Abstract
Antimicrobial peptides (AMPs) collectively constitute a key component of the host innate immune system. They span a diverse space of sequences and can be α-helical, β-sheet, or unfolded in structure. Despite a wealth of knowledge about them from decades of experiments, it remains difficult to articulate general principles governing such peptides. How are they different from other molecules that are also cationic and amphiphilic? What other functions, in immunity and otherwise, are enabled by these simple sequences? In this short review, we present some recent work that engages these questions using methods not usually applied to AMP studies, such as machine learning. We find that not only do AMP-like sequences confer membrane remodeling activity to an unexpectedly broad range of protein classes, their cationic and amphiphilic signature also allows them to act as meta-antigens and self-assemble with immune ligands into nanocrystalline complexes for multivalent presentation to Toll-like receptors.
Keywords: Antimicrobial peptides, Membrane curvature, Machine learning, Toll-like receptors, Innate immunity
Introduction
Antimicrobial peptides (AMPs, or host defense peptides or innate immune peptides) have been systematically investigated for four decades. AMPs are generally characterized by their short amino acid sequences (<50 amino acids), net cationic charge ( 2 to 9), and amphiphilicity [1–6]. These seemingly simple molecules also span an enormous diversity of sequences and secondary structures and can be α-helical [7], β-sheet [8], or extended linear peptides [9,10]. Well over 2000 natural and artificial AMPs have been characterized and information about them are organized into well-maintained databases [11–13]. However, given this knowledge, how do we think about them as a class of molecules? They are known to be cationic amphiphilic molecules but how are they different from other amphiphilic cationic molecules? They are known to be important components of innate host defense [1–6] since they preferentially permeate bacterial membranes over eukaryotic membranes. However, are there other functions in innate immunity besides membrane activity that are enabled by their cationic charge and amphiphilicity? What do the answers of these questions imply about undiscovered functions of AMPs and how we ultimately define AMPs? We and others have done some recent work that suggests such questions may be answerable and offer a biased perspective on some emerging directions of inquiry by incorporating machine learning methods. By way of passing, we provide an early critical assessment of machine learning in the context of AMPs by giving examples of what machine learning discovered unexpectedly and what it failed to anticipate, at least with the currently available data and data structures.
Amphiphilicity and membrane activity
It is thought that AMPs generally function by preferentially permeating microbial membranes, which leads to cell death due to the loss of electrochemical gradients, vulnerability to osmotic stress, loss of cellular contents, and disruption of metabolic processes [2,6]. This type of antimicrobial activity has be ascribed to interactions between amphiphilic AMPs and membranes [14]. Such interactions have been studied using experimental techniques such as X-ray scattering, nuclear magnetic resonance (NMR), fluorescence microscopy, and electron microscopy [6], and historically different models of membrane permeation have been proposed [15–18]. One of the canonical features of AMPs is their selectivity for bacterial membranes over eukaryotic membranes, which is a result of their differential activity on membranes of different compositions [1,2,6]. Early work has identified one necessary condition: bacteria membranes contain large amounts of exposed anionic lipids (e.g. phosphatidylglycerol) on the outer leaflet, while eukaryotic membranes contain mostly zwitterionic lipids (e.g. phosphatidylcholine and sphingomyelin) [19–21]. However, bacterial membranes also contain large amounts of negative intrinsic curvature lipids, such as phosphatidylethanolamine, which makes such membranes vulnerable to permeation [22–27]. The majority of early experimental studies agree that membrane permeation underlies the primary mechanism of action of AMPs, and that such action is a natural consequence of AMP amphiphilicity.
Machine learning of AMP sequences
In this age of “big data”, it is not surprising that molecular data science has been applied to AMPs. Machine learning studies of AMPs are typically based on the development of quantitative structure-active relationship (QSAR) models, which seek to use physicochemical descriptors to predict levels of biological activity. Such studies have employed a broad range of machine-learning approaches including artificial neural networks (ANN), support vector machines (SVM), quantitative matrices (QM), hidden Markov models (HMM), random forests (RF), knearest neighbors (k-NN), and self-organized maps (SOM) [28–38]. The preponderance of this work has employed data-driven learning with the primary goal of the discovery and design AMPs with enhanced microbial potency. In 2016, we employed machine learning with an additional goal: We built an SVM classifier trained on ahelical AMP sequences with the dual intents of achieving high predictive accuracy and teaching us something about what physicochemical and sequence-level properties make a peptide antimicrobial [39–41]. To achieve this aim, we combined a linear SVM as a relatively “white box” machine-learning technique together with principled feature selection to establish a 12-descriptor QSAR model that identified the key properties governing AMP activity and exposed the learned classification procedure in a human-comprehensible manner. Despite the simplicity of our approach, we nevertheless achieved high classification accuracy and thus established an accurate and interpretable model of α-helical AMP activity. The trained model also allowed us to probe how increasing degrees of “antimicrobial-ness” in a peptide e as defined by the machine learning classifier e ultimately correlate to that peptide’s increasing ability to mediate specific physical processes, and how those processes might in turn correlate to antimicrobial activity.
We constructed and trained an SVM classifier on a dataset of 286 α-helical antimicrobial peptides and 286 “decoy” α-helical non-antimicrobial peptides. We began with a panel of 1588 physicochemical descriptors [42–46], which included simple peptide metrics of length, charge, hydrophobicity, as well as more complicated metrics based on autocorrelation and sequence order. From these descriptors, we identified the 12 most predictive descriptors using the L1 feature selection approach of Bi et al. [47]. The linear SVM constructed in this way possessed a 91.9% classification accuracy on an 86-peptide hold-out test set, demonstrating as good or better performance than more complex models employing the entire descriptor ensemble and/or nonlinear SVMs. Each peptide sequence occupies a point in the 12-dimensional space spanned by these descriptors, and the linear SVM classifier draws the optimal 11-dimensional hyperplane dividing the antimicrobial sequences from the non-antimicrobial. From an arbitrary peptide sequence as input, the SVM outputs a single parameter s, the distance between that peptide sequence to separating hyperplane. Intuitively, a larger positive value of σ denotes higher confidence of antimicrobial activity, and a large negative value denotes a higher confidence of lack of antimicrobial activity.
