White D. A., Lennarz W. J., and Schnaitman C. A. (1972) Distribution of lipids in the wall and cytoplasmic membrane subfractions of the cell envelope of Escherichia coli. J. Bacteriol. 109, 686–690 10.1128/JB.109.2.686-690.1972 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155. Ishinaga M., Kanamoto R., and Kito M. (1979) Distribution of phospholipid molecular species in outer and cytoplasmic membrane of Escherichia coli. J. Biochem. 86, 161–165 10.1093/oxfordjournals.jbchem.a132497 383707 [DOI] [Google Scholar]
  • 156. Jarosławski S., Duquesne K., Sturgis J. N., and Scheuring S. (2009) High-resolution architecture of the outer membrane of the Gram-negative bacteria Roseobacter denitrificans. Mol. Microbiol. 74, 1211–1222 10.1111/j.1365-2958.2009.06926.x [DOI] [PubMed] [Google Scholar]
  • 157. Lessen H. J., Fleming P. J., Fleming K. G., and Sodt A. J. (2018) Building blocks of the outer membrane: calculating a general elastic energy model for β-barrel membrane proteins. J. Chem. Theory Comput. 14, 4487–4497 10.1021/acs.jctc.8b00377 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158. Oddershede L., Dreyer J. K., Grego S., Brown S., and Berg-Sørensen K. (2002) The motion of a single molecule, the λ-receptor, in the bacterial outer membrane. Biophys. J. 83, 3152–3161 10.1016/S0006-3495(02)75318-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159. Kumar M., Mommer M. S., and Sourjik V. (2010) Mobility of cytoplasmic, membrane, and DNA-binding proteins in Escherichia coli. Biophys. J. 98, 552–559 10.1016/j.bpj.2009.11.002 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160. Ritchie K., Lill Y., Sood C., Lee H., and Zhang S. (2013) Single-molecule imaging in live bacteria cells. Philos. Trans. R. Soc. Lond. B. Biol. Sci. 368, 20120355 10.1098/rstb.2012.0355 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161. Oh D., Yu Y., Lee H., Wanner B. L., and Ritchie K. (2014) Dynamics of the serine chemoreceptor in the Escherichia coli inner membrane: a high-speed single-molecule tracking study. Biophys. J. 106, 145–153 10.1016/j.bpj.2013.09.059 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162. Oswald F., Varadarajan A., Lill H., Peterman E. J. G., and Bollen Y. J. M. (2016) MreB-dependent organization of the E. coli cytoplasmic membrane controls membrane protein diffusion. Biophys. J. 110, 1139–1149 10.1016/j.bpj.2016.01.010 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163. Gibbs K. A., Isaac D. D., Xu J., Hendrix R. W., Silhavy T. J., and Theriot J. A. (2004) Complex spatial distribution and dynamics of an abundant Escherichia coli outer membrane protein, LamB. Mol. Microbiol. 53, 1771–1783 10.1111/j.1365-2958.2004.04242.x [DOI] [PubMed] [Google Scholar]
  • 164. Spector J., Zakharov S., Lill Y., Sharma O., Cramer W. A., and Ritchie K. (2010) Mobility of BtuB and OmpF in the Escherichia coli outer membrane: implications for dynamic formation of a translocon complex. Biophys. J. 99, 3880–3886 10.1016/j.bpj.2010.10.029 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165. Rothenberg E., Sepúlveda L. A., Skinner S. O., Zeng L., Selvin P. R., and Golding I. (2011) Single-virus tracking reveals a spatial receptor-dependent search mechanism. Biophys. J. 100, 2875–2882 10.1016/j.bpj.2011.05.014 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166. Kleanthous C., Rassam P., and Baumann C. G. (2015) Protein-protein interactions and the spatiotemporal dynamics of bacterial outer membrane proteins. Curr. Opin. Struct. Biol. 35, 109–115 10.1016/j.sbi.2015.10.007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167. Deich J., Judd E. M., McAdams H. H., and Moerner W. E. (2004) Visualization of the movement of single histidine kinase molecules in live Caulobacter cells. Proc. Natl. Acad. Sci. U. S. A. 101, 15921–15926 10.1073/pnas.0404200101 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168. Mullineaux C. W., Nenninger A., Ray N., and Robinson C. (2006) Diffusion of green fluorescent protein in three cell environments in Escherichia coli. J. Bacteriol. 188, 3442–3448 10.1128/JB.188.10.3442-3448.2006 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169. Leake M. C., Chandler J. H., Wadhams G. H., Bai F., Berry R. M., and Armitage J. P. (2006) Stoichiometry and turnover in single, functioning membrane protein complexes. Nature 443, 355–358 10.1038/nature05135 [DOI] [PubMed] [Google Scholar]
  • 170. Leake M. C., Greene N. P., Godun R. M., Granjon T., Buchanan G., Chen S., Berry R. M., Palmer T., and Berks B. C. (2008) Variable stoichiometry of the TatA component of the twin-arginine protein transport system observed by in vivo single-molecule imaging. Proc. Natl. Acad. Sci. U. S. A. 105, 15376–15381 10.1073/pnas.0806338105 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171. Szczepaniak J., Holmes P., Rajasekar K., Kaminska R., Samsudin F., Inns P. G., Rassam P., Khalid S., Murray S. M., Redfield C., and Kleanthous C. (2020) The lipoprotein Pal stabilises the bacterial outer membrane during constriction by a mobilisation-and-capture mechanism. Nat. Commun. 11, 1305 10.1038/s41467-020-15083-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172. Mühlradt P. F., Menzel J., Golecki J. R., and Speth V. (1974) Lateral mobility and surface density of lipopolysaccharide in the outer membrane of Salmonella typhimurium. Eur. J. Biochem. 43, 533–539 10.1111/j.1432-1033.1974.tb03440.x [DOI] [PubMed] [Google Scholar]
  • 173. Schindler M., Osborn M. J., and Koppel D. E. (1980) Lateral diffusion of lipopolysaccharide in the outer membrane of Salmonella typhimurium. Nature 285, 261–263 10.1038/285261a0 [DOI] [PubMed] [Google Scholar]
  • 174. Nenninger A., Mastroianni G., Robson A., Lenn T., Xue Q., Leake M. C., and Mullineaux C. W. (2014) Independent mobility of proteins and lipids in the plasma membrane of Escherichia coli. Mol. Microbiol. 92, 1142–1153 10.1111/mmi.12619 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175. Mika J. T., Thompson A. J., Dent M. R., Brooks N. J., Michiels J., Hofkens J., and Kuimova M. K. (2016) Measuring the viscosity of the Escherichia coli plasma membrane using molecular rotors. Biophys. J. 111, 1528–1540 10.1016/j.bpj.2016.08.020 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176. Elowitz M. B., Surette M. G., Wolf P. E., Stock J. B., and Leibler S. (1999) Protein mobility in the cytoplasm of Escherichia coli. J. Bacteriol. 181, 197–203 10.1128/JB.181.1.197-203.1999 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177. Cluzel P., Surette M., and Leibler S. (2000) An ultrasensitive bacterial motor revealed by monitoring signaling proteins in single cells. Science 287, 1652–1655 10.1126/science.287.5458.1652 [DOI] [PubMed] [Google Scholar]
  • 178. Konopka M. C., Shkel I. A., Cayley S., Record M. T., and Weisshaar J. C. (2006) Crowding and confinement effects on protein diffusion in vivo. J. Bacteriol. 188, 6115–6123 10.1128/JB.01982-05 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179. Potma E. O., de Boeij W. P., Bosgraaf L., Roelofs J., van Haastert P. J., and Wiersma D. A. (2001) Reduced protein diffusion rate by cytoskeleton in vegetative and polarized dictyostelium cells. Biophys. J. 81, 2010–2019 10.1016/S0006-3495(01)75851-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 180. Terry B. R., Matthews E. K., and Haseloff J. (1995) Molecular characterisation of recombinant green fluorescent protein by fluorescence correlation microscopy. Biochem. Biophys. Res. Commun. 217, 21–27 10.1006/bbrc.1995.2740 [DOI] [PubMed] [Google Scholar]
  • 181. Gunasinghe S. D., Shiota T., Stubenrauch C. J., Schulze K. E., Webb C. T., Fulcher A. J., Dunstan R. A., Hay I. D., Naderer T., Whelan D. R., Bell T. D. M., Elgass K. D., Strugnell R. A., and Lithgow T. (2018) The WD40 protein BamB mediates coupling of BAM complexes into assembly precincts in the bacterial outer membrane. Cell Rep. 23, 2782–2794 10.1016/j.celrep.2018.04.093 [DOI] [PubMed] [Google Scholar]
  • 182. Lill Y., Jordan L. D., Smallwood C. R., Newton S. M., Lill M. A., Klebba P. E., and Ritchie K. (2016) Confined mobility of TonB and FepA in Escherichia coli membranes. PLoS ONE 11, e0160862 10.1371/journal.pone.0160862 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183. Gold V. A. M., Ieva R., Walter A., Pfanner N., van der Laan M., and Kühlbrandt W. (2014) Visualizing active membrane protein complexes by electron cryotomography. Nat. Commun. 5, 4129 10.1038/ncomms5129 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184. Casuso I., Khao J., Chami M., Paul-Gilloteaux P., Husain M., Duneau J.-P., Stahlberg H., Sturgis J. N., and Scheuring S. (2012) Characterization of the motion of membrane proteins using high-speed atomic force microscopy. Nat. Nanotechnol. 7, 525–529 10.1038/nnano.2012.109 [DOI] [PubMed] [Google Scholar]
  • 185. Yamashita H., Taoka A., Uchihashi T., Asano T., Ando T., and Fukumori Y. (2012) Single-molecule imaging on living bacterial cell surface by high-speed AFM. J. Mol. Biol. 422, 300–309 10.1016/j.jmb.2012.05.018 [DOI] [PubMed] [Google Scholar]
  • 186. Goose J. E., and Sansom M. S. P (2013) Reduced lateral mobility of lipids and proteins in crowded membranes. PLoS Comput. Biol. 9, e1003033 10.1371/journal.pcbi.1003033 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187. Chavent M., Duncan A. L., Rassam P., Birkholz O., Hélie J., Reddy T., Beliaev D., Hambly B., Piehler J., Kleanthous C., and Sansom M. S. P. (2018) How nanoscale protein interactions determine the mesoscale dynamic organisation of bacterial outer membrane proteins. Nat. Commun. 9, 2846 10.1038/s41467-018-05255-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188. Smit J., and Nikaido H. (1978) Outer membrane of Gram-negative bacteria. XVIII. Electron microscopic studies on porin insertion sites and growth of cell surface of Salmonella typhimurium. J. Bacteriol. 135, 687–702 10.1128/JB.135.2.687-702.1978 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 189. Arunmanee W., Pathania M., Solovyova A. S., Le Brun A. P., Ridley H., Baslé A., van den Berg B., and Lakey J. H. (2016) Gram-negative trimeric porins have specific LPS binding sites that are essential for porin biogenesis. Proc. Natl. Acad. Sci. U. S. A. 113, E5034–E5043 10.1073/pnas.1602382113 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190. Verhoeven G. S., Dogterom M., and den Blaauwen T. (2013) Absence of long-range diffusion of OmpA in E. coli is not caused by its peptidoglycan binding domain. BMC Microbiol. 13, 66 10.1186/1471-2180-13-66 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191. Fowler P. W., Hélie J., Duncan A., Chavent M., Koldsø H., and Sansom M. S. P. (2016) Membrane stiffness is modified by integral membrane proteins. Soft Matter 12, 7792–7803 10.1039/c6sm01186a [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192. Heberle F. A., and Feigenson G. W. (2011) Phase separation in lipid membranes. Cold Spring Harb. Perspect. Biol. 3, a004630 10.1101/cshperspect.a004630 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193. Bigay J., and Antonny B. (2012) Curvature, lipid packing, and electrostatics of membrane organelles: defining cellular territories in determining specificity. Dev. Cell 23, 886–895 10.1016/j.devcel.2012.10.009 [DOI] [PubMed] [Google Scholar]
  • 194. Budin I., de Rond T., Chen Y., Chan L. J. G., Petzold C. J., and Keasling J. D. (2018) Viscous control of cellular respiration by membrane lipid composition. Science 362, 1186–1189 10.1126/science.aat7925 [DOI] [PubMed] [Google Scholar]
  • 195. Jackson M. B., and Cronan J. E. (1978) An estimate of the minimum amount of fluid lipid required for the growth of Escherichia coli. Biochim. Biophys. Acta 512, 472–479 10.1016/0005-2736(78)90157-8 [DOI] [PubMed] [Google Scholar]
  • 196. Janoff A. S., Haug A., and McGroarty E. J. (1979) Relationship of growth temperature and thermotropic lipid phase changes in cytoplasmic and outer membranes from Escherichia coli K12. Biochim. Biophys. Acta 555, 56–66 10.1016/0005-2736(79)90071-3 [DOI] [PubMed] [Google Scholar]
  • 197. Nakayama H., Mitsui T., Nishihara M., and Kito M. (1980) Relation between growth temperature of E. coli and phase transition temperatures of its cytoplasmic and outer membranes. Biochim. Biophys. Acta 601, 1–10 10.1016/0005-2736(80)90508-8 [DOI] [PubMed] [Google Scholar]
  • 198. Janoff A. S., Gupte S., and McGroarty E. J. (1980) Correlation between temperature range of growth and structural transitions in membranes and lipids of Escherichia coli K12. Biochim. Biophys. Acta 598, 641–646 10.1016/0005-2736(80)90043-7 [DOI] [PubMed] [Google Scholar]
  • 199. Souzu H. (1982) Escherichia coli B membrane stability related to cell growth phase. Measurement of temperature dependent physical state change of the membrane over a wide range. Biochim. Biophys. Acta 691, 161–170 10.1016/0005-2736(82)90225-5 [DOI] [PubMed] [Google Scholar]
  • 200. van den Brink-van der Laan E., Killian J. A., and de Kruijff B. (2004) Nonbilayer lipids affect peripheral and integral membrane proteins via changes in the lateral pressure profile. Biochim. Biophys. Acta 1666, 275–288 10.1016/j.bbamem.2004.06.010 [DOI] [PubMed] [Google Scholar]
  • 201. Kranenburg M., and Smit B. (2005) Phase behavior of model lipid bilayers. J. Phys. Chem. B 109, 6553–6563 10.1021/jp0457646 [DOI] [PubMed] [Google Scholar]
  • 202. Brown D. A., and London E. (1998) Structure and origin of ordered lipid domains in biological membranes. J. Membr. Biol. 164, 103–114 10.1007/s002329900397 [DOI] [PubMed] [Google Scholar]
  • 203. M'Baye G., Mély Y., Duportail G., and Klymchenko A. S. (2008) Liquid ordered and gel phases of lipid bilayers: fluorescent probes reveal close fluidity but different hydration. Biophys. J. 95, 1217–1225 10.1529/biophysj.107.127480 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 204. Quinn P. J., and Wolf C. (2009) The liquid-ordered phase in membranes. Biochim. Biophys. Acta 1788, 33–46 10.1016/j.bbamem.2008.08.005 [DOI] [PubMed] [Google Scholar]
  • 205. Mouritsen O. G. (2010) The liquid-ordered state comes of age. Biochim. Biophys. Acta 1798, 1286–1288 10.1016/j.bbamem.2010.02.017 [DOI] [PubMed] [Google Scholar]
  • 206. Sodt A. J., Sandar M. L., Gawrisch K., Pastor R. W., and Lyman E. (2014) The molecular structure of the liquid-ordered phase of lipid bilayers. J. Am. Chem. Soc. 136, 725–732 10.1021/ja4105667 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 207. Boscia A. L., Treece B. W., Mohammadyani D., Klein-Seetharaman J., Braun A. R., Wassenaar T. A., Klösgen B., and Tristram-Nagle S. (2014) X-ray structure, thermodynamics, elastic properties and MD simulations of cardiolipin/dimyristoylphosphatidylcholine mixed membranes. Chem. Phys. Lipids 178, 1–10 10.1016/j.chemphyslip.2013.12.010 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 208. Bramkamp M., and Lopez D. (2015) Exploring the existence of lipid rafts in bacteria. Microbiol. Mol. Biol. Rev. 79, 81–100 10.1128/MMBR.00036-14 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209. Shearer J., and Khalid S. (2018) Communication between the leaflets of asymmetric membranes revealed from coarse-grain molecular dynamics simulations. Sci. Rep. 8, 1805 10.1038/s41598-018-20227-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 210. Shearer J., Jefferies D., and Khalid S. (2019) Outer membrane proteins OmpA, FhuA, OmpF, EstA, BtuB, and OmpX have unique lipopolysaccharide fingerprints. J. Chem. Theory Comput. 15, 2608–2619 10.1021/acs.jctc.8b01059 [DOI] [PubMed] [Google Scholar]
  • 211. Thanassi D. G., Stathopoulos C., Karkal A., and Li H. (2005) Protein secretion in the absence of ATP: the autotransporter, two-partner secretion and chaperone/usher pathways of Gram-negative bacteria (review). Mol. Membr. Biol. 22, 63–72 10.1080/09687860500063290 [DOI] [PubMed] [Google Scholar]
  • 212. Sinensky M. (1974) Homeoviscous adaptation—a homeostatic process that regulates the viscosity of membrane lipids in Escherichia coli. Proc. Natl. Acad. Sci. U. S. A. 71, 522–525 10.1073/pnas.71.2.522 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 213. Velázquez J. B., and Fernández M. S. (2006) GPS, the slope of Laurdan generalized polarization spectra, in the study of phospholipid lateral organization and Escherichia coli lipid phases. Arch. Biochem. Biophys. 455, 163–174 10.1016/j.abb.2006.09.005 [DOI] [PubMed] [Google Scholar]
  • 214. Demchenko A. P., Mély Y., Duportail G., and Klymchenko A. S. (2009) Monitoring biophysical properties of lipid membranes by environment-sensitive fluorescent probes. Biophys. J. 96, 3461–3470 10.1016/j.bpj.2009.02.012 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215. Leung S. S. W., Brewer J., Bagatolli L. A., and Thewalt J. L. (2019) Measuring molecular order for lipid membrane phase studies: linear relationship between Laurdan generalized polarization and deuterium NMR order parameter. Biochim. Biophys. Acta. Biomembr. 1861, 183053 10.1016/j.bbamem.2019.183053 [DOI] [PubMed] [Google Scholar]
  • 216. Davis J. H., Nichol C. P., Weeks G., and Bloom M. (1979) Study of the cytoplasmic and outer membranes of Escherichia coli by deuterium magnetic resonance. Biochemistry 18, 2103–2112 10.1021/bi00577a041 [DOI] [PubMed] [Google Scholar]
  • 217. Nichol C. P., Davis J. H., Weeks G., and Bloom M. (1980) Quantitative study of the fluidity of Escherichia coli membranes using deuterium magnetic resonance. Biochemistry 19, 451–457 10.1021/bi00544a008 [DOI] [PubMed] [Google Scholar]
  • 218. Souzu H. (1986) Fluorescence polarization studies on Escherichia coli membrane stability and its relation to the resistance of the cell to freeze-thawing. I. Membrane stability in cells of differing growth phase. Biochim. Biophys. Acta 861, 353–360 10.1016/0005-2736(86)90438-4 [DOI] [PubMed] [Google Scholar]
  • 219. Melchior D. L., and Steim J. M. (1976) Thermotropic transitions in biomembranes. Annu. Rev. Biophys. Bioeng. 5, 205–238 10.1146/annurev.bb.05.060176.001225 [DOI] [PubMed] [Google Scholar]
  • 220. Vanounou S., Pines D., Pines E., Parola A. H., and Fishov I. (2002) Coexistence of domains with distinct order and polarity in fluid bacterial membranes. Photochem. Photobiol. 76, 1–11 [DOI] [PubMed] [Google Scholar]
  • 221. Fishov I., and Woldringh C. L. (1999) Visualization of membrane domains in Escherichia coli. Mol. Microbiol. 32, 1166–1172 10.1046/j.1365-2958.1999.01425.x [DOI] [PubMed] [Google Scholar]
  • 222. Sperka-Gottlieb C. D. M., Hermetter A., Paltauf F., and Daum G. (1988) Lipid topology and physical properties of the outer mitochondrial membrane of the yeast, Saccharomyces cerevisiae. Biochim. Biophys. Acta 946, 227–234 10.1016/0005-2736(88)90397-5 [DOI] [PubMed] [Google Scholar]
  • 223. Liu P., Duan W., Wang Q., and Li X. (2010) The damage of outer membrane of Escherichia coli in the presence of TiO2 combined with UV light. Colloids Surf. B Biointerfaces 78, 171–176 10.1016/j.colsurfb.2010.02.024 [DOI] [PubMed] [Google Scholar]
  • 224. Wu C.-H., Bialecka-Fornal M., and Newman D. K. (2015) Methylation at the C-2 position of hopanoids increases rigidity in native bacterial membranes. Elife 4, e05663 10.7554/eLife.05663 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 225. Wu Y., Stefl M., Olzyńska A., Hof M., Yahioglu G., Yip P., Casey D. R., Ces O., Humpolíčková J., and Kuimova M. K. (2013) Molecular rheometry: direct determination of viscosity in Lo and Ld lipid phases via fluorescence lifetime imaging. Phys. Chem. Chem. Phys. 15, 14986–14993 10.1039/c3cp51953h [DOI] [PubMed] [Google Scholar]
  • 226. Nikaido H., Takeuchi Y., Ohnishi S. I., and Nakae T. (1977) Outer membrane of Salmonella typhimurium: electron spin resonance studies. Biochim. Biophys. Acta 465, 152–164 10.1016/0005-2736(77)90363-7 [DOI] [PubMed] [Google Scholar]
  • 227. Labischinski H., Barnickel G., Bradaczek H., Naumann D., Rietschel E. T., and Giesbrecht P. (1985) High state of order of isolated bacterial lipopolysaccharide and its possible contribution to the permeation barrier property of the outer membrane. J. Bacteriol. 162, 9–20 10.1128/JB.162.1.9-20.1985 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 228. Brandenburg K., and Blume A. (1987) Investigations into the thermotropic phase behaviour of natural membranes extracted from Gram-negative bacteria and artificial membrane systems made from lipopolysaccharides and free lipid A. Thermochim. Acta 119, 127–142 10.1016/0040-6031(87)88014-0 [DOI] [Google Scholar]
  • 229. Brandenburg K., and Seydel U. (1990) Investigation into the fluidity of lipopolysaccharide and free lipid A membrane systems by Fourier-transform infrared spectroscopy and differential scanning calorimetry. Eur. J. Biochem. 191, 229–236 10.1111/j.1432-1033.1990.tb19114.x [DOI] [PubMed] [Google Scholar]
  • 230. Conde-Álvarez R., Arce-Gorvel V., Iriarte M., Manček-Keber M., Barquero-Calvo E., Palacios-Chaves L., Chacón-Díaz C., Chaves-Olarte E., Martirosyan A., von Bargen K., Grilló M.-J., Jerala R., Brandenburg K., Llobet E., Bengoechea J. A., et al. (2012) The lipopolysaccharide core of Brucella abortus acts as a shield against innate immunity recognition. PLoS Pathog. 8, e1002675 10.1371/journal.ppat.1002675 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 231. Brandenburg K., and Seydel U. (1984) Physical aspects of structure and function of membranes made from lipopolysaccharides and free lipid A. Biochim. Biophys. Acta Biomembr. 775, 225–238 10.1016/0005-2736(84)90174-3 [DOI] [Google Scholar]
  • 232. Seydel U., Koch M. H. J., and Brandenburg K. (1993) Structural polymorphisms of rough mutant lipopolysaccharides Rd to Ra from Salmonella minnesota. J. Struct. Biol. 110, 232–243 10.1006/jsbi.1993.1026 [DOI] [PubMed] [Google Scholar]
  • 233. Paracini N., Clifton L. A., Skoda M. W. A., and Lakey J. H. (2018) Liquid crystalline bacterial outer membranes are critical for antibiotic susceptibility. Proc. Natl. Acad. Sci. U. S. A. 115, E7587–E7594 10.1073/pnas.1803975115 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 234. Hughes A. V., Patel D. S., Widmalm G., Klauda J. B., Clifton L. A., and Im W. (2019) Physical properties of bacterial outer membrane models: neutron reflectometry & molecular simulation. Biophys. J. 116, 1095–1104 10.1016/j.bpj.2019.02.001 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 235. Gally H. U., Pluschke G., Overath P., and Seelig J. (1980) Structure of Escherichia coli membranes: fatty acyl chain order parameters of inner and outer membranes and derived liposomes. Biochemistry 19, 1638–1643 10.1021/bi00549a018 [DOI] [PubMed] [Google Scholar]
  • 236. Mühlradt P. F., and Golecki J. R. (1975) Asymmetrical distribution and artifactual reorientation of lipopolysaccharide in the outer membrane bilayer of Salmonella typhimurium. Eur. J. Biochem. 51, 343–352 10.1111/j.1432-1033.1975.tb03934.x [DOI] [PubMed] [Google Scholar]
  • 237. Sciandrone B., Forti F., Perego S., Falchi F., and Briani F. (2019) Temperature-dependent regulation of the Escherichia coli lpxT gene. Biochim. Biophys. Acta Gene Regul. Mech. 1862, 786–795 10.1016/j.bbagrm.2019.06.010 [DOI] [PubMed] [Google Scholar]
  • 238. Carty S. M., Sreekumar K. R., and Raetz C. R. H. (1999) Effect of cold shock on lipid A biosynthesis in Escherichia coli: induction at 12 °C of an acyltransferase specific for palmitoleoyl-acyl carrier protein. J. Biol. Chem. 274, 9677–9685 10.1074/jbc.274.14.9677 [DOI] [PubMed] [Google Scholar]
  • 239. Li Y., Powell D. A., Shaffer S. A., Rasko D. A., Pelletier M. R., Leszyk J. D., Scott A. J., Masoudi A., Goodlett D. R., Wang X., Raetz C. R. H., and Ernst R. K. (2012) LPS remodeling is an evolved survival strategy for bacteria. Proc. Natl. Acad. Sci. U. S. A. 109, 8716–8721 10.1073/pnas.1202908109 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 240. Reinés M., Llobet E., Dahlström K. M., Pérez-Gutiérrez C., Llompart C. M., Torrecabota N., Salminen T. A., and Bengoechea J. A. (2012) Deciphering the acylation pattern of Yersinia enterocolitica lipid A. PLoS Pathog. 8, e1002978 10.1371/journal.ppat.1002978 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 241. Ernst R. K., Adams K. N., Moskowitz S. M., Kraig G. M., Kawasaki K., Stead C. M., Trent M. S., and Miller S. I. (2006) The Pseudomonas aeruginosa lipid A deacylase: selection for expression and loss within the cystic fibrosis airway. J. Bacteriol. 188, 191–201 10.1128/JB.188.1.191-201.2006 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 242. Needham B. D., and Trent M. S. (2013) Fortifying the barrier: the impact of lipid A remodelling on bacterial pathogenesis. Nat. Rev. Microbiol. 11, 467–481 10.1038/nrmicro3047 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 243. Schnaitman C. A., and Klena J. D. (1993) Genetics of lipopolysaccharide biosynthesis in enteric bacteria. Microbiol. Rev. 57, 655–682 10.1128/MMBR.57.3.655-682.1993 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 244. Missiakas D., Betton J. M., and Raina S. (1996) New components of protein folding in extracytoplasmic compartments of Escherichia coli SurA, FkpA and Skp/OmpH. Mol. Microbiol. 21, 871–884 10.1046/j.1365-2958.1996.561412.x [DOI] [PubMed] [Google Scholar]
  • 245. Klein G., Lindner B., Brabetz W., Brade H., and Raina S. (2009) Escherichia coli K-12 suppressor-free mutants lacking early glycosyltransferases and late acyltransferases: minimal lipopolysaccharide structure and induction of envelope stress response. J. Biol. Chem. 284, 15369–15389 10.1074/jbc.M900490200 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 246. López C. A., Zgurskaya H., and Gnanakaran S. (2020) Molecular characterization of the outer membrane of Pseudomonas aeruginosa. Biochim. Biophys. Acta Biomembr. 1862, 183151 10.1016/j.bbamem.2019.183151 [DOI] [PubMed] [Google Scholar]
  • 247. Patel D. S., Re S., Wu E. L., Qi Y., Klebba P. E., Widmalm G., Yeom M. S., Sugita Y., and Im W. (2016) Dynamics and interactions of OmpF and LPS: influence on pore accessibility and ion permeability. Biophys. J. 110, 930–938 10.1016/j.bpj.2016.01.002 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 248. Lee J., Patel D. S., Kucharska I., Tamm L. K., and Im W. (2017) Refinement of OprH-LPS interactions by molecular simulations. Biophys. J. 112, 346–355 10.1016/j.bpj.2016.12.006 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 249. Wu E. L., Fleming P. J., Yeom M. S., Widmalm G., Klauda J. B., Fleming K. G., and Im W. (2014) E. coli outer membrane and interactions with OmpLA. Biophys. J. 106, 2493–2502 10.1016/j.bpj.2014.04.024 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 250. Ma H., Irudayanathan F. J., Jiang W., and Nangia S. (2015) Simulating Gram-negative bacterial outer membrane: a coarse grain model. J. Phys. Chem. B 119, 14668–14682 10.1021/acs.jpcb.5b07122 [DOI] [PubMed] [Google Scholar]
  • 251. Petrov A. G., Gawrisch K., Brezesinski G., Klose G., and Möps A. (1982) Optical detection of phase transitions in simple and mixed lipid-water phases. Biochim. Biophys. Acta 690, 1–7 10.1016/0005-2736(82)90231-0 [DOI] [PubMed] [Google Scholar]
  • 252. Li A., Schertzer J. W., and Yong X. (2018) Molecular dynamics modeling of Pseudomonas aeruginosa outer membranes. Phys. Chem. Chem. Phys. 20, 23635–23648 10.1039/c8cp04278k [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 253. Ma H., Khan A., and Nangia S. (2018) Dynamics of OmpF trimer formation in the bacterial outer membrane of Escherichia coli. Langmuir 34, 5623–5634 10.1021/acs.langmuir.7b02653 [DOI] [PubMed] [Google Scholar]
  • 254. Rottem S., and Leive L. (1977) Effect of variations in lipopolysaccharide on the fluidity of the outer membrane of Escherichia coli. J. Biol. Chem. 252, 2077–2081 [PubMed] [Google Scholar]
  • 255. Müller-Loennies S., Holst O., and Brade H. (1994) Chemical structure of the core region of Escherichia coli J-5 lipopolysaccharide. Eur. J. Biochem. 224, 751–760 10.1111/j.1432-1033.1994.t01-1-00751.x [DOI] [PubMed] [Google Scholar]
  • 256. Bello G., Bodin A., Lawrence M. J., Barlow D., Mason A. J., Barker R. D., and Harvey R. D. (2016) The influence of rough lipopolysaccharide structure on molecular interactions with mammalian antimicrobial peptides. Biochim. Biophys. Acta 1858, 197–209 10.1016/j.bbamem.2015.11.007 [DOI] [PubMed] [Google Scholar]
  • 257. Takeuchi Y., and Nikaido H. (1981) Persistence of segregated phospholipid domains in phospholipid-lipopolysaccharide mixed bilayers: studies with spin-labeled phospholipids. Biochemistry 20, 523–529 10.1021/bi00506a013 [DOI] [PubMed] [Google Scholar]
  • 258. Takeuchi Y., and Nikaido H. (1984) Physical interaction between lipid A and phospholipids: a study with spin-labeled phospholipids. Rev. Infect. Dis. 6, 488–492 10.1093/clinids/6.4.488 [DOI] [PubMed] [Google Scholar]
  • 259. Niemelä P. S., Miettinen M. S., Monticelli L., Hammaren H., Bjelkmar P., Murtola T., Lindahl E., and Vattulainen I. (2010) Membrane proteins diffuse as dynamic complexes with lipids. J. Am. Chem. Soc. 132, 7574–7575 10.1021/ja101481b [DOI] [PubMed] [Google Scholar]
  • 260. Javanainen M., Hammaren H., Monticelli L., Jeon J., Miettinen M. S., Martinez-Seara H., Metzler R., and Vattulainen I. (2013) Anomalous and normal diffusion of proteins and lipids in crowded lipid membranes. Faraday Discuss. 161, 397–417, discussion 419–459 10.1039/c2fd20085f [DOI] [PubMed] [Google Scholar]
  • 261. Guigas G., and Weiss M. (2016) Effects of protein crowding on membrane systems. Biochim. Biophys. Acta 1858, 2441–2450 10.1016/j.bbamem.2015.12.021 [DOI] [PubMed] [Google Scholar]
  • 262. Kirschner K. N., Lins R. D., Maass A., and Soares T. A. (2012) A glycam-based force field for simulations of lipopolysaccharide membranes: Parametrization and validation. J. Chem. Theory Comput. 8, 4719–4731 10.1021/ct300534j [DOI] [PubMed] [Google Scholar]
  • 263. Ortiz-Suarez M. L., Samsudin F., Piggot T. J., Bond P. J., and Khalid S. (2016) Full-length OmpA: structure, function, and membrane interactions predicted by molecular dynamics simulations. Biophys. J. 111, 1692–1702 10.1016/j.bpj.2016.09.009 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 264. Balusek C., and Gumbart J. C. (2016) Role of the native outer-membrane environment on the transporter BtuB. Biophys. J. 111, 1409–1417 10.1016/j.bpj.2016.08.033 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 265. Matthias K. A., Strader M. B., Nawar H. F., Gao Y. S., Lee J., Patel D. S., Im W., and Bash M. C. (2017) Heterogeneity in non-epitope loop sequence and outer membrane protein complexes alters antibody binding to the major porin protein PorB in serogroup B Neisseria meningitidis. Mol. Microbiol. 105, 934–953 10.1111/mmi.13747 [DOI] [PubMed] [Google Scholar]
  • 266. Lee J., Pothula K. R., Kleinekathöfer U., and Im W. (2018) Simulation study of Occk5 functional properties in Pseudomonas aeruginosa outer membranes. J. Phys. Chem. B 122, 8185–8192 10.1021/acs.jpcb.8b07109 [DOI] [PubMed] [Google Scholar]
  • 267. Saunders G. M., Bruce Macdonald H. E., Essex J. W., and Khalid S. (2018) Prediction of the closed conformation and insights into the mechanism of the membrane enzyme LpxR. Biophys. J. 115, 1445–1456 10.1016/j.bpj.2018.09.002 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 268. Kesireddy A., Pothula K. R., Lee J., Patel D. S., Pathania M., van den Berg B., Im W., and Kleinekathöfer U. (2019) Modeling of specific lipopolysaccharide binding sites on a Gram-negative porin. J. Phys. Chem. B 123, 5700–5708 10.1021/acs.jpcb.9b03669 [DOI] [PubMed] [Google Scholar]
  • 269. Samsudin F., and Khalid S. (2019) Movement of arginine through OprD: the energetics of permeation and the role of lipopolysaccharide in directing arginine to the protein. J. Phys. Chem. B 123, 2824–2832 10.1021/acs.jpcb.9b00063 [DOI] [PubMed] [Google Scholar]
  • 270. Lundquist K. P., and Gumbart J. C. (2020) Presence of substrate aids lateral gate separation in LptD. Biochim. Biophys. Acta Biomembr. 1862, 183025 10.1016/j.bbamem.2019.07.013 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 271. Kurisu G., Zakharov S. D., Zhalnina M. V., Bano S., Eroukova V. Y., Rokitskaya T. I., Antonenko Y. N., Wiener M. C., and Cramer W. A. (2003) The structure of BtuB with bound colicin E3 R-domain implies a translocon. Nat. Struct. Biol. 10, 948–954 10.1038/nsb997 [DOI] [PubMed] [Google Scholar]
  • 272. Straatsma T. P., and Soares T. A. (2009) Characterization of the outer membrane protein OprF of Pseudomonas aeruginosa in a lipopolysaccharide membrane by computer simulation. Proteins 74, 475–488 10.1002/prot.22165 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 273. Piggot T. J., Holdbrook D. A., and Khalid S. (2013) Conformational dynamics and membrane interactions of the E. coli outer membrane protein FecA: a molecular dynamics simulation study. Biochim. Biophys. Acta 1828, 284–293 10.1016/j.bbamem.2012.08.021 [DOI] [PubMed] [Google Scholar]
  • 274. Holdbrook D. A., Piggot T. J., Sansom M. S. P., and Khalid S. (2013) Stability and membrane interactions of an autotransport protein: MD simulations of the Hia translocator domain in a complex membrane environment. Biochim. Biophys. Acta 1828, 715–723 10.1016/j.bbamem.2012.09.002 [DOI] [PubMed] [Google Scholar]
  • 275. Schindler H., and Rosenbusch J. P. (1978) Matrix protein from Escherichia coli outer membranes forms voltage-controlled channels in lipid bilayers. Proc. Natl. Acad. Sci. U. S. A. 75, 3751–3755 10.1073/pnas.75.8.3751 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 276. Rocque W. J., Coughlin R. T., and McGroarty E. J. (1987) Lipopolysaccharide tightly bound to porin monomers and trimers from Escherichia coli K-12. J. Bacteriol. 169, 4003–4010 10.1128/jb.169.9.4003-4010.1987 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 277. Holzenburg A., Engel A., Kessler R., Manz H. J., Lustig A., and Aebi U. (1989) Rapid isolation of OmpF porin-LPS complexes suitable for structure-function studies. Biochemistry 28, 4187–4193 10.1021/bi00436a010 [DOI] [PubMed] [Google Scholar]
  • 278. Ferguson A. D., Welte W., Hofmann E., Lindner B., Holst O., Coulton J. W., and Diederichs K. (2000) A conserved structural motif for lipopolysaccharide recognition by procaryotic and eucaryotic proteins. Structure 8, 585–592 10.1016/S0969-2126(00)00143-X [DOI] [PubMed] [Google Scholar]
  • 279. DiRienzo J. M., Nakamura K., and Inouye M. (1978) The outer membrane proteins of Gram-negative bacteria: biosynthesis, assembly, and functions. Annu. Rev. Biochem. 47, 481–532 10.1146/annurev.bi.47.070178.002405 [DOI] [PubMed] [Google Scholar]
  • 280. Loos M. S., Ramakrishnan R., Vranken W., Tsirigotaki A., Tsare E.-P., Zorzini V., Geyter J. D., Yuan B., Tsamardinos I., Klappa M., Schymkowitz J., Rousseau F., Karamanou S., and Economou A. (2019) Structural basis of the subcellular topology landscape of Escherichia coli. Front. Microbiol. 10, 1670–1622 10.3389/fmicb.2019.01670 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 281. Rand R. P., Fuller N., Parsegian V. A., and Rau D. C. (1988) Variation in hydration forces between neutral phospholipid bilayers: evidence for hydration attraction. Biochemistry 27, 7711–7722 10.1021/bi00420a021 [DOI] [PubMed] [Google Scholar]
  • 282. Petrache H. I., Dodd S. W., and Brown M. F. (2000) Area per lipid and acyl length distributions in fluid phosphatidylcholines determined by (2)H NMR spectroscopy. Biophys. J. 79, 3172–3192 10.1016/S0006-3495(00)76551-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 283. Rappolt M., Hickel A., Bringezu F., and Lohner K. (2003) Mechanism of the lamellar/inverse hexagonal phase transition examined by high resolution x-ray diffraction. Biophys. J. 84, 3111–3122 10.1016/S0006-3495(03)70036-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 284. Leekumjorn S., and Sum A. K. (2007) Molecular studies of the gel to liquid-crystalline phase transition for fully hydrated DPPC and DPPE bilayers. Biochim. Biophys. Acta 1768, 354–365 10.1016/j.bbamem.2006.11.003 [DOI] [PubMed] [Google Scholar]
  • 285. Piggot T. J., Holdbrook D. A., and Khalid S. (2011) Electroporation of the E. coli and S. aureus membranes: molecular dynamics simulations of complex bacterial membranes. J. Phys. Chem. B 115, 13381–13388 10.1021/jp207013v [DOI] [PubMed] [Google Scholar]
  • 286. Pierucci O. (1978) Dimensions of Escherichia coli at various growth rates: model for envelope growth. J. Bacteriol. 135, 559–574 10.1128/JB.135.2.559-574.1978 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 287. Trueba F. J., and Woldringh C. L. (1980) Changes in cell diameter during the division cycle of Escherichia coli. J. Bacteriol. 142, 869–878 10.1128/JB.142.3.869-878.1980 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 288. Reshes G., Vanounou S., Fishov I., and Feingold M. (2008) Timing the start of division in E. coli: a single-cell study. Phys. Biol. 5, 046001 10.1088/1478-3975/5/4/046001 [DOI] [PubMed] [Google Scholar]
  • 289. Mathelié-Guinlet M., Asmar A. T., Collet J., and Dufrêne Y. F. (2020) Lipoprotein Lpp regulates the mechanical properties of the E. coli cell envelope. Nat. Commun. 11, 1789 10.1038/s41467-020-15489-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 290. Rao S., Bates G. T., Matthews C. R., Newport T. D., Vickery O. N., and Stansfeld P. J. (2020) Characterizing membrane association and periplasmic transfer of bacterial lipoproteins through molecular dynamics simulations. Structure 28, 475–487.e3 10.1016/j.str.2020.01.012 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 291. Clifton L. A., Skoda M. W. A., Daulton E. L., Hughes A. V., Le Brun A. P., Lakey J. H., and Holt S. A. (2013) Asymmetric phospholipid: lipopolysaccharide bilayers; a Gram-negative bacterial outer membrane mimic. J. R. Soc. Interface 10, 20130810 10.1098/rsif.2013.0810 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 292. Clifton L. A., Skoda M. W. A., Le Brun A. P., Ciesielski F., Kuzmenko I., Holt S. A., and Lakey J. H. (2015) Effect of divalent cation removal on the structure of Gram-negative bacterial outer membrane models. Langmuir 31, 404–412 10.1021/la504407v [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 293. Markones M., Fippel A., Kaiser M., Drechsler C., Hunte C., and Heerklotz H. (2020) Stairway to asymmetry: five steps to lipid-asymmetric proteoliposomes. Biophys. J. 118, 294–302 10.1016/j.bpj.2019.10.043 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 294. Markones M., Drechsler C., Kaiser M., Kalie L., Heerklotz H., and Fiedler S. (2018) Engineering asymmetric lipid vesicles: Accurate and convenient control of the outer leaflet lipid composition. Langmuir 34, 1999–2005 10.1021/acs.langmuir.7b03189 [DOI] [PubMed] [Google Scholar]
  • 295. Drechsler C., Markones M., Choi J.-Y., Frieling N., Fiedler S., Voelker D. R., Schubert R., and Heerklotz H. (2018) Preparation of asymmetric liposomes using a phosphatidylserine decarboxylase. Biophys. J. 115, 1509–1517 10.1016/j.bpj.2018.08.036 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 296. Gurtovenko A. A., and Vattulainen I. (2008) Membrane potential and electrostatics of phospholipid bilayers with asymmetric transmembrane distribution of anionic lipids. J. Phys. Chem. B 112, 4629–4634 10.1021/jp8001993 [DOI] [PubMed] [Google Scholar]
  • 297. Esteban-Martín S., Risselada H. J., Salgado J., and Marrink S. J. (2009) Stability of asymmetric lipid bilayers assessed by molecular dynamics simulations. J. Am. Chem. Soc. 131, 15194–15202 10.1021/ja904450t [DOI] [PubMed] [Google Scholar]
  • 298. Park S., Beaven A. H., Klauda J. B., and Im W. (2015) How tolerant are membrane simulations with mismatch in area per lipid between leaflets? J. Chem. Theory Comput. 11, 3466–3477 10.1021/acs.jctc.5b00232 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 299. Heberle F. A., Marquardt D., Doktorova M., Geier B., Standaert R. F., Heftberger P., Kollmitzer B., Nickels J. D., Dick R. A., Feigenson G. W., Katsaras J., London E., and Pabst G. (2016) Subnanometer structure of an asymmetric model membrane: interleaflet coupling influences domain properties. Langmuir 32, 5195–5200 10.1021/acs.langmuir.5b04562 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 300. Doktorova M., and Weinstein H. (2018) Accurate in silico modeling of asymmetric bilayers based on biophysical principles. Biophys. J. 115, 1638–1643 10.1016/j.bpj.2018.09.008 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 301. Franklin M. W., and Slusky J. S. G. (2018) Tight turns of outer membrane proteins: an analysis of sequence, structure, and hydrogen bonding. J. Mol. Biol. 430, 3251–3265 10.1016/j.jmb.2018.06.013 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 302. Steinkühler J., Sezgin E., Urbančič I., Eggeling C., and Dimova R. (2019) Mechanical properties of plasma membrane vesicles correlate with lipid order, viscosity and cell density. Commun. Biol. 2, 337 10.1038/s42003-019-0583-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 303. Mogensen J. E., Kleinschmidt J. H., Schmidt M. A., and Otzen D. E. (2005) Misfolding of a bacterial autotransporter. Protein Sci. 14, 2814–2827 10.1110/ps.051628705 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 304. Chaturvedi D., and Mahalakshmi R. (2018) Folding determinants of transmembrane β-barrels using engineered OMP chimeras. Biochemistry 57, 1987–1996 10.1021/acs.biochem.8b00012 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 305. Hong H., Park S., Jiménez R. H. F., Rinehart D., and Tamm L. K. (2007) Role of aromatic side chains in the folding and thermodynamic stability of integral membrane proteins. J. Am. Chem. Soc. 129, 8320–8327 10.1021/ja068849o [DOI] [PubMed] [Google Scholar]
  • 306. Sanchez K. M., Gable J. E., Schlamadinger D. E., and Kim J. E. (2008) Effects of tryptophan microenvironment, soluble domain, and vesicle size on the thermodynamics of membrane protein folding: lessons from the transmembrane protein OmpA. Biochemistry 47, 12844–12852 10.1021/bi800860k [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 307. Andersen K. K., Wang H., and Otzen D. E. (2012) A kinetic analysis of the folding and unfolding of OmpA in urea and guanidinium chloride: single and parallel pathways. Biochemistry 51, 8371–8383 10.1021/bi300974y [DOI] [PubMed] [Google Scholar]
  • 308. Pocanschi C. L., Popot J.-L., and Kleinschmidt J. H. (2013) Folding and stability of outer membrane protein A (OmpA) from Escherichia coli in an amphipathic polymer, amphipol A8-35. Eur. Biophys. J. 42, 103–118 10.1007/s00249-013-0887-z [DOI] [PubMed] [Google Scholar]
  • 309. Moon C. P., and Fleming K. G. (2011) Side-chain hydrophobicity scale derived from transmembrane protein folding into lipid bilayers. Proc. Natl. Acad. Sci. U. S. A. 108, 10174–10177 10.1073/pnas.1103979108 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 310. Moon C. P., Kwon S., and Fleming K. G. (2011) Overcoming hysteresis to attain reversible equilibrium folding for outer membrane phospholipase A in phospholipid bilayers. J. Mol. Biol. 413, 484–494 10.1016/j.jmb.2011.08.041 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 311. McDonald S. K., and Fleming K. G. (2016) Aromatic side chain water-to-lipid transfer free energies show a depth dependence across the membrane normal. J. Am. Chem. Soc. 138, 7946–7950 10.1021/jacs.6b03460 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 312. Chaturvedi D., and Mahalakshmi R. (2013) Methionine mutations of outer membrane protein X influence structural stability and β-barrel unfolding. PLoS ONE 8, e79351 10.1371/journal.pone.0079351 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 313. Schweizer M., Hindennach I., Garten W., and Henning U. (1978) Major proteins of the Escherichia coli outer cell envelope membrane. Interaction of protein II with lipopolysaccharide. Eur. J. Biochem. 82, 211–217 10.1111/j.1432-1033.1978.tb12013.x [DOI] [PubMed] [Google Scholar]
  • 314. Dornmair K., Kiefer H., and Jähnig F. (1990) Refolding of an integral membrane protein. OmpA of Escherichia coli. J. Biol. Chem. 265, 18907–18911 [PubMed] [Google Scholar]
  • 315. Garavito R. M., and Rosenbusch J. P. (1986) Isolation and crystallization of bacterial porin. Methods Enzymol. 125, 309–328 10.1016/s0076-6879(86)25027-2 [DOI] [PubMed] [Google Scholar]
  • 316. Surrey T., and Jähnig F. (1995) Kinetics of folding and membrane insertion of a β-barrel membrane protein. J. Biol. Chem. 270, 28199–28203 10.1074/jbc.270.47.28199 [DOI] [PubMed] [Google Scholar]
  • 317. Dekker N., Merck K., Tommassen J., and Verheij H. M. (1995) In vitro folding of Escherichia coli outer-membrane phospholipase A. Eur. J. Biochem. 232, 214–219 10.1111/j.1432-1033.1995.tb20801.x [DOI] [PubMed] [Google Scholar]
  • 318. Kleinschmidt J. H., Wiener M. C., and Tamm L. K. (1999) Outer membrane protein A of E. coli folds into detergent micelles, but not in the presence of monomeric detergent. Protein Sci. 8, 2065–2071 10.1110/ps.8.10.2065 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 319. Ebie Tan A., Burgess N. K., DeAndrade D. S., Marold J. D., and Fleming K. G. (2010) Self-association of unfolded outer membrane proteins. Macromol. Biosci. 10, 763–767 10.1002/mabi.200900479 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 320. Schnaitman C. A. (1973) Outer membrane proteins of Escherichia coli. I. Effect of preparative conditions on the migration of protein in polyacrylamide gels. Arch. Biochem. Biophys. 157, 541–552 10.1016/0003-9861(73)90673-5 [DOI] [PubMed] [Google Scholar]
  • 321. Ohnishi S., and Kameyama K. (2001) Escherichia coli OmpA retains a folded structure in the presence of sodium dodecyl sulfate due to a high kinetic barrier to unfolding. Biochim. Biophys. Acta 1515, 159–166 10.1016/S0005-2736(01)00410-2 [DOI] [PubMed] [Google Scholar]
  • 322. Anfinsen C. B. (1973) Principles that govern the folding of protein chains. Science 181, 223–230 10.1126/science.181.4096.223 [DOI] [PubMed] [Google Scholar]
  • 323. Horne J. E., and Radford S. E. (2016) A growing toolbox of techniques for studying β-barrel outer membrane protein folding and biogenesis. Biochem. Soc. Trans. 44, 802–809 10.1042/BST20160020 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 324. Clark A. C. (2008) Protein folding: are we there yet?. Arch. Biochem. Biophys. 469, 1–3 10.1016/j.abb.2007.10.007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 325. Bartlett A. I., and Radford S. E. (2009) An expanding arsenal of experimental methods yields an explosion of insights into protein folding mechanisms. Nat. Struct. Mol. Biol. 16, 582–588 10.1038/nsmb.1592 [DOI] [PubMed] [Google Scholar]
  • 326. Gruebele M., Dave K., and Sukenik S. (2016) Globular protein folding in vitro and in vivo. Annu. Rev. Biophys. 45, 233–251 10.1146/annurev-biophys-062215-011236 [DOI] [PubMed] [Google Scholar]
  • 327. Muñoz V., and Cerminara M. (2016) When fast is better: protein folding fundamentals and mechanisms from ultrafast approaches. Biochem. J. 473, 2545–2559 10.1042/BCJ20160107 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 328. Kleinschmidt J. H., and Tamm L. K. (1996) Folding intermediates of a β-barrel membrane protein: kinetic evidence for a multi-step membrane insertion mechanism. Biochemistry 35, 12993–13000 10.1021/bi961478b [DOI] [PubMed] [Google Scholar]
  • 329. Nugent S. G., Kumar D., Rampton D. S., and Evans D. F. (2001) Intestinal luminal pH in inflammatory bowel disease: possible determinants and implications for therapy with aminosalicylates and other drugs. Gut 48, 571–577 10.1136/gut.48.4.571 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 330. Kleinschmidt J. H., and Tamm L. K. (2002) Secondary and tertiary structure formation of the β-barrel membrane protein OmpA is synchronized and depends on membrane thickness. J. Mol. Biol. 324, 319–330 10.1016/S0022-2836(02)01071-9 [DOI] [PubMed] [Google Scholar]
  • 331. Pocanschi C. L., Apell H.-J., Puntervoll P., Høgh B., Jensen H. B., Welte W., and Kleinschmidt J. H. (2006) The major outer membrane protein of Fusobacterium nucleatum (FomA) folds and inserts into lipid bilayers via parallel folding pathways. J. Mol. Biol. 355, 548–561 10.1016/j.jmb.2005.10.060 [DOI] [PubMed] [Google Scholar]
  • 332. Huysmans G. H. M., Radford S. E., Baldwin S. A., and Brockwell D. J. (2012) Malleability of the folding mechanism of the outer membrane protein PagP: parallel pathways and the effect of membrane elasticity. J. Mol. Biol. 416, 453–464 10.1016/j.jmb.2011.12.039 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 333. Kleinschmidt J. H., and Tamm L. K. (1999) Time-resolved distance determination by tryptophan fluorescence quenching: probing intermediates in membrane protein folding. Biochemistry 38, 4996–5005 10.1021/bi9824644 [DOI] [PubMed] [Google Scholar]
  • 334. Kleinschmidt J. H., den Blaauwen T., Driessen A. J., and Tamm L. K. (1999) Outer membrane protein A of Escherichia coli inserts and folds into lipid bilayers by a concerted mechanism. Biochemistry 38, 5006–5016 10.1021/bi982465w [DOI] [PubMed] [Google Scholar]
  • 335. Rodionova N. A., Tatulian S. A., Surrey T., Jähnig F., and Tamm L. K. (1995) Characterization of two membrane-bound forms of OmpA. Biochemistry 34, 1921–1929 10.1021/bi00006a013 [DOI] [PubMed] [Google Scholar]
  • 336. Kleinschmidt J. H., Bulieris P. V., Qu J., Dogterom M., and den Blaauwen T. (2011) Association of neighboring β-strands of outer membrane protein A in lipid bilayers revealed by site-directed fluorescence quenching. J. Mol. Biol. 407, 316–332 10.1016/j.jmb.2011.01.021 [DOI] [PubMed] [Google Scholar]
  • 337. Popot J. L., and Engelman D. M. (2000) Helical membrane protein folding, stability, and evolution. Annu. Rev. Biochem. 69, 881–922 10.1146/annurev.biochem.69.1.881 [DOI] [PubMed] [Google Scholar]
  • 338. Engelman D. M., Chen Y., Chin C.-N., Curran A. R., Dixon A. M., Dupuy A. D., Lee A. S., Lehnert U., Matthews E. E., Reshetnyak Y. K., Senes A., and Popot J.-L. (2003) Membrane protein folding: beyond the two stage model. FEBS Lett. 555, 122–125 10.1016/S0014-5793(03)01106-2 [DOI] [PubMed] [Google Scholar]
  • 339. Hong H., and Tamm L. K. (2004) Elastic coupling of integral membrane protein stability to lipid bilayer forces. Proc. Natl. Acad. Sci. U. S. A. 101, 4065–4070 10.1073/pnas.0400358101 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 340. Pocanschi C. L., Patel G. J., Marsh D., and Kleinschmidt J. H. (2006) Curvature elasticity and refolding of OmpA in large unilamellar vesicles. Biophys. J. 91, L75–7 10.1529/biophysj.106.091439 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 341. Marsh D., Shanmugavadivu B., and Kleinschmidt J. H. (2006) Membrane elastic fluctuations and the insertion and tilt of β-barrel proteins. Biophys. J. 91, 227–232 10.1529/biophysj.105.079004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 342. Maurya S. R., Chaturvedi D., and Mahalakshmi R. (2013) Modulating lipid dynamics and membrane fluidity to drive rapid folding of a transmembrane barrel. Sci. Rep. 3, 1989 10.1038/srep01989 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 343. Danoff E. J., and Fleming K. G. (2015) Membrane defects accelerate outer membrane β-barrel protein folding. Biochemistry 54, 97–99 10.1021/bi501443p [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 344. Nielsen L. K., Bjørnholm T., and Mouritsen O. G. (2000) Fluctuations caught in the act. Nature 404, 352 10.1038/35006162 [DOI] [PubMed] [Google Scholar]
  • 345. Enders O., Ngezahayo A., Wiechmann M., Leisten F., and Kolb H.-A. (2004) Structural calorimetry of main transition of supported DMPC bilayers by temperature-controlled AFM. Biophys. J. 87, 2522–2531 10.1529/biophysj.104.040105 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 346. Blok M. C., van der Neut-Kok E. C., van Deenen L. L., and de Gier J. (1975) The effect of chain length and lipid phase transitions on the selective permeability properties of liposomes. Biochim. Biophys. Acta 406, 187–196 10.1016/0005-2736(75)90003-6 [DOI] [PubMed] [Google Scholar]
  • 347. Blok M. C., van Deenen L. L., and De Gier J. (1976) Effect of the gel to liquid crystalline phase transition on the osmotic behaviour of phosphatidylcholine liposomes. Biochim. Biophys. Acta 433, 1–12 10.1016/0005-2736(76)90172-3 [DOI] [PubMed] [Google Scholar]
  • 348. van Hoogevest P., de Gier J., and de Kruijff B. (1984) Determination of the size of the packing defects in dimyristoylphosphatidylcholine bilayers, present at the phase transition temperature. FEBS Lett. 171, 160–164 10.1016/0014-5793(84)80479-2 [DOI] [Google Scholar]
  • 349. Hays L. M., Crowe J. H., Wolkers W., and Rudenko S. (2001) Factors affecting leakage of trapped solutes from phospholipid vesicles during thermotropic phase transitions. Cryobiology 42, 88–102 10.1006/cryo.2001.2307 [DOI] [PubMed] [Google Scholar]
  • 350. Heimburg T. (2007) Thermal Biophysics of Membranes, pp. 289–300, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, Germany [Google Scholar]
  • 351. Danoff E. J., and Fleming K. G. (2017) Novel kinetic intermediates populated along the folding pathway of the transmembrane β-barrel OmpA. Biochemistry 56, 47–60 10.1021/acs.biochem.6b00809 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 352. Huysmans G. H. M., Baldwin S. A., Brockwell D. J., and Radford S. E. (2010) The transition state for folding of an outer membrane protein. Proc. Natl. Acad. Sci. U. S. A. 107, 4099–4104 10.1073/pnas.0911904107 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 353. Bond P. J., and Sansom M. S. P. (2006) Insertion and assembly of membrane proteins via simulation. J. Am. Chem. Soc. 128, 2697–2704 10.1021/ja0569104 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 354. Moon C. P., Zaccai N. R., Fleming P. J., Gessmann D., and Fleming K. G. (2013) Membrane protein thermodynamic stability may serve as the energy sink for sorting in the periplasm. Proc. Natl. Acad. Sci. U. S. A. 110, 4285–4290 10.1073/pnas.1212527110 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 355. Noinaj N., Kuszak A. J., Gumbart J. C., Lukacik P., Chang H., Easley N. C., Lithgow T., and Buchanan S. K. (2013) Structural insight into the biogenesis of β-barrel membrane proteins. Nature 501, 385–390 10.1038/nature12521 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 356. Noinaj N., Kuszak A. J., Balusek C., Gumbart J. C., and Buchanan S. K. (2014) Lateral opening and exit pore formation are required for BamA function. Structure 22, 1055–1062 10.1016/j.str.2014.05.008 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 357. Gu Y., Li H., Dong H., Zeng Y., Zhang Z., Paterson N. G., Stansfeld P. J., Wang Z., Zhang Y., Wang W., and Dong C. (2016) Structural basis of outer membrane protein insertion by the BAM complex. Nature 531, 64–69 10.1038/nature17199 [DOI] [PubMed] [Google Scholar]
  • 358. Fleming P. J., Patel D. S., Wu E. L., Qi Y., Yeom M. S., Sousa M. C., Fleming K. G., and Im W. (2016) BamA POTRA domain interacts with a native lipid membrane surface. Biophys. J. 110, 2698–2709 10.1016/j.bpj.2016.05.010 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 359. Schiffrin B., Calabrese A. N., Higgins A. J., Humes J. R., Ashcroft A. E., Kalli A. C., Brockwell D. J., and Radford S. E. (2017) Effects of periplasmic chaperones and membrane thickness on BamA-catalyzed outer-membrane protein folding. J. Mol. Biol. 429, 3776–3792 10.1016/j.jmb.2017.09.008 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 360. Lundquist K., Bakelar J., Noinaj N., and Gumbart J. C. (2018) C-terminal kink formation is required for lateral gating in BamA. Proc. Natl. Acad. Sci. U. S. A. 115, E7942–E7949 10.1073/pnas.1722530115 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 361. Ellena J. F., Lackowicz P., Mongomery H., and Cafiso D. S. (2011) Membrane thickness varies around the circumference of the transmembrane protein BtuB. Biophys. J. 100, 1280–1287 10.1016/j.bpj.2011.01.055 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 362. Holdbrook D. A., Huber R. G., Piggot T. J., Bond P. J., and Khalid S. (2016) Dynamics of crowded vesicles: local and global responses to membrane composition. PLoS ONE 11, e0156963 10.1371/journal.pone.0156963 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 363. Ramakrishnan M., Pocanschi C. L., Kleinschmidt J. H., and Marsh D. (2004) Association of spin-labeled lipids with β-barrel proteins from the outer membrane of Escherichia coli. Biochemistry 43, 11630–11636 10.1021/bi048858e [DOI] [PubMed] [Google Scholar]
  • 364. Anbazhagan V., Qu J., Kleinschmidt J. H., and Marsh D. (2008) Incorporation of outer membrane protein OmpG in lipid membranes: protein-lipid interactions and beta-barrel orientation. Biochemistry 47, 6189–6198 10.1021/bi800203g [DOI] [PubMed] [Google Scholar]
  • 365. Anbazhagan V., Vijay N., Kleinschmidt J. H., and Marsh D. (2008) Protein-lipid interactions with Fusobacterium nucleatum major outer membrane protein FomA: spin-label EPR and polarized infrared spectroscopy. Biochemistry 47, 8414–8423 10.1021/bi800750s [DOI] [PubMed] [Google Scholar]
  • 366. Ni D., Wang Y., Yang X., Zhou H., Hou X., Cao B., Lu Z., Zhao X., Yang K., and Huang Y. (2014) Structural and functional analysis of the β-barrel domain of BamA from Escherichia coli. FASEB J. 28, 2677–2685 10.1096/fj.13-248450 [DOI] [PubMed] [Google Scholar]
  • 367. Albrecht R., Schütz M., Oberhettinger P., Faulstich M., Bermejo I., Rudel T., Diederichs K., and Zeth K. (2014) Structure of BamA, an essential factor in outer membrane protein biogenesis. Acta Crystallogr. D Biol. Crystallogr. 70, 1779–1789 10.1107/S1399004714007482 [DOI] [PubMed] [Google Scholar]
  • 368. Bakelar J., Buchanan S. K., and Noinaj N. (2016) The structure of the β-barrel assembly machinery complex. Science 351, 180–186 10.1126/science.aad3460 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 369. Han L., Zheng J., Wang Y., Yang X., Liu Y., Sun C., Cao B., Zhou H., Ni D., Lou J., Zhao Y., and Huang Y. (2016) Structure of the BAM complex and its implications for biogenesis of outer-membrane proteins. Nat. Struct. Mol. Biol. 23, 192–196 10.1038/nsmb.3181 [DOI] [PubMed] [Google Scholar]
  • 370. Gu Y., Zeng Y., Wang Z., and Dong C. (2017) BamA β16C strand and periplasmic turns are critical for outer membrane protein insertion and assembly. Biochem. J. 474, 3951–3961 10.1042/BCJ20170636 [DOI] [PubMed] [Google Scholar]
  • 371. Hartmann J.-B., Zahn M., Burmann I. M., Bibow S., and Hiller S. (2018) Sequence-specific solution NMR assignments of the β-barrel insertase BamA to monitor its conformational ensemble at the atomic level. J. Am. Chem. Soc. 140, 11252–11260 10.1021/jacs.8b03220 [DOI] [PubMed] [Google Scholar]
  • 372. Lessen H. J., Majumdar A., and Fleming K. G. (2020) Backbone hydrogen bond energies in membrane proteins are insensitive to large changes in local water concentration. J. Am. Chem. Soc. 142, 6227–6235 10.1021/jacs.0c00290 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 373. Doerner P. A., and Sousa M. C. (2017) Extreme dynamics in the BamA β-barrel seam. Biochemistry 56, 3142–3149 10.1021/acs.biochem.7b00281 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 374. Marsh D. (2008) Energetics of hydrophobic matching in lipid-protein interactions. Biophys. J. 94, 3996–4013 10.1529/biophysj.107.121475 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 375. Yin F., and Kindt J. T. (2012) Hydrophobic mismatch and lipid sorting near OmpA in mixed bilayers: atomistic and coarse-grained simulations. Biophys. J. 102, 2279–2287 10.1016/j.bpj.2012.04.005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 376. Katira S., Mandadapu K. K., Vaikuntanathan S., Smit B., and Chandler D. (2016) Pre-transition effects mediate forces of assembly between transmembrane proteins. Elife. 5, e13150 10.7554/eLife.13150 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 377. Ricci D. P., Hagan C. L., Kahne D., and Silhavy T. J. (2012) Activation of the Escherichia coli β-barrel assembly machine (Bam) is required for essential components to interact properly with substrate. Proc. Natl. Acad. Sci. U. S. A. 109, 3487–3491 10.1073/pnas.1201362109 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 378. Rigel N. W., Schwalm J., Ricci D. P., and Silhavy T. J. (2012) BamE modulates the Escherichia coli β-barrel assembly machine component BamA. J. Bacteriol. 194, 1002–1008 10.1128/JB.06426-11 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 379. Rigel N. W., Ricci D. P., and Silhavy T. J. (2013) Conformation-specific labeling of BamA and suppressor analysis suggest a cyclic mechanism for β-barrel assembly in Escherichia coli. Proc. Natl. Acad. Sci. U. S. A. 110, 5151–5156 10.1073/pnas.1302662110 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 380. McCabe A. L., Ricci D., Adetunji M., and Silhavy T. J. (2017) Conformational changes that coordinate the activity of BamA and BamD allowing β-barrel assembly. J. Bacteriol. 199, 1–11 10.1128/jb.00373-17 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 381. Hart E. M., Gupta M., Wühr M., and Silhavy T. J. (2019) The synthetic phenotype of ΔbamB ΔbamE double mutants results from a lethal jamming of the Bam complex by the lipoprotein RcsF. MBio 10, 1–12 10.1128/mBio.00662-19 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 382. Tata M., and Konovalova A. (2019) Improper coordination of BamA and BamD results in Bam complex jamming by a lipoprotein substrate. MBio 10, e00660–19 10.1128/mBio.00660-19 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 383. Warner L. R., Gatzeva-Topalova P. Z., Doerner P. A., Pardi A., and Sousa M. C. (2017) Flexibility in the periplasmic domain of BamA is important for function. Structure 25, 94–106 10.1016/j.str.2016.11.013 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 384. Höhr A. I. C., Lindau C., Wirth C., Qiu J., Stroud D. A., Kutik S., Guiard B., Hunte C., Becker T., Pfanner N., and Wiedemann N. (2018) Membrane protein insertion through a mitochondrial β-barrel gate. Science 359, eaah6834 10.1126/science.aah6834 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 385. Kutik S., Stojanovski D., Becker L., Becker T., Meinecke M., Krüger V., Prinz C., Meisinger C., Guiard B., Wagner R., Pfanner N., and Wiedemann N. (2008) Dissecting membrane insertion of mitochondrial β-barrel proteins. Cell 132, 1011–1024 10.1016/j.cell.2008.01.028 [DOI] [PubMed] [Google Scholar]
  • 386. Paramasivam N., Habeck M., and Linke D. (2012) Is the C-terminal insertional signal in Gram-negative bacterial outer membrane proteins species-specific or not? BMC Genomics 13, 510 10.1186/1471-2164-13-510 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 387. Gruss F., Zähringer F., Jakob R. P., Burmann B. M., Hiller S., and Maier T. (2013) The structural basis of autotransporter translocation by TamA. Nat. Struct. Mol. Biol. 20, 1318–1320 10.1038/nsmb.2689 [DOI] [PubMed] [Google Scholar]
  • 388. Noinaj N., Rollauer S. E., and Buchanan S. K. (2015) The β-barrel membrane protein insertase machinery from Gram-negative bacteria. Curr. Opin. Struct. Biol. 31, 35–42 10.1016/j.sbi.2015.02.012 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 389. Schiffrin B., Brockwell D. J., and Radford S. E. (2017) Outer membrane protein folding from an energy landscape perspective. BMC Biol. 15, 123 10.1186/s12915-017-0464-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 390. Wang Y., Wang R., Jin F., Liu Y., Yu J., Fu X., and Chang Z. (2016) A supercomplex spanning the inner and outer membranes mediates the biogenesis of β-barrel outer membrane proteins in bacteria. J. Biol. Chem. 291, 16720–16729 10.1074/jbc.M115.710715 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 391. Alvira S., Watkins D. W., Troman L., Lorriman J., Daum B., Gold V. A. M., and Collinson I. (2019) Trans-membrane association of the Sec and BAM complexes for bacterial outer-membrane biogenesis. bioRxiv 10.1101/589077 10.1101/589077 [DOI] [PMC free article] [PubMed]
  • 392. Carlson M. L., Stacey R. G., Young J. W., Wason I. S., Zhao Z., Rattray D. G., Scott N., Kerr C. H., Babu M., Foster L. J., and Duong Van Hoa F. (2019) Profiling the Escherichia coli membrane protein interactome captured in Peptidisc libraries. Elife 8, e46615 10.7554/eLife.46615 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 393. Thoma J., Manioglu S., Kalbermatter D., Bosshart P. D., Fotiadis D., and Müller D. J. (2018) Protein-enriched outer membrane vesicles as a native platform for outer membrane protein studies. Commun. Biol. 1, 23 10.1038/s42003-018-0027-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 394. Thoma J., Sun Y., Ritzmann N., and Müller D. J. (2018) POTRA domains, extracellular lid, and membrane composition modulate the conformational stability of the β-barrel assembly factor BamA. Structure 26, 987–996.e3 10.1016/j.str.2018.04.017 [DOI] [PubMed] [Google Scholar]
  • 395. Thoma J., and Burmann B. M. (2020) High-resolution in situ NMR spectroscopy of bacterial envelope proteins in outer membrane vesicles. Biochemistry 59, 1656–1660 10.1021/acs.biochem.9b01123 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 396. Yang J., and Zhang Y. (2015) I-TASSER server: new development for protein structure and function predictions. Nucleic Acids Res. 43, W174–W181 10.1093/nar/gkv342 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 397. Chorev D. S., Baker L. A., Wu D., Beilsten-Edmands V., Rouse S. L., Zeev-Ben-Mordehai T., Jiko C., Samsudin F., Gerle C., Khalid S., Stewart A. G., Matthews S. J., Grünewald K., and Robinson C. V. (2018) Protein assemblies ejected directly from native membranes yield complexes for mass spectrometry. Science 362, 829–834 10.1126/science.aau0976 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 398. Booth P. J., and Curnow P. (2009) Folding scene investigation: membrane proteins. Curr. Opin. Struct. Biol. 19, 8–13 10.1016/j.sbi.2008.12.005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 399. Robert V., Volokhina E. B., Senf F., Bos M. P., Van Gelder P., and Tommassen J. (2006) Assembly factor Omp85 recognizes its outer membrane protein substrates by a species-specific C-terminal motif. PLos Biol. 4, e377 10.1371/journal.pbio.0040377 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 400. Arora A., Abildgaard F., Bushweller J. H., and Tamm L. K. (2001) Structure of outer membrane protein A transmembrane domain by NMR spectroscopy. Nat. Struct. Biol. 8, 334–338 10.1038/86214 [DOI] [PubMed] [Google Scholar]
  • 401. Ishida H., Garcia-Herrero A., and Vogel H. J. (2014) The periplasmic domain of Escherichia coli outer membrane protein A can undergo a localized temperature dependent structural transition. Biochim. Biophys. Acta 1838, 3014–3024 10.1016/j.bbamem.2014.08.008 [DOI] [PubMed] [Google Scholar]
  • 402. Baltoumas F. A., Hamodrakas S. J., and Iconomidou V. A. (2019) The Gram-negative outer membrane modeler: automated building of lipopolysaccharide-rich bacterial outer membranes in four force fields. J. Comput. Chem. 40, 1727–1734 10.1002/jcc.25823 [DOI] [PubMed] [Google Scholar]
  • 403. Tomasek D., Rawson S., and Lee J., Wzorek J. S., Harrison S. C., Li Z., and Kahne D. (2020) Structure of a nascent membrane protein as it folds on the BAM complex. Nature 10.1038/s41586-020-2370-1 [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from The Journal of Biological Chemistry are provided here courtesy of American Society for Biochemistry and Molecular Biology

RESOURCES