Abstract
Membrane–peptide interactions play critical roles in many cellular and organismic functions, including protection from infection, remodeling of membranes, signaling, and ion transport. Peptides interact with membranes in a variety of ways: some associate with membrane surfaces in either intrinsically disordered conformations or well-defined secondary structures. Peptides with sufficient hydrophobicity can also insert vertically as transmembrane monomers, and many associate further into membrane-spanning helical bundles. Indeed, some peptides progress through each of these stages in the process of forming oligomeric bundles. In each case, the structure of the peptide and the membrane represent a delicate balance between peptide–membrane and peptide–peptide interactions. We will review this literature from the perspective of several biologically important systems, including antimicrobial peptides and their mimics, α-synuclein, receptor tyrosine kinases, and ion channels. We also discuss the use of de novo design to construct models to test our understanding of the underlying principles and to provide useful leads for pharmaceutical intervention of diseases.
Peptide–membrane interactions play important roles in all organisms and viruses, performing a diverse array of functions. To name just a few, antimicrobial peptides (AMPs) and certain classes of antibiotics disrupt the membranes of their targets, apolipoproteins stabilize lipoprotein particles and bar-code them for internalization and processing, fusion peptides promote vesicle budding and fusion, and ion channel-forming peptides are found in both antibiotics and viruses. Indeed, membrane-interactive peptides comprise such a fundamental facet of the functional fabric of living systems that it is difficult to overstate the importance of improving our understanding of how their structures and dynamics lead to their remarkable properties. Here, we categorize distinct conformational themes and functional mechanisms of peptide–membrane interactions, drawing from peptides with important biological functions. We also discuss the use of de novo design to construct peptides that test hypotheses concerning the structures and functions of peptides. We will also discuss one example of the translation of an AMP mimetic to the clinic. The literature of peptide–membrane interactions is so immense that any review will necessarily be incomplete. To provide a modicum of focus, we will primarily limit our scope to ribosomally expressed peptides and proteins with native peptide backbones, rather than those synthesized via synthetases or extensive post-translational modification. We apologize in advance for drawing disproportionately from examples of work from our lab; they have been taken to illustrate principles rather than to claim any priority over many other contributions to this vast literature.
We first must define what a “peptide” is in this context. We will adopt the definition that membrane peptides are relatively short sequences (mostly < 50 residues) that form autonomous units capable of interacting with membranes. In the case of larger proteins, we will consider a membrane-interactive domain as a “peptide” if it preserves the membrane-interacting function of the larger protein when studied as an isolated peptide. For example, fusion peptides of viral fusion proteins would be considered as a class of membrane-interacting peptides. Also, the protein α-synuclein, which has a large number of tandemly repeated membrane-binding units, will be included in our discussion.
Intrinsically disordered membrane peptides.
Membrane-interacting peptides can be subdivided based on their conformations and localization in the membrane (Figure 1). One prominent class of membrane peptides are “natively disordered” in solution and on membranes. They interact with the membrane surfaces through highly dynamic electrostatic and hydrophobic interactions. Frequently, the peptides are rich in cationic residues, Lys and Arg, which are attracted to anionic groups in the headgroups of phospholipids, as well as aromatic residues that dip deeper to the acyl glycerol region of phospholipids.1–3 Such conformational states have been identified as discrete kinetic intermediates in the process of insertion of helical proteins into membranes.4 In other cases, this dynamic disordered conformation represents the native mode of interaction observed at equilibrium. For example, a highly positively charged 25-residue peptide domain tethers the MARCKS protein to negatively charged lipids,5 and similar motifs are found in GTPases such as RAS, which helps localize the protein to phases rich in phosphatidylinositol lipids. It is likely that some non-helical AMPs similarly interact with membranes.6 Intrinsically disordered proteins can also form networks that interact with lipid bilayers through liquid–liquid phase separation to remodel membranes.7, 8
Fig. 1: Illustrations of the different ways peptides can associate with membranes.

Peptides can interact with membranes in their (A) intrinsically disordered conformations or through formation of secondary structure. (B) Amphiphilic helices, which have hydrophobic and hydrophilic faces, bind to membrane surfaces such that the helix lies parallel to the interface. Peptide-membrane interactions can influence the insertion of transmembrane peptides (C) and their assembly into helical bundles (D).
The amphiphilic helix.
The amphiphilic helix is a ubiquitous motif that reprises frequently in the structures of membrane-interactive peptides. In this motif, hydrophobic and polar residues segregate onto opposing faces of the helix, imparting amphiphilic character ideally suited for binding to apolar/water interfaces such as those found at membrane surfaces. Such helices are also often termed “amphipathic helices”. However, following Tanford,9 we prefer the more euphonious and friendly term amphiphilic, etymologically stemming from “loving both” as opposed to “suffering both” when traced to the Greek roots.
As the apolipoproteins were first sequenced their potential to form amphiphilic helices became clear from physical models or when their sequences were drawn on helical wheels (a popular way to appreciate the orientation of sidechains when viewed axially down a helix).10, 11 The distribution of residues was striking, with apolar residues lining one face of the helix, flanked by Asp, Glu, Lys and Arg, arranged so that the positive charges flank the hydrophobic sector, and the acidic residues are furthest from the apolar region. This configuration was expected to foster a close orientation of the zwitterionic and acidic head groups in ion pairs, and the interaction of the apolar residues with the acyl-glycerols and acyl chains of phospholipids. Segrest and coworkers demonstrated the validity of this hypothesis through the design of peptides representing single repeats, as well as consensus sequences.12–16 Depending on the apolipoprotein, such peptides bound to membrane surfaces, and also were able to stabilize disk-like assemblies (now designated nanodiscs) by binding to the exposed apolar regions of the phospholipids.17, 18
Kaiser and Kézdy took this strategy one step further by designing peptides that idealize the structural and physicochemical properties proposed to be responsible for activity, rather than focusing on any natural sequence.19–23 The resulting peptides were found to bind to and stabilize small unilamellar vesicles of a size and shape similar to that of high-density lipoproteins, and the concept was soon expanded to other peptides and proteins.20, 24, 25
Cytotoxic peptides and antimicrobial peptides (AMPs)
As a graduate student working in the labs of Kaiser and Kézdy, one of us (WFD) expanded this concept to test the structural basis for the cytotoxic function of the bee venom peptide, melittin (although spelled with a single l, melittin is named for the honey bee, Apis mellifera, often causing confusion about its spelling in the literature). At the time it was unknown whether melittin elicited its toxic effects by a purely physical or receptor-based biochemical mechanism. Examining its structure on a helical wheel revealed an amphiphilic structure, but one that was distinct from apolipoproteins; in melittin 2/3 of the arc was composed of hydrophobic residues and the remaining 1/3 primarily small neutral polar residues. A highly basic hexapeptide resides C-terminal to the helical region, and melittin is devoid of any acid residues. A peptide designed to idealize melittin’s amphiphilic structure was several-fold more potent than melittin in its ability to lyse red cells and disrupt vesicles.26, 27 Also, as was the case for melittin, the peptide was in a random coil configuration in dilute solution, but became helical when allowed to self-associate or bind to apolar-water interfaces. Subsequently, sequences of a family of 14-residue peptides from wasp venom, the mastoparans,28 were discovered to have an even simpler helical motif;28, 29 again about 2/3 of the residues were hydrophobic, but the positively-charged residues located to the polar side of the helix, rather in a distinct segment C-terminal to the helix as in melittin. Thus, a minimal structural feature for disruption of mammalian cells appeared to be a short, positively charged amphiphilic helix.
At about this time, Boman discovered the peptide cecropin30 in the course of studying innate immunity in insects.31 Cecropin was the first ribosomally synthesized AMP that was not toxic to eukaryotic cells. Cecropin had two basic amphiphilic helices; both were less hydrophobic than those found in melittin or the mastoparans. The labs of Merrifield32, 33 and DeGrado34 noticed the amphiphilic character and showed the first helix alone to have antimicrobial activity. DeGrado also designed the first minimal synthetic model for AMPs with the sequence (LKKLLKL)2.29, 34 Soon thereafter, Michael Zasloff made his seminal discovery of the first AMP, magainin.35 Propelled by the importance of innate immunity to human health and Michael’s charismatic leadership, the field rapidly grew. Soon AMPs were found in many organisms including humans. Many were found to form amphiphilic helices, but others were small disulfide-rich mini-proteins (e.g., defensins), Pro/Arg-rich peptides, Trp-rich peptides, anionic AMPs, and β-hairpins.36–44 Because many AMPs work by a biophysical mechanism, rather than by targeting a specific protein or ribosomal target, they have less tendency to develop resistance. The field of both designed and natural antimicrobial peptides has exploded, and there are now books and entire conferences devoted to the subject.45, 46 AMPs have diverse targets, but most target membranes as at least a part of their mechanisms of action.39, 47–49
Much biophysical work has been devoted to understanding the mechanism by which AMPs disrupt bacterial membranes, and what differentiates AMPs from more toxic peptides such as melittin.36–40 A reading of the literature highlights a diversity of mechanisms, ranging from channel formation to more generalized membrane disruption. While many papers in this volume discuss very detailed measurements of AMP–membrane interactions, here we present a less detailed understanding of the principles of AMP–membrane interactions, which guided the design of both peptide and the non-peptidic mimics of AMPs described in the subsequent section.
AMPs disrupt membranes at relatively high peptide:lipid ratios, when the surface concentration exceeds a certain threshold,50 at which the peptides essentially carpet the membrane surface, leading to both generalized membrane disruption51 and heterogeneous pore formation.6, 50, 52–63 The asymmetric accumulation of such a large amount of a peptide on only one leaflet of the bilayer results in a large imbalance in chemical potential. For example, at a peptide/lipid ratio of 1:100 (corresponding to 1:50 if the AMPs localize to only the outer leaflet), there is a strong thermodynamic driving force to equalize the concentration on both sides of the bilayer. The gradient is roughly equivalent to the osmotic gradient created by introducing a solute at 1 M on just one side of a water-permeable membrane (in which case the mole fraction of the solute to water would be approximately 1:50 for a small molecule or ion). In addition to this strong thermodynamic driving force, the physical properties of the peptide contribute to lowering the kinetic barrier for membrane translocation. Insertion leads to expansion, disordering and thinning of the bilayer.61, 64 Moreover, asymmetric insertion leads to surface distortions such as buckling. In each case the bilayer is stressed, helping to reduce the energy of activation for AMP translocation and equilibration across the bilayer. Simulations suggest that the translocation step can also occur with a transient increase in ion permeability.65, 66
AMPs also decrease the stability of bilayers and enhance the stability of nonlamellar phases including lipidic cubic phases and cylindrical micelles, with head groups projecting into a water channel and the apolar groups radiating outward; a similar structure might be formed as a defect within a bilayer, creating a pore that is lined by both headgroups and surface-absorbed peptides.52, 53, 55–60, 67–70 The resulting toroidal pore has a shape similar to the inside of a donut with a topology designated as negative Gaussian curvature (a shape in which there is the curvature is negative in one direction and positive in the orthogonal dimension). Gerard Wong has shown that antimicrobial peptides of many classes stabilize negative Gaussian curvature.71 His group has shown that Lys, Arg, and aromatic amino acids bias towards phases with negative Gaussian curvature and used this principle together with machine learning to design highly effective de novo AMPs.72
Fully peptide-lined pores can also be formed by peptides such as alamethicin, which are sufficiently hydrophobic and long enough to span the bilayer (at least about 18–20 residues in α-helical peptides). For example, melittin binds initially as a random coil to acidic membranes, next it forms a laterally inserted helix, and ultimately forming a membrane-spanning helix at sufficiently high peptide:lipid ratios.4, 73, 74 The vertically inserted helix can next associate in a variety of stoichiometries to form heterogenous transient pores, which then disrupt ionic gradients required for cell viability.
In summary, AMPs have a common feature of being able to bind to the surface of the bilayers, by using a combination of electrostatics and hydrophobic interactions. They insert aromatic or hydrophobic groups sufficiently deep the bilayer to lead to a surface pressure imbalance, and modulation of the properties of the bilayer. The destabilization of the bilayer also decreases the activation energy for subsequent transitions of peptides into and/or across the bilayer. The accumulation of peptides on one side of the bilayer also leads to an imbalance in chemical potential providing a thermodynamic driving force for AMP translocation, which can occur with concomitant diffusion of ions and disruption of the membrane potential. Full-fledged pore formation can occur through the formation of channels that are either lined solely by peptides or a combination of peptides and lipid head groups. Different AMPs can access different mechanisms.
The selectivity of AMPs for bacterial membranes contrasts with toxins like melittin, which disrupt membranes of both mammalian and bacterial cells – a phenomenon that has been widely attributed to differences in the lipid compositions of different cell types. Bacterial cells are rich in acidic lipids compared to mammalian cells, and bacteria also lack cholesterol. In many cases,36 subtle effects relating to phospholipid composition are involved; primary phospholipids in mammalian membranes include zwitterionic phosphatidylcholine (PC) and negatively charged phosphatidylserine (PS) headgroups, while bacterial membranes are rich in zwitterionic phosphatidylethanolamine (PE) and negatively charged phosphatidylglycerol (PG) lipids including cardiolipin. The differences in the physical properties of these various lipids can be as important as the overall charge of the membrane.75–78
A number of features of a peptide define its selectivity. We have already seen that cytotoxic peptides tend to have greater hydrophobicity, leading to more non-discriminate binding than selective AMPs. For helical AMPs, the insertion of the peptides is thermodynamically linked with helix formation, so peptides that strongly favor the helical state are intrinsically predisposed to binding more tightly than closely related peptides with similar charge and hydrophobicity. Thus, for each scaffold there exists a delicate balance between charge, hydrophobicity and rigidity that must be fine-tuned to optimize both potency and selectivity. Finally, the response of a bilayer to bound AMPs likely varies with respect to differences in the lipid composition, so changes to not only binding affinity, but also biological activity likely contribute to selectivity.76
Foldamers and polymers that mimic AMPs
Given the above mechanistic understanding of the structural features and mechanisms of action of AMPs, it became apparent that it should be possible to mimic their activities using molecules other than peptides composed of α-amino acids. Because the activities of many AMPs depends primarily on their overall physicochemical properties—rather than the fine details of their precise amino acid sequences—several groups designed “coarse-grained” molecules idealizing the amphiphilicity of natural AMPs.29, 37, 79, 80 Modifications on natural peptides have included the introduction of d-amino acids,81, 82 long acyl chains,83, 84 and cyclization.81, 85 The availability of β-peptides provided another avenue to probe the requirements for activity, as they can adopt distinct secondary structures such as “12-helices” and “14-helices”.86–90 Our group91, 92 as well as those of Gellman93 and Seebach94 independently showed that β-peptides capable of forming amphiphilic 14- or 12-helices had potent antimicrobial activity. Oligomers with lengths of 10–15 residues that achieved an appropriate hydrophilic/lipophilic balance were selective for killing bacteria vs. mammalian cells. Oligomers that were too short were inactive, and these short sequences also failed to adopt the desired conformation. Moreover, foldamers that were too long or hydrophobic were toxic towards mammalian cells.95 These studies were extended to a variety of different helical types.76–78, 96, 97 In a similar manner, Patch and Barron designed amphiphilic, helical, antibacterial N-substituted glycine oligomers (peptoids),98 and Guichard and coworkers have synthesized antimicrobial foldamers based on a urea backbone.99 Thus, by the early 2000s it had been well established that medium-sized molecules that adopt amphiphilic secondary structures were capable of acting as highly potent and selective antimicrobial agents.100–102
Going one step further, the groups of DeGrado, Tew, and Klein collaborated on the design of antimicrobial oligomers that were more akin to small molecules than earlier foldamers.103–108 Using a rigid arylamide as a framework they systematically introduced cationic groups (cat. 1 and cat. 2 in Figure 2) and hydrophobic groups to maximize activity against Staphylococcus aureus, while minimizing toxicity towards mammalian cells in vitro and in mouse models.107 109
Fig. 2: Generic structure of a rigid platform used to design antimicrobial peptide mimics.

By systematically varying the cationic group (cat. 1 and 2) and the hydrophobic group (Hb), it was possible to maximize potency and efficacy:safety ratio. Brilacidin is undergoing clinical trials for bacterial infections and SARS-CoV-2.
Brilacidin has undergone two phase II clinical trials for S. aureus infections and was shown to have an efficacy on par with daptomycin.109 There is currently great interest in AMPs as potential antiviral agents against enveloped viruses,110–113 and brilacidin has similarly been shown to be active against SARS-CoV-2.114 A phase II clinical trial for COVID is currently in progress (http://www.ipharminc.com/brilacidin-1).
The antibacterial mechanisms of action of brilacidin and closely related compounds have been studied by probing the transcriptional response and leakage of bacterial content in E. coli, as a Gram negative organism,115 and S. aureus as a Gram positive.116 These compounds cause significant changes in the permeability of the outer membrane of E. coli, similar to that observed in polymyxin B and nisin.117, 118 They cause comparatively little permeabilization of the inner membrane, although they lead to a loss of membrane integrity, reaching critical levels corresponding with the time required to bring about bacterial cell death. Transcriptional profiling of E. coli treated with sub-inhibitory concentrations of the arylamides also showed induction of genes related to membrane and oxidative stresses. The induction of membrane-stress response regulons coupled with morphological changes at the membrane observed by electron microscopy indicate that the activity of the arylamides at the membrane is the primary mechanism of action.115
In S. aureus brilacidin also causes membrane depolarization, to an extent comparable to that caused by the lipopeptidic drug, daptomycin.116 However, there was little leakage of cellular contents, ruling out mechanisms that involve large pores. Transcriptional profiling showed that the global response to brilacidin treatment is well-correlated with the corresponding response to daptomycin and a derivative of the cationic antimicrobial peptide, LL37, and induced regulons that respond to perturbation of cell wall and membrane function. These stress responses were mainly orchestrated by three two-component systems: GraSR, VraSR and NsaSR, which have been implicated in virulence and drug resistance against other clinically available antibiotics.
Coarse-grain MD allowed simulations with a relevant number of arylamides per bilayer65 at coarse-grained oligomer/phospholipid (CGO/PL) from 1:256 to 1:14, spanning surface concentrations that experimentally give rise to from no lysis to very rapid vesicle lysis. At low CGO/PL ratios the antimicrobial inserts into the bilayer with its long axis parallel to the membrane surface and its apolar groups penetrating into the membrane. By contrast, at the high CGO/phospholipid ratios required for lysis of bilayers in experimental systems, the drugs initially insert with an ensemble of different angles. The very large imbalance in the concentration of the oligomers between the proximal versus the distal leaflets of the bilayer leads to a metastable state with pronounced buckling of the proximal leaflet of the bilayer. At longer times, the molecules diffuse to the opposite side of the bilayer. The CGOs frequently diffuse in pairs across the bilayer, often with accompanying ions and water molecules. Following the translocation step the oligomers are oriented predominantly parallel to the bilayer. However, they are less well ordered than in the simulations at low antimicrobial:phospholipid ratios. In the final configuration, the polarity gradient of the membrane is altered, and there is an increased permeability to water in the simulations. These simulations provide pictorial detail to prior models described above.
We hope the above section is helpful in defining the principles required for design of AMP mimetics. There are currently a number of AMPs and AMP mimetics undergoing clinical trials for various indications.119, 120 In addition to understanding the principles for design, which has been the primary focus here, it is also essential to understanding the biophysical bases for their action. The reader is pointed to other papers in this volume for a wealth of information on AMP-membrane interactions.
Dynamics, structure, and mechanism of membrane-bound α-synuclein
One of the most prominent examples of membrane-binding peptides, α-synuclein (Figure 3A), is also one the most complex.121 Although synuclein is sufficiently long to be considered as a protein, it has a number of repeating peptide units, and different portions of its sequence interact differently with the membrane that show the diversity of ways peptides can interact with membranes.
Fig. 3: Sequence and structures of α-synuclein.

(A) Helical wheel depicting the amphiphilic nature of α-synuclein’s 11-residue consensus sequence, which mediates membrane binding. (B) Alignment of the 11-residue consensus sequences of α-synuclein (α-syn) and apolipoprotein A1 (apoA1).129 (C) Structural model of α-synuclein showing C-terminal dynamics in an extended, membrane-bound helix, which is supported by data from FRET,130 EPR,131, 132 NMR,133, 134 and DMS.135 (D) Cartoon model of α-synuclein tethering nearby vesicles.136 (E) Hypothetical model of α-synuclein interacting with membranes of negative Gaussian curvature, such as those in fusion pores.137
Originally discovered by screening cDNA expression libraries for proteins localized to synapses using antibodies raised against synaptic preparations from electric rays, α-synuclein was shown by electron microscopy to localize to the surface of synaptic vesicles, suggesting a direct peptide–membrane interaction.122 This hypothesis was further supported by similarity between the sequences of α-synuclein and apolipoproteins (Figure 3B),123 which use amphiphilic helices featuring unique 11-residue repeating segments, to sequester and transport lipids.124 α-Synuclein localizes specifically to synaptic vesicles, rather than the plasma membrane, and disperses from nerve terminals after stimulation,125 suggesting that it might participate in synaptic vesicle trafficking. Indeed, early experiments on knockout animals revealed changes in stimulated dopamine release,126 suggesting that α-synuclein plays a key role in neurotransmission. These early studies, as well as genetic127 and pathological128 links to Parkinson’s disease, motivated detailed study of the interactions of α-synuclein with lipid membranes, which have revealed a series of remarkable biophysical features that contribute directly to α-synuclein’s physiological and pathological activities.
Plotting the distribution of residues on a helical wheel revealed a facial separation of hydrophilic and hydrophobic residues, with cationic seams in between to potentially support membrane binding (Figure 3A).123 Early experiments on purified α-synuclein confirmed this hypothesis, showing by circular dichroism that α-synuclein, which lacks defined structure in isolation,138 adopts helical secondary structure upon association with lipid membranes.139 Although these behaviors are shared with other membrane-binding peptides, α-synuclein’s length and hydrophobicity make it unique, and these features likely contribute to both its physiological and pathological activities.
Compared to other amphiphilic helices, α-synuclein is only modestly hydrophobic.140 Its hydrophobic face consists primarily of smaller residues like Gly, Ala, and Val, which are interrupted by hydrophilic Thr residues. In contrast, the hydrophobic face of APLS (amphipathic lipid packing sensor) motifs, for example, often feature uninterrupted stretches of larger, greasier residues like Leu, Ile, Met, and Phe.141 α-Synuclein’s modest hydrophobicity tempers its affinity for lipid membranes. Unsurprisingly, for example, substituting the Thr residues with bulkier hydrophobic residues increases affinity, which in turn increases toxicity in cellular and animal models,142 suggesting that the affinity of α-synuclein for lipid membranes is tightly optimized. Moreover, the prevalence of Gly residues in this helix further attenuates membrane-binding affinity by increasing the entropic cost of helix formation, though the alignment of these Gly residues along one face of the helix suggests additional or more nuanced roles, perhaps mediating interactions with specific proteins or lipids. This tuning of α-synuclein’s membrane affinity not only increases the rate of exchange between the membrane surface and solution, which could be involved in its reported roles in synaptic vesicle trafficking;143 it also enables α-synuclein to discriminate membranes with different physical and chemical properties.
For example, α-synuclein has a higher affinity for membranes of increasingly negative charge,139 likely owing to the relative importance of electrostatic interactions, rather than strictly hydrophobic association, to drive membrane binding.144 α-Synuclein’s modest membrane affinity has also been suggested to contribute to its ability to discriminate membranes with different packing or curvature.139, 145 The modest affinity likely provides insufficient energy to disrupt the packing of planar bilayers but could be sufficient to bind to the gaps between lipid molecules in highly curved membranes, such as those of synaptic vesicles. Moreover, theoretical models146 and course-grained MD simulations147 demonstrate that the partitioning of α-synuclein to a specific depth within the membrane, as tuned by its physicochemical properties, is sufficient to drive curvature formation.
α-Synuclein is also distinctive because of its length.148 Preliminary sequence analysis suggested that the first 90–100 residues could form a continuous amphiphilic helix, far longer than most examples, the most notable exception being the perilipins that bind to lipid droplets.149 Preliminary NMR experiments confirmed the membrane-binding region as the first 90–100 residues,150 which include the 11-residue segments predicted to mediate lipid-binding; the remainder of the 140-residue protein remains disordered and unbound from the membrane. Subsequent studies disagreed about whether this region forms a long, single helix,131 or whether the membrane-binding domain is broken into two segments.151 The discrepancy highlights the effect that the membrane or membrane mimetic can have on peptide and protein structure. The first high-resolution NMR structure, solved in SDS micelles, suggested such a break.152 In contrast, studies by EPR,132 FRET,130 and solution-133 and solid-state NMR,134 conducted with protein reconstituted in bilayers, instead favored a single, extended helix bound to the membrane. Consensus has emerged that this discrepancy likely arises from the curvature of the membrane mimetic; highly-curved species like SDS micelles induce breaking of the helix, while less curved membranes such as LUVs support helix formation across the entire membrane-binding domain.
The diversity of structures formed by α-synuclein has raised important questions about the functional and pathological roles of each individual species. To begin dissecting these roles, we adapted deep mutational scanning (DMS) to reveal the structure associated with a specific activity. In DMS, a pooled library of protein missense variants is screened for relative activity, which is inferred from changes in the frequency of each variant before and after selection.153 By subjecting comprehensive libraries of missense variants to relatively weak selective pressure, such that most variants are retained in the population after selection, the relative effects of thousands of mutations can be determined and used to infer structural features of the protein that are responsible for activity.
We sought to determine the structural basis for a relatively simple activity of α-synuclein: its toxicity when ectopically expressed in yeast cells. By identifying mutations that interfere with formation of the toxic species and rescue the growth rate of yeast cells, we discovered that α-synuclein toxicity in yeast is driven by a dynamic, membrane-bound helix with increasing dynamics toward its C terminus.135 For example, mutations that introduce polar groups onto the hydrophobic face reduce both membrane binding and toxicity, but this effect becomes less pronounced toward the dynamic C terminus. Similarly, helix-breaking proline residues reduce toxicity, again most potently at the N terminus. This result is in excellent agreement with models proposed to mediate vesicle clustering,136 implicating this activity in toxicity, at least in yeast.
We then developed a simple structural model for this state (Figure 3C),135 which we used to make thermodynamic predictions of the effect of each mutation on membrane binding, based on the known energy associated with positioning each amino acid to varying depths in the lipid bilayer.154 These predictions were remarkably successful at predicting the effects in our cellular assay, providing strong support for the role of this species in toxicity. Our model also demonstrates that the entire helix contacts the membrane to similar depth, while the C terminus engages in increasingly frequent dynamic dissociation from the membrane surface. Again, these results from functional measurements are in remarkable agreement with inferences made from complementary spectroscopic methods, specifically NMR133, 134 and EPR.131, 132 The dynamics at the C-terminal end of the helix might result in part from the sequence of α-synuclein’s uniquely long amphiphilic helix. Although each of the repeated 11-residue binding segments encodes nearly three turns of an a-helix, the slight mismatch between the sequence and its encoded structure causes the hydrophobic face of the helix to gradually wrap around helix.155 The energy associated with overwinding the helix to align the hydrophobic groups could prevent both ends from binding tightly.
The length of this helix creates interesting dynamical properties. In order to align the hydrophobic groups on one face of the helix, α-synuclein would have to overwind, and although the energy required to do so is small for individual residues or helical turns, that energy becomes significant for a helix of α-synuclein’s length. This energetic penalty to adopting the structure that would optimize the hydrophobic groups for membrane binding prevents the protein from fully associating with the membrane surface across its entire length.
Dynamics in the membrane-bound helix are hypothesized to have a variety of physiological roles. One putative role is the tethering of membrane surfaces to one another, release of the C-terminal region from the surface of one membrane allows it to associate with nearby surfaces,136 which might mediate vesicle clustering (Figure 3D), one known roles of α-synuclein. Intriguingly, aberrant protein-rich inclusions in the brains of Parkinson’s patients called Lewy bodies are enriched in both α-synuclein and membrane-bound organelles and vesicles,156 suggesting that this clustering activity might contribute to pathology, although this hypothesis is controversial.157 Alternatively, dynamics in the membrane-bound state could support protein–protein interactions with other membrane-remodeling machinery, such as SNARE proteins.158
Dynamics in the membrane-bound state can also promote membrane curvature;159, 160 as the surface concentration of α-synuclein increases, clashes between monomers involving the dynamic end of the helix and the disordered tail would be relieved upon increased curvature. Electrostatic repulsion between the acid tails of each monomer could further drive repulsion and increased curvature. These biophysical mechanisms might underlie α-synuclein’s ability to tubulate membranes.131, 161 Alternatively, the inability of the protein to associate with the membrane surface across its entire length might create an energetic driving force for the association with or formation of membranes with negative Gaussian curvature (Figure 3E), which has been observed in course-grained simulations,162 such as those that form during vesicle endocytosis or exocytosis. The energy release from binding of α-synuclein to such a surface could contribute to its observed ability accelerate fusion pore dilation.137
Membrane-binding has a strong influence on the aggregation of α-synuclein,163 which is implicated in Parkinson’s and related diseases. At low protein:lipid ratios, membrane binding suppresses α-synuclein aggregation by distracting individual monomers from self-association.164 At higher concentrations, however, membrane-binding induces aggregation by concentrating monomers on the membrane surface. Moreover, the dynamic release of the C terminus of the amphiphilic helix from the membrane surface exposes the aggregation-prone NAC (non-amyloid component) region, which is then poised for self-association at high protein:lipid ratios.165 Dysregulation of either the concentration of α-synuclein or its membrane affinity could therefore drive pathological aggregation.166 In turn, α-synuclein aggregation can perturb its interaction with lipid membranes by either concealing167 or revealing membrane-binding regions.168
In summary, α-synuclein exemplifies many of the features of peptide-lipid interactions considered in this volume. In one single protein, we see surface helices that specifically recognize and bind to membranes with high curvature, using a combination of peptide-lipid interactions as well as generalized crowding effects. We also see multiple types of interactions exemplified in a single protein: α-synuclein adopts (1) a highly stable helix that is inserted into the bilayer surface near the N terminus; (2) a dynamic helix, which adopts a dynamic equilibrium between inserted helical conformations and unfolded structures; and (3) a fully unstructured C-terminal tail tethered to the membrane surface.
INSERTION, ASSEMBLY, AND DESIGN OF SINGLE-SPAN TM HELICES
Insertion of transmembrane helices
Membrane proteins are generally inserted into membranes via the translocon, and they complete the folding process in the membrane environment.169, 170 The helical conformation is stabilized in a membrane because the backbone amide-carbonyl hydrogen bonds are satisfied within the helix and sequestered away from the nonpolar environment of the bilayer.171, 172 In the two-stage model of membrane protein folding,173–175 once inserted in the bilayer, the protein is then able to fold via the coalescence of the helices to form the native tertiary structure. The study of designed and natural transmembrane peptides has greatly informed our understanding of the forces that stabilize both the insertion and coalescence of helices in membrane proteins. Thus, TM peptides are excellent models for multi-span membrane protein folding. For the biophysicist, self-associating TM helices allow investigation of unconstrained inter-helical interactions with a clear unfolded state—a monomeric α-helix—where conformational specificity and thermodynamics can be simply evaluated by the oligomeric distribution. From a more biological perspective, half of all human membrane proteins have a single transmembrane helix.176, 177 Moreover, TM helices often do more than anchor proteins in membranes; they self-associate or interact with other membrane proteins to play vital roles from signaling to ion conduction, and often the aberrant assembly of TM helices is central to devastating diseases from cancer to Alzheimer’s disease.178, 179 Thus, the study of TM peptides is a fruitful endeavor for biophysicists and biologists alike.
The features influencing the insertion of TM helices into membranes has been extensively. Sufficient hydrophobicity is required to enter the translocon and stabilize the helix in an apolar environment,180 dictating the placement of apolar residues at most positions in the helix.154, 170, 181–183 Additionally, aromatic and basic residues are well accommodated in the headgroup region of a bilayer, where they can contribute to stability.184–187 The residues in the headgroup region also define the transverse shift of a helix in a membrane, as does the presence of polar residues within the hydrophobic region.188 Charged residues in the hydrophobic region of a TM peptide also have a large effect on the ability of peptides to insert into bilayers. Early work from Caputo and London189, 190 focused on three model TM Leu-rich peptides, each containing Asp at different positions in their hydrophobic core. When the Asp residue was protonated at low pH, the peptides inserted vertically into vesicles composed of dioleoylphosphatidylcholine (DOPC). When the Asp residues was ionized at neutral or high pH, the topography was altered in a manner that would allow the charged Asp residues to reside near the bilayer surface. More recently, Engelman and coworkers have developed a “pHLIP” peptide, which is a 36-aa peptide derived from the bacteriorhodopsin C helix with an acidic residue near the center of the TM helix. This peptide exists in three states: soluble in aqueous solution, bound to the surface of a membrane, and inserted across the membrane as an α-helix; decreasing pH stabilizes the inserted state.191, 192 An “ATRAM” peptide from the Barrera lab has similar physical characteristics.193, 194 Both insert unidirectionally into bilayers, and are capable of bringing cargos into cells. Given the acidic environment of tumors they show promise for targeted delivery and diagnosis of tumors.195–197
There is generally a minimum length of about 20 residues (30 Å) to stably span the bilayer in a vertical orientation, and there are significant effects of a hydrophobic mismatch between the membrane and the peptide.198–201 Shorter TM helices will necessarily experience a hydrophobic mismatch with the bilayer, leading to destabilization of the membrane lamellar phase.202 On the other hand, TM helices with hydrophobic lengths significantly longer than 20 residues will tend to insert at an angle relative to the membrane normal, as this orientation allows more efficient burial of the TM helix.203, 204 Finally, while the rotation of a helix about its own axis does not affect the depth of insertion of an amino acid residue, rotation of a tilted about its axis places a sidechain into different positions in the bilayer, which can differ significantly near the headgroup region.205–207 Thus, the membrane environment has a very significant impact on “setting up” a helix for favorable helix-helix interactions.208
Structural aspects of the assembly of transmembrane helices
Peptide–membrane interactions also play an essential role in defining helix–helix interactions in membranes.209, 210 The hydrophobic effect provided by burial of apolar side chains in a protein’s interior represents the predominant driving force for helix–helix association and folding in water, yet it is negligible in lipid membranes. What then drives assembly and folding of membrane proteins? Lipid-specific effects such as “solvophobic” exclusion likely contribute.211, 212 Oligomerization also relies on matching of the hydrophobic thickness of the TMD with that of the surrounding lipid. If the lipid bilayer is too thin or too thick, oligomerization can be reduced in model membranes.213, 214 Moreover, there can be very large effects when comparing native cell membranes with model membranes. These effects can even overwhelm stabilizing contributions from favorable helix–helix interactions, lateral packing pressure and excluded volume.215 Thus, while the following section focuses on stabilizing helix–helix interactions, it is important to keep in mind that the association of helices represents a balance between membrane–protein and protein–protein interactions, which is sensitive to the membrane environment.
Structural informatics, peptide design, and site-directed mutagenesis of natural proteins have greatly advanced our understanding of the mechanisms of association of TM helices. We identify three general categories of interactions that stabilize protein structures. In some packings such as GX3G motifs, the backbones of two helices approach closely forming favorable weakly polar interactions and efficient van der Waals packing. In other cases, electrostatic interactions and hydrogen bonds involving polar or charged amino acid sidechains help drive association. Finally, tight packing of apolar sidechains can be a strong driving force for association. We will consider each separately.
GX3G (and SmallX3Small) and Small-X6-Small motifs.
In the GX3G motif, first identified in glycophorin, Gly residues spaced four residues apart mediate a very close association of the backbone of two helices stabilized in part by C–H hydrogen bonds between the alpha CH groups and carbonyl groups of adjacent helices.216–218 This non-classical hydrogen bonded interaction is favorable, and it also allows very efficient van der Waals packing interactions between proximal groups in the interacting helices.219–221 In the classical GX3G motif, the two interacting helices are parallel, and they have a right-handed crossing angle of about +40°. Variations on the GX3G motif are seen in which one or both of the Gly residues are instead small side chains, such as Ala. It is noteworthy that these residues lose no side-chain conformational entropy when the helices associate, as is the case with larger, more flexible side chains. By contrast, the release of lipid tails to the bulk will increase the entropy of the lipids. The GX3G motif mediates critical TM interactions in many biologically important single-pass proteins that include receptor tyrosine kinases. An extensive literature attests to the importance of the subject, and the interested reader is directed to a number of classical reviews from the labs of Engelman, Fleming, MacKenzie, Langosch, Akin, Senes, Schneider, and others.210, 217, 222–225 A more recent comprehensive review by Westerfield and Barrera is particularly insightful in its thoughtful treatment of the role of GX3G and related motifs in receptor association.226
There are several variations on the GX3G motif,227 in both TM peptides and larger protein structures.228 The classical GX3G motif is C2 symmetric, and the interactions are identical on each helix, as in glycophorin A.229 Additionally, small amino acids can occur repetitively every four residues to create extended “glycine zippers”.230 In other cases, an asymmetric interaction utilizing the GX3G motif on just one of two helices is seen in heterodimers. One example occurs in the integrins. In these heterodimeric proteins the TM helices of their alpha and beta subunits interact tightly in the resting state, but dissociate when they are activated for binding to extracellular proteins. Thus, the interaction strength between the two helices is finely balanced to allow the conformational change to occur in response to intracellular and extracellular signals. The interface of the αIIbβ3 TM domain heterodimer231–233 consists of a tightly packed structure in which small and large residues interdigitate by efficient van der Waals packing along the heterodimer interface. Thus, a sequence motif in the αIIb TM domain, G-X3-G-X3-L, packs in a reciprocal manner with the β3 TM domain sequence V-X3-I-X3-G such that bulky residues from one TM helix contact a hole formed by a small Gly residue on the neighboring helix.227 A similar interaction occurs between the TM helices of a second β3 integrin, αvβ3, but using a different face of its TM helix to bind the αv TM helix the one used to interact with the αIIb subunit.234
The GX3G motifs discussed above mediate interactions between parallel helices. An analogous type of interaction occurs between antiparallel helices with small residues spaced every 7 residues apart,154, 230, 235 designated here a Small-X6-Small motif. The interaction is mediated by a mix of favorable electrostatic interactions between the peptide backbones, efficient van der Waals packing, and other interactions that depend on the sequence context. In this motif, the antiparallel helices have a left-handed crossing angle of about −10° to −30°, similar to classical coiled coils. Coiled coils have a characteristic 7-residue repeat, whose positions are designated “a” through “g” within a single heptad. By convention, the residues at the “a” and “d” positions of the heptad pack in the core of a coiled coil. The stability of water-soluble coiled coils scales with the size and hydrophobicity of the side chains at the “a” position increasing over the series Gly < Ala <Val < Ile.236 Just the opposite trend was seen in membrane-soluble coiled-coil peptides, Gly > Ala >> Val > Ile.237 Furthermore, the peptides in which the “a” residues were either Gly or Ala had a strong preference to form antiparallel “Ala-coil” structures (Figure 4) when in a membrane,237, 238 which appeared to be associated with alignment of dipoles from the amides of adjacent helices. Similar results are seen with Ser at the “a” position. Interestingly, Ser can stabilize both antiparallel packing as seen in protein structures239 and model peptides (Figure 4),240 as well as a parallel helical dimerization motif found in the murine erythropoietin receptor.241, 242
Fig. 4: Inner tetramer of inactive channel form of designed ion channel from pdb 6yb1.
(A) Helical wheel diagram depicting key interactions, namely the Ser and Ala zippers, that stabilize the tetrameric assembly. (B) The inner tetramer is stabilized by a hydrogen-bonding network formed by Ser (purple), Thr (blue), and waters (not shown). Views of the Ala-coil and the Ser-coil interfaces reveal favorable interactions that mediate the formation of the tetramer.
Polar interactions.
Polar interactions, including salt bridges and hydrogen bonds provide a significant driving force for assembly of TM helices.243–248 Numerous oncogenic mutations of receptor-mediated tyrosine kinases and other receptors involve the introduction of polar residues in the TM helices, which leads to enhanced association or alteration of the geometry of the TM helix-helix interaction (for reviews see 224, 241, 249, 250). Mutants with polar substitutions in their TM helices also figure largely in membrane protein misfolding diseases.251
Studies with model TM peptides have shown that Asn, Gln, Glu, Asp (and to a lesser extent His) enhance TM helix association.2, 244, 245, 247, 248, 252 Presumably, the Glu and Asp residues are protonated when inserted deep into a bilayer.253 Thus, each of these residues that induce strong association have a common feature of having both hydrogen bond donors and acceptors, capable of mediating inter-helical interactions. Ser and Thr are exceptions, however. These residues generally form an intramolecular hydrogen bond to a carbonyl three or four residues above in the helix, so there is a reduced driving force for forming hydrogen bonds with adjacent TM helices.171, 254, 255 The position of side-chain hydrogen bonds in a TM helix and their surrounding sequence have significant effects on stability.256–259 As compared to an interaction between apolar residues, Asn–Asn interactions range from highly favorable in the apolar region of the bilayer to unfavorable when buried in the interior of a water-soluble structure.260 Because Asn is capable of inducing or enhancing TM association, Asn residues have been scanned through the TM helices of several receptors, including the thrombopoietin receptor,249 the erythropoietin receptor,242 and the integrin αIIbβ3,261, 262 leading to both activation and inhibition of signaling.
Intramembrane salt bridges can also be a potent driving force for the association of TM helices. For example, signaling complexes of immune cells employ multi-component protein receptors composed of distinct binding and signaling subunits263–265 that assemble via helix–helix interactions of their transmembrane domains. The binding subunits associate with extracellular ligands, while the signaling subunit typically contains immunoreceptor tyrosine-based activation motifs (ITAMs), which couple to downstream phosphorylation and signaling pathways. The interaction between a pair of acidic Asp residues on the signaling dimer and a basic residue (Lys or Arg) on the ligand-binding coreceptor are required for complex assembly within the otherwise nonpolar membrane.266, 267 This 2-to-1 acidic to basic polar association motif is ubiquitous in multi-component receptor families including Fc receptors268 and T-cell receptor-CD3 complexes.269 For example, in the T-cell receptor complex, the positively charged Arg and Lys residues on the TCRαβ TMs interact with negatively charged Asp and Glu residues on the CD3δε and CD3 δε heterodimers, and the ζζ-homodimer.263 These interactions are necessary for proper assembly of the complex in the ER and subsequent transport to the cell surface.263 Furthermore, the 2-to-1 acidic to basic association is observed in a family of more than 20 receptors that contain the signaling subunit DNAX-activation protein 12 (DAP12), a disulfide-linked homodimer with a minimal extracellular region.265, 267, 270 Recent, MD simulations and density functional theory calculations of the transmembrane helices with differing protonation states of the Asp-Asp-Lys triad identified a structurally stable interaction in which a singly protonated Asp-Asp pair forms a hydrogen-bonded carboxyl-carboxylate clamp that clasps onto a charged Lys sidechain.271 This polar motif was also frequently observed in a search of the Protein Data Bank-wide search, indicating that it is a widely used motif.
van der Waals packing of apolar sidechains.
In water the burial of apolar sidechains in a protein interior gives rise to a large hydrophobic driving force that is not operative in a membrane. Thus, membrane proteins utilize many of the interactions and motifs described above. Nevertheless, the great majority of contacts in membrane proteins involve interactions between apolar sidechains.211 Structural informatics suggests that these hydrophobic residues pack more efficiently in membrane proteins,228, 257, 272–274 and the resulting van der Waals (vdW) interactions provide a primary driving force for folding. On the other hand, apolar side-chains in the folded state might provide little net stabilization, because lipids interact with these side-chains similarly in the unfolded state. In this view, vdW packing is secondary to stronger than the forces and structural restraints, including conventional and CH-mediated hydrogen bonds,218, 219, 244, 248, 256 topology (i.e., loops, water-soluble domains, α-helical insertion),275 and weakly polar interactions.210, 276 Furthermore, experimental studies using site-directed mutagenesis suggest that apolar packing contributes similarly to stability in water-soluble and membrane proteins.257 Finally, attempts to design proteins using membrane-solubilized versions of peptides that assembly favorably in water, such as the leucine zippers resulted in very rather weak or structurally uncharacterized association in the absence of polar interactions.244, 277
Recent work has, however, shown that van der Waals interactions can play a dominant role in the assembly of a natural membrane homopentameric peptide, phospholamban. Mravic and coworkers278 deciphered a steric code involving tight and regular packing that stabilizes natural membrane proteins, and used this insight to design self-associating pentameric peptides de novo. The structures of the resulting pentamers were solved to high resolution. Indeed, the assemblies were so stable that they remained partially associated after boiling in strong denaturant solutions. These packing motifs are frequently observed in numerous unrelated membrane protein families. These results reveal a substantial and widespread role for apolar packing in the folding, stabilization, and evolution of membrane proteins. Interestingly, a comparison of membrane-soluble and water-soluble sequences indicated that on average, the steric code was more stringent and the helices packed more tightly in the TM motifs versus water-soluble structures. Thus, achieving tight and specific packing would appear to be more essential in membrane proteins, which cannot rely on the hydrophobic effect.
A second recent example of the importance of van der Waals packing has been discovered through a random screening approach. DiMaio and coworkers have used clever genetic screens to discover “traptamers”279 that interact in a sequence-specific manner with receptor tyrosine kinases, including the erythropoietin receptor and platelet-derived growth factor Beta (PDGFβR).280 Their work began with small single-span TM protein, E5, a peptide from a tumor virus that targets and activates PDGFβR in a key step of oncogenesis.281 E5 is a disulfide-bonded dimer and the dimeric conformation is required for activation PDGFβR. A single transmembrane Gln residue is required for activity in the context of the native sequence, but genetic selections282, 283 identified entirely hydrophobic sequences that are similarly able to activate PDGFβR in a highly specific manner.284–287 Going one step further, DiMaio’s group has designed traptamers that composed of only Leu and Ile that target TM helices,288 again in sequence and target-specific manners. Finally, they have extended this approach to activator of the erythropoietin receptor289 and other TM proteins.287 It seems reasonable to assume van der Waals forces provide the a major driving force for assembly in these cases.
Combining a diversity of forces and motifs.
In the above sections, we have focused on examples that illustrate individual forces and motifs, but in typical membrane peptides and proteins, a number of stabilizing interactions are combined to achieve stability and function.211, 290–292 We have already seen the intimate interconnection between non-conventional CH···O=C hydrogen bonds and van der Waals packing in GX3G motifs.221 Also, π-stacking293 and hydrogen-bonded interactions can augment this motif. For example, strong hydrogen bonds between a Ser and His sidechain augments the affinity of a GX3G motif in the transmembrane homodimer of BNIP.291, 292 Similarly, His37 forms multiple polar interactions in the homo-tetrameric four-helix bundle, M2 from influenza A virus.294 While these interactions help stabilize the tetramer,295 the tetrameric structure is retained in the H37A mutant, as is ion channel activity although the ion selectivity is altered by this substitution.296
Multiple types of zippers of the SmallX6Small family can combine in protein assemblies larger than a dimer, when they are placed at two, non-overlapping positions of a helix. A recent crystal structure of a designed channel peptide in a non-conducting state illustrates the concept. Its structure displays two types of helix–helix interfaces, one featuring an Ala-coil and the other a Ser-coil. The designed sequence has a sequence repeat with small Ala or Ser residues of the heptad repeat (Figure 6). As mentioned above, small residues spaced 7 residues apart are the signatures of Ala and Ser zippers,228, 239, 297, 298 in which two helices wrap around one another in an antiparallel coiled coil. In this structure, the packing of small residues provides a strong driving force for assembly in a phospholipid bilayer, while the Ser residues make polar interactions. Both Ala coils and Ser zippers are observed widely in membrane proteins and have been used in membrane protein design.237, 240, 299
Fig. 6: Designed ion channels.

Models of (A) the homotetrameric LS2, and (B) the homohexameric LS3 channels show pore-facing Ser residues with Leu at well-packed helix–helix interfaces. (C) Designed Zn2+/H+ antiporter Rocker (pdb 2MUZ) binds zinc through two 4-Glu, 4-His sites in the channel.
Analysis and design of functional membrane-spanning TM helices and helical bundles
Modulation of protein–protein interactions in the membrane.
The association of TM helices is often a critical feature in the assembly and function of single-span TM proteins. Peptides and small molecules that disrupt or modulate association provide very useful tools to probe the contribution of TM association to the function of membrane proteins—and they can translate to useful therapeutics. For example, Eltrombopag was the first small molecule drug approved that targets the TM domain of a single-pass membrane protein, and it is clinically approved for multiple indications.300 This drug targets the TM domain of the thrombopoietin receptor near His99.249, 301 There are several approaches to design peptides and small molecules that modulate receptor function, ranging from random screening/synthetic optimization to computational design of TM peptides. Many excellent reviews of the field are available, so we will only introduce the topic briefly.226, 302
The simplest way to design peptide modulators of TM proteins is to synthesize or express a peptide spanning the sequence of the TM helix of interest. This was first demonstrated for glycophorin A in 1976 by Furthmayr and Marchesi,303, 304 who showed that a TM peptide was able to shift the equilibrium of the full-length protein from a dimer to a monomer by hetero-association. Since then, this approach has been widely used to study association of numerous proteins, as reviewed recently.226, 302 Going a step further, Barrera and coworkers have developed the introduced a Glu residue into the TM helix of EphA2, rendering it water-soluble, but still capable of inserting into membranes in a controllable manner.305–307 This peptide activates EphA2 by interacting with the TM helix and its juxtamembrane domain. Attesting to its specificity, this peptide only inserts into membranes at neutral pH in the presence of the TM region of EphA2.
A second approach involves computational design of peptides that are designed to target the TM domains of proteins in a sequence- and structure-dependent manner. The GX3G motif has been used as a basis to design “Computed Helical Anti-Membrane Protein” (CHAMP) peptides, which bind to the TM helices of single-span membrane proteins that have this motif in their sequence.308 A second designed CHAMP helix is docked to make good interactions with the peptide, then its sequence is optimized to recognize the specific sequence surrounding the GX3G motif of the target. This method has been used to specifically recognize the αIIb, αv, and β1 subunits of integrins. They stabilize an activated form of the integrin by blocking the interaction of the alpha and beta TM helices278, 308–310 and have been helpful in determining the contribution of TM helix heterodimerization to overall activation of the integrin αIIb.311 Similar computational methods have now been expanded to a growing number of targets as reviewed recently.226, 302
The third approach involves genetic screening to identify modulators of a given activity. A classical method to determine associations is through the construction of trans dominant negative mutations. In early work, this approach was elegantly employed by Pinto and Lamb to determine the tetrameric association state of the M2 proton channel from influenza A virus, through quantitative analysis of the effects of a dominant negative mutant on the proton channel activity of the protein. The elegant work of DiMaio and coworkers is also covered above in the section on van der Waals interactions.
Irrespective of the method used to design TM peptides, there are a number of technical issues that need to be carefully considered and appropriate controls conducted. If TM peptides are synthesized and added exogenously, one must tackle the problem of their poor solubility and tendency to aggregate in solution. While adding short poly-Lys sequences has been employed, it is important to conduct appropriate controls with mutants or scrambled peptides as the addition of a poly-Lys tag can be perturbing. An alternative is to use a short polyethylene glycol sequence of defined length308 in place of the Lys residues. Even after tagging, the peptides can have poor properties, which necessitates addition of organic co-solvents or micelles. Again, rigorous controls for these variable need to be conducted. An alternative approach is to express the peptide genetically for biosynthetic incorporation into membranes. Again, however, care needs to be taken to assure that the peptides are expressed appropriately in the desired target membrane, and to assure that they are not triggering an unfolded protein response in the ER or other off-pathway effects.
Analysis and design of self-assembling peptide channels.
Ion channels are essential for all electrical activity in the central nervous system, and for the translocation of ions, water, and small molecules across cellular membranes. Given our advanced understanding of the TM helices and helical bundles, we should now be able to design functional membrane proteins from first principles. Here, we focus primarily on designed peptide channels that transport protons and ions, highlighting natural model systems that have illuminated principles underlying channel activity and defining key considerations in peptide-membrane interactions that allow us to fine-tune ion conduction.
Our understanding of the detailed molecular mechanisms of ion conduction has progressed in a number of distinct stages.312 Before the sequences and structures of ion channel proteins were available, detailed studies of the structural basis of ion conduction focused on simple peptides, primarily gramicidin A (GA) and alamethicin. Next, as the sequences, but not yet atomic-level structures, of channels began to emerge, the de novo design of model peptides that mimic the pores of helical ion channels became an active area of research,313, 314 and studies of GA and alamethicin continued apace. As structures of natural ion channel proteins became increasingly available, they received increasing attention. Finally, we have reached a level of understanding of both membrane protein folding and ion channel function that it should now be possible to design channel proteins from first principles. The first forays into the field of de novo design have focused on testing and sharpening principles of folding and conduction, but applications315 for sensing, water purification, and ion separation could follow.
Several milestones in the field of de novo design of ion channels have already been achieved. In early work, the design of peptides with well-defined rates of ion conduction and selectivity were reported, but high-resolution structures were not available.313, 314, 316 Very recently, the design of peptides and proteins that assemble into channels have been reported by Baker and coworkers, but channel recordings were limited to macroscopic conductance measurements and the per/protein or per/channel rate of conductance was not reported.317 The design of channels with well-defined channel properties and high-resolution structures is imminent, and we will probably see papers on this subject emerge in the next year. However, for now, the only published report focuses on a peptide that uses a dynamic rocking mechanism to transport Zn(II).299, 318 Below, we briefly summarize work on the natural peptide channels, gramicidin A and alamethicin, as well as a natural four-helical proton channel. We then highlight progress in the design of peptides that assemble into ion-conducting channels.
Gramicidin A and Alamethicin.
Gramicidin A from Bracilus brevis, a bacteriostatic peptide comprised of alternating d- and l-amino acids, forms head-to-head dimers in lipid membranes to selectively transport monovalent cations, including H+, K+, and Na+.319–323 Carbonyl groups of the peptide bonds line the pore, and although they are hydrogen-bonded to amide NH groups, they are able to tilt to stabilize transient cations. The conformational dynamics of the β-helices drive ion transport, and conduction properties are affected by its interactions with the surrounding lipid.324–326 Inspired by β-helix structure of gramicidin, early peptide designs focused on synthetic polypeptides that formed similar helical architectures that open to allow for the permeation of ions.327, 328 Taking a different approach, Ghadiri and colleagues similarly designed cyclic tetra- and hexapeptides that assemble into cation-selective channels.329
Unlike gramicidin, alamethicin, an antibiotic peptide from Trichoderma viride, forms “barrel stave” pores lined by α-helices, which more closely resemble ion channel proteins.300 This 20-residue peptide consisting of interspersed α-aminoisobutyric acid residues is postulated to associate into n = 4–8 α-helical bundles with its polar face, which contains the Gln and Glu residues, pointed to the center of the pore.73, 300, 330–337 While a structure exists for the monomer, there is no structure available for the alamethicin channel.338 Nevertheless, MD simulations of various oligomeric states (n = 5–8) reveals stability in all of the oligomeric states which corroborates findings on the multiplicity of conductance levels observed for this channel.339 Also, a lower-resolution model of the channel has been determined by atomic force microscopy and is consistent with the barrel stave model.340
M2, an example from nature that illustrates the requirements for selective proton transport.
The proton channel from influenza A virus is a small membrane protein, and the founding member of the viroporin family or proteins.341–346 M2’s proton channel is essential to the life cycle of the virus; it mediates the acidification of the interior of the virus,347–349 necessary for release of viral RNA. Following the discovery of M2, a large number of viruses have been found to also express small membrane proteins, many of which have ion channel activity.341–346 Most topically, the structure of the E-protein, a virporin from SARS-CoV-2 has been solved at moderate resolution by solids NMR,350 and is an area of very active investigation.
The 24-residue TM helix of the M2 proton channel from the influenza A virus is the minimal construct of the M2 polypeptide that still retains its proton channel activity.351 Like M2 it assembles into a parallel homotetramer with a water-filled channel lumen composed of apolar residues that leads to the proton-shuttling His37 tetrad (Figure 5). Very high-resolution crystal structures of M2 (<1.1 Å),294, 352, 353 together with solution354 and solids NMR structures,355, 356 have revealed the interactions that stabilize the protein, and the path taken by protons passing through the channel. Protons enter through a narrow hydrophobic sphincter and then travel along “water wires” to protonate a critical His37 residue. High-resolution crystallographic structures352, 353 reveal that the backbone carbonyls of the pore-facing hydrophobic residues, along with the interlumenal waters, constitute the extensive hydrogen-bonding network that both lowers the energy barrier for conductance and stabilizes excess positive charge during proton conduction.357 Based on different solids NMR measurements there is currently an unresolved question of whether or not the His37 residues are hydrogen bonded to one another.358–360 In every independent crystallographic or solution NMR structure of the channel that we have solved, the His37 residues are either fully hydrated in one conducting state or indirectly involved in edge-on aromatic interactions with strong water-mediated hydrogen bonds in a second conformational state,294, 352, 353, 361–365 which agrees well with the model of M. Hong.358 The conformation of the protein varies with the degree of protonation of His37, and in the proton-conduction cycle, this region of the channel becomes more and less open as protons come on and off.366 Voth’s group has used classical MD, reactive MD, and QM calculations to evaluate the conduction mechanism, which are able to quantitatively predict the conductance of the channel as well as the mechanism of binding to small molecule drugs.357, 367–369
Fig. 5: Key structural features that define the proton channel activity for M2.

The TM domain of M2 forms a proton-conductive channel. Here, we see the drug, amantadine, binding to the channel in the hydrophobic gasket with its ammonium group pointed towards the water-filled pore. The ammonium group acts as a proton and locks the interlumenal waters near the gating His in place, essentially blocking proton conduction.
Just below the His37 residue, a ring of four Trp41 residues form tight aromatic–aromatic or aromatic–cation interactions with His37, which is further stabilized by a solvent-mediated hydrogen bond to Asp44.370 This gate is interrupted when the His37 residues reach a critical protonation state of approximately +3 and a proton can move into the interior. Thus, M2 has a transporter-like mechanism.351, 371–374
The contribution of each residue to the free energy of tetramerization of the channel has been evaluated in a series of mutants in micelles and bilayers. No single residue is absolutely required for tetramerization, although some substitutions result in a loss of thermodynamic stability,295, 375, 376 particularly when Ala is substituted for His37. On the other hand, these mutants have substantially altered conductance and/or proton selectivity.296 It is noteworthy that all the hydrogen bonds within the TM region are water-mediated rather than direct, with the only direct interaction being a salt bridge at the hydrated exit of the channel. Even the polar ammonium group of channel-blocking drugs form indirect, water-mediate hydrogen bonds to the channel (Figure 5).362–364, 377 This finding underscores the importance of considering water in the stability and function of this channel.
Finally, no discussion of M2 would be complete without mentioning some early debates, which are now resolved. Early work questioned whether a second, cytoplasmic helix was needed for the structural integrity of the protein, but this was addressed thoroughly by deletion mutagenesis of M2 expressed in mammalian cells351 and by solving crystallographic362 and NMR structures378 with and without C-terminal extensions. Instead, the C-terminal cytoplasmic helix was found to be important for budding of the virus.71, 379 Also in early work, Chou et al. suggested that the channel-blocker amantadine (which was clearly observed in a crystal structures361, 364, 377) might instead bind on the outside of the channel.354 This debate was also resolved when Chou was able to solve a structure with drug bound in the interior pore.380
De novo designed ion channels.
Designed peptide channels offer the opportunity to decipher the complexities of channel function in simple, minimalist systems. Beyond the considerations for peptide–peptide and peptide–membrane interactions in the design of membrane-soluble helical bundles, in designing functional ion channels, we need to account for water- and ion-accessibility of the pore. In the late 1980s, DeGrado and coworkers designed the first TM α-helical peptide channel using a repeating sequence of Leu and Ser: H2N-(Leu-Ser-Leu-Leu-Leu-Ser-Leu)3-CONH2 and H2N-(Leu-Ser-Ser-Leu-Leu-Ser-Leu)3-CONH2, or LS2 and LS3, respectively.313, 316 The Leu residues were chosen to pack well at helix–helix interfaces, while the Ser residues are important for both packing and the stabilization of a proton-conducting pore. The packing of small Ser residues near the core and larger Leu sidechains at the helical interfaces dictated a tetrameric arrangement for the proton channel, LS2. The Ser hydroxyls, together with water molecules, appeared to create a proton conduction pathway via a water-hopping mechanism. LS3 formed hexameric channels with a pore large enough to accommodate a solvated ion in the hexameric bundle. While crystallographic structures were not available at the time, a large body of subsequent data, including the incorporation of LS2 peptides onto a rigid tetrameric scaffold,381 supported the hypothetical structures and conduction model.313, 314, 316, 382–384
Recently, Woolfson and colleagues approached the design of α-helical channels by converting water-soluble coiled coils into membrane-spanning assemblies.385 Unlike previous designs which borrowed from the D4 domain of E. coli polysaccharide transporter Wza,386 these designs were purely de novo. Their work began with a careful consideration of the features required to pack helices into various sized helical bundles of various sizes.387 They showed that the packing of small residues at positions designated “e” and “g” of a coiled coil dictate the formation oligomeric bundles with sizes ranging from 5 to 8 helices, depending on the nature of the small residues and other features. In the initial work, the residues in the core at positions “a” and “d” were large Leu and Ile residues, respectively. They reasoned at least some of these Leu and Ile residues could be changed to combinations of Ser and Thr to form water-filled pores, which was validated by crystal structures of the water-soluble peptides. Next, they converted the peptides to membrane peptides by introducing apolar residues at the exterior positions.385 Single-channel recordings were consistent with the designs, suggesting that the designed structures had been realized. However, the peptides failed to crystallize in the active channel-forming states. Nevertheless, an unanticipated structure of a non-conducting antiparallel tetrameric state (Figure 4) illustrated important design principles of de novo membrane protein design.
The most ambitious functional membrane protein designed to date is a TM four-helix bundle, Rocker (Figure 6), that transports first-row transition metal ions Zn2+ in exchange for protons.299, 318 Rocker has two 4Glu, 4-His Zn2+ binding sites, which can alternately accommodate protons and metal ions. A computational design algorithm was used to stabilize two energetically degenerate asymmetric states of the protein while destabilizing a competing fully symmetrical state which might otherwise bind metal ions too tightly and impede motions required for ion transport. The computed TM bundle formed a dimer of dimers with two non-equivalent helix–helix interfaces (Figure 6). A “tight interface” had a small inter-helical distance (8.9 Å) stabilized by efficient packing of small Ala residues in an Ala-coil motif, which are seen in the crystal structure of the protein. The “loose interface” had a larger interhelical distance of 12.0 Å and was less well packed. Solids NMR was thus used to define the structure of the dimer of dimers. The resulting membrane-spanning four-helical bundle transported first-row transition metal ions Zn2+ and Co2+, but not Ca2+ across membranes. Vesicle flux experiments show that as Zn2+ ions diffuse down their concentration gradients, protons were antiported. These experiments illustrate the feasibility of designing membrane proteins with predefined structural and dynamic properties. These studies illustrate the how multiple forces, including structural motifs and diverse polar interactions can be combined to build functional assemblies.
OUTLOOK
This review has given only the smallest glimpse of the vibrant and growing field of peptide–membrane interactions. Peptides have proven to be outstanding systems for understanding larger proteins, particularly for deciphering the rules of membrane peptide/protein association, insertion, and folding. This field has also spawned antimicrobial peptides and peptide mimetics currently in the clinic, while also inspiring new areas of antimicrobial polymer science.103, 106, 388–395 Clearly, the future is very bright.
FUNDING
B.M. and W.F.D were supported by NIH grant R35-GM122603. H.T.K was supported by NIH grant K99-GM138753. R.W.N. was supported by NIH grant K99-NS116679.
REFERENCES
- 1.White SH, J. Gen. Physiol, 2007, 129, 363–369. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Meindl-Beinker NM, Lundin C, Nilsson I, White SH and von Heijne G, EMBO Rep, 2006, 7, 1111–1116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Wimley WC and White SH, Biochemistry, 2000, 39, 4432–4442. [DOI] [PubMed] [Google Scholar]
- 4.Wolfe C, Cladera J and O’Shea P, Mol. Membr. Biol, 1998, 15, 221–227. [DOI] [PubMed] [Google Scholar]
- 5.McLaughlin S and Aderem A, Trends Biochem. Sci, 1995, 20, 272–276. [DOI] [PubMed] [Google Scholar]
- 6.Nguyen LT, Haney EF and Vogel HJ, Trends Biotechnol, 2011, 29, 464–472. [DOI] [PubMed] [Google Scholar]
- 7.Day KJ, Kago G, Wang L, Richter JB, Hayden CC, Lafer EM and Stachowiak JC, Nat. Cell Biol, 2021, 23, 366–376. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Yuan F, Alimohamadi H, Bakka B, Trementozzi AN, Day KJ, Fawzi NL, Rangamani P and Stachowiak JC, Proc. Natl. Acad. Sci. U.S.A, 2021, 118, e2017435118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Tanford C, Science, 1978, 200, 1012–1018. [DOI] [PubMed] [Google Scholar]
- 10.Segrest JP, Jackson RL, Morrisett JD and Gotto AM Jr., FEBS Lett, 1974, 38, 247–258. [DOI] [PubMed] [Google Scholar]
- 11.Jackson RL, Morrisett JD, Gotto AM Jr. and Segrest JP, Mol. Cell Biochem, 1975, 6, 43–50. [DOI] [PubMed] [Google Scholar]
- 12.Segrest JP, Chung BH, Brouillette CG, Kanellis P and McGahan R, J. Biol. Chem, 1983, 258, 2290–2295. [PubMed] [Google Scholar]
- 13.Epand RM, Shai YC, Segrest JP and Anantharamaiah GM, Biopolymers, 1995, 37, 319–338. [DOI] [PubMed] [Google Scholar]
- 14.Mishra VK, Palgunachari MN, Segrest JP and Anantharamaiah GM, J. Biol. Chem, 1994, 269, 7185–7191. [PubMed] [Google Scholar]
- 15.Segrest JP, Jones MK, De Loof H and Dashti N, J. Lipid Res, 2001, 42, 1346–1367. [PubMed] [Google Scholar]
- 16.Segrest JP, Jones MK, Klon AE, Sheldahl CJ, Hellinger M, De Loof H and Harvey SC, J. Biol. Chem, 1999, 274, 31755–31758. [DOI] [PubMed] [Google Scholar]
- 17.Chung BH, Anatharamaiah GM, Brouillette CG, Nishida T and Segrest JP, J. Biol. Chem, 1985, 260, 10256–10262. [PubMed] [Google Scholar]
- 18.Anantharamaiah GM, Jones JL, Brouillette CG, Schmidt CF, Chung BH, Hughes TA, Bhown AS and Segrest JP, J. Biol. Chem, 1985, 260, 10248–10255. [PubMed] [Google Scholar]
- 19.Kaiser ET and Kezdy FJ, Annu. Rev. Biophys. Biophys. Chem, 1987, 16, 561–581. [DOI] [PubMed] [Google Scholar]
- 20.Kaiser ET and Kezdy FJ, Science, 1984, 223, 249–255. [DOI] [PubMed] [Google Scholar]
- 21.Lau SH, Rivier J, Vale W, Kaiser ET and Kezdy FJ, Proc. Natl. Acad. Sci. U.S.A, 1983, 80, 7070–7074. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Kaiser ET and Kezdy FJ, Proc. Natl. Acad. Sci. U.S.A, 1983, 80, 1137–1143. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Fukushima D, Yokoyama S, Kezdy FJ and Kaiser ET, Proc. Natl. Acad. Sci. U.S.A, 1981, 78, 2732–2736. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Kornmueller K, Lehofer B, Leitinger G, Amenitsch H and Prassl R, Nano Res, 2018, 11, 913–928. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Dufourc EJ, Biochim. Biophys. Acta, 2021, 1863, 183478. [DOI] [PubMed] [Google Scholar]
- 26.DeGrado WF, Kèzdy FJ and Kaiser ET, J. Am. Chem. Soc, 1981, 103, 679–681. [Google Scholar]
- 27.DeGrado WF, Musso GF, Lieber M, Kaiser ET and Kézdy FJ, Biophys. J, 1982, 37, 329–338. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Hirai Y, Yasuhara T, Yoshida H, Nakajima T, Fujino M and Kitada C, Chem. Pharm. Bull, 1979, 27, 1942–1944. [DOI] [PubMed] [Google Scholar]
- 29.DeGrado WF, Adv. Prot. Chem, 1988, 39, 51–124. [DOI] [PubMed] [Google Scholar]
- 30.Hultmark D, Engstrom A, Bennich H, Kapur R and Boman HG, Eur. J. Biochem, 1982, 127, 207–217. [DOI] [PubMed] [Google Scholar]
- 31.Boman HG, Faye I, Gan R, Gudmundsson GH, Lidholm DA, Lee JY and Xanthopoulos KG, Mem. Inst. Oswaldo Cruz, 1987, 82, 115–124. [DOI] [PubMed] [Google Scholar]
- 32.Andreu D, Merrifield RB, Steiner H and Boman HG, Proc. Natl. Acad. Sci. U.S.A, 1983, 80, 6475–6479. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Merrifield RB, Vizioli LD and Boman HG, Biochemistry, 1982, 21, 5020–5031. [DOI] [PubMed] [Google Scholar]
- 34.DeGrado WF, Peptides: Structure and Function. Proceedings of the eighth American Peptide Symposium., 1983, 1983. [Google Scholar]
- 35.Zasloff M, Proc. Natl. Acad. Sci. U.S.A, 1987, 84, 5449–5453. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Zasloff M, Nature, 2002, 415, 389–395. [DOI] [PubMed] [Google Scholar]
- 37.Tossi A, Sandri L and Giangaspero A, Biopolymers, 2000, 55, 4–30. [DOI] [PubMed] [Google Scholar]
- 38.Bishop JL and Finlay BB, Trends Mol. Med, 2006, 12, 3–6. [DOI] [PubMed] [Google Scholar]
- 39.Brogden KA, Nat. Rev. Microbiol, 2005, 3, 238–250. [DOI] [PubMed] [Google Scholar]
- 40.Shai Y, Curr. Pharm. Des, 2002, 8, 715–725. [DOI] [PubMed] [Google Scholar]
- 41.Sarah RD, Frederick H, Manuela M and David AP, Curr. Protein Pept. Sci, 2018, 19, 823–838.29484989 [Google Scholar]
- 42.Shagaghi N, Palombo EA, Clayton AHA and Bhave M, World J. Microbiol. Biotechnol, 2016, 32, 31. [DOI] [PubMed] [Google Scholar]
- 43.Chan DI, Prenner EJ and Vogel HJ, Biochim. Biophys. Acta, 2006, 1758, 1184–1202. [DOI] [PubMed] [Google Scholar]
- 44.Arias M, Piga KB, Hyndman ME and Vogel HJ, Biomolecules, 2018, 8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Wang G, Antimicrobial peptides: discovery, design and novel therapeutic strategies, CABI Publishing, 2010. [Google Scholar]
- 46.Matsuzaki K, ed., Antimicrobial Peptides: Basics for Clinical Application, Springer, Singapore, 2019. [Google Scholar]
- 47.Park CB, Kim HS and Kim SC, Biochem. Biophys. Res. Commun, 1998, 244, 253–257. [DOI] [PubMed] [Google Scholar]
- 48.Kragol G, Lovas S, Varadi G, Condie BA, Hoffmann R and Otvos L Jr., Biochemistry, 2001, 40, 3016–3026. [DOI] [PubMed] [Google Scholar]
- 49.del Castillo FJ, del Castillo I and Moreno F, J. Bacteriol, 2001, 183, 2137–2140. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Huang HW, Biochim. Biophys. Acta, 2006, 1758, 1292–1302. [DOI] [PubMed] [Google Scholar]
- 51.Nielsen JE, Bjornestad VA, Pipich V, Jenssen H and Lund R, J. Colloid Interface Sci, 2021, 582, 793–802. [DOI] [PubMed] [Google Scholar]
- 52.Shai Y, Biopolymers, 2002, 66, 236–248. [DOI] [PubMed] [Google Scholar]
- 53.Shai Y, Biochim. Biophys. Acta, 1999, 1462, 55–70. [DOI] [PubMed] [Google Scholar]
- 54.Roversi D, Luca V, Aureli S, Park Y, Mangoni ML and Stella L, ACS Chem. Biol, 2014, 9, 2003–2007. [DOI] [PubMed] [Google Scholar]
- 55.Matsuzaki K, Biochim. Biophys. Acta, 2009, 1788, 1687–1692. [DOI] [PubMed] [Google Scholar]
- 56.Matsuzaki K, Biochim. Biophys. Acta, 1999, 1462, 1–10. [DOI] [PubMed] [Google Scholar]
- 57.Matsuzaki K, Sugishita K, Ishibe N, Ueha M, Nakata S, Miyajima K and Epand RM, Biochemistry, 1998, 37, 11856–11863. [DOI] [PubMed] [Google Scholar]
- 58.Matsuzaki K, Biochim. Biophys. Acta, 1998, 1376, 391–400. [DOI] [PubMed] [Google Scholar]
- 59.Matsuzaki K, Murase O, Fujii N and Miyajima K, Biochemistry, 1996, 35, 11361–11368. [DOI] [PubMed] [Google Scholar]
- 60.Matsuzaki K, Sugishita K, Fujii N and Miyajima K, Biochemistry, 1995, 34, 3423–3429. [DOI] [PubMed] [Google Scholar]
- 61.Sato H and Feix JB, Biochim. Biophys. Acta, 2006, 1758, 1245–1256. [DOI] [PubMed] [Google Scholar]
- 62.Wimley WC and Hristova K, Aust. J. Chem, 2020, 73, 96–103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Guha S, Ghimire J, Wu E and Wimley WC, Chem. Rev, 2019, 119, 6040–6085. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Huang HW, Biochemistry, 2000, 39, 8347–8352. [DOI] [PubMed] [Google Scholar]
- 65.Lopez C, Nielsen S, Srinivas G, WF D and MK K, J. Chem. Theory Comp, 2006, 2, 649–655. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Tieleman DP, Biophys. J, 2017, 113, 1–3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Yang LH, Gordon VD, Mishra A, Sorn A, Purdy KR, Davis MA, Tew GN and Wong GCL, J. Am. Chem. Soc, 2007, 129, 12141–12147. [DOI] [PubMed] [Google Scholar]
- 68.Yang LH, Gordon VD, Trinkle DR, Schmidt NW, Davis MA, DeVries C, Som A, Cronan JE, Tew GN and Wong GCL, Proc. Natl. Acad. Sci. U.S.A, 2008, 105, 20595–20600. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Schmidt NW, Mishra A, Lai GH, Davis M, Sanders LK, Tran D, Garcia A, Tai KP, McCray PB, Ouellette AJ, Selsted ME and Wong GCL, J. Am. Chem. Soc, 2011, 133, 6720–6727. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Schmidt NW and Wong GCL, Curr. Opin. Solid State Mater. Sci, 2013, 17, 151–163. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Schmidt NW, Mishra A, Wang J, DeGrado WF and Wong GC, J. Am. Chem. Soc, 2013, 135, 13710–13719. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Lee EY, Fulan BM, Wong GCL and Ferguson AL, Proc. Natl. Acad. Sci. U.S.A, 2016, 113, 13588. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Bechinger B, J. Membr. Biol, 1997, 156, 197–211. [DOI] [PubMed] [Google Scholar]
- 74.Pal S, Chakraborty H and Chattopadhyay A, J. Phys. Chem. B, 2021, 125, 8450–8459. [DOI] [PubMed] [Google Scholar]
- 75.Epand RF, Mowery BP, Lee SE, Stahl SS, Lehrer RI, Gellman SH and Epand RM, J. Mol. Biol, 2008, 379, 38–50. [DOI] [PubMed] [Google Scholar]
- 76.Epand RF, Schmitt MA, Gellman SH and Epand RM, Biochim. Biophys. Acta, 2006. [DOI] [PubMed] [Google Scholar]
- 77.Epand RF, Schmitt MA, Gellman SH, Sen A, Auger M, Hughes DW and Epand RM, Mol. Membr. Biol, 2005, 22, 457–469. [DOI] [PubMed] [Google Scholar]
- 78.Epand RF, Raguse TL, Gellman SH and Epand RM, Biochemistry, 2004, 43, 9527–9535. [DOI] [PubMed] [Google Scholar]
- 79.Shai Y and Oren Z, Peptides, 2001, 22, 1629–1641. [DOI] [PubMed] [Google Scholar]
- 80.Chen L, Yin H, Farooqi B, Sebti S, Hamilton AD and Chen J, Mol. Cancer Ther, 2005, 4, 1019–1025. [DOI] [PubMed] [Google Scholar]
- 81.Prenner EJ, Kiricsi M, Jelokhani-Niaraki M, Lewis RN, Hodges RS and McElhaney RN, J. Biol. Chem, 2005, 280, 2002–2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Braunstein A, Papo N and Shai Y, Antimicrob. Agents Chemother, 2004, 48, 3127–3129. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Shai Y, Makovitzky A and Avrahami D, Curr. Protein Pept. Sci, 2006, 7, 479–486. [DOI] [PubMed] [Google Scholar]
- 84.Makovitzki A, Baram J and Shai Y, Biochemistry, 2008, 47, 10630–10636. [DOI] [PubMed] [Google Scholar]
- 85.Fernandez-Lopez S, Kim HS, Choi EC, Delgado M, Granja JR, Khasanov A, Kraehenbuehl K, Long G, Weinber DA, wilcoxen KM and Ghadiri MR, Nature, 2001, 412, 452–455. [DOI] [PubMed] [Google Scholar]
- 86.Horne WS and Gellman SH, Acc. Chem. Res, 2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Cheng RP, Gellman SH and DeGrado WF, Chem. Rev, 2001, 101, 3219–3232. [DOI] [PubMed] [Google Scholar]
- 88.Gellman SH, Acc. Chem. Res, 1998, 31, 173–180. [Google Scholar]
- 89.Seebach D, Hook DF and Glattli A, Biopolymers, 2006, 84, 23–37. [DOI] [PubMed] [Google Scholar]
- 90.Seebach D, Beck AK and Bierbaum DJ, Chem. Biodivers, 2004, 1, 1111–1239. [DOI] [PubMed] [Google Scholar]
- 91.Hamuro Y, Schneider JP and DeGrado WF, J. Am. Chem. Soc, 1999, 121, 12200–12201. [Google Scholar]
- 92.Liu D and DeGrado W, J. Am. Chem. Soc, 2001, 123, 7553–7559. [DOI] [PubMed] [Google Scholar]
- 93.Porter EA, Wang X, Lee HS, Weisblum B and Gellman SH, Nature, 2000, 404, 565. [DOI] [PubMed] [Google Scholar]
- 94.Arvidsson PI, Ryder NS, Weiss HM, Hook DF, Escalante J and Seebach D, Chem. Biodivers, 2005, 2, 401–420. [DOI] [PubMed] [Google Scholar]
- 95.Schmitt MA, Weisblum B and Gellman SH, J. Am. Chem. Soc, 2007, 129, 417–428. [DOI] [PubMed] [Google Scholar]
- 96.Porter EA, Weisblum B and Gellman SH, J. Am. Chem. Soc, 2005, 127, 11516–11529. [DOI] [PubMed] [Google Scholar]
- 97.Schmitt MA, Weisblum B and Gellman SH, J. Am. Chem. Soc, 2004, 126, 6848–6849. [DOI] [PubMed] [Google Scholar]
- 98.Patch JA and Barron AE, J. Am. Chem. Soc, 2003, 125, 12092–12093. [DOI] [PubMed] [Google Scholar]
- 99.Violette A, Fournel S, Lamour K, Chaloin O, Frisch B, Briand JP, Monteil H and Guichard G, Chem. Biol, 2006, 13, 531–538. [DOI] [PubMed] [Google Scholar]
- 100.Kuroki A, Sangwan P, Qu Y, Peltier R, Sanchez-Cano C, Moat J, Dowson CG, Williams EGL, Locock KES, Hartlieb M and Perrier S, ACS Appl. Mater. Interfaces, 2017, 9, 40117–40126. [DOI] [PubMed] [Google Scholar]
- 101.Takahashi H, Caputo GA and Kuroda K, Biomater. Sci, 2021, 9, 2758–2767. [DOI] [PubMed] [Google Scholar]
- 102.Bhat R, Foster LL, Rani G, Vemparala S and Kuroda K, RSC Adv, 2021, 11, 22044–22056. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Tew GN, Liu D, Chen B, Doerksen RJ, Kaplan J, Carroll PJ, Klein ML and DeGrado WF, Proc. Natl. Acad. Sci. U.S.A, 2002, 99, 5110–5114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Liu D, Choi S, Chen B, Doerksen RJ, Clements DJ, Winkler JD, Klein ML and DeGrado WF, Angew. Chem. Int. Ed. Engl, 2004, 43, 1158–1162. [DOI] [PubMed] [Google Scholar]
- 105.Ivanov I, Vemparala S, Pophristic V, Kuroda K, DeGrado WF, McCammon JA and Klein ML, J. Am. Chem. Soc, 2006, 128, 1778–1779. [DOI] [PubMed] [Google Scholar]
- 106.Tew GN, Scott RW, Klein ML and DeGrado WF, Acc. Chem. Res, 2010, 43, 30–39. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Scott RW, DeGrado WF and Tew GN, Curr. Opin. Biotechnol, 2008, 19, 620–627. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Choi S, Isaacs A, Clements D, Liu D, Kim H, Scott RW, Winkler JD and DeGrado WF, Proc. Natl. Acad. Sci. U.S.A, 2009, 106, 6968–6973. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Scott RW and Tew GN, Curr. Top. Med. Chem, 2017, 17, 576–589. [DOI] [PubMed] [Google Scholar]
- 110.Zhao H, To KKW, Sze KH, Yung TT, Bian M, Lam H, Yeung ML, Li C, Chu H and Yuen KY, Nat. Commun, 2020, 11, 4252. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Barlow PG, Svoboda P, Mackellar A, Nash AA, York IA, Pohl J, Davidson DJ and Donis RO, PLoS One, 2011, 6, e25333. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Salata C, Calistri A, Parolin C, Baritussio A and Palu G, Expert Rev. Anti-Infect. Ther, 2017, 15, 483–492. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Gunesch AP, Zapatero-Belinchon FJ, Pinkert L, Steinmann E, Manns MP, Schneider G, Pietschmann T, Bronstrup M and von Hahn T, Antimicrob. Agents Chemother, 2020, 64. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Bakovic A, Risner K, Bhalla N, Alem F, Chang TL, Weston WK, Harness JA and Narayanan A, Viruses, 2021, 13. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Mensa B, Kim YH, Choi S, Scott R, Caputo GA and DeGrado WF, Antimicrob. Agents Chemother, 2011, 55, 5043–5053. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Mensa B, Howell GL, Scott R and DeGrado WF, Antimicrob. Agents Chemother, 2014, 58, 5136–5145. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Goode A, Yeh V and Bonev BB, Faraday Discuss, 2021, DOI: 10.1039/d1fd00036e. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Bonev BB, Breukink E, Swiezewska E, De Kruijff B and Watts A, FASEB J, 2004, 18, 1862–1869. [DOI] [PubMed] [Google Scholar]
- 119.Mahlapuu M, Håkansson J, Ringstad L and Björn C, Front. Cell. Infect. Microbiol, 2016, 6, 194. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120.Chen CH and Lu TK, Antibiotics, 2020, 9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Pfefferkorn CM, Jiang Z and Lee JC, Biochim. Biophys. Acta, 2012, 1818, 162–171. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122.Maroteaux L, Campanelli JT and Scheller RH, J. Neurosci, 1988, 8, 2804–2815. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.George JM, Jin H, Woods WS and Clayton DF, Neuron, 1995, 15, 361–372. [DOI] [PubMed] [Google Scholar]
- 124.Segrest JP, Loof HD, Dohlman JG, Brouillette CG and Anantharamaiah GM, Proteins, 1990, 8, 103–117. [DOI] [PubMed] [Google Scholar]
- 125.Fortin DL, Nemani VM, Voglmaier SM, Anthony MD, Ryan TA and Edwards RH, J. Neurosci, 2005, 25, 10913–10921. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126.Abeliovich A, Schmitz Y, Farinas I, Choi-Lundberg D, Ho W-H, Verdugo JMG, Armanini M, Ryan A, Hynes M, Phillips H, Sulzer D and Rosenthal A, Neuron, 2000, 25, 239–252. [DOI] [PubMed] [Google Scholar]
- 127.Polymeropoulos MH, Lavedan C, Leroy E, Ide SE, Dehejia A, Dutra A, Pike B, Root H, Rubenstein J, Boyer R, Stenroos ES, Chandrasekharappa S, Athanassiadou A, Papapetropoulos T, Johnson WG, Lazzarini AM, Duvoisin RC, Di Iorio G, Golbe LI and Nussbaum RL, Science, 1997, 276, 2045–2047. [DOI] [PubMed] [Google Scholar]
- 128.Spillantini MG, Schmidt ML, Lee VM, Trojanowski JQ, Jakes R and Goedert M, Nature, 1997, 388, 839–840. [DOI] [PubMed] [Google Scholar]
- 129.McLachlan AD, Nature, 1977, 267, 465–466. [DOI] [PubMed] [Google Scholar]
- 130.Trexler AJ and Rhoades E, Biochemistry, 2009, 49, 2304–2306. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131.Jao CC, Der-Sarkissian A, Chen J and Langen R, Proc. Natl. Acad. Sci. U.S.A, 2004, 101, 8331–8336. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.Jao CC, Hegde BG, Chen J, Haworth IS and Langen R, Proc. Natl. Acad. Sci. U.S.A, 2008, 105, 19666–19671. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133.Bodner CR, Dobson CM and Bax A, J. Mol. Biol, 2009, 390, 775–790. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134.Fusco G, De Simone A, Gopinath T, Vostrikov V, Vendruscolo M, Dobson CM and Veglia G, Nat. Commun, 2014, 5, 3827. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135.Newberry RW, Arhar T, Costello J, Hartoularos GC, Maxwell AM, Naing ZZC, Pittman M, Reddy NR, Schwarz DMC, Wassarman DR, Wu TS, Barrero D, Caggiano C, Catching A, Cavazos TB, Estes LS, Faust B, Fink EA, Goldman MA, Gomez YK, Gordon MG, Gunsalus LM, Hoppe N, Jaime-Garza M, Johnson MC, Jones MG, Kung AF, Lopez KE, Lumpe J, Martyn C, McCarthy EE, Miller-Vedam LE, Navarro EJ, Palar A, Pellegrino J, Saylor W, Stephens CA, Strickland J, Torosyan H, Wankowicz SA, Wong DR, Wong G, Redding S, Chow ED, DeGrado WF and Kampmann M, ACS Chem. Biol, 2020, 15, 2137–2153. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 136.Fusco G, Pape T, Stephens AD, Mahou P, Costa AR, Kaminski CF, Kaminski Schierle GS, Vendruscolo M, Veglia G, Dobson CM and De Simone A, Nat. Commun, 2016, 7, 12563. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137.Logan T, Bendor J, Toupin C, Thorn K and Edwards RH, Nat. Neurosci, 2017, 20, 681–689. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 138.Weinreb PH, Zhen W, Poon AW, Conway KA and Lansbury PT Jr., Biochemistry, 1996, 35, 13709–13715. [DOI] [PubMed] [Google Scholar]
- 139.Davidson WS, Jonas A, Clayton DF and George JM, J. Biol. Chem, 1998, 16, 9443–9449. [DOI] [PubMed] [Google Scholar]
- 140.Antonny B, Annu. Rev. Biochem, 2011, 80, 101–123. [DOI] [PubMed] [Google Scholar]
- 141.Drin G and Antonny B, FEBS Lett, 2010, 584, 1840–1847. [DOI] [PubMed] [Google Scholar]
- 142.Dettmer U, Front. Neurosci, 2018, 12, 623. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 143.Auluck PK, Caraveo G and Lindquist S, Annu. Rev. Cell Dev. Biol, 2010, 26, 211–233. [DOI] [PubMed] [Google Scholar]
- 144.Pranke IM, Morello V, Bigay J, Gibson K, Verbavatz JM, Antonny B and Jackson CL, J. Cell Biol, 2011, 194, 89–103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145.Middleton ER and Rhoades E, Biophys. J, 2010, 99, 2279–2288. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 146.Zemel A, Ben-Shaul A and May S, J. Phys. Chem. B, 2008, 112, 6988–6996. [DOI] [PubMed] [Google Scholar]
- 147.Braun AR, Lacy MM, Ducas VC, Rhoades E and Sachs JN, J. Am. Chem. Soc, 2014, 136, 9962–9972. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148.Braun AR, Lacy MM, Ducas VC, Rhoades E and Sachs JN, J. Membr. Biol, 2017, 250, 183–193. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 149.Copic A, Antoine-Bally S, Gimenez-Andres M, La Torre Garay C, Antonny B, Manni MM, Pagnotta S, Guihot J and Jackson CL, Nat. Commun, 2018, 9, 1332. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 150.Eliezer D, Kutluay E, Bussell R Jr. and Browne G, J. Mol. Biol, 2001, 307, 1061–1073. [DOI] [PubMed] [Google Scholar]
- 151.Chandra S, Chen X, Rizo J, Jahn R and Südhof TC, J. Biol. Chem, 2003, 278, 15313–15318. [DOI] [PubMed] [Google Scholar]
- 152.Ulmer TS, Bax A, Cole NB and Nussbaum RL, J. Biol. Chem, 2005, 280, 9595–9603. [DOI] [PubMed] [Google Scholar]
- 153.Fowler DM and Fields S, Nat. Methods, 2014, 11, 801–807. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154.Schramm CA, Hannigan BT, Donald JE, Keasar C, Saven JG, Degrado WF and Samish I, Structure, 2012, 20, 924–935. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155.Bussell R Jr. and Eliezer D, J. Mol. Biol, 2003, 329, 763–778. [DOI] [PubMed] [Google Scholar]
- 156.Shahmoradian SH, Lewis AJ, Genoud C, Hench J, Moors TE, Navarro PP, Castaño-Díez D, Schweighauser G, Graff-Meyer A, Goldie KN, Sütterlin R, Huisman E, Ingrassia A, Gier Y, Rozemuller AJM, Wang J, Paepe A, Erny J, Staempfli A, Hoernschemeyer J, Grosserüschkamp F, Niedieker D, El-Mashtoly SF, Quadri M, Van IJcken WFJ, Bonifati V, Gerwert K, Bohrmann B, Frank S, Britschgi M, Stahlberg H, Van de Berg WDJ and Lauer ME, Nat. Neurosci, 2019, 22, 1099–1109. [DOI] [PubMed] [Google Scholar]
- 157.Trinkaus VA, Riera-Tur I, Martinez-Sanchez A, Bauerlein FJB, Guo Q, Arzberger T, Baumeister W, Dudanova I, Hipp MS, Hartl FU and Fernandez-Busnadiego R, Nat. Commun, 2021, 12, 2110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 158.Burré J, Sharma M, Tsetsenis T, Buchman V, Etherton MR and Südhof TC, Science, 2010, 329, 1663–1667. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 159.Fakhree MAA, Engelbertink SAJ, van Leijenhorst-Groener KA, Blum C and Claessens M, Biomacromolecules, 2019, 20, 1217–1223. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 160.Zeno WF, Thatte AS, Wang L, Snead WT, Lafer EM and Stachowiak JC, J. Am. Chem. Soc, 2019, 141, 10361–10371. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 161.Jo E, McLaurin J, Yip CM, St George-Hyslop P and Fraser PE, J. Biol. Chem, 2000, 275, 34328–34334. [DOI] [PubMed] [Google Scholar]
- 162.Braun AR, Sevcsik E, Chin P, Rhoades E, Tristram-Nagle S and Sachs JN, J. Am. Chem. Soc, 2012, 134, 2613–2620. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 163.Galvagnion C, Buell AK, Meisl G, Michaels TCT, Vendruscolo M, Knowles TPJ and Dobson CM, Nat. Chem. Biol, 2015, 11, 229–234. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 164.Zhu M and Fink AL, J. Biol. Chem, 2003, 278, 16873–16877. [DOI] [PubMed] [Google Scholar]
- 165.Snead D and Eliezer D, J. Biol. Chem, 2019, 294, 3325–3342. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166.Newberry RW, Leong JT, Chow ED, Kampmann M and DeGrado WF, Nat. Chem. Biol, 2020, 16, 653–659. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 167.Bartels T, Choi JG and Selkoe DJ, Nature, 2011, 477, 107–110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 168.Fusco G, Chen SW, Williamson PTF, Cascella R, Perni M, Jarvis JA, Cecchi C, Vendruscolo M, Chiti F, Cremades N, Ying L, Dobson CM and Simone AD, Science, 2017, 358, 1440–1443. [DOI] [PubMed] [Google Scholar]
- 169.Von Heijne G, Adv. Protein Chem, 2003, 63, 1–18. [DOI] [PubMed] [Google Scholar]
- 170.White SH and von Heijne G, Curr. Opin. Struct. Biol, 2004, 14, 397–404. [DOI] [PubMed] [Google Scholar]
- 171.Engelman DM and Steitz TA, Cell, 1981, 23, 411–422. [DOI] [PubMed] [Google Scholar]
- 172.White SH and Wimley WC, Annu. Rev. Biophys. Biomol. Struct, 1999, 28, 319–365. [DOI] [PubMed] [Google Scholar]
- 173.Engelman DM, Chen Y, Chin CN, Curran AR, Dixon AM, Dupuy AD, Lee AS, Lehnert U, Matthews EE, Reshetnyak YK, Senes A and Popot JL, FEBS Lett, 2003, 555, 122–125. [DOI] [PubMed] [Google Scholar]
- 174.Popot JL and Engelman DM, Annu. Rev. Biochem, 2000, 69, 881–922. [DOI] [PubMed] [Google Scholar]
- 175.Popot J-L and Engelman DM, Biochemistry, 1990, 29, 4031–4037. [DOI] [PubMed] [Google Scholar]
- 176.Lomize AL, Lomize MA, Krolicki SR and Pogozheva ID, Nucleic Acids Res, 2017, 45, D250–D255. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 177.Kirrbach J, Krugliak M, Ried CL, Pagel P, Arkin IT and Langosch D, Bioinformatics, 2013, 29, 1623–1630. [DOI] [PubMed] [Google Scholar]
- 178.Partridge AW, Liu S, Kim S, Bowie JU and Ginsberg MH, J. Biol. Chem, 2005, 28, 7294–7300. [DOI] [PubMed] [Google Scholar]
- 179.Schlebach JP and Sanders CR, Q. Rev. Biophys, 2015, 48, 1–34. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 180.White SH and von Heijne G, Biochem. Soc. Trans, 2005, 33, 1012–1015. [DOI] [PubMed] [Google Scholar]
- 181.Hessa T, Meindl-Beinker NM, Bernsel A, Kim H, Sato Y, Lerch-Bader M, Nilsson I, White SH and von Heijne G, Nature, 2007, 450, 1026–1030. [DOI] [PubMed] [Google Scholar]
- 182.Senes A, Chadi DC, Law PB, Walters RF, Nanda V and Degrado WF, J. Mol. Biol, 2007, 366, 436–448. [DOI] [PubMed] [Google Scholar]
- 183.Worch R, Bokel C, Hofinger S, Schwille P and Weidemann T, Proteomics, 2010, 10, 4196–4208. [DOI] [PubMed] [Google Scholar]
- 184.Lerch-Bader M, Lundin C, Kim H, Nilsson I and von Heijne G, Proc. Natl. Acad. Sci. U.S.A, 2008, 105, 4127–4132. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 185.Lundin C, Kim H, Nilsson I, White SH and von Heijne G, Proc. Natl. Acad. Sci. U.S.A, 2008, 105, 15702–15707. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 186.Elazar A, Weinstein J, Biran I, Fridman Y, Bibi E and Fleishman SJ, Elife, 2016, 5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 187.Wimley WC, Creamer TP and White SH, Biochemistry, 1996, 35, 5109–5124. [DOI] [PubMed] [Google Scholar]
- 188.Krishnakumar SS and London E, J. Mol. Biol, 2007, 374, 1251–1269. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 189.Caputo GA and London E, Biochemistry, 2004, 43, 8794–8806. [DOI] [PubMed] [Google Scholar]
- 190.Lew S, Caputo GA and London E, Biochemistry, 2003, 42, 10833–10842. [DOI] [PubMed] [Google Scholar]
- 191.Andreev OA, Dupuy AD, Segala M, Sandugu S, Serra DA, Chichester CO, Engelman DM and Reshetnyak YK, Proc. Natl. Acad. Sci. U.S.A, 2007, 104, 7893–7898. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 192.Barrera FN, Weerakkody D, Anderson M, Andreev OA, Reshetnyak YK and Engelman DM, J. Mol. Biol, 2011, 413, 359–371. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 193.Nguyen VP, Dixson AC and Barrera FN, Biophys. J, 2019, 117, 659–667. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 194.Nguyen VP, Alves DS, Scott HL, Davis FL and Barrera FN, Biochemistry, 2015, 54, 6567–6575. [DOI] [PubMed] [Google Scholar]
- 195.Andreev OA, Engelman DM and Reshetnyak YK, Front. Physiol, 2014, 5, 97. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 196.Wyatt LC, Lewis JS, Andreev OA, Reshetnyak YK and Engelman DM, Trends Biotechnol, 2018, 36, 1300. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 197.Reshetnyak YK, Moshnikova A, Andreev OA and Engelman DM, Front. Bioeng. Biotechnol, 2020, 8, 335. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 198.Strandberg E, Esteban-Martin S, Ulrich AS and Salgado J, Biochim. Biophys. Acta, 2012, 1818, 1242–1249. [DOI] [PubMed] [Google Scholar]
- 199.Ren J, Lew S, Wang J and London E, Biochemistry, 1999, 38, 5905–5912. [DOI] [PubMed] [Google Scholar]
- 200.Caputo GA and London E, Biochemistry, 2003, 42, 3275–3285. [DOI] [PubMed] [Google Scholar]
- 201.Kandasamy SK and Larson RG, Biophys. J, 2006, 90, 2326–2343. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 202.Morein S, Koeppe IR, Lindblom G, de Kruijff B and Killian JA, Biophys. J, 2000, 78, 2475–2485. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 203.Stefansson A, Armulik A, Nilsson I, von Heijne G and Johansson S, J. Biol. Chem, 2004, 279, 21200–21205. [DOI] [PubMed] [Google Scholar]
- 204.Killian JA, FEBS Lett, 2003, 555, 134–138. [DOI] [PubMed] [Google Scholar]
- 205.Afrose F and Koeppe RE II, Biomolecules, 2020, 10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 206.McKay MJ, Martfeld AN, De Angelis AA, Opella SJ, Greathouse DV and Koeppe RE II, Biophys. J, 2018, 114, 2617–2629. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 207.Holt A, Koehorst RBM, Rutters-Meijneke T, Gelb MH, Rijkers DTS, Hemminga MA and Killian JA, Biophys. J, 2009, 97, 2258–2266. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 208.Soubias O, Teague Walter E. Jr., Hines Kirk G. and Gawrisch K, Biophys. J, 2015, 108, 1125–1132. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 209.Grasberger B, Minton AP, DeLisi C and Metzger H, Proc. Natl. Acad. Sci. U.S.A, 1986, 83, 6258–6262. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 210.Langosch D and Arkin IT, Protein Sci, 2009, 18, 1343–1358. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 211.Hong H, Arch. Biochem. Biophys, 2014, 564, 297–313. [DOI] [PubMed] [Google Scholar]
- 212.Nelson JC, Saven JG, Moore JS and Wolynes PG, Science, 1997, 277, 1793–1796. [DOI] [PubMed] [Google Scholar]
- 213.Anbazhagan V and Schneider D, Biochim. Biophys. Acta, 2010, 1798, 1899–1907. [DOI] [PubMed] [Google Scholar]
- 214.Cristian L, Lear JD and DeGrado WF, Proc. Natl. Acad. Sci. U.S.A, 2003, 100, 14772–14777. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 215.Hong H and Bowie JU, J. Am. Chem. Soc, 2011, 133, 11389–11398. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 216.Senes A, Gerstein M and Engelman DM, J. Mol. Biol, 2000, 296, 921–936. [DOI] [PubMed] [Google Scholar]
- 217.Senes A, Engel DE and DeGrado WF, Curr. Opin. Struct. Biol, 2004, 14, 465–479. [DOI] [PubMed] [Google Scholar]
- 218.Teese MG and Langosch D, Biochemistry, 2015, 54, 5125–5135. [DOI] [PubMed] [Google Scholar]
- 219.Senes A, Ubarretxena-Belandia I and Engelman DM, Proc. Natl. Acad. Sci. U.S.A, 2001, 98, 9056–9061. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 220.Mueller BK, Subramaniam S and Senes A, Proc. Natl. Acad. Sci. U.S.A, 2014, 111, E888–E895. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 221.Anderson SM, Mueller BK, Lange EJ and Senes A, J. Am. Chem. Soc, 2017, 139, 15774–15783. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 222.MacKenzie KR and Fleming KG, Curr. Opin. Struct. Biol, 2008, 18, 412–419. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 223.Cymer F and Schneider D, Cell Adh. Migr, 2010, 4, 299–312. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 224.Moore DT, Berger BW and DeGrado WF, Structure, 2008, 16, 991–1001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 225.Lemmon MA and Engelman DM, Q. Rev. Biophys, 1994, 27, 157–218. [DOI] [PubMed] [Google Scholar]
- 226.Westerfield JM and Barrera FN, J. Biol. Chem, 2020, 295, 1792–1814. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 227.Berger BW, Kulp DW, Span LM, DeGrado JL, Billings PC, Senes A, Bennett JS and DeGrado WF, Proc. Natl. Acad. Sci. U.S.A, 2010, 107, 703–708. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 228.Zhang SQ, Kulp DW, Schramm CA, Mravic M, Samish I and DeGrado WF, Structure, 2015, 23, 527–541. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 229.MacKenzie KR, Prestegard JH and Engelman DM, Science, 1997, 276, 131–133. [DOI] [PubMed] [Google Scholar]
- 230.Kim S, Jeon TJ, Oberai A, Yang D, Schmidt JJ and Bowie JU, Proc. Natl. Acad. Sci. U.S.A, 2005, 102, 14278–14283. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 231.Yang J, Ma YQ, Page RC, Misra S, Plow EF and Qin J, Proc. Natl. Acad. Sci. U.S.A, 2009, 106, 17729–17734. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 232.Zhu J, Luo BH, Barth P, Schonbrun J, Baker D and Springer TA, Mol. Cell, 2009, 34, 234–249. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 233.Metcalf DG, Kulp DW, Bennett JS and DeGrado WF, J. Mol. Biol, 2009, 392, 1087–1101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 234.Litvinov RI, Mravic M, Zhu H, Weisel JW, DeGrado WF and Bennett JS, Proc. Natl. Acad. Sci. U.S.A, 2019, 116, 12295–12300. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 235.Walters RF and DeGrado WF, Proc. Natl. Acad. Sci. U.S.A, 2006, 103, 13658–13663. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 236.Acharya A, Rishi V and Vinson C, Biochemistry, 2006, 45, 11324–11332. [DOI] [PubMed] [Google Scholar]
- 237.Zhang Y, Kulp DW, Lear JD and DeGrado WF, J. Am. Chem. Soc, 2009, 131, 11341–11343. [DOI] [PubMed] [Google Scholar]
- 238.Lear JD, Stouffer AL, Gratkowski H, Nanda V and Degrado WF, Biophys. J, 2004, 87, 3421–3429. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 239.Adamian L and Liang J, Proteins, 2002, 47, 209–218. [DOI] [PubMed] [Google Scholar]
- 240.North B, Cristian L, Fu Stowell X, Lear JD, Saven JG and DeGrado WF, J. Mol. Biol, 2006, 359, 930–939. [DOI] [PubMed] [Google Scholar]
- 241.Kubatzky KF, Ruan W, Gurezka R, Cohen J, Ketteler R, Watowich SS, Neumann D, Langosch D and Klingmuller U, Curr. Biol, 2001, 11, 110–115. [DOI] [PubMed] [Google Scholar]
- 242.Ruan W, Becker V, Klingmuller U and Langosch D, J. Biol. Chem, 2004, 279, 3273–3279. [DOI] [PubMed] [Google Scholar]
- 243.Bowie JU, Nat. Struct. Biol, 2000, 7, 91–94. [DOI] [PubMed] [Google Scholar]
- 244.Choma C, Gratkowski H, Lear JD and DeGrado WF, Nat. Struct. Biol, 2000, 7, 161–166. [DOI] [PubMed] [Google Scholar]
- 245.Gratkowski H, Lear JD and DeGrado WF, Proc. Natl. Acad. Sci. U.S.A, 2001, 98, 880–885. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 246.Pasternak A, Kaplan J, Lear JD and Degrado WF, Protein Sci, 2001, 10, 958–969. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 247.Zhou FX, Cocco MJ, Russ WP, Brunger AT and Engelman DM, Nat. Struct. Biol, 2000, 7, 154–160. [DOI] [PubMed] [Google Scholar]
- 248.Zhou FX, Merianos HJ, Brunger AT and Engelman DM, Proc. Natl. Acad. Sci. U.S.A, 2001, 98, 2250–2255. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 249.Leroy E, Defour JP, Sato T, Dass S, Gryshkova V, Shwe MM, Staerk J, Constantinescu SN and Smith SO, J. Biol. Chem, 2016, 291, 2974–2987. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 250.Partridge AW, Therien AG and Deber CM, Proteins, 2004, 54, 648–656. [DOI] [PubMed] [Google Scholar]
- 251.Ng DP, Poulsen BE and Deber CM, Biochim. Biophys. Acta, 2012, 1818, 1115–1122. [DOI] [PubMed] [Google Scholar]
- 252.Hermansson M and von Heijne G, J. Mol. Biol, 2003, 334, 803–809. [DOI] [PubMed] [Google Scholar]
- 253.Choe S, Hecht KA and Grabe M, J. Gen. Physiol, 2008, 131, 563–573. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 254.Baker EN and Hubbard RE, Prog. Biophys. Molec. Biol, 1984, 44, 97–179. [DOI] [PubMed] [Google Scholar]
- 255.Mravic M, Thomaston JL, Tucker M, Solomon PE, Liu L and DeGrado WF, Science, 2019, 363, 1418–1423. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 256.Joh NH, Min A, Faham S, Whitelegge JP, Yang D, Woods VL and Bowie JU, Nature, 2008, 453, 1266–1270. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 257.Joh NH, Oberai A, Yang D, Whitelegge JP and Bowie JU, J. Am. Chem. Soc, 2009, 131, 10846–10847. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 258.Tatko CD, Nanda V, Lear JD and Degrado WF, J. Am. Chem. Soc, 2006, 128, 4170–4171. [DOI] [PubMed] [Google Scholar]
- 259.Nordholm J, da Silva DV, Damjanovic J, Dou D and Daniels R, J. Biol. Chem, 2013, 288, 10652–10660. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 260.Lear JD, Gratkowski H, Adamian L, Liang J and DeGrado WF, Biochemistry, 2003, 42, 6400–6407. [DOI] [PubMed] [Google Scholar]
- 261.Li R, Mitra N, Gratkowski H, Vilaire G, Litvinov R, Nagasami C, Weisel JW, Lear JD, DeGrado WF and Bennett JS, Science, 2003, 300, 795–798. [DOI] [PubMed] [Google Scholar]
- 262.Li W, Metcalf DG, Gorelik R, Li R, Mitra N, Nanda V, Law PB, Lear JD, Degrado WF and Bennett JS, Proc. Natl. Acad. Sci. U.S.A, 2005, 102, 1424–1429. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 263.Call ME, Pyrdol J, Wiedmann M and Wucherpfennig KW, Cell, 2002, 111, 967–979. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 264.Call ME and Wucherpfennig KW, Nat. Rev. Immunol, 2007, 7, 841–850. [DOI] [PubMed] [Google Scholar]
- 265.Lanier LL, Immunol. Rev, 2009, 227, 150–160. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 266.Feng J, Garrity D, Call ME, Moffett H and Wucherpfennig KW, Immunity, 2005, 22, 427–438. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 267.Call ME, Wucherpfennig KW and Chou JJ, Nat. Immunol, 2010, 11, 1023–1029. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 268.Blazquez-Moreno A, Park S, Im W, Call MJ, Call ME and Reyburn HT, Proc. Natl. Acad. Sci. U.S.A, 2017, 114, E5645–E5654. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 269.Adams MW, Jenney FE Jr., Clay MD and Johnson MK, J. Biol. Inorg. Chem, 2002, 7, 647–652. [DOI] [PubMed] [Google Scholar]
- 270.Call ME, Schnell JR, Xu C, Lutz RA, Chou JJ and Wucherpfennig KW, Cell, 2006, 127, 355–368. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 271.Fong LK, Chalkley MJ, Tan SK, Grabe M and DeGrado WF, Proc. Natl. Acad. Sci. U.S.A, 2021, 118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 272.Eilers M, Shekar SC, Shieh T, Smith SO and Fleming PJ, Proc. Natl. Acad. Sci. U.S.A, 2000, 97, 5796–5801. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 273.Adamian L and Liang J, J. Mol. Biol, 2001, 311, 891–907. [DOI] [PubMed] [Google Scholar]
- 274.Langosch D and Heringa J, Proteins, 1998, 31, 150–159. [DOI] [PubMed] [Google Scholar]
- 275.Cymer F, von Heijne G and White SH, J. Mol. Biol, 2015, 427, 999–1022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 276.Johnson RM, Hecht K and Deber CM, Biochemistry, 2007, 46, 9208–9214. [DOI] [PubMed] [Google Scholar]
- 277.Gurezka R and Langosch D, J. Biol. Chem, 2001, 276, 45580–45587. [DOI] [PubMed] [Google Scholar]
- 278.Mravic M, Hu H, Lu Z, Bennett JS, Sanders CR, Orr AW and DeGrado WF, Protein Eng. Des. Sel, 2018, 31, 181–190. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 279.Xie J and DiMaio D, FEBS J, 2021, DOI: 10.1111/febs.15775. [DOI] [Google Scholar]
- 280.Talbert-Slagle K and DiMaio D, Virology, 2009, 384, 345–351. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 281.DiMaio D and Petti LM, Virology, 2013, 445, 99–114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 282.Freeman-Cook LL, Dixon AM, Frank JB, Xia Y, Ely L, Gerstein M, Engelman DM and DiMaio D, J. Mol. Biol, 2004, 338, 907–920. [DOI] [PubMed] [Google Scholar]
- 283.Marlatt SA, Kong Y, Cammett TJ, Korbel G, Noonan JP and Dimaio D, Protein Eng. Des. Sel, 2011, 24, 311–320. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 284.Talbert-Slagle K, Marlatt S, Barrera FN, Khurana E, Oates J, Gerstein M, Engelman DM, Dixon AM and Dimaio D, J. Virol, 2009, 83, 9773–9785. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 285.Ptacek JB, Edwards AP, Freeman-Cook LL and DiMaio D, Proc. Natl. Acad. Sci. U.S.A, 2007, 104, 11945–11950. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 286.Cammett TJ, Jun SJ, Cohen EB, Barrera FN, Engelman DM and Dimaio D, Proc. Natl. Acad. Sci. U.S.A, 2010, 107, 3447–3452. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 287.Petti LM, Talbert-Slagle K, Hochstrasser ML and DiMaio D, J. Biol. Chem, 2013, 288, 27273–27286. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 288.Heim EN, Marston JL, Federman RS, Edwards AP, Karabadzhak AG, Petti LM, Engelman DM and DiMaio D, Proc. Natl. Acad. Sci. U.S.A, 2015, 112, E4717–E4725. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 289.He L, Cohen EB, Edwards APB, Xavier-Ferrucio J, Bugge K, Federman RS, Absher D, Myers RM, Kragelund BB, Krause DS and DiMaio D, iScience, 2019, 17, 167–181. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 290.Guo R, Gaffney K, Yang Z, Kim M, Sungsuwan S, Huang X, Hubbell WL and Hong H, Nat. Chem. Biol, 2016, 12, 353–360. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 291.Sulistijo ES and Mackenzie KR, Biochemistry, 2009, 48, 5106–5120. [DOI] [PubMed] [Google Scholar]
- 292.Lawrie CM, Sulistijo ES and MacKenzie KR, J. Mol. Biol, 2010, 396, 924–936. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 293.Deber CM, Liu LP and Wang C, J. Pept. Res, 1999, 54, 200–205. [DOI] [PubMed] [Google Scholar]
- 294.Acharya R, Carnevale V, Fiorin G, Levine BG, Polishchuk AL, Balannik V, Samish I, Lamb RA, Pinto LH, DeGrado WF and Klein ML, Proc. Natl. Acad. Sci. U.S.A, 2010, 107, 15075–15080. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 295.Howard KP, Lear JD and DeGrado WF, Proc. Natl. Acad. Sci. U.S.A, 2002, 99, 8568–8572. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 296.Mould JA, Li H-C, Dudlak CS, Lear JD, Pekosz A, Lamb RA and Pinto LH, J. Biol. Chem, 2000, 275, 8592–8599. [DOI] [PubMed] [Google Scholar]
- 297.Gernert KM, Surles MC, Labean TH, Richardson JS and Richardson DC, Protein Sci, 1995, 4, 2252–2260. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 298.North B, Summa CM, Ghirlanda G and DeGrado WF, J. Mol. Biol, 2001, 311, 1081–1090. [DOI] [PubMed] [Google Scholar]
- 299.Joh NH, Wang T, Bhate MP, Acharya R, Wu Y, Grabe M, Hong M, Grigoryan G and DeGrado WF, Science, 2014, 346, 1520–1524. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 300.Cafiso DS, Annu. Rev. Biophys. Biomol. Struct, 1994, 23, 141–165. [DOI] [PubMed] [Google Scholar]
- 301.Kim MJ, Park SH, Opella SJ, Marsilje TH, Michellys PY, Seidel HM and Tian SS, J. Biol. Chem, 2007, 282, 14253–14261. [DOI] [PubMed] [Google Scholar]
- 302.Zeng X, Wu P, Yao C, Liang J, Zhang S and Yin H, Biochemistry, 2017, 56, 2076–2085. [DOI] [PubMed] [Google Scholar]
- 303.Furthmayr H and Marchesi VT, Biochemistry, 1976, 15, 1137–1144. [DOI] [PubMed] [Google Scholar]
- 304.Bormann B-J, Knowles WJ and Marchesi VT, J. Biol. Chem, 1989, 264, 4033–4037. [PubMed] [Google Scholar]
- 305.Westerfield JM, Sahoo AR, Alves DS, Grau B, Cameron A, Maxwell M, Schuster JA, Souza PCT, Mingarro I, Buck M and Barrera FN, J. Mol. Biol, 2021, 433, 167144. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 306.Stefanski KM, Russell CM, Westerfield JM, Lamichhane R and Barrera FN, J. Biol. Chem, 2021, 296, 100149. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 307.Alves DS, Westerfield JM, Shi X, Nguyen VP, Stefanski KM, Booth KR, Kim S, Morrell-Falvey J, Wang BC, Abel SM, Smith AW and Barrera FN, Elife, 2018, 7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 308.Yin H, Slusky JS, Berger BW, Walters RS, Vilaire G, Litvinov RI, Lear JD, Caputo GA, Bennett JS and DeGrado WF, Science, 2007, 315, 1817–1822. [DOI] [PubMed] [Google Scholar]
- 309.Caputo GA, Litvinov RI, Li W, Bennett JS, DeGrado WF and Yin H, Biochemistry, 2008, 47, 8600–8606. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 310.Shandler SJ, Korendovych IV, Moore DT, Smith-Dupont KB, Streu CN, Litvinov RI, Billings PC, Gai F, Bennett JS and DeGrado WF, J. Am. Chem. Soc, 2011, 133, 12378–12381. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 311.Fong KP, Zhu H, Span LM, Moore DT, Yoon K, Tamura R, Yin H, DeGrado WF and Bennett JS, J. Biol. Chem, 2016, 291, 11706–11716. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 312.Hille B, Ionic Channels of Excitable Membranes, Sinauer Associates, Sunderland, MA, 3rd edn., 2001. [Google Scholar]
- 313.DeGrado WF, Wasserman ZR and Lear JD, Science, 1989, 243, 622–628. [DOI] [PubMed] [Google Scholar]
- 314.Åkerfeldt KS, Lear JD, Waserman ZR, Chung LA and DeGrado WF, Acc. Chem. Res, 1993, 26, 191–197. [Google Scholar]
- 315.Nguyen TK and Ueno T, Curr. Opin. Struct. Biol, 2018, 51, 1–8. [DOI] [PubMed] [Google Scholar]
- 316.Lear JD, Wasserman ZR and DeGrado WF, Science, 1988, 240, 1177–1181. [DOI] [PubMed] [Google Scholar]
- 317.Xu C, Lu P, Gamal El-Din TM, Pei XY, Johnson MC, Uyeda A, Bick MJ, Xu Q, Jiang D, Bai H, Reggiano G, Hsia Y, Brunette TJ, Dou J, Ma D, Lynch EM, Boyken SE, Huang PS, Stewart L, DiMaio F, Kollman JM, Luisi BF, Matsuura T, Catterall WA and Baker D, Nature, 2020, 585, 129–134. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 318.Joh NH, Grigoryan G, Wu Y and DeGrado WF, Philos. Trans. R. Soc. Lond. B Biol. Sci, 2017, 372. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 319.Veatch W and Stryer L, J. Mol. Biol, 1977, 113, 89–102. [DOI] [PubMed] [Google Scholar]
- 320.Urry DW, Goodall MC, Glickson JD and Mayers DF, Proc. Natl. Acad. Sci. U.S.A, 1971, 68, 1907–1911. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 321.Hladky SB and Haydon DA, Biochim. Biophys. Acta, 1972, 274, 294. [DOI] [PubMed] [Google Scholar]
- 322.Ketchem RR, Hu W and Cross TA, Science, 1993, 261, 1457–1460. [DOI] [PubMed] [Google Scholar]
- 323.Berl V, Huc I, Khoury RG, Krische MJ and Lehn JM, Nature, 2000, 407, 720–723. [DOI] [PubMed] [Google Scholar]
- 324.Roux B and Karplus M, Annu. Rev. Biophys. Biomol. Struct, 1994, 23, 731–761. [DOI] [PubMed] [Google Scholar]
- 325.Roux B and Karplus M, J. Phys. Chem, 1991, 95, 4856–4868. [Google Scholar]
- 326.Roux B and Karplus M, Biophys. J, 1988, 53, 297–309. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 327.Goodall MC and Urry DW, Biochim. Biophys. Acta, 1973, 291, 317–320. [DOI] [PubMed] [Google Scholar]
- 328.Kennedy SJ, Roeske RW, Freeman AR, Watanabe AM and Besche HR Jr., Science, 1977, 196, 1341–1342. [DOI] [PubMed] [Google Scholar]
- 329.Ghadiri MR, Granja JR, Milligan RA, McRee DE and Khazanovich N, Nature, 1993, 366, 324–327. [DOI] [PubMed] [Google Scholar]
- 330.Vogel H, Biochemistry, 1987, 26, 4562–4572. [DOI] [PubMed] [Google Scholar]
- 331.Lee MT, Chen FY and Huang HW, Biochemistry, 2004, 43, 3590–3599. [DOI] [PubMed] [Google Scholar]
- 332.Biggin PC and Sansom MS, Biophys. Chem, 1999, 76, 161–183. [DOI] [PubMed] [Google Scholar]
- 333.Boheim G, J. Membr. Biol, 1974, 19, 277–303. [DOI] [PubMed] [Google Scholar]
- 334.Nagaraj R and Balaram P, Acc. Chem. Res, 1981, 14, 356–362. [Google Scholar]
- 335.Fox RO and Richards FM, Nature, 1982, 300, 325–330. [DOI] [PubMed] [Google Scholar]
- 336.Hall JE, Vodyanoy I, Balasubramanian TM and Marshall GR, Biophys. J, 1984, 45, 233–247. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 337.Woolley GA and Wallace BA, J. Membr. Biol, 1992, 129, 109–136. [DOI] [PubMed] [Google Scholar]
- 338.Cantor RS, Biophys. J, 2002, 82, 2520–2525. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 339.Tieleman DP, Breed J, Berendsen HJ and Sansom MS, Faraday Discuss, 1998, DOI: 10.1039/a806266h, 209–223; discussion 225–246. [DOI] [PubMed] [Google Scholar]
- 340.Pieta P, Mirza J and Lipkowski J, Proc. Natl. Acad. Sci. U.S.A, 2012, 109, 21223–21237. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 341.DiMaio D, Annu. Rev. Microbiol, 2014, 68, 21–43. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 342.Scott C and Griffin S, J. Gen. Virol, 2015, 96, 2000–2027. [DOI] [PubMed] [Google Scholar]
- 343.To J, Surya W and Torres J, Adv. Protein Chem. Struct. Biol, 2016, 104, 307–355. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 344.Schoeman D and Fielding BC, Virol. J, 2019, 16, 69. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 345.Farag NS, Breitinger U, Breitinger HG and El Azizi MA, Int. J. Biochem. Cell Biol, 2020, 122, 105738. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 346.Schoeman D and Fielding BC, Front. Microbiol, 2020, 11, 2086. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 347.Hay AJ, Wolstenholme AJ, Skehel JJ and Smith MH, EMBO J, 1985, 4, 3021–3024. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 348.Hay AJ, Zambon MC, Wolstenholme AJ, Skehel JJ and Smith MH, J. Antimicrob. Chemother, 1986, 18 Suppl B, 19–29. [DOI] [PubMed] [Google Scholar]
- 349.Ciampor F, Thompson CA, Grambas S and Hay AJ, Virus Res, 1992, 22, 247–258. [DOI] [PubMed] [Google Scholar]
- 350.Mandala VS, McKay MJ, Shcherbakov AA, Dregni AJ, Kolocouris A and Hong M, Nat. Struct. Mol. Biol, 2020, 27, 1202–1208. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 351.Ma C, Polishchuk AL, Ohigashi Y, Stouffer AL, Schon A, Magavern E, Jing X, Lear JD, Freire E, Lamb RA, DeGrado WF and Pinto LH, Proc. Natl. Acad. Sci. U.S.A, 2009, 106, 12283–12288. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 352.Thomaston JL, Alfonso-Prieto M, Woldeyes RA, Fraser JS, Klein ML, Fiorin G and DeGrado WF, Proc. Natl. Acad. Sci. U.S.A, 2015, 112, 14260–14265. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 353.Thomaston JL, Woldeyes RA, Nakane T, Yamashita A, Tanaka T, Koiwai K, Brewster AS, Barad BA, Chen Y, Lemmin T, Uervirojnangkoorn M, Arima T, Kobayashi J, Masuda T, Suzuki M, Sugahara M, Sauter NK, Tanaka R, Nureki O, Tono K, Joti Y, Nango E, Iwata S, Yumoto F, Fraser JS and DeGrado WF, Proc. Natl. Acad. Sci. U.S.A, 2017, 114, 13357–13362. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 354.Schnell JR and Chou JJ, Nature, 2008, 451, 591–595. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 355.Cady SD, Schmidt-Rohr K, Wang J, Soto CS, DeGrado WF and Hong M, Nature, 2010, 463, 689–692. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 356.Sharma M, Yi M, Dong H, Qin H, Peterson E, Busath DD, Zhou HX and Cross TA, Science, 2010, 330, 509–512. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 357.Liang R, Swanson JM, Madsen JJ, Hong M, DeGrado WF and Voth GA, Proc. Natl. Acad. Sci. U.S.A, 2016, DOI: 10.1073/pnas.1615471113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 358.Williams JK, Shcherbakov AA, Wang J and Hong M, J. Biol. Chem, 2017, 292, 17876–17884. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 359.Fu R, Miao Y, Qin H and Cross TA, J. Am. Chem. Soc, 2020, 142, 2115–2119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 360.Fu R, Miao Y, Qin H and Cross TA, J. Am. Chem. Soc, 2016, 138, 15801–15804. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 361.Stouffer AL, Acharya R, Salom D, Levine AS, Di Costanzo L, Soto CS, Tereshko V, Nanda V, Stayrook S and DeGrado WF, Nature, 2008, 451, 596–599. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 362.Thomaston JL, Samways ML, Konstantinidi A, Ma C, Hu Y, Bruce Macdonald HE, Wang J, Essex JW, DeGrado WF and Kolocouris A, Biochemistry, 2021, 60, 2471–2482. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 363.Thomaston JL, Konstantinidi A, Liu L, Lambrinidis G, Tan J, Caffrey M, Wang J, Degrado WF and Kolocouris A, Biochemistry, 2020, 59, 627–634. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 364.Thomaston JL, Wu Y, Polizzi N, Liu L, Wang J and DeGrado WF, J. Am. Chem. Soc, 2019, 141, 11481–11488. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 365.Thomaston JL and DeGrado WF, Protein Sci, 2016, 25, 1551–1554. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 366.Hong M and DeGrado WF, Protein Sci, 2012, 21, 1620–1633. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 367.Watkins LC, DeGrado WF and Voth GA, J. Am. Chem. Soc, 2020, 142, 17425–17433. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 368.Watkins LC, Liang R, Swanson JMJ, DeGrado WF and Voth GA, J. Am. Chem. Soc, 2019, 141, 11667–11676. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 369.Liang R, Li H, Swanson JM and Voth GA, Proc. Natl. Acad. Sci. U.S.A, 2014, 111, 9396–9401. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 370.Ma C, Fiorin G, Carnevale V, Wang J, Lamb RA, Klein ML, Wu Y, Pinto LH and DeGrado WF, Structure, 2013, 21, 2033–2041. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 371.Ivanovic T, Rozendaal R, Floyd DL, Popovic M, van Oijen AM and Harrison SC, PLoS One, 2012, 7, e31566. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 372.Hu F, Schmidt-Rohr K and Hong M, J. Am. Chem. Soc, 2012, 134, 3703–3713. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 373.Khurana E, Dal Peraro M, DeVane R, Vemparala S, DeGrado WF and Klein ML, Proc. Natl. Acad. Sci. U.S.A, 2009, 106, 1069–1074. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 374.Busath DD, in Advances in Planar Lipid Bilayers and Liposomes, eds. Leitmannova Liu A and Iglič A, Academic Press, Burlington, 2009, vol. 10, pp. 161–201. [Google Scholar]
- 375.Stouffer AL, Ma C, Cristian L, Ohigashi Y, Lamb RA, Lear JD, Pinto LH and DeGrado WF, Structure, 2008, 16, 1067–1076. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 376.Stouffer AL, Nanda V, Lear JD and DeGrado WF, J. Mol. Biol, 2005, 347, 169–179. [DOI] [PubMed] [Google Scholar]
- 377.Thomaston JL, Polizzi NF, Konstantinidi A, Wang J, Kolocouris A and DeGrado WF, J. Am. Chem. Soc, 2018, 140, 15219–15226. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 378.Wang J, Wu Y, Ma C, Fiorin G, Wang J, Pinto LH, Lamb RA, Klein ML and DeGrado WF, Proc. Natl. Acad. Sci. U.S.A, 2013, 110, 1315–1320. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 379.Rossman JS, Jing X, Leser GP and Lamb RA, Cell, 2010, 142, 902–913. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 380.Pielak RM, Oxenoid K and Chou JJ, Structure, 2011, 19, 1655–1663. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 381.Åkerfeldt K, Kim RM, Camac D, Groves JT, Lear JD and DeGrado WF, J. Am. Chem. Soc, 1992, 114, 9656–9657. [Google Scholar]
- 382.Dieckmann GR, Lear JD, Zhong Q, Klein ML, DeGrado WF and Sharp KA, Biophys. J, 1999, 76, 618–630. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 383.Nguyen TH, Liu Z and Moore PB, Biophys. J, 2013, 105, 1569–1580. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 384.Zhong Q, Moore PB, Newns DM and Klein ML, FEBS Lett, 1998, 427, 267–270. [DOI] [PubMed] [Google Scholar]
- 385.Scott AJ, Niitsu A, Kratochvil HT, Lang EJM, Sengel JT, Dawson WM, Mahendran KR, Mravic M, Thomson AR, Brady RL, Liu L, Mulholland AJ, Bayley H, DeGrado WF, Wallace MI and Woolfson DN, Nat. Chem, 2021, 13, 643–650. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 386.Mahendran KR, Niitsu A, Kong L, Thomson AR, Sessions RB, Woolfson DN and Bayley H, Nat. Chem, 2017, 9, 411–419. [DOI] [PubMed] [Google Scholar]
- 387.Beesley JL and Woolfson DN, Curr. Opin. Biotechnol, 2019, 58, 175–182. [DOI] [PubMed] [Google Scholar]
- 388.Kuroda K, Caputo GA and DeGrado WF, Chemistry, 2009, 15, 1123–1133. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 389.Mowery BP, Lee SE, Kissounko DA, Epand RF, Epand RM, Weisblum B, Stahl SS and Gellman SH, J. Am. Chem. Soc, 2007, 129, 15474–15476. [DOI] [PubMed] [Google Scholar]
- 390.Mowery BP, Lindner AH, Weisblum B, Stahl SS and Gellman SH, J. Am. Chem. Soc, 2009, 131, 9735–9745. [DOI] [PubMed] [Google Scholar]
- 391.Munoz-Bonilla A and Fernandez-Garcia M, Prog. Polym. Sci, 2012, 37, 281–339. [Google Scholar]
- 392.Campoccia D, Montanaro L and Arciola CR, Biomaterials, 2013, 34, 8533–8554. [DOI] [PubMed] [Google Scholar]
- 393.Fraternali F, Marzuoli I and da Cruz CHB, Faraday Discuss, 2021, DOI: 10.1039/d1fd00041a. [DOI] [PubMed] [Google Scholar]
- 394.Marzuoli I, Cruz CHB, Lorenz CD and Fraternali F, Nanoscale, 2021, 13, 10342–10355. [DOI] [PubMed] [Google Scholar]
- 395.Castelletto V, de Santis E, Alkassem H, Lamarre B, Noble JE, Ray S, Bella A, Burns JR, Hoogenboom BW and Ryadnov MG, Chem. Sci, 2016, 7, 1707–1711. [DOI] [PMC free article] [PubMed] [Google Scholar]