Although the “antimicrobial-ness” of the peptide sequence is described by the metric s, calibrating experiments exposed the initially puzzling finding that σ did not correlate with traditional metrics of antimicrobial activity such as minimum inhibitory concentrations (MIC) [39]. We reasoned that this lack of correlation was due to the multiple modalities by which AMPs in the training data mediated their antimicrobial activity, including disruption of DNA synthesis, inhibition of protein production, and modulation of immune and defense responses. In other words, the trained classifier learned not to distinguish these diverse modes of antimicrobial-ness directly, but rather identified a unifying characteristic underpinning these various mechanisms. This led us to hypothesize that membrane activity may have been learned as the discriminant between AMPs and decoys as a prerequisite to antimicrobial activity by numerous pathways. We tested this hypothesis using synthesized peptides with diverse σ scores assigned by the classifier, and characterized their activity using antimicrobial assays and small-angle X-ray scattering (SAXS) with artificial membranes. The SVM σ score was found to correlate strongly with a peptide’s ability to generate negative Gaussian membrane curvature (NGC) [39], the type of curvature topologically required for membrane restructuring processes, such as pore formation, blebbing, budding, vesicularization. Indeed, AMPs were found to induce NGC preferentially in membranes with bacterial lipid compositions [48,49]. The ability to generate NGC has also been observed for both natural and synthetic AMPs [50,51], cytokines [52], cell-penetrating peptides [53], and viral fusion proteins [54,55]. Taken together, these results suggest that the generation of NGC is not only a feature of AMPs, but a far more general and common root mechanism for membrane restructuring processes in general.
Using the SVM classifier as a screening tool, we then proceeded to explore peptide sequence space with the intention of finding new AMP-like sequences in two specific regions of unknown sequence space: peptides close in homology to known AMP sequences, and those far in homology e and therefore far in evolutionary distance e to known AMPs. Using a Monte Carlo approach, we screened peptides with lengths of 20e25 amino acids, the most common lengths of AMPs, and focused our search on AMP-like candidates with large positive values of s. To identify optimal candidates, we applied multi-objective optimization to identify sequences lying on a “Pareto frontier” that compromises between s, homology criteria, and peptide helicity. This analysis led us to discover interesting surprises in peptides very dissimilar to known AMPs. One might intuitively expect that peptide sequences far in homology from known AMPs, membrane permeating peptides par excellence, should have lower or negative values of s. Surprisingly, not just new peptides but a number of peptide families and domains within proteins were predicted by our classifier to be membrane active at the same level as AMPs [39]. That these peptides with AMP-like membrane activity have little sequence homology with known AMPs highlights an interesting and important difference between our methods and traditional bioinformatics methods based on sequence comparison such as BLAST. The sequences with AMP-like activity include neuropeptides, amyloids, and viral fusion proteins. In other words, the SVM classifier enabled us to efficiently identify and discover sequences in both peptides and diverse proteins families that generate NGC to remodel membranes for permeation or transduction. We describe a few of these examples below.
Neuropeptides
Neuropeptides are used by neurons to communicate [56] and mediate functions ranging from endocrine signaling and homeostatic regulation to immune signaling, pain modulation, and circadian rhythm maintenance. At present, over 100 neuropeptides are known to be released from neurons in mammals and are not only present in the central nervous system (CNS) [57], but also are found in the enteric nervous system, peripheral nervous system, and within immune organs. Neuropeptides are known to primarily exert their biological function by binding to their cognate receptor (usually a G-coupled protein receptor), triggering a signal transduction pathway that leads to a functional change in the target cell [56]. Neuropeptides are typically thought of as neurotransmitters or endocrine/ paracrine hormones, but recent work has illuminated their potential role as integral components of the innate immune system. In fact, neuropeptides are central to modulation of neuroinflammation [58] and triggering of the innate [59,60] and adaptive [61] immune responses, and facilitating communication between the immune system and nervous system [62].
A significant number of positive hits predicted to have membrane remodeling activity were found to be members of the neuropeptide family. These included vasoactive intestinal peptide (VIP), neuropeptide Y (NPY), α-melanocyte stimulating hormone (α-MSH), and pituitary adenylate cyclase activating peptide (PACAP). Another peptide, substance P, not included in the initial search, was also found to have predicted antimicrobial activity. Interestingly, previous work has preliminarily shown that substance P, VIP, and NPY possess moderate to substantial antimicrobial activity against a range of organisms [63–65]. Moreover, consistent with our predictions, the neuropeptide PACAP was very recently found to possess potent in vitro antimicrobial activity [66]. A small number of neuropeptides possess similar amino acid content and secondary structure to AMPs cathelicidins and defensins [67], but our results suggest that antimicrobial activity and membrane activity may be far more common than previously thought, even in sequences without obvious sequence homology. These findings lead to many questions that may attract and reward attention: Why do neuropeptides possess direct antimicrobial activity? Are neuropeptides induced at sites of infection? Is this observed activity relevant to how the CNS defends itself? Additional studies are needed to show biological relevance.
Viral membrane fusion and fission machinery
Enveloped viruses, such as human immunodeficiency virus (HIV) and influenza, enter host cells to initiate infection via virusecell fusion. This process involves viral membrane fusion proteins, which mediate the fusion between the viral lipid envelope and the host cell plasma membrane. These proteins typically feature an N-terminal fusion peptide (FP) and a C-terminal transmembrane domain (TMD) [68]. While previous studies have demonstrated that the FPs of HIV, influenza, and paramyxovirus adopt a partially inserted topology in the membrane to induce nonlamellar structures [69–73] the structure and role of the TMD in viral membrane fusion are less not well understood. In fact, the TMD has been traditionally regarded as a passive anchor to the virus envelope. Recent work using solid-state NMR and SAXS on the parainfluenza virus 5 (PIV5) fusion protein has found that the TMD changes its conformation in response to the membrane lipid composition to play an active role in viralecell fusion by mediating the necessary membrane topological changes [55]. More specifically, in membranes that have high concentrations of phosphatidylethanolamine, the TMD adopts a β-strand-rich conformation and generates NGC, the type of curvature required for the formation of hemifusion intermediates and fusion pores. Aside from the key role of proteins, this finding additionally underscores the influence of constituent lipids of membranes in viralecell fusion. Lipid-dependent conformational changes, which have been found to occur for both the FP [73,74] and TMD [55], suggest that the intact fusion protein can adapt to the different lipid environments of the virus envelope and host cell membrane [75].
To spread infection, many enveloped viruses complete their replication cycle by budding progeny virions from infected host cells. As viruses engage membrane fusion to enter a cell, viruses exiting a cell undergo the reciprocal process in which viral particles assemble, bud, and pinch off from the host membrane. This budding process involves membrane deformation and fission, which require membrane curvature generation. A large number of enveloped viruses recruit the host endosomal sorting complex required for transport (ESCRT) machinery to assist in budding and virion release [76–78], however, influenza has been implicated to utilize an ESCRTindependent mechanism [77,79]. M2 from the influenza A virus is a multifunctional protein that assembles into a homotetramer to function as a proton channel in the membrane [80,81]. The protein has also been recognized for other roles during the viral life cycle, including mediating budding and virion release from cells [82–86]. Interestingly, not only is M2 predominantly localized at the necks of budding virions, but experiments using electron microscopy and SAXS have found the protein capable of generating NGC in model membranes and that this membrane activity is primarily attributed to its cytoplasmic C-terminal α-helix [54]. Indeed, this region of the protein received a high σ score from the SVM classifier [87].
Mitochondrial remodeling machinery
The morphology and intracellular distribution of mitochondria are crucial in maintaining normal cell function and are governed by a balance between the antagonistic processes of mitochondrial fission and fusion [88–91]. Excessive fusion leads to elongated mitochondria that form highly interconnective networks, while uninhibited fission leads to increased mitochondrial fragmentation [89,92–94]. The proteins that control mitochondrial fission and fusion play important roles in health and disease, as the dysregulation of these dynamic processes is associated with a variety of developmental disorders and neurodegenerative diseases [94–96]. The major essential protein involved in mitochondrial fission, Dnm1 in yeast [88,91,93,97] and Drp1 in humans [92,98], is a highly conserved cytosolic dynamin-related GTPase. One prominent model describes Dnm1/Drp1 as a molecular motor that selfassembles into ring-like oligomeric structures that encircle the outer mitochondrial membrane at sites of fission. GTP hydrolysis then leads to conformational changes in Dnm1/Drp1 that cause constriction and pinching of the membrane to drive membrane scission [99–103]. Our recent work using machine learning predicted a conserved α-helical domain in Dnm1 and Drp1 (and protein relatives with fission activity) to be capable of remodeling membranes by generating NGC. SAXS measurements revealed that the full Dnm1 protein restructures model mitochondrial membranes into phases rich in NGC and can induce a membrane fission neck diameter of 12.6 nm, which is smaller than the observed diameters achieved from mechanical constriction. These findings together suggest that Dnm1-induced membrane curvature and molecular motor driven mechanochemical forces function synergistically to efficiently drive mitochondrial fission. In fact, when members of the dynamin superfamily are individually examined using the SVM classifier, it is found that their machine learning σ scores decreased as phylogenetic distances from classical dynamin Dyn1 increased, which suggests that the dynamin superfamily GTPases likely evolved the ability to generate membrane curvature to optimize their membrane remodeling roles [104].
New architectures of cell penetrating peptides
Cell-penetrating peptides (CPPs) are recognized for their ability to efficiently translocate across cell membranes, and therefore, often utilized for mediating the uptake of conjugated cargos [105–107]. Similar to AMPs, CPPs are generally short, cationic peptides that can also be amphiphilic. In fact, there are marked similarities in the amino acid content of AMPs and CPPs in the form of a correlation between the lysine/arginine ratio and the hydrophobicity, which has been rationalized in terms of patterns of H-bonding required to induce NGC in target membranes [49,108]. Naturally derived CPPs, which typically exist as linear amino acid sequences, have since inspired numerous research groups to explore synthetic mimetics with often more complex architectures but still feature the key characteristics of cationic charge, hydrophobicity, and amphiphilicity. The diversity of these synthetic designs has ranged from circular peptides to side-chain-rich comb, brush, and dendrimer structures [109–113]. More recently, unique architectures composed of long flexible side chains surrounding a core have introduced the new attributes of radial amphiphilicity and metaphilic surfaces [51,114], which can lead to greater membrane activity than that of natural AMPs and CPPs.
Beyond membrane activity: AMPs as selfassembling amphiphilic meta-antigens that organize innate immune ligands for presentation
Here we describe an aspect of AMP activity that has not been predicted by machine learning, that of immune modulation. There have been many examples of AMPinduced immune in the last 15 years (especially with human cathelicidin LL37 and defensins [115]), but we focus on cases where the cationic charge and amphiphilicity of AMPs, the very characteristics that drive membrane activity according to both machine learning and traditional forms of scientific inquiry, may form the basis of unanticipated immune activity. Cationic AMPs can structurally organize and scaffold anionic immune ligands into spatially periodic complexes, and that the crystallinity of such complexes can determine the degree of immune amplification via multivalent presentation [116–118].
AMP–dsDNA complexes
Toll-like receptor 9 (TLR9) is an innate immune sensor for viral and bacterial CpG DNA [119]. In 2007, Lande et al. demonstrated that human cathelicidin LL37 could form complexes with human genomic dsDNA and potently activate plasmacytoid dendritic cells (pDCs) by binding to TLR9 [120,121]. In fact, overexpression of LL37 in autoimmune diseases like lupus [122] and psoriasis [123] has been linked to TLR9-mediated inflammation in both pDCs and keratinocytes [124]. Moreover, the ability to enable immune recognition of DNA and induce TLR9 activation in immune cells is not limited to LL37. Subsequent studies have shown that the AMPs human β-defensin 2 (hbD2) and human β- defensin 3 (hbD3) also co-assemble with dsDNA to activate pDCs [125]. It is clear that certain natural AMPs can induce TLR9 signaling through dsDNA binding, but many other molecules that also bind to dsDNA cannot. To delineate the necessary and sufficient criteria for immune activation by dsDNA complexes, high-resolution synchrotron SAXS was used to correlate the structures of complexes formed between dsDNA and AMPs and measurements of pDC IFN-a production induced by these complexes [116]. Cationic AMP molecules such as LL37 can form columnar nanocrystalline complexes with dsDNA, and present the DNA at an optimum range of inter-DNA spacing (dw3.5 nm) that can promote multivalent binding with clusters of TLR9 [126,127].
AMP–dsRNA complexes
The textbook role of Toll-like receptor 3 (TLR3) is to sense viral dsRNAs in infections [128]. Recently, it was shown that TLR3, which is expressed at high levels in human keratinocytes [129], also plays a critical role in sensing skin injury by binding to non-coding self-dsRNA [130,131]. Similarly, keratinocytes also produce large amounts of AMPs which are important for microbial defense in the skin. Interestingly, LL37 and other cationic peptides have been shown to enhance or inhibit TLR3 signaling by viral dsRNAs [132–134]. In addition, in autoimmune diseases like psoriasis, IL-6 production by keratinocytes, which is downstream from TLR3, play a role in aberrant cytokine production in response to LL37 [131,135,136]. We showed that the structural ordering of dsRNAs by AMPs can influence TLR3 activation, similar to how scaffolding of dsDNA by AMPs modulates TLR9 activation. Again, AMPe dsRNA complexes with inter-dsRNA spacings commensurate with the steric size of TLR3 lead to a w5e10-fold amplification in IL-6 production from keratinocytes [117,137], but complexes with interdsRNA spacings much smaller or larger than the steric size of TLR3 lead to low levels of activation.
In both LL37edsDNA complexes and LL37edsRNA complexes, the inter-nucleic acid spacing is significantly larger than the diameter of the nucleic acid and too large to be spanned by the diameter of a single α-helical LL37 mol. There has been a recent surge of interest in understanding the assembly of amphiphilic or chemically “patchy”, anisotropic objects. These systems can assemble into structures considerably more complex than micelles [138]. Indeed, systems such as colloids with directional bonding [139] and programmed assembly of “Janus” particles into a Kagome lattice [140] have been realized. Moreover, shape and chemical heterogeneity can exert different demands on the assembly process and lead to counterintuitive results [141]. It is interesting to see how these types of organizing principles influence immune modulation. Although AMPs are known to aggregate under special conditions [142], it is not clear how curved cationic amphiphiles such as AMPs electrostatically assemble with anionic DNA, much less how the resultant self-assembled structures connect to their immunomodulatory behavior.
Outlook
In this review, we discussed two different aspects of AMP activity mediated by their amphiphilicity and cationic charge: their well-known membrane remodeling activity, and their ability to assemble with and present immune ligands: AMPs can kill microbes through direct action, including membrane permeation, disruption of electrochemical gradients, and inhibition of metabolic processes. However, AMPs can also orchestrate host immune responses by communicating with the innate and adaptive immune systems, leading to downstream responses such as chemotaxis, differentiation/maturation, and cytokine production. We have shown that the application of machine learning methods can lead to a powerful forms of sequence analysis beyond algorithms for sequence comparison such as BLAST. (The SVM classifier can identify AMP-like behavior in sequences that have little homology with natural AMP sequences [87].) On the other hand, we have also highlighted some limitations of machine learning: Whereas membrane activity can be successfully “predicted” by machine learning in arbitrary sequences, the immunological activity is completely unanticipated, even though both types of activity are related to amphiphilicity and cationic charge. Finally, the picture of AMPs presented above represents a generalization of the central paradigm in immunology: AMPs are not just effector molecules but can also act as meta-antigens that activate innate immunity by organizing immune ligands. Moreover, innate immune receptors such as Toll-like receptors (TLRs) recognize not just pathogenassociated molecular patterns of single ligand molecules, but also recognize nanocrystalline arrangements of AMPs and ligands.
Acknowledgements
G.C.L.W. and M.W.L. acknowledge support from the National Science Foundation (DMR1808459). E.Y.L. acknowledges support from the Systems and Integrative Biology Training Program (T32GM008185), the Medical Scientist Training Program (T32GM008042), and the Dermatology Scientist Training Program (T32AR071307) at UCLA. E.Y.L. and G.C.L.W. acknowledge an Early Career Research Grant and a Discovery Grant from the National Psoriasis Foundation, respectively.
Footnotes
Conflict of interest statement
Nothing declared.
References
- 1.Zasloff M: Antimicrobial peptides of multicellular organisms. Nature 2002, 415:389–395, 10.1038/415389a. [DOI] [PubMed] [Google Scholar]
- 2.Shai Y: Mechanism of the binding, insertion and destabilization of phospholipid bilayer membranes by α-helical antimicrobial and cell non-selective membrane-lytic peptides. Biochim Biophys Acta Biomembr 1999, 1462:55–70, 10.1016/S0005-2736(99)00200-X. [DOI] [PubMed] [Google Scholar]
- 3.Brogden KA: Antimicrobial peptides: pore formers or metabolic inhibitors in bacteria? Nat Rev Microbiol 2005, 3: 238–250, 10.1038/nrmicro1098. [DOI] [PubMed] [Google Scholar]
- 4.Hancock REW, Lehrer R: Cationic peptides: a new source of antibiotics. Trends Biotechnol 1998, 16:82–88, 10.1016/S0167-7799(97)01156-6. [DOI] [PubMed] [Google Scholar]
- 5.Hancock REW, Sahl H-G: Antimicrobial and host-defense peptides as new anti-infective therapeutic strategies. Nat Biotechnol 2006, 24:1551–1557, 10.1038/nbt1267. [DOI] [PubMed] [Google Scholar]
- 6.Yeaman MR, Yount NY: Mechanisms of antimicrobial peptide action and resistance. Pharmacol Rev 2003, 55:27–55, 10.1124/pr.55.1.2. [DOI] [PubMed] [Google Scholar]
- 7.Zasloff M: Magainins, a class of antimicrobial peptides from Xenopus skin: isolation, characterization of two active forms, and partial cDNA sequence of a precursor. Proc Natl Acad Sci U S A 1987, 84:5449–5453, 10.1073/pnas.84.15.5449. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Ganz T: Defensins: antimicrobial peptides of innate immunity. Nat Rev Immunol 2003, 3:710–720, 10.1038/nri1180. [DOI] [PubMed] [Google Scholar]
- 9.Selsted ME, Novotny MJ, Morris WL, Tang YQ, Smith W, Cullor JS: Indolicidin, a novel bactericidal tridecapeptide amide from neutrophils. J Biol Chem 1992, 267:4292–4295. [PubMed] [Google Scholar]
- 10.Agerberth B, Lee JY, Bergman T, Carlquist M, Boman HG, Mutt V, et al. : Amino acid sequence of PR-39. Isolation from pig intestine of a new member of the family of prolinearginine-rich antibacterial peptides. Eur J Biochem 1991, 202:849–854, 10.1111/j.1432-1033.1991.tb16442.x. [DOI] [PubMed] [Google Scholar]
- 11.Wang Z, Wang G: APD: the antimicrobial peptide database. Nucleic Acids Res 2004, 32:D590–D592, 10.1093/nar/gkh025. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Wang G, Li X, Wang Z: APD2: the updated antimicrobial peptide database and its application in peptide design. Nucleic Acids Res 2009, 37:D933–D937, 10.1093/nar/gkn823. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Wang G, Li X, Wang Z: APD3: the antimicrobial peptide database as a tool for research and education. Nucleic Acids Res 2016, 44:D1087–D1093, 10.1093/nar/gkv1278. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Hancock REW, Rozek A: Role of membranes in the activities of antimicrobial cationic peptides. FEMS (Fed Eur Microbiol Soc) Microbiol Lett 2002, 206:143–149, 10.1111/j.1574-6968.2002.tb11000.x. [DOI] [PubMed] [Google Scholar]
- 15.Yang L, Harroun TA, Weiss TM, Ding L, Huang HW: Barrel-stave model or toroidal model? A case study on melittin pores. Biophys J 2001, 81:1475–1485, 10.1016/S00063495(01)75802-X. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Bechinger B, Kim Y, Chirlian LE, Gesell J, Neumann JM, Montal M, et al. : Orientations of amphipathic helical peptides in membrane bilayers determined by solid-state NMR spectroscopy. J Biomol NMR 1991, 1:167–173, 10.1007/BF01877228. [DOI] [PubMed] [Google Scholar]
- 17.Pouny Y, Rapaport D, Mor A, Nicolas P, Shai Y: Interaction of antimicrobial dermaseptin and its fluorescently labeled analogs with phospholipid membranes. Biochemistry 1992, 31: 12416–12423, 10.1021/bi00164a017. [DOI] [PubMed] [Google Scholar]
- 18.Matsuzaki K, Murase O, Fujii N, Miyajima K: An antimicrobial peptide, magainin 2, induced rapid flip-flop of phospholipids coupled with pore formation and peptide translocation. Biochemistry 1996, 35:11361–11368, 10.1021/bi960016v. [DOI] [PubMed] [Google Scholar]
- 19.Epand RM, Epand RF: Bacterial membrane lipids in the action of antimicrobial agents. J Pept Sci 2011, 17:298–305, 10.1002/psc.1319. [DOI] [PubMed] [Google Scholar]
- 20.Zachowski A: Phospholipids in animal eukaryotic membranes transverse asymmetry and movement. Biochem J 1993, 294: 1–14, 10.1042/bj2940001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.van Meer G, Voelker DR, Feigenson GW: Membrane lipids: where they are and how they behave. Nat Rev Mol Cell Biol 2008, 9:112–124, 10.1038/nrm2330. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Epand RF, Savage PB, Epand RM: Bacterial lipid composition and the antimicrobial efficacy of cationic steroid compounds (Ceragenins). Biochim Biophys Acta 2007, 1768:2500–2509, 10.1016/j.bbamem.2007.05.023. [DOI] [PubMed] [Google Scholar]
- 23.Zimmerberg J, Kozlov MM: How proteins produce cellular membrane curvature. Nat Rev Mol Cell Biol 2006, 7:9–19, 10.1038/nrm1784. [DOI] [PubMed] [Google Scholar]
- 24.Siegel DP, Kozlov MM: The Gaussian curvature elastic modulus of N-monomethylated dioleoylphosphatidylethanolamine: relevance to membrane fusion and lipid phase behavior. Biophys J 2004, 87:366–374, 10.1529/biophysj.104.040782. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Som A, Yang L, Wong GCL, Tew GN: Divalent metal ion triggered activity of a synthetic antimicrobial in cardiolipin membranes. J Am Chem Soc 2009, 131:15102–15103, 10.1021/ja9067063. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Yang L, Gordon VD, Trinkle DR, Schmidt NW, Davis MA, DeVries C, et al. : Mechanism of a prototypical synthetic membrane-active antimicrobial: efficient hole-punching via interaction with negative intrinsic curvature lipids. Proc Natl Acad Sci U S A 2008, 105:20595–20600, 10.1073/pnas.0806456105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Yang L, Gordon VD, Mishra A, Som A, Purdy KR, Davis MA, et al. : Synthetic antimicrobial oligomers induce a composition-dependent topological transition in membranes. J Am Chem Soc 2007, 129:12141–12147, 10.1021/ja072310o. [DOI] [PubMed] [Google Scholar]
- 28.Lata S, Sharma BK, Raghava G: Analysis and prediction of antibacterial peptides. BMC Bioinf 2007, 8:1, 10.1186/1471-2105-8-263. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Fjell CD, Jenssen H, Fries P, Aich P, Griebel P, Hilpert K, et al. : Identification of novel host defense peptides and the absence of α-defensins in the bovine genome. Proteins: Structure, Function, and Bioinformatics 2008, 73:420–430, 10.1002/prot.22059. [DOI] [PubMed] [Google Scholar]
- 30.Fjell CD, Jenssen H, Hilpert K, Cheung WA, Panté N, Hancock REW, et al. : Identification of novel antibacterial peptides by chemoinformatics and machine learning. J Med Chem 2009, 52:2006–2015, 10.1021/jm8015365. [DOI] [PubMed] [Google Scholar]
- 31.Cherkasov A, Hilpert K, Jenssen H, Fjell CD, Waldbrook M, Mullaly SC, et al. : Use of artificial intelligence in the design of small peptide antibiotics effective against a broad spectrum of highly antibiotic-resistant superbugs. ACS Chem Biol 2008, 4:65–74, 10.1021/cb800240j. [DOI] [PubMed] [Google Scholar]
- 32.Wang P, Hu L, Liu G, Jiang N, Chen X, Xu J, et al. : Prediction of Antimicrobial Peptides Based on Sequence Alignment and Feature Selection Methods. PloS One 2011, 6:e18476, 10.1371/journal.pone.0018476. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Torrent M, Andreu D, Nogués VM, Boix E: Connecting peptide physicochemical and antimicrobial properties by a rational prediction model. PloS One 2011, 6:e16968, 10.1371/journal.pone.0016968. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Xiao X, Wang P, Lin W-Z, Jia J-H, Chou K-C: iAMP-2L: a twolevel multi-label classifier for identifying antimicrobial peptides and their functional types. Anal Biochem 2013, 436: 168–177, 10.1016/j.ab.2013.01.019. [DOI] [PubMed] [Google Scholar]
- 35.Maccari G, Di Luca M, Nifosí R, Cardarelli F, Signore G, Boccardi C, et al. : Antimicrobial Peptides Design by Evolutionary Multiobjective Optimization. PLoS Comput Biol 2013, 9 e1003212, 10.1371/journal.pcbi.1003212. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Giguère S, Laviolette F, Marchand M, Tremblay D, Moineau S, Liang X, et al. : Machine learning assisted design of highly active peptides for drug discovery. PLoS Comput Biol 2015, 11 e1004074, 10.1371/journal.pcbi.1004074. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Schneider P, Müller AT, Gabernet G, Button AL, Posselt G, Wessler S, et al. : Hybrid network model for “deep learning” of chemical data: application to antimicrobial peptides. Molecular Informatics 2017, 36:1600011, 10.1002/minf.201600011. [DOI] [PubMed] [Google Scholar]
- 38.Rondon-Villarreal P, Sierra DA, Torres R: Machine learning in the rational design of antimicrobial peptides. Curr Comput Aided Drug Des 2014, 10:183–190, 10.2174/1573409910666140624124807. [DOI] [PubMed] [Google Scholar]
- 39.Lee EY, Fulan BM, Wong GCL, Ferguson AL: Mapping membrane activity in undiscovered peptide sequence space using machine learning. Proc Natl Acad Sci U S A 2016, 113: 13588–13593, 10.1073/pnas.1609893113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Lee EY, Wong GCL, Ferguson AL: Machine learning-enabled discovery and design of membrane-active peptides. Bioorg Med Chem 2018, 26:2708–2718, 10.1016/j.bmc.2017.07.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Lee EY, Lee MW, Fulan BM, Ferguson AL, Wong GCL: What can machine learning do for antimicrobial peptides, and what can antimicrobial peptides do for machine learning? Interface Focus 2017, 7:20160153, 10.1098/rsfs.2016.0153. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Hilpert K, Fjell CD, Cherkasov A: Short linear cationic antimicrobial peptides: screening, optimizing, and prediction. Methods Mol Biol 2008, 494:127–159, 10.1007/978-1-59745-419-3_8. [DOI] [PubMed] [Google Scholar]
- 43.Porto WF, Pires ÁS, Franco OL, CS-AMPPred: An updated SVM model for antimicrobial activity prediction in cysteinestabilized peptides. PloS One 2012, 7:e51444, 10.1371/journal.pone.0051444. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Mauri A, Ballabio D, Consonni V, Manganaro A, Todeschini R: Peptides multivariate characterisation using a molecular descriptor based approach. Match Commun Math Comp Chem 2008, 60:671–690. [Google Scholar]
- 45.Li ZR, Lin HH, Han LY, Jiang L, Chen X, Chen YZ: PROFEAT: a web server for computing structural and physicochemical features of proteins and peptides from amino acid sequence. Nucleic Acids Res 2006, 34:W32–W37, 10.1093/nar/gkl305. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Cao D-S, Xu Q-S, Liang Y-Z: propy: a tool to generate various modes of Chou’s PseAAC. Bioinformatics 2013, 29:960–962, 10.1093/bioinformatics/btt072. [DOI] [PubMed] [Google Scholar]
- 47.Bi J, Bennett K, Embrechts M, Breneman C, Song M: Dimensionality reduction via sparse support vector machines.J Mach Learn Res 2003, 3:1229–1243. [Google Scholar]
- 48.Schmidt NW, Tai KP, Kamdar K, Mishra A, Lai GH, Zhao K, et al. : Arginine in α-defensins: differential effects on bactericidal activity correspond to geometry of membrane curvature generation and peptide-lipid phase behavior. J Biol Chem 2012, 287:21866–21872, 10.1074/jbc.M112.358721. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Schmidt NW, Mishra A, Lai GH, Davis M, Sanders LK, Tran D, et al. : Criterion for amino acid composition of defensins and antimicrobial peptides based on geometry of membrane destabilization. J Am Chem Soc 2011, 133:6720–6727, 10.1021/ja200079a. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Lee MW, Chakraborty S, Schmidt NW, Murgai R, Gellman SH, Wong GCL: Two interdependent mechanisms of antimicrobial activity allow for efficient killing in nylon-3-based polymeric mimics of innate immunity peptides. Biochim Biophys Acta 2014, 1838:2269–2279, 10.1016/j.bbamem.2014.04.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Xiong M, Lee MW, Mansbach RA, Song Z, Bao Y, Peek RM, et al. : Helical antimicrobial polypeptides with radial amphiphilicity. Proc Natl Acad Sci U S A 2015, 112:13155–13160, 10.1073/pnas.1507893112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Kaplan A, Lee MW, Wolf AJ, Limon JJ, Becker CA, Ding M, et al. : Direct antimicrobial activity of IFN-b. J Immunol 2017, 198: 4036–4045, 10.4049/jimmunol.1601226. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Schmidt NW, Mishra A, Lai GH, Wong GCL: Arginine-rich cellpenetrating peptides. FEBS Lett 2010, 584:1806–1813, 10.1016/j.febslet.2009.11.046. [DOI] [PubMed] [Google Scholar]
- 54.Schmidt NW, Mishra A, Wang J, DeGrado WF, Wong GCL: Influenza virus a M2 protein generates negative Gaussian membrane curvature necessary for budding and scission. J Am Chem Soc 2013, 135:13710–13719, 10.1021/ja400146z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Yao H, Lee MW, Waring AJ, Wong GCL, Hong M: Viral fusion protein transmembrane domain adopts β-strand structure to facilitate membrane topological changes for virus-cell fusion. Proc Natl Acad Sci U S A 2015, 112:10926–10931, 10.1073/pnas.1501430112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.van den Pol AN: Neuropeptide transmission in brain circuits. Neuron 2012, 76:98–115, 10.1016/j.neuron.2012.09.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Wang Y, Wang M, Yin S, Jang R, Wang J, Xue Z, et al. NeuroPep: a comprehensive resource of neuropeptides Database, vol. 2015 Oxford; 2015. bav038, 10.1093/database/bav038. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Mykicki N, Herrmann AM, Schwab N, Deenen R, Sparwasser T, Limmer A, et al. : Melanocortin-1 receptor activation is neuroprotective in mouse models of neuroinflammatory disease. Sci Transl Med 2016, 8:362ra146, 10.1126/scitranslmed.aaf8732. [DOI] [PubMed] [Google Scholar]
- 59.Zugasti O, Bose N, Squiban B, Belougne J, Kurz CL, Schroeder FC, et al. : Activation of a G protein-coupled receptor by its endogenous ligand triggers the innate immune response of Caenorhabditis elegans. Nat Immunol 2014, 15: 833–838, 10.1038/ni.2957. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Cardoso V, Chesné J, Ribeiro H, García-Cassani B, Carvalho T, Bouchery T, et al. : Neuronal regulation of type 2 innate lymphoid cells via neuromedin U. Nature 2017, 549:277–281, 10.1038/nature23469. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Klose CSN, Mahlakõiv T, Moeller JB, Rankin LC, Flamar A-L, Kabata H, et al. : The neuropeptide neuromedin U stimulates innate lymphoid cells and type 2 inflammation. Nature 2017, 549:282–286, 10.1038/nature23676. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Wallrapp A, Riesenfeld SJ, Burkett PR, Abdulnour R-EE, Nyman J, Dionne D, et al. : The neuropeptide NMU amplifies ILC2-driven allergic lung inflammation. Nature 2017, 549: 351–356, 10.1038/nature24029. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Gonzalez Rey E, Chorny A, VIP Delgado M: An agent with license to kill infective parasites. Ann N Y Acad Sci 2006, 1070: 303–308, 10.1196/annals.1317.032. [DOI] [PubMed] [Google Scholar]
- 64.Karim El IA, Linden GJ, Orr DF, Lundy FT: Antimicrobial activity of neuropeptides against a range of micro-organisms from skin, oral, respiratory and gastrointestinal tract sites. J Neuroimmunol 2008, 200:11–16, 10.1016/j.jneuroim.2008.05.014. [DOI] [PubMed] [Google Scholar]
- 65.Kowalska K, Carr DB, Lipkowski AW: Direct antimicrobial properties of substance P. Life Sci 2002, 71:747–750. [DOI] [PubMed] [Google Scholar]
- 66.Starr CG, Maderdrut JL, He J, Coy DH, Wimley WC: Pituitary adenylate cyclase-activating polypeptide is a potent broadspectrum antimicrobial peptide: structure-activity relationships. Peptides 2018, 104:35–40, 10.1016/j.peptides.2018.04.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Brogden KA, Guthmiller JM, Salzet M, Zasloff M: The nervous system and innate immunity: the neuropeptide connection. Nat Immunol 2005, 6:558–564, 10.1038/ni1209. [DOI] [PubMed] [Google Scholar]
- 68.Lamb RA, Jardetzky TS: Structural basis of viral invasion: lessons from paramyxovirus F. Curr Opin Struct Biol 2007, 17: 427–436, 10.1016/j.sbi.2007.08.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Tamm LK, Crane J, Kiessling V: Membrane fusion: a structural perspective on the interplay of lipids and proteins. Curr Opin Struct Biol 2003, 13:453–466, 10.1016/S0959440X(03)00107-6. [DOI] [PubMed] [Google Scholar]
- 70.Qiang W, Sun Y, Weliky DP: A strong correlation between fusogenicity and membrane insertion depth of the HIV fusion peptide. Proc Natl Acad Sci U S A 2009, 106:15314–15319, 10.1073/pnas.0907360106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Lai AL, Moorthy AE, Li Y, Tamm LK: Fusion activity of HIV gp41 fusion domain is related to its secondary structure and depth of membrane insertion in a cholesterol-dependent fashion. J Mol Biol 2012, 418:3–15, 10.1016/j.jmb.2012.02.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Lorieau JL, Louis JM, Bax A: The complete influenza hemagglutinin fusion domain adopts a tight helical hairpin arrangement at the lipid:water interface. Proc Natl Acad Sci U S A 2010, 107:11341–11346, 10.1073/pnas.1006142107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Yao H, Hong M: Conformation and lipid interaction of the fusion peptide of the paramyxovirus PIV5 in anionic and negative-curvature membranes from solid-state NMR. J Am Chem Soc 2014, 136:2611–2624, 10.1021/ja4121956. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Yao H, Hong M: Membrane-dependent conformation, dynamics, and lipid interactions of the fusion peptide of the paramyxovirus PIV5 from solid-state NMR. J Mol Biol 2013, 425:563–576, 10.1016/j.jmb.2012.11.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Gerl MJ, Sampaio JL, Urban S, Kalvodova L, Verbavatz J-M, Binnington B, et al. : Quantitative analysis of the lipidomes of the influenza virus envelope and MDCK cell apical membrane. J Cell Biol 2012, 196:213–221, 10.1083/jcb.201108175. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Carlton JG, Martin-Serrano J: The ESCRT machinery: new functions in viral and cellular biology. Biochem Soc Trans 2009, 37:195–199, 10.1042/BST0370195. [DOI] [PubMed] [Google Scholar]
- 77.Chen BJ, Lamb RA: Mechanisms for enveloped virus budding: can some viruses do without an ESCRT? Virology 2008, 372: 221–232, 10.1016/j.virol.2007.11.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Pornillos O, Garrus JE, Sundquist WI: Mechanisms of enveloped RNA virus budding. Trends Cell Biol 2002, 12:569–579, 10.1016/S0962-8924(02)02402-9. [DOI] [PubMed] [Google Scholar]
- 79.Bruce EA, Medcalf L, Crump CM, Noton SL, Stuart AD, Wise HM, et al. : Budding of filamentous and non-filamentous influenza A virus occurs via a VPS4 and VPS28-independent pathway. Virology 2009, 390:268–278, 10.1016/j.virol.2009.05.016. [DOI] [PubMed] [Google Scholar]
- 80.Stewart SM, Pekosz A: Mutations in the membrane-proximal region of the influenza A virus M2 protein cytoplasmic tail have modest effects on virus replication. J Virol 2011, 85: 12179–12187, 10.1128/JVI.05970-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Schnell JR, Chou JJ: Structure and mechanism of the M2 proton channel of influenza A virus. Nature 2008, 451: 591–595, 10.1038/nature06531. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Takeda M, Pekosz A, Shuck K, Pinto LH, Lamb RA: Influenza a virus M2 ion channel activity is essential for efficient replication in tissue culture. J Virol 2002, 76:1391–1399, 10.1128/JVI.76.3.1391-1399.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Watanabe T, Watanabe S, Ito H, Kida H, Kawaoka Y: Influenza A virus can undergo multiple cycles of replication without M2 ion channel activity. J Virol 2001, 75:5656–5662, 10.1128/JVI.75.12.5656-5662.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Iwatsuki-Horimoto K, Horimoto T, Noda T, Kiso M, Maeda J, Watanabe S, et al. : The cytoplasmic tail of the influenza A virus M2 protein plays a role in viral assembly. J Virol 2006, 80:5233–5240, 10.1128/JVI.00049-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Rossman JS, Jing X, Leser GP, Balannik V, Pinto LH, Lamb RA: Influenza virus m2 ion channel protein is necessary for filamentous virion formation. J Virol 2010, 84:5078–5088, 10.1128/JVI.00119-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Rossman JS, Jing X, Leser GP, Lamb RA: Influenza virus M2 protein mediates ESCRT-independent membrane scission. Cell 2010, 142:902–913, 10.1016/j.cell.2010.08.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Lee MW, Lee EY, Wong GCL: What can pleiotropic proteins in innate immunity teach us about bioconjugation and molecular design? Bioconjug Chem 2018, 29:2127–2139, 10.1021/acs.bioconjchem.8b00176. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Shaw JM, Nunnari J: Mitochondrial dynamics and division in budding yeast. Trends Cell Biol 2002, 12:178–184. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Westermann B: Mitochondrial fusion and fission in cell life and death. Nat Rev Mol Cell Biol 2010, 11:872–884, 10.1038/nrm3013. [DOI] [PubMed] [Google Scholar]
- 90.Sesaki H, Jensen RE: Division versus fusion: Dnm1p and Fzo1p antagonistically regulate mitochondrial shape. J Cell Biol 1999, 147:699–706. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Bleazard W, McCaffery JM, King EJ, Bale S, Mozdy A, Tieu Q, et al. : The dynamin-related GTPase Dnm1 regulates mitochondrial fission in yeast. Nat Cell Biol 1999, 1:298–304, 10.1038/13014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Smirnova E, Shurland DL, Ryazantsev SN, van der Bliek AM:A human dynamin-related protein controls the distribution of mitochondria. J Cell Biol 1998, 143:351–358. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Otsuga D, Keegan BR, Brisch E, Thatcher JW, Hermann GJ, Bleazard W, et al. : The dynamin-related GTPase, Dnm1p, controls mitochondrial morphology in yeast. J Cell Biol 1998, 143:333–349. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Chan DC: Mitochondrial fusion and fission in mammals. Annu Rev Cell Dev Biol 2006, 22:79–99, 10.1146/annurev.cellbio.22.010305.104638. [DOI] [PubMed] [Google Scholar]
- 95.Itoh K, Nakamura K, Iijima M, Sesaki H: Mitochondrial dynamics in neurodegeneration. Trends Cell Biol 2013, 23: 64–71, 10.1016/j.tcb.2012.10.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Knott AB, Perkins G, Schwarzenbacher R, Bossy-Wetzel E: Mitochondrial