Skip to main content
Protein Science : A Publication of the Protein Society logoLink to Protein Science : A Publication of the Protein Society
. 2020 Oct 24;29(12):2348–2362. doi: 10.1002/pro.3973

How bilayer properties influence membrane protein folding

Karolina Corin 1,, James U Bowie 1
PMCID: PMC7679971  PMID: 33058341

Abstract

The question of how proteins manage to organize into a unique three‐dimensional structure has been a major field of study since the first protein structures were determined. For membrane proteins, the question is made more complex because, unlike water‐soluble proteins, the solvent is not homogenous or even unique. Each cell and organelle has a distinct lipid composition that can change in response to environmental stimuli. Thus, the study of membrane protein folding requires not only understanding how the unfolded chain navigates its way to the folded state, but also how changes in bilayer properties can affect that search. Here we review what we know so far about the impact of lipid composition on bilayer physical properties and how those properties can affect folding. A better understanding of the lipid bilayer and its effects on membrane protein folding is not only important for a theoretical understanding of the folding process, but can also have a practical impact on our ability to work with and design membrane proteins.

Keywords: lipids, membrane insertion, packing pressure, phospholipids, reconstitution, stability, topology


Abbreviations

CHAPSO

3‐([3‐Cholamidopropyl]dimethylammonio)‐2‐hydroxy‐1‐propanesulfonate

DGDG

digalactosyldiglyceride

DHPC

l‐α‐1,2‐dihexanoyl‐sn‐glycero‐3‐phosphocholine

DMPC

l‐α‐1,2‐dimyristoyl‐sn‐glycero‐3‐phophocholine

DMPE

1,2‐dimyristoyl‐sn‐glycero‐3‐phosphoethanolamine

DMPG

1,2‐dimyristoyl‐sn‐glycero‐3‐phophoglycerol

DOPA

1,2‐dioleoyl‐sn‐glycero‐3‐phosphate

DOPC

l‐α‐1,2‐dioleoyl‐sn‐glycero‐3‐phophocholine

DOPE

l‐α‐1,2‐dioleoyl‐sn‐glycero‐3‐phosphoethanolamine

DOPG

l‐α‐1,2‐dioleoyl‐sn‐glycero‐3‐phosphoglycerol

DOPS

l‐α‐1,2‐dioleoyl‐sn‐glycero‐3‐phospho‐L‐serine

DOTAP

1,2‐dioleoyl‐3‐trimethylammonium‐propane

DPoPC

L‐α‐1,2‐dipalmitoleoylphosphatidylcholine

DPoPE

l‐α‐1,2‐ dipalmitoleoylphosphatidylethanolamine

GlcGlcDAG

diglucosyl diacylglycerol

lysoOPC

1‐oleoyl‐2‐dioleoyl‐sn‐glycero‐3‐phosphocholine

lysoPPC

1‐palmitoyl‐2‐hydroxy‐sn‐glycero‐3‐phosphocholine

MGDG

monogalactosyldiglyceride

MGlcDAG

monoglucosyldiacylglycerol

POPC

1‐palmitoyl‐2‐oleoyl‐glycero‐3‐phosphocholine

POPG

1‐palmitoyl‐2‐oleoyl‐glycero‐3‐phosphoglycerol

POPS

1‐palmitoyl‐2‐oleoyl‐sn‐glycero‐3‐phospho‐L‐serine

1. INTRODUCTION

The lipid bilayer is a complex system comprised of many different lipids. The specific composition and relative abundance of each type of lipid depends on the particular organism and organelle. For example, Escheria coli membranes are primarily comprised of phosphatidylethanolamine (PE), phosphatidylglycerol (PG), and cardiolipin (CL), while mammalian cytoplasmic membranes primarily consist of phosphatidylcholine (PC), phosphatidylethanolamine (PE), phosphatidylserine (PS), sphingomyelin, and cholesterol 1 , 2 (Table 1, Figure 1). The acyl chains of each lipid can range in length from 12–26 carbons, and can be variably saturated. 3 They can also be asymmetrically attached to the glycerol backbone (Figure 1). As a result, bacterial membranes can contain thousands of different kinds of lipids, while eukaryotic membranes can have orders of magnitude more. 4 , 5 Adding to this complexity is the fact that the bilayer lipid composition is dynamic. Bacteria can adjust the composition of their membranes in response to changes in pH, pressure, or temperature, 6 , 7 while acyl chain composition in humans can be affected by diet. 8 , 9

TABLE 1.

Lipid composition in various organisms and organelles 1 , 63 , 112 , 113 , 114

PE PG PC PS PI CL Ch SM MGlcDAG DGlcDAG DPG BmP
Acholeplasma laidlawii a 10–20 20–43 31–56 9‐17 b
Escheria coli 70–75 20–25 5–10
Proteus mirabilis 80 10 5
Pseudomonas aeruginosa 60 21 11
Bacillus polymyxa 60 3 8
Bacillus subtilis 12 70 4
Streptococcus pneumoniae 50 50
Plasma membrane 11 23 8 34 17
Golgi 21 36 6 12 18 6
ER 20 54 11 8
Mitochondria 31 37 6 22
Endosomes/lysosomes 11 30 7 30 15 7

Note: All values listed are percentages of the total lipid content.

Abbreviations: BmP, bis(monoacylglycero)phosphate; Ch, cholesterol; CL, cardiolipin; DGDG, diglucosyldiglyceride, DPG, diphosphatidylglycerol; MGDG, monoglucosyldiglyceride; PC, phosphatidylcholine; PE, phosphatidylethanolamine; PG, phosphatidylglycerol; PI, phosphatidylinositol; PS, phosphatidylserine; SM: sphingomyelin.

a

The values listed are the range in which most tested strains fell.

b

Several strains only had trace amounts.

FIGURE 1.

FIGURE 1

The structures, properties, and bilayer‐forming tendencies of common membrane lipids. Lipids can have neutral (green), zwitterionic (blue), anionic (orange), or cationic (grey) head groups, and acyl chains of varying length (denoted R1 and R2). The glycerol backbone is colored red. Reprinted from Biochim Biophys Acta 1818(4), Dowhan W and Bogdanov M, Molecular genetic and biochemical approaches for defining lipid‐dependent membrane protein folding, 1,097–1,107, Copyright (2012), with permission from Elsevier 113

Early studies with bacteria suggested that bilayer properties serve a physiologically functional and vital role. In response to environmental cues, E. coli and Acholeplasma laidlawii were found to adjust the ratio of bilayer and nonbilayer lipids to keep the membrane just below the point of phase instability. 6 , 10 , 11 , 12 In response to a change in acyl chain saturation, A. laidlawii altered its lipid head group composition to maintain a stable membrane curvature. 13 When individual lipids were removed, E. coli was shown to alter the lipid composition of its membrane to compensate. 14 These experiments suggested that certain global properties of the bilayer need to be maintained for proper physiological functions.

Cells may control lipid composition partly because of its effect on membrane protein structure and folding. A particularly striking example is illustrated by the M. tuberculosis large‐conductance mechanosensitive channel (MscL), which opens upon membrane deformation induced by osmotic stress. Extensive distance measurements by electron paramagnetic resonance methods reveal a remarkable conformational change from the closed state when MscL is solubilized in asymmetric lipid bilayers instead of detergent 15 , 16 , 17 (Figure 2). Changes in lipid composition have also been shown to result in other extraordinary conformational changes, like flipping entire domains of various permeases within the membrane (see Figure 2 and below). 3 Furthermore, defects in membrane protein folding and insertion were seen in mutant E. coli that could not synthesize certain lipids. 18 In some cases, lipids with similar properties could be substituted to assure proper protein folding and function. 19 , 20 , 21 Clearly, an understanding of membrane protein structure and folding must include lipid composition.

FIGURE 2.

FIGURE 2

Examples of dramatic structural changes upon alterations in lipid composition. (a) Structural changes in MscL in response to changes in lipid composition. When MscL is reconstituted into 18:1 dioleoylphosphatidylcholine, the closed state (left) is observed. When lysophosphatidylcholine is added to one side of the bilayer the open conformation (right) is stabilized. 16 , 17 Adapted from Curr Opin Struct Biol 13(4), Perozo E and Rees DC, Structure and mechanism in prokaryotic mechanosensitive channels, 432–42, Copyright (2003), with permission from Elsevier. 17 (b) Topological flipping of LacY, PheP, and GabP in the presence and absence of PE. 47 , 48 , 54 , 57 The approximate locations of positively charged residues are shown with red dots, and the locations of negatively charged residues are shown with blue dots. Figure adapted from Refs. 48, 53, 54, 55, 56, 57, 113

Given the complex and diverse composition of lipid bilayers, how can we even begin to elucidate the relationship between lipid properties and membrane protein folding? While lipids can bind specifically to membrane proteins, stabilizing their structure in a way unique to the individual protein, 22 , 23 , 24 , 25 , 26 , 27 the global physical and chemical properties of a membrane can also play a critical role. Indeed, it has been shown that electrostatic interactions can modulate protein topology, while mechanical stresses can influence membrane protein insertion, folding, or activity. It is these general properties that will be the focus of this review. Although it is well known that bilayer properties can have a profound effect on β‐barrel membrane protein folding and stability, 28 , 29 we will concentrate on helix‐bundle membrane proteins.

2. LIPID AND BILAYER PROPERTIES

The electrostatic and mechanical properties 1 of membranes are determined by the lipid headgroup size and charge, as well as acyl chain length and saturation. The headgroup can be either charged (cationic or anionic) or uncharged (neutral or zwitterionic), and can vary in size (Figure 1). Acyl chains have varying lengths and can be saturated or unsaturated. Saturated chains can pack more densely and reduce fluidity in the bilayer. Lipids with saturated acyl chains or those with an overall cylindrical shape tend to form planar bilayers. Lipids with unsaturated chains or with conical shapes tend to curve toward water. Lipids with single chains or with inverted conical shapes tend to curve away from water (Figure 3).

FIGURE 3.

FIGURE 3

Phase‐forming tendencies of lipids. Lipids with a cylindrical shape tend to form planar bilayer structures. Examples include phosphatidylcholine and phosphatidylserine. Lipids with an inverted cone shape tend to curve away from water and form micelles. Examples include lysophosphatidylcholine and other lysophospholipids. Lipids with a conical shape tend to curve toward water and form the inverted hexagonal phase. Examples include phosphatidylethanolamine and phosphatidic acid. Reprinted from FEBS Lett 593(17), Zhukovsky MA, Filograna A, Luini A, Corda D, Valente C, Phophatidic acid in membrane arrangements, 2,428–2,451, Copyright (2019), with permission from Wiley 115

All of these properties affect the pressure profile throughout the lipid bilayer 30 (Figure 4). Electrostatic charges, steric hindrance, or hydration effects modulate the pressure in the headgroup area, often leading to positive pressure. Collisions among acyl chains lead to positive pressure in the chain area. The pressure at the polar/apolar interface is typically negative due to expansion to exclude water from the hydrocarbon core. 31 When nonbilayer hexagonal phase lipids are added to bilayer lipids, the pressure in the chain region becomes more positive as a result of more chain collisions. Because the net lateral pressure in the bilayer must remain zero to prevent the membrane from stretching, compressing, or distorting, the increase in chain pressure is balanced by a pressure decrease in the headgroup region or at the polar/apolar interface 32 , 33 , 34 (Figure 4). As shown in Figure 5, acyl chain lateral pressure may help shape membrane proteins by favoring conformations that are thinner in the middle of the bilayer. 28 , 35

FIGURE 4.

FIGURE 4

A typical lateral pressure profile through a bilayer. Adding nonbilayer lipids to bilayer lipids increases the pressure in the chain region while decreasing the pressure in the head group region. Adding lipids with longer acyl chains can also increase the lateral chain pressure, while lysophospholipids can decrease the pressure. Figure adapted from Ref. 34

FIGURE 5.

FIGURE 5

How bilayer pressure can shape membrane proteins or cause the bilayer to adapt to them. (a) Nonbilayer lipids can make an hourglass shape more stable. (b) The drive to shield hydrophobic side chains can either alter the protein conformation, the bilayer conformation, or both. Lipid shapes can facilitate or hinder bilayer adjustments

Acyl chain length and saturation can also affect bilayer thickness. This can lead to instances where the thickness is much greater or smaller than the length of the transmembrane hydrophobic domain (Figure 5). To minimize any mismatch, lipids may respond by altering their packing or degree of acyl chain ordering. 36 , 37 , 38 , 39 Protein helices may respond by tilting, bending, or stretching. 38 , 39 , 40 , 41 Nonbilayer cone‐shaped lipids can make it easier to locally distort the bilayer outward to shield a thicker hydrophobic protein domain, or make it more difficult to accommodate a thinner one. 12

But which properties of the bilayer are most critical for membrane protein folding and how do they affect it? Properties like membrane curvature, bilayer rigidity, transmembrane lateral pressure, bilayer thickness, elastic curvature stress, and even charge are interrelated and determined by the lipid composition. Although the ways in which they influence membrane protein structure or folding seem to be disparate and numerous, they can all likely be ascribed to electrostatic and mechanical forces.

3. LIPID ELECTROSTATIC EFFECTS ON MEMBRANE PROTEIN FOLDING

3.1. Lipid charge and membrane protein topology

Pioneering work on E. coli lactose permease (LacY) suggested that bilayer mechanical properties affect membrane protein folding, 42 , 43 and also yielded the first clues that protein‐lipid electrostatic interactions could play a role. 43 , 44 LacY activity was found to be consistently optimal in E. coli lipids or synthetic mixtures that approximated the lipid composition of E. coli membranes. 42 , 44 , 45 Bilayer‐forming PE lipids with saturated acyl chains supported proper folding, 46 but nonbilayer‐forming PE with unsaturated acyl chains did not. 43 However, if nonbilayer‐forming PE lipids were supplemented with another lamellar or bilayer‐forming phospholipid like PG, LacY renaturation could be observed. 43 While this evidence suggested that proper LacY folding depends on lipid mechanical and phase properties, other experiments indicated that electrostatic interactions could play a critical role. In particular, negatively charged PS, a lipid that also contains a free amine in the headgroup, could substitute for PE. 43 , 44 Moreover, increasing methylation of the PE headgroup amine led to progressively decreasing LacY activity 44 or renaturation. 43 Indeed, PC, whose amino group is completely methylated, did not significantly support either LacY activity or renaturation. Together, this suggested that the ionizable nature or hydrogen‐bonding ability of the free primary amine might be critical for proper membrane protein folding.

These early studies led the Dowhan group on a surprising and fascinating journey. To elucidate the mechanism through which PE influences LacY folding, it was critical to understand the nature of the folding defect. Bogdanov et al. showed that PE can radically alter the topological organization of LacY. 47 By labeling individual cysteines in the extracellular loops between transmembrane domains, they showed that the six N‐terminal transmembrane helices were inverted in PE‐deficient membranes (Figure 2). Remarkably, this inversion appeared to be reversible 2 as it could be corrected by inducing PE expression to normal levels. Even more surprising, when LacY was reconstituted into liposomes, similar results were observed, regardless of whether the protein had originally been expressed in a PE‐expressing or PE‐deficient strain. 49 Moreover, the quantity of PE in the liposomes directly correlated with the amount of topologically correct LacY. When the PE content was low, almost all reconstituted LacY displayed an inverted topology. But as the PE content increased, the amount of topologically correct LacY progressively increased, reaching normal levels at PE compositions above 50%. 50 Remarkably, these extensive topological rearrangements occurred within minutes. 51 These results firmly demonstrate that lipid composition in general, and PE specifically, can determine whether the N‐terminal half of LacY is inverted. Moreover, the rearrangements can occur on a physiologically relevant timescale.

The role of lipid‐protein electrostatic interactions in LacY inversion became clearer following the observation that LacY's cytoplasmic loops between TMs I‐VI had many negatively charged residues. 48 Replacing a single negative residue (Asp, Glu) with its neutral amide (Asn, Gln) in any of the three N‐terminal cytoplasmic loops was sufficient to prevent inversion of the N‐terminal helices in PE‐deficient cells. Removing both a positive and a negative charge from one loop, thereby keeping the net charge the same, resulted in inversion in PE‐deficient cells. Replacing positively charged residues in each loop with negatively charged residues induced inversion even in PE‐expressing cells. These experiments demonstrated that the charge across the cytoplasmic surface of LacY can determine its topology, and is modulated by lipid interactions. A sufficiently high negative charge across the cytoplasmic surface results in topological inversion. The presence of PE diminishes this effect, perhaps through pKa shifts that alter the charge state of ionizable side chains. 52

Charge‐modulated topological inversion has been observed in other E. coli proteins, indicating that it is a more generalized phenomenon. Like LacY, phenylalanine permease (PheP) contains many acidic residues in its N‐terminal cytoplasmic loops. 53 In the presence of PE, the effects of these charges are dampened and the loops favor a cytoplasmic orientation. In the absence of PE, TMIII acts as a hinge to invert TMI and TMII 54 (Figure 2). When the net charge of the N‐terminus and second extracellular loop is engineered to be more positive, this inversion does not occur. 55 The γ‐aminobutyric acid permease (GabP) has a similar charge distribution in its extracellular loops. 56 Like PheP, in the absence of PE, the GabP TMIII acts as a topological hinge 57 (Figure 2). Sucrose permease (CscB) also has a significant number of acidic residues in its cytoplasmic loops. Because of the higher hydrophobicity of its putative hinge domain, the N‐terminal TM helices do not invert when PE is absent. However, when an additional salt bridge is added to this region, the N‐terminal helices can invert when the cytoplasmic loops become sufficiently negative. 55 The amount of topologically correct CscB is directly proportional to the amount of PE in the membrane, nearing approximately 100% at near‐physiological levels. 58

Intriguingly, even charges from posttranslational modifications can affect protein topology. When phosphorylation sites were introduced into LacY, inversion occurred upon addition of kinases. 59 However, because cholesterol and sphingomyelin also affect LacY topology, and because these components affect the electrostatic and mechanical properties of the bilayer, direct correlations between topology and charge were not drawn. 59 Nevertheless, this work raises the possibility of signal transduction events flipping membrane protein topologies, although natural systems like this have, to our knowledge, not yet been discovered.

While several studies suggest that direct interaction between charged residues and lipid headgroups controls membrane protein function and topology, many other studies show that the regulatory nature of these electrostatic interactions is broader and more complex. When LacY was expressed in cells where PE was replaced with PC, it was able to assume a near wild‐type topological organization capable of uphill transport. 60 However, this was only observed with a mixture of 70% PC, 27% CL, and 3% PG, where PC had at least one saturated alkyl chain or a Δ9 trans‐fatty acid. When reconstituted into liposomes with dioleoyl‐PC, native topology was observed, but uphill transport was not. 44 , 45 , 49 When LacY was expressed in cells where PE was replaced with monoglucosyldiacylglycerol (MGlcDAG), normal topology and uphill transport occurred. 19 , 21 Yet diglucosyl diacylglycerol (GlcGlcDAG) only restored topology, but not active transport. 20 The ability of PE, PC, and MGlcDAG to be used interchangeably to support native or near‐native topology and function of LacY suggests that the ethanolamine headgroup of PE is not a specific requirement. PC is a bilayer lipid whose head group cannot act as a hydrogen bond donor, while PE and GlcDAG are nonbilayer lipids that can potentially form hydrogen bonds. PE and PC are zwitterionic, while MGlcDAG is nonionic. This diversity suggests that other more global properties play a role in the interactions between lipids and membrane proteins.

Generally, it appears that there must be sufficient neutral or zwitterionic lipids to favor correct topology. It is possible that these neutral lipids effectively dilute the charged lipids and reduce the membrane surface charge density that the protein encounters. The high quantities of zwitterionic PE and neutral MGlcDAG in bacterial membranes, 14 , 19 , 21 , 61 , 62 , 63 and the need for up to 70% of both in engineered systems 19 , 50 , 58 , 64 , 65 suggests that these net neutral lipids indeed play a crucial role in dampening the negative surface charge of anionic lipids. 20 , 48 Consistent with this idea, increasing the anionic lipid content increases the fraction of inverted LacY 64 and can induce the translocation of Pf3 coat protein. 66 Moreover, single molecule force spectroscopy experiments showed that LacY adopts its native conformation when reconstituted into PE, but is equally likely to assume an aberrant fold when reconstituted into PG. 67 Altogether, studies overwhelmingly suggest that anionic lipids and acidic residues can drive changes in topology while net neutral lipids suppress this effect, perhaps by lowering repulsion between anionic lipids and the protein, or by altering protein charge by modulating the pKa of ionizable side chains. 67

3.2. Lipid charge and in vitro reconstitution

Lipid charge is an important consideration in membrane protein reconstitution experiments as it influences membrane protein insertion and refolding yields. For example, MscL insertion and diacylglycerol kinase (DGK) folding yields were optimal in pure negatively charged DOPG liposomes. 68 , 69 Similarly, GlpG refolding rates improved as the negatively charged DMPG concentration in DMPC/CHAPSO bicelles increased, reaching an approximately sevenfold increase at a DMPG fraction of 30% compared to 0% DMPG. 70

Membrane charge can also play a practical role in defining the orientation of proteins in artificial liposomes. The N‐terminus of proteorhodopsin has many negatively charged residues, and the C terminus has many positively charged residues. Insertion in positively charged DOPC‐DOTAP liposomes favors an outward‐facing N‐terminal orientation, while insertion into negatively charged POPC‐POPG liposomes favors an outward‐facing C‐terminal orientation. 71 In zwitterionic vesicles, neither orientation is favored, and pH, but not alkyl chain saturation, can be used to control proteorhodopsin orientation. Moreover, bacteriorhodopsin, which has a similar structure and approximately 24% sequence homology but does not have asymmetrically charged termini, does not show preferential topological organization. 71 Taken together, these results suggest that pR orientation is determined by its charged termini and the liposome lipid headgroups.

3.3. Lipid charge and membrane protein structure and stability

Negatively charged lipids can alter the structure and stability of the folded state. For example, dimer formation of E. coli MraY requires an anionic lipid. 72 PG is contained within the light‐harvesting complex of photosystem II, and may help stabilize or facilitate trimerization. 73 Increasing the DOPG fraction in DOPC/DOPG vesicles increases the unfolding free energy of LeuT. 74 KcsA tetramer stability also increases with negative charge, with DOPE/DOPG liposomes optimally stabilizing the KcsA tetramer. 75 Replacing DOPG with DOPC decreases KcsA's melting temperature, while replacement with the negatively charged DOPA has no effect. In contrast, adding negatively charged PG, PS, or cardiolipin to PC bilayers destabilized the glycophorin A transmembrane helix dimer. 76 This was attributed to interactions with positively charged residues that could possibly alter the tilt angle of the transmembrane helices. Indeed, when the charged residues at the end of the TM helix were replaced with neutral polar ones, negatively charged lipids did not affect dimer stability. Finally, ATR‐FTIR experiments showed that anionic lipids could alter the helical tilt of EmrE with respect to the bilayer 77 : increasing the POPG fraction in POPG/POPE or POPG/POPC vesicles decreased the tilt angle, but no changes were observed in the net neutral POPE/POPC vesicles. 77 These results clearly demonstrate that lipid charge can help define protein conformation.

3.4. π‐Interactions between membrane proteins and lipids

A prominent characteristic of membrane proteins is the so‐called “aromatic belt,” where aromatic side chains preferentially occupy the interfacial region of membrane bilayers. 78 , 79 , 80 , 81 , 82 Aromatic residues may be preferred in this region because of favorable electrostatic interactions with the polar lipid head groups. These π‐interactions have been directly observed between PE or PC and aromatic residues in cytochrome c oxidase 24 and in the human phosphatidyl transfer protein. 83 Moreover, experiments with the aspartate transporter from Pyrococcus horikoshii (GltPh) suggest that a cation‐π interaction between its Y33 residue and lipid headgroups can regulate aspartate transport and proper transmembrane orientation. 84 These interactions can be quite energetically significant, as aromatic side chains in the interfacial region of the β‐barrel protein OmpA were found to contribute up to 2.6 kcal/mol to the unfolding free energy. 85

3.5. The effects of pKa modulation by direct protein‐lipid hydrogen bonding

The bilayer environment can alter side chain pKa's through general solvent and electrostatic effects. 86 , 87 , 88 Membrane lipids can also interact with proteins directly via hydrogen bonds, 27 thereby modulating residue pKas. For example, when HorA was reconstituted into DOPE vesicles, it could hydrolyze ATP and transport the substrate Hoechst 33342. Reconstitution in DOPC abolished its transport activity, and decreased the mean angle of its transmembrane helices. This change in structure and activity was attributed to PE being a stronger hydrogen donor than PC. In support of this hypothesis, interactions between negatively charged residues and the PE head group have been observed in crystal structures. 89 , 90 Similarly, when LmpR was reconstituted in PC instead of PE, it suffered impaired activity. 91 Increasing methylation of PE, which progressively removes hydrogen bonding potential, stabilized the outward‐open conformation in a pH‐dependent manner, suggesting hydrogen bond interactions. 92 A single D68C mutation mimicked the effects of reconstituting wild‐type LmpR in PC, indicating that hydrogen bonding between D68 and PE could be critical for activity. 91 Further, when LmrP was reconstituted in liposomes containing PE, the mean carboxylate residue pKa shifted to 6.5. When PE was absent, the mean pKa returned to 4.6. 52 , 92

4. BILAYER MECHANICAL PROPERTIES AND MEMBRANE PROTEIN FOLDING

General principles governing the effects of bilayer stresses on membrane protein folding have begun to emerge. Lower pressure in the head group region or low curvature stress can favor engagement with the membrane surface. Higher pressure in the center of the bilayer or a high curvature stress can impede insertion through the bilayer, but can stabilize the structure of the folded protein. Increasing the stress can increase the folding rate by facilitating helix packing. But, if the stress or pressure is too high, the folding rate can be slowed. Nonbilayer lipids can also influence the ease with which the bilayer can deform to accommodate hydrophobic mismatch (Figure 5). The importance of these various influences will depend on the protein and the specific circumstances. In the following sections, we provide examples of how mechanical properties influence insertion and folding in a variety of experimental systems.

4.1. Bilayer mechanics and membrane protein insertion

Numerous studies support the notion that high hydrocarbon chain lateral pressure can hinder uncatalyzed protein insertion through the bilayer. One of the first studies to directly demonstrate this link attempted to refold bR into various PC/PE vesicles. 93 The regeneration yield of bR decreased as the PE content, and hence hydrocarbon chain pressure, increased. Moreover, at a fixed PE mole fraction of 0.16, the regeneration yield decreased as the alkyl chain length, and hence chain pressure or bilayer thickness, increased. Subsequent mechanistic work revealed that bilayer stress influenced insertion, rather than folding after insertion. 94

Studies with other proteins mirrored the results observed with bR. The activation energy of insertion for a designed transmembrane helix was observed to increase as the fraction of DOPE in DOPC/DOPE vesicles increased. 95 EmrE reconstitution into DOPC/DOPE liposomes likewise decreased as the DOPE fraction increased, though the observed decrease was found to depend on the reconstitution method. 96 LacY reconstitution into DOPC/DOPE liposomes followed the same trend. 64 When LacY was refolded into DOPE/DOPC liposomes, increasing the DOPE concentration decreased the yield of folded protein. 64 Indeed, optimal reconstitution occurred at high DOPG fractions, which would likely reduce the bilayer bending rigidity 97 while adding negative charge. Finally, DsbB preferentially inserted into DMPC liposomes, and had lower insertion rates in liposomes with high PE content. 98

Although most studies suggest that high lipid chain pressure can impede protein insertion, it is likely to be protein‐dependent. GlpG had the highest insertion rates in 1:1 PG:PE bilayers, and the lowest insertion rates in pure DMPC liposomes, 98 the exact opposite of DsbB. In single molecule folding experiments, GlpG refolding rates increased from approximately 15% to approximately 60% when the bilayer curvature was increased by reducing vesicle size, thereby presumably reducing the barrier to insertion into the head group region. 70 GalP also preferred higher lateral pressures: optimal GalP refolding in DOPE/DOPC vesicles was observed as the DOPE fraction was increased up to 60%. 99 Similarly, incorporation of KcsA into DOPE/DOPC vesicles increased as the fraction of PE increased. The authors postulated that the smaller head size of PE facilitated insertion, presumably at the stage of interaction with the membrane surface. 75 Thermostabilized turkey β1AR also seemed to prefer an environment with high lateral chain pressure: solubilization in PC or PS lipids increased as the acyl chain length or degree of unsaturation increased, 100 with DOPC (18:1/18:1 chains) being preferred over POPC (16:0/18:1) or DMPC (14:0/14:0), and DOPS (18:1/18:1) over POPS (16:0/18:1). Moreover, its activity increased as the acyl chain length of PC, PG, or PS lipids increased. However, the effects of bilayer thickness cannot be excluded, and the charged DMPG, POPG, and DOPG lipids solubilized β1AR nearly as well as DOPC.

Refolding experiments with diacylglycerol kinase (DGK) suggest that curvature elastic stress also affects the ability to insert into membranes. Adding increasing fractions of DMPC or lysoOPC to DOPC decreased the final yield of refolded protein by slowing the rate of formation of protein‐vesicle complexes. Both DMPC and lysoOPC decrease the stored curvature elastic stress, which lowers lateral chain pressure and increases pressure in the head group region. This result is thus consistent with the idea that higher pressure in the head group region can hinder the ability of proteins to engage the membrane surface. Interestingly, DOPE had little effect on the folding rate, and only slightly increased the final yield. DOPG had the biggest effect on both the folding rate and yield of functionally refolded protein. Charge was unlikely to be the primary factor since increasing the ionic strength of the buffers or substituting another anionic lipid had little effect on DGK folding rates or yields. Rather, the effect of PG was attributed to headgroup interactions. 69

4.2. Effects of lipids on membrane protein folding rates

While relatively few studies have investigated the effects of lipid composition on the kinetics of membrane protein folding, those that have support the notion that membrane pressure can affect folding rates in addition to the insertion process. When EmrE was refolded or reconstituted into DOPC/DOPE and DOPG/DOPE vesicles, the rate of folding increased as the percentage of PE increased. However, the amount of functional protein recovered decreased. This is consistent with the hypothesis that an increase in lateral pressure can inhibit insertion but facilitate packing once transmembrane helices are inserted. 77 A series of experiments with bR likewise demonstrated that membrane rigidity can control folding rates. When bR was folded into DMPC/DHPC micelles, the rate constants for a rate‐limiting folding intermediate decreased as the DMPC concentration (and thus lateral pressure) was increased. 31 Experiments in DPoPC bilayers, which are more stressed than DMPC/DHPC bilayers, suggested that bilayer lateral pressure directly affects the rate of formation and relative population of two intermediate bR folding states. 101 When the curvature stress of DPoPC vesicles was increased by adding DPoPE, the rate constants for noncovalent retinal binding and final folding to bR slowed. 94

4.3. Lipid bilayer properties and membrane protein activity

Early studies suggested that lipid phase behavior was critical for proper folding and activity of membrane proteins. Both monogalactosyldiglyceride (MGDG) and PE lipids tend to form hexagonal II structures, and increasing the fraction of either lipid in DOPC vesicles increased Ca2+ uptake by Ca2+ ‐ATPase. Methylation of DOPE or glycosylation of MGDG, which decreases their ability to form the hexagonal II phase and increases their tendency to form bilayers, decreased ATP‐dependent Ca2+ uptake. Activity in the bilayer lipids PS, PI, DOPC, or DOPG was minimal. 102 Similarly, leucine transport by the Branched‐Chain Amino Acid Transport System of S. cremoris increased as the PE fraction in PE/PC membranes increased, and as the glycolipid (GL) content in GL/PC membranes increased. Higher transport was observed with MGDG than digalactosyldiglyceride (DGDG), and leucine transport decreased as PE was increasingly methylated. 62 These results validated those with the Ca2+ ‐ATPase, indicating that nonbilayer lipids impart properties to the membrane critical for optimal protein activity. Because high activity for the Branched‐Chain Transporter was also observed in POPC/POPS vesicles, and because POPS can increase the bending rigidity of POPC bilayers, 97 the effect of these nonbilayer lipids on membrane protein activity is likely mechanical in nature.

Several studies support the notion that variations in protein activity observed upon changes in lipid composition may be due to structural alterations. 35 When alamethicin was reconstituted into DOPC/DOPE vesicles, the probability of higher conductance levels increased as the DOPE mole fraction increased. The increase in probability was proportional to an increase in the curvature stress. Because high conductance states are associated with multimeric channel aggregates, this suggests that the increased curvature stress stabilized alamethicin in a multimeric form. 103 Further, when DGK was reconstituted into PC liposomes, optimal activity was observed with 18 carbon acyl chains, which minimize the degree of hydrophobic mismatch. Longer or shorter chains resulted in lower activities. 104 Similar results were observed for the Ca2+‐ATPase, 105 and for MelB. 106 Tilting of the transmembrane helices or altered helix packing likely explains the effect of bilayer thickness on protein activity. 104 , 105

4.4. Bilayer mechanics and membrane protein stability

High bilayer curvature stress and high lateral pressure have been shown to increase membrane protein stability. Bacteriorhodopsin was more stable to irreversible thermal denaturation in mixed DMPC/DOPC vesicles than in DMPC vesicles, 94 presumably due to the lower curvature stress of DMPC. The unfolding free energy of LeuT was found to increase as the PE content in PC liposomes increased. 74 Increasing the DOPE fraction in DOPE/DOPC vesicles increased the lifetime of alamethicin channels and the denaturation temperature for LacY. 64 , 103 Adding PE to PC vesicles stabilized the GpATM dimer, while adding lysoPC (and thus lowering chain lateral pressure) destabilized it. 76 KcsA tetramer formation in DOPC/DOPE vesicles increased as the PE fraction increased to approximately 40%. 75 Furthermore, a higher fraction of the denaturant trifluoroethanol (TFE) was required to induce tetramer dissociation in vesicles containing PE compared to those without PE. In this case, PE may have stabilized the protein directly, or countered the bilayer destabilization and pressure‐reducing effects of TFE.

Lateral pressure can also influence thermodynamic stability, as suggested by studies with the peptide alamethicin. 12 Alamethicin insertion is reversible, allowing free energy measurements between the water‐solubilized and membrane‐inserted states. The free energy of alamethicin insertion into DOPC/DOPE vesicles increased with increasing DOPE. Moreover, the free energy decreased with each methyl group added to the DOPE headgroup, or as the alkyl chains were increasingly saturated, thereby reducing curvature stress. 12 These results suggest that the inserted state is destabilized relative to the common water‐solubilized reference state, and is directly correlated with bilayer mechanics. Alamethicin has a relatively short hydrophobic segment, implying that the bilayer must thin to accommodate the inserted form. It was therefore suggested that nonbilayer lipids destabilize the inserted state by making it harder to locally distort the bilayer (see Figure 5).

5. CONCLUSION

The lipid bilayer is a complex system that interacts with membrane proteins and affects their structure and function. Indeed, the bilayer's electrostatic and mechanical properties can have diverse effects on membrane protein insertion, folding, stability, and activity. Although most studies focused on the effects of electrostatic or mechanical interactions in isolation, a growing number demonstrate that both are critical. 20 , 64 , 74 , 75 , 100 , 107 , 108 , 109

Lipid mixtures may be necessary to balance the disparate effects of electromechanical properties on membrane proteins. The bilayer must contain enough bilayer lipids to maintain structural integrity, but enough nonbilayer lipids to allow for essential destabilizing events like budding, mitosis, and fusion. 10 , 30 The membrane must be relaxed enough to allow for insertion, but stressed enough for optimal folding and stability. 12 , 94 , 103 There must be enough neutral lipids to dilute negative surface charge to promote correct topology, but enough charged lipids to stabilize tertiary structures and enhance activity. It is no surprise then, that mixtures of lipids are generally more favorable for folding than vesicles made from pure lipids. 45 , 64 , 67 , 75 , 107 The composition cells choose likely facilitates all critical functions without necessarily optimizing them 64 (Figure 6). Many studies found that the optimal lipid content for in vitro folding studies was near physiological values, 50 , 62 , 65 , 67 , 75 , 107 , 109 , 110 , 111 which could perhaps be expected since the protein and bilayer composition must have evolved together.

FIGURE 6.

FIGURE 6

Lipid mixtures are necessary to simultaneously ensure proper folding, stability, insertion, and activity. LacY reconstitution, refolding, topology, and activity were assayed in different compositions of DOPE, DOPG, and DOPC, and the optimal concentrations for each individual parameter, as well as for all of them together, are shown. (a) Legend showing the relative fractions of DOPE, DOPG, and DOPC on each gridline. The corners labeled DOPE, DOPG, and DOPC represent samples with 100% of each lipid, respectively. The edges represent bilayer mixtures. For example, the edge between DOPE and DOPC represents samples made of those two lipids in varying percentages, where the orange numbers along the edge state the DOPE fraction. Each point or corner in the center of the large triangle represents samples with a mixture of all three lipids, with the numbers along the edge describing the exact composition at each point. (b) The reconstitution of LacY is optimal at high DOPG fractions, and impeded at low DOPE fractions. (c) The refolding of LacY is inhibited at high DOPE and DOPC fractions. (d) High DOPG fractions impede correct topological organization of LacY. (e) LacY transport activity is optimal at high DOPE fractions. (f) The lipid fractions at which LacY reconstitution, refolding, topology, and activity are all favorable, even if they are not all optimal. Only a subset of lipid compositions can support both proper protein folding and function. In (b)‐(e), each colored dot represents the results of an experiment at the indicated lipid composition. The color of the dot shows how well each lipid composition supports proper LacY structure and activity, with green dots representing optimal function or folding and red dots illustrating when the protein is most compromised. Adapted from Sci Rep 7(1):13056, Findlay HE and Booth PJ, The folding, stability, and function of lactose permease differ in their dependence on bilayer lipid composition, Copyright (2017) 64

Although some general principles governing lipid‐protein interactions have begun to emerge, most studies on lipid composition and membrane protein folding have been anecdotal and hard to generalize. Indeed, lipids can have disparate and even opposite effects on different proteins. This may reflect differences in the energetic barriers to folding. For example, the differential effects of lateral packing pressure on membrane protein insertion may reflect the relative importance of initial engagement of the protein with the bilayer interface versus insertion through the bilayer. To better understand the interplay between membrane proteins and bilayer properties, systematic studies varying both lipid composition and protein sequence will likely be necessary. It may also be necessary to characterize the mechanics and phase behavior of additional lipid mixtures. This information can accelerate future membrane protein studies, which may aid in the understanding of misfolding diseases and the development of drugs.

CONFLICT OF INTEREST

K.C. declares no conflict of interest. J.U.B. has founded a company involved in the production of natural chemicals that bears no direct relation to this work.

AUTHOR CONTRIBUTIONS

Karolina Corin wrote the initial draft and Karolina Corin and James U. Bowie worked together on revisions.

ACKNOWLEDGEMENTS

This work was supported by NIH grant R01GM063919 to J.U.B.

Corin K, Bowie JU. How bilayer properties influence membrane protein folding. Protein Science. 2020;29:2348–2362. 10.1002/pro.3973

James Bowie is the winner of the 2020 Stein & Moore Award.

Endnotes

1

Each of these properties can be, and have been, adjusted experimentally. For example, bilayer thickness can be adjusted by modifying the number of carbons in the alkyl chains. DHPC (6 saturated carbons) will provide a thinner layer than DMPC (14 saturated carbons), and DPoPC (16 carbons, Δ9‐Cis unsaturation) will make a thinner layer than DOPC (18 carbons, Δ9‐Cis unsaturation). The lateral chain pressure can be increased by adding PE lipids (non‐bilayer) to a PC (bilayer) background, or by increasing the acyl chain length. This pressure can be decreased by adding lysophospholipids. Membrane charge can be controlled via the headgroup. Phosphatidylserine and phosphatidylglycerol headgroups are negatively charged, while the zwitterionic phosphatidylcholine and phosphatidylethanolamine are net neutral.

2

Initially, only the orientation of the loop between TMVI and TMVII was shown to be reversible. However, because activity was restored, and because active transport depends on the interaction between TMII, TMVII, and TMXI, it implied that the orientation of additional helices must have been reversible. Subsequent experiments demonstrated that the orientations of TMIII‐TMVI are reversible.49

REFERENCES

  • 1. Casares D, Escribá PV, Rosselló CA. Membrane lipid composition: Effect on membrane and organelle structure, function and compartmentalization and therapeutic avenues. Int J Mol Sci. 2019;20:2167. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2. Sohlenkamp C, Geiger O. Bacterial membrane lipids: Diversity in structures and pathways. FEMS Microbiol Rev. 2016;40:133–159. [DOI] [PubMed] [Google Scholar]
  • 3. Dowhan W, Vitrac H, Bogdanov M. Lipid‐assisted membrane protein folding and topogenesis. Protein J. 2019;38:274–288. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4. Harayama T, Riezman H. Understanding the diversity of membrane lipid composition. Nat Rev Mol Cell Biol. 2018;19:281–296. [DOI] [PubMed] [Google Scholar]
  • 5. Shevchenko A, Simons K. Lipidomics: Coming to grips with lipid diversity. Nat Rev Mol Cell Biol. 2010;11:593–598. [DOI] [PubMed] [Google Scholar]
  • 6. Morein S, Andersson A, Rilfors L, Lindblom G. Wild‐type Escherichia coli cells regulate the membrane lipid composition in a “window” between gel and non‐lamellar structures. J Biol Chem. 1996;271:6801–6809. [DOI] [PubMed] [Google Scholar]
  • 7. Hazel JR, Williams EE. The role of alterations in membrane lipid composition in enabling physiological adaptation of organisms to their physical environment. Prog Lipid Res. 1990;29:167–227. [DOI] [PubMed] [Google Scholar]
  • 8. Brenna JT, Salem N, Sinclair AJ, Cunnane SC, International Society for the Study of Fatty Acids and Lipids, ISSFAL . Alpha‐Linolenic acid supplementation and conversion to n‐3 long‐chain polyunsaturated fatty acids in humans. Prostaglandins Leukot Essent Fatty Acids. 2009;80:85–91. [DOI] [PubMed] [Google Scholar]
  • 9. Ryan AS, Astwood JD, Gautier S, Kuratko CN, Nelson EB, Salem N. Effects of long‐chain polyunsaturated fatty acid supplementation on neurodevelopment in childhood: A review of human studies. Prostaglandins Leukot Essent Fatty Acids. 2010;82:305–314. [DOI] [PubMed] [Google Scholar]
  • 10. Lindblom G, Hauksson JB, Rilfors L, Bergenståhl B, Wieslander A, Eriksson PO. Membrane lipid regulation in Acholeplasma laidlawii grown with saturated fatty acids. Biosynthesis of a triacylglucolipid forming reversed micelles. J Biol Chem. 1993;268:16198–16207. [PubMed] [Google Scholar]
  • 11. Vikström S, Li L, Karlsson OP, Wieslander A. Key role of the diglucosyldiacylglycerol synthase for the nonbilayer‐bilayer lipid balance of Acholeplasma laidlawii membranes. Biochemistry. 1999;38:5511–5520. [DOI] [PubMed] [Google Scholar]
  • 12. Lewis JR, Cafiso DS. Correlation between the free energy of a channel‐forming voltage‐gated peptide and the spontaneous curvature of bilayer lipids. Biochemistry. 1999;38:5932–5938. [DOI] [PubMed] [Google Scholar]
  • 13. Osterberg F, Rilfors L, Wieslander A, Lindblom G, Gruner SM. Lipid extracts from membranes of Acholeplasma laidlawii A grown with different fatty acids have a nearly constant spontaneous curvature. Biochim Biophys Acta. 1995;1257:18–24. [DOI] [PubMed] [Google Scholar]
  • 14. Rietveld AG, Killian JA, Dowhan W, de Kruijff B. Polymorphic regulation of membrane phospholipid composition in Escherichia coli . J Biol Chem. 1993;268:12427–12433. [PubMed] [Google Scholar]
  • 15. Chang G, Spencer RH, Lee AT, Barclay MT, Rees DC. Structure of the MscL homolog from Mycobacterium tuberculosis: A gated mechanosensitive ion channel. Science. 1998;282:2220–2226. [DOI] [PubMed] [Google Scholar]
  • 16. Perozo E, Cortes DM, Sompornpisut P, Kloda A, Martinac B. Open channel structure of MscL and the gating mechanism of mechanosensitive channels. Nature. 2002;418:942–948. [DOI] [PubMed] [Google Scholar]
  • 17. Perozo E, Rees DC. Structure and mechanism in prokaryotic mechanosensitive channels. Curr Opin Struct Biol. 2003;13:432–442. [DOI] [PubMed] [Google Scholar]
  • 18. Dowhan W, Mileykovskaya E, Bogdanov M. Diversity and versatility of lipid‐protein interactions revealed by molecular genetic approaches. Biochim Biophys Acta. 2004;1666:19–39. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19. Wikström M, Xie J, Bogdanov M, et al. Monoglucosyldiacylglycerol, a foreign lipid, can substitute for phosphatidylethanolamine in essential membrane‐associated functions in Escherichia coli . J Biol Chem. 2004;279:10484–10493. [DOI] [PubMed] [Google Scholar]
  • 20. Wikström M, Kelly AA, Georgiev A, et al. Lipid‐engineered Escherichia coli membranes reveal critical lipid headgroup size for protein function. J Biol Chem. 2009;284:954–965. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21. Xie J, Bogdanov M, Heacock P, Dowhan W. Phosphatidylethanolamine and monoglucosyldiacylglycerol are interchangeable in supporting topogenesis and function of the polytopic membrane protein lactose permease. J Biol Chem. 2006;281:19172–19178. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22. Lee AG. Lipid‐protein interactions in biological membranes: A structural perspective. Biochim Biophys Acta. 2003;1612:1–40. [DOI] [PubMed] [Google Scholar]
  • 23. Zaydman MA, Silva JR, Delaloye K, et al. Kv7.1 ion channels require a lipid to couple voltage sensing to pore opening. Proc Natl Acad Sci USA. 2013;110:13180–13185. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24. Shinzawa‐Itoh K, Aoyama H, Muramoto K, et al. Structures and physiological roles of 13 integral lipids of bovine heart cytochrome c oxidase. EMBO J. 2007;26:1713–1725. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25. Li L, Shi X, Guo X, Li H, Xu C. Ionic protein‐lipid interaction at the plasma membrane: What can the charge do? Trends Biochem Sci. 2014;39:130–140. [DOI] [PubMed] [Google Scholar]
  • 26. Laganowsky A, Reading E, Allison TM, et al. Membrane proteins bind lipids selectively to modulate their structure and function. Nature. 2014;510:172–175. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27. Wiener MC. A census of ordered lipids and detergents in X‐ray crystal structures of integral membrane proteins Protein–lipid interactions, Weinheim, Germany: John Wiley & Sons, Ltd, 2006; p. 95–117. [Google Scholar]
  • 28. Hong H, Tamm LK. Elastic coupling of integral membrane protein stability to lipid bilayer forces. Proc Natl Acad Sci USA. 2004;101:4065–4070. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29. Burgess NK, Dao TP, Stanley AM, Fleming KG. Beta‐barrel proteins that reside in the Escherichia coli outer membrane in vivo demonstrate varied folding behavior in vitro. J Biol Chem. 2008;283:26748–26758. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30. Gruner SM. Intrinsic curvature hypothesis for biomembrane lipid composition: A role for nonbilayer lipids. Proc Natl Acad Sci USA. 1985;82:3665–3669. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31. Booth PJ, Riley ML, Flitsch SL, et al. Evidence that bilayer bending rigidity affects membrane protein folding. Biochemistry. 1997;36:197–203. [DOI] [PubMed] [Google Scholar]
  • 32. Bezrukov SM. Functional consequences of lipid packing stress. Curr Opin Colloid Interface Sci. 2000;5:237–243. [Google Scholar]
  • 33. Booth PJ. Sane in the membrane: Designing systems to modulate membrane proteins. Curr Opin Struct Biol. 2005;15:435–440. [DOI] [PubMed] [Google Scholar]
  • 34. van den Brink‐van der Laan E, Killian JA, de Kruijff B. Nonbilayer lipids affect peripheral and integral membrane proteins via changes in the lateral pressure profile. Biochim Biophys Acta. 2004;1666:275–288. [DOI] [PubMed] [Google Scholar]
  • 35. Cantor RS. Lateral pressures in cell membranes: A mechanism for modulation of protein function. J Phys Chem B. 1997;101:1723–1725. [Google Scholar]
  • 36. Andersen OS, Koeppe RE. Bilayer thickness and membrane protein function: An energetic perspective. Annu Rev Biophys Biomol Struct. 2007;36:107–130. [DOI] [PubMed] [Google Scholar]
  • 37. Weiss TM, van der Wel PCA, Killian JA, Koeppe RE, Huang HW. Hydrophobic mismatch between helices and lipid bilayers. Biophys J. 2003;84:379–385. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38. de Jesus AJ, Allen TW. The determinants of hydrophobic mismatch response for transmembrane helices. Biochim Biophys Acta. 2013;1828:851–863. [DOI] [PubMed] [Google Scholar]
  • 39. Killian JA. Hydrophobic mismatch between proteins and lipids in membranes. Biochim Biophys Acta. 1998;1376:401–415. [DOI] [PubMed] [Google Scholar]
  • 40. Williamson IM, Alvis SJ, East JM, Lee AG. Interactions of phospholipids with the potassium channel KcsA. Biophys J. 2002;83:2026–2038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41. Powl AM, East JM, Lee AG. Lipid‐protein interactions studied by introduction of a tryptophan residue: The mechanosensitive channel MscL. Biochemistry. 2003;42:14306–14317. [DOI] [PubMed] [Google Scholar]
  • 42. Page MG, Rosenbusch JP, Yamato I. The effects of pH on proton sugar symport activity of the lactose permease purified from Escherichia coli . J Biol Chem. 1988;263:15897–15905. [PubMed] [Google Scholar]
  • 43. Bogdanov M, Umeda M, Dowhan W. Phospholipid‐assisted refolding of an integral membrane protein. Minimum structural features for phosphatidylethanolamine to act as a molecular chaperone. J Biol Chem. 1999;274:12339–12345. [DOI] [PubMed] [Google Scholar]
  • 44. Chen CC, Wilson TH. The phospholipid requirement for activity of the lactose carrier of Escherichia coli . J Biol Chem. 1984;259:10150–10158. [PubMed] [Google Scholar]
  • 45. Seto‐Young D, Chen CC, Wilson TH. Effect of different phospholipids on the reconstitution of two functions of the lactose carrier of Escherichia coli . J Membr Biol. 1985;84:259–267. [DOI] [PubMed] [Google Scholar]
  • 46. Bogdanov M, Dowhan W. Phosphatidylethanolamine is required for in vivo function of the membrane‐associated lactose permease of Escherichia coli . J Biol Chem. 1995;270:732–739. [DOI] [PubMed] [Google Scholar]
  • 47. Bogdanov M, Heacock PN, Dowhan W. A polytopic membrane protein displays a reversible topology dependent on membrane lipid composition. EMBO J. 2002;21:2107–2116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48. Bogdanov M, Xie J, Heacock P, Dowhan W. To flip or not to flip: Lipid‐protein charge interactions are a determinant of final membrane protein topology. J Cell Biol. 2008;182:925–935. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49. Wang X, Bogdanov M, Dowhan W. Topology of polytopic membrane protein subdomains is dictated by membrane phospholipid composition. EMBO J. 2002;21:5673–5681. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50. Vitrac H, Bogdanov M, Dowhan W. In vitro reconstitution of lipid‐dependent dual topology and postassembly topological switching of a membrane protein. Proc Natl Acad Sci USA. 2013;110:9338–9343. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51. Vitrac H, MacLean DM, Jayaraman V, Bogdanov M, Dowhan W. Dynamic membrane protein topological switching upon changes in phospholipid environment. Proc Natl Acad Sci USA. 2015;112:13874–13879. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52. Gbaguidi B, Hakizimana P, Vandenbussche G, Ruysschaert J‐M. Conformational changes in a bacterial multidrug transporter are phosphatidylethanolamine‐dependent. Cell Mol Life Sci. 2007;64:1571–1582. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53. Pi J, Pittard AJ. Topology of the phenylalanine‐specific permease of Escherichia coli . J Bacteriol. 1996;178:2650–2655. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54. Zhang W, Bogdanov M, Pi J, Pittard AJ, Dowhan W. Reversible topological organization within a polytopic membrane protein is governed by a change in membrane phospholipid composition. J Biol Chem. 2003;278:50128–50135. [DOI] [PubMed] [Google Scholar]
  • 55. Vitrac H, Bogdanov M, Heacock P, Dowhan W. Lipids and topological rules of membrane protein assembly: Balance between long and short range lipid‐protein interactions. J Biol Chem. 2011;286:15182–15194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56. Hu LA, King SC. Membrane topology of the Escherichia coli gamma‐aminobutyrate transporter: Implications on the topography and mechanism of prokaryotic and eukaryotic transporters from the APC superfamily. Biochem J. 1998;336:69–76. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57. Zhang W, Campbell HA, King SC, Dowhan W. Phospholipids as determinants of membrane protein topology. Phosphatidylethanolamine is required for the proper topological organization of the gamma‐aminobutyric acid permease (GabP) of Escherichia coli . J Biol Chem. 2005;280:26032–26038. [DOI] [PubMed] [Google Scholar]
  • 58. Vitrac H, Dowhan W, Bogdanov M. Effects of mixed proximal and distal topogenic signals on the topological sensitivity of a membrane protein to the lipid environment. Biochim Biophys Acta Biomembr. 2017;1859:1291–1300. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59. Vitrac H, MacLean DM, Karlstaedt A, et al. Dynamic lipid‐dependent modulation of protein topology by post‐translational phosphorylation. J Biol Chem. 2017;292:1613–1624. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60. Bogdanov M, Heacock P, Guan Z, Dowhan W. Plasticity of lipid‐protein interactions in the function and topogenesis of the membrane protein lactose permease from Escherichia coli . Proc Natl Acad Sci USA. 2010;107:15057–15062. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61. Raetz CR. Enzymology, genetics, and regulation of membrane phospholipid synthesis in Escherichia coli . Microbiol Rev. 1978;42:614–659. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62. Driessen AJ, Zheng T, In't Veld G, Op den Kamp JA, Konings WN. Lipid requirement of the branched‐chain amino acid transport system of Streptococcus cremoris . Biochemistry. 1988;27:865–872. [DOI] [PubMed] [Google Scholar]
  • 63. Steinick LE, Wieslander A, Johansson KE, Liss A. Membrane composition and virus susceptibility of Acholeplasma laidlawii . J Bacteriol. 1980;143:1200–1207. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64. Findlay HE, Booth PJ. The folding, stability and function of lactose permease differ in their dependence on bilayer lipid composition. Sci Rep. 2017;7:13056. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65. Barrera FN, Renart ML, Poveda JA, de Kruijff B, Killian JA, González‐Ros JM. Protein self‐assembly and lipid binding in the folding of the potassium channel KcsA. Biochemistry. 2008;47:2123–2133. [DOI] [PubMed] [Google Scholar]
  • 66. Ridder AN, Kuhn A, Killian JA, de Kruijff B. Anionic lipids stimulate sec‐independent insertion of a membrane protein lacking charged amino acid side chains. EMBO Rep. 2001;2:403–408. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67. Serdiuk T, Sugihara J, Mari SA, Kaback HR, Müller DJ. Observing a lipid‐dependent alteration in single lactose permeases. Structure. 2015;23:754–761. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68. Harris NJ, Charalambous K, Findlay HE, Booth PJ. Lipids modulate the insertion and folding of the nascent chains of alpha helical membrane proteins. Biochem Soc Trans. 2018;46:1355–1366. [DOI] [PubMed] [Google Scholar]
  • 69. Seddon AM, Lorch M, Ces O, Templer RH, Macrae F, Booth PJ. Phosphatidylglycerol lipids enhance folding of an alpha helical membrane protein. J Mol Biol. 2008;380:548–556. [DOI] [PubMed] [Google Scholar]
  • 70. Choi H‐K, Min D, Kang H, et al. Watching helical membrane proteins fold reveals a common N‐to‐C‐terminal folding pathway. Science. 2019;366:1150–1156. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71. Tunuguntla R, Bangar M, Kim K, Stroeve P, Ajo‐Franklin CM, Noy A. Lipid bilayer composition can influence the orientation of proteorhodopsin in artificial membranes. Biophys J. 2013;105:1388–1396. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72. Henrich E, Peetz O, Hein C, et al. Analyzing native membrane protein assembly in nanodiscs by combined non‐covalent mass spectrometry and synthetic biology. eLife. 2017;6:e20954. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73. Nussberger S, Dörr K, Wang DN, Kühlbrandt W. Lipid‐protein interactions in crystals of plant light‐harvesting complex. J Mol Biol. 1993;234:347–356. [DOI] [PubMed] [Google Scholar]
  • 74. Sanders MR, Findlay HE, Booth PJ. Lipid bilayer composition modulates the unfolding free energy of a knotted α‐helical membrane protein. Proc Natl Acad Sci USA. 2018;115:E1799–E1808. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75. van Dalen A, Hegger S, Killian JA, de Kruijff B. Influence of lipids on membrane assembly and stability of the potassium channel KcsA. FEBS Lett. 2002;525:33–38. [DOI] [PubMed] [Google Scholar]
  • 76. Hong H, Bowie JU. Dramatic destabilization of transmembrane helix interactions by features of natural membrane environments. J Am Chem Soc. 2011;133:11389–11398. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77. Miller D, Charalambous K, Rotem D, Schuldiner S, Curnow P, Booth PJ. In vitro unfolding and refolding of the small multidrug transporter EmrE. J Mol Biol. 2009;393:815–832. [DOI] [PubMed] [Google Scholar]
  • 78. Chamberlain AK, Faham S, Yohannan S, Bowie JU. Construction of helix‐bundle membrane proteins. Adv Protein Chem. 2003;63:19–46. [DOI] [PubMed] [Google Scholar]
  • 79. Wimley WC, White SH. Experimentally determined hydrophobicity scale for proteins at membrane interfaces. Nat Struct Biol. 1996;3:842–848. [DOI] [PubMed] [Google Scholar]
  • 80. Killian JA, von Heijne G. How proteins adapt to a membrane‐water interface. Trends Biochem Sci. 2000;25:429–434. [DOI] [PubMed] [Google Scholar]
  • 81. Schramm CA, Hannigan BT, Donald JE, et al. Knowledge‐based potential for positioning membrane‐associated structures and assessing residue‐specific energetic contributions. Structure. 2012;20:924–935. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82. Senes A, Chadi DC, Law PB, Walters RFS, Nanda V, Degrado WF. E(z), a depth‐dependent potential for assessing the energies of insertion of amino acid side‐chains into membranes: Derivation and applications to determining the orientation of transmembrane and interfacial helices. J Mol Biol. 2007;366:436–448. [DOI] [PubMed] [Google Scholar]
  • 83. Roderick SL, Chan WW, Agate DS, et al. Structure of human phosphatidylcholine transfer protein in complex with its ligand. Nat Struct Biol. 2002;9:507–511. [DOI] [PubMed] [Google Scholar]
  • 84. McIlwain BC, Vandenberg RJ, Ryan RM. Transport rates of a glutamate transporter homologue are influenced by the lipid bilayer. J Biol Chem. 2015;290:9780–9788. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85. Hong H, Park S, Jiménez RHF, Rinehart D, Tamm LK. Role of aromatic side chains in the folding and thermodynamic stability of integral membrane proteins. J Am Chem Soc. 2007;129:8320–8327. [DOI] [PubMed] [Google Scholar]
  • 86. Kyrychenko A, Vasquez‐Montes V, Ulmschneider MB, Ladokhin AS. Lipid headgroups modulate membrane insertion of pHLIP peptide. Biophys J. 2015;108:791–794. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87. Ladokhin AS, White SH. Interfacial folding and membrane insertion of a designed helical peptide. Biochemistry. 2004;43:5782–5791. [DOI] [PubMed] [Google Scholar]
  • 88. Vasquez‐Montes V, Vargas‐Uribe M, Pandey NK, Rodnin MV, Langen R, Ladokhin AS. Lipid‐modulation of membrane insertion and refolding of the apoptotic inhibitor Bcl‐xL. Biochim Biophys Acta Proteins Proteomics. 2019;1867:691–700. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89. Nogi T, Fathir I, Kobayashi M, Nozawa T, Miki K. Crystal structures of photosynthetic reaction center and high‐potential iron‐sulfur protein from Thermochromatium tepidum: Thermostability and electron transfer. Proc Natl Acad Sci USA. 2000;97:13561–13566. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90. Lange C, Nett JH, Trumpower BL, Hunte C. Specific roles of protein‐phospholipid interactions in the yeast cytochrome bc1 complex structure. EMBO J. 2001;20:6591–6600. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91. Hakizimana P, Masureel M, Gbaguidi B, Ruysschaert J‐M, Govaerts C. Interactions between phosphatidylethanolamine headgroup and LmrP, a multidrug transporter: A conserved mechanism for proton gradient sensing? J Biol Chem. 2008;283:9369–9376. [DOI] [PubMed] [Google Scholar]
  • 92. Martens C, Stein RA, Masureel M, et al. Lipids modulate the conformational dynamics of a secondary multidrug transporter. Nat Struct Mol Biol. 2016;23:744–751. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93. Curran AR, Templer RH, Booth PJ. Modulation of folding and assembly of the membrane protein bacteriorhodopsin by intermolecular forces within the lipid bilayer. Biochemistry. 1999;38:9328–9336. [DOI] [PubMed] [Google Scholar]
  • 94. Allen SJ, Curran AR, Templer RH, Meijberg W, Booth PJ. Controlling the folding efficiency of an integral membrane protein. J Mol Biol. 2004;342:1293–1304. [DOI] [PubMed] [Google Scholar]
  • 95. Meijberg W, Booth PJ. The activation energy for insertion of transmembrane alpha‐helices is dependent on membrane composition. J Mol Biol. 2002;319:839–853. [DOI] [PubMed] [Google Scholar]
  • 96. Curnow P, Lorch M, Charalambous K, Booth PJ. The reconstitution and activity of the small multidrug transporter EmrE is modulated by non‐bilayer lipid composition. J Mol Biol. 2004;343:213–222. [DOI] [PubMed] [Google Scholar]
  • 97. Doktorova M, Harries D, Khelashvili G. Determination of bending rigidity and tilt modulus of lipid membranes from real‐space fluctuation analysis of molecular dynamics simulations. Phys Chem Chem Phys. 2017;19:16806–16818. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98. Harris NJ, Reading E, Ataka K, et al. Structure formation during translocon‐unassisted co‐translational membrane protein folding. Sci Rep. 2017;7:8021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99. Findlay HE, Rutherford NG, Henderson PJF, Booth PJ. Unfolding free energy of a two‐domain transmembrane sugar transport protein. Proc Natl Acad Sci USA. 2010;107:18451–18456. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100. Rues R‐B, Dötsch V, Bernhard F. Co‐translational formation and pharmacological characterization of beta1‐adrenergic receptor/nanodisc complexes with different lipid environments. Biochim Biophys Acta. 2016;1858:1306–1316. [DOI] [PubMed] [Google Scholar]
  • 101. Allen SJ, Curran AR, Templer RH, Meijberg W, Booth PJ. Folding kinetics of an alpha helical membrane protein in phospholipid bilayer vesicles. J Mol Biol. 2004;342:1279–1291. [DOI] [PubMed] [Google Scholar]
  • 102. Navarro J, Toivio‐Kinnucan M, Racker E. Effect of lipid composition on the calcium/adenosine 5′‐triphosphate coupling ratio of the Ca2+‐ATPase of sarcoplasmic reticulum. Biochemistry. 1984;23:130–135. [DOI] [PubMed] [Google Scholar]
  • 103. Keller SL, Bezrukov SM, Gruner SM, Tate MW, Vodyanoy I, Parsegian VA. Probability of alamethicin conductance states varies with nonlamellar tendency of bilayer phospholipids. Biophys J. 1993;65:23–27. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104. Pilot JD, East JM, Lee AG. Effects of bilayer thickness on the activity of diacylglycerol kinase of Escherichia coli . Biochemistry. 2001;40:8188–8195. [DOI] [PubMed] [Google Scholar]
  • 105. Lee AG. How lipids interact with an intrinsic membrane protein: The case of the calcium pump. Biochim Biophys Acta. 1998;1376:381–390. [DOI] [PubMed] [Google Scholar]
  • 106. Dumas F, Tocanne JF, Leblanc G, Lebrun MC. Consequences of hydrophobic mismatch between lipids and melibiose permease on melibiose transport. Biochemistry. 2000;39:4846–4854. [DOI] [PubMed] [Google Scholar]
  • 107. Charalambous K, Miller D, Curnow P, Booth PJ. Lipid bilayer composition influences small multidrug transporters. BMC Biochem. 2008;9:31. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108. Gustot A, Null S, Ruysschaert J‐M, McHaourab H, Govaerts C. Lipid composition regulates the orientation of transmembrane helices in HorA, an ABC multidrug transporter. J Biol Chem. 2010;285:14144–14151. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109. van der Does C, Swaving J, van Klompenburg W, Driessen AJ. Non‐bilayer lipids stimulate the activity of the reconstituted bacterial protein translocase. J Biol Chem. 2000;275:2472–2478. [DOI] [PubMed] [Google Scholar]
  • 110. Hoischen C, Gura K, Luge C, Gumpert J. Lipid and fatty acid composition of cytoplasmic membranes from Streptomyces hygroscopicus and its stable protoplast‐type L form. J Bacteriol. 1997;179:3430–3436. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111. Schauner C, Dary A, Lebrihi A, Leblond P, Decaris B, Germain P. Modulation of lipid metabolism and spiramycin biosynthesis in Streptomyces ambofaciens unstable mutants. Appl Environ Microbiol. 1999;65:2730–2737. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112. Bogdanov M, Mileykovskaya E, Dowhan W. Lipids in the assembly of membrane proteins and organization of protein supercomplexes: Implications for lipid‐linked disorders. Subcell Biochem. 2008;49:197–239. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113. Dowhan W, Bogdanov M. Molecular genetic and biochemical approaches for defining lipid‐dependent membrane protein folding. Biochim Biophys Acta. 2012;1818:1097–1107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114. Epand RF, Savage PB, Epand RM. Bacterial lipid composition and the antimicrobial efficacy of cationic steroid compounds (Ceragenins). Biochim Biophys Acta. 2007;1768:2500–2509. [DOI] [PubMed] [Google Scholar]
  • 115. Zhukovsky MA, Filograna A, Luini A, Corda D, Valente C. Phosphatidic acid in membrane rearrangements. FEBS Lett. 2019;593:2428–2451. [DOI] [PubMed] [Google Scholar]

Articles from Protein Science : A Publication of the Protein Society are provided here courtesy of The Protein Society

RESOURCES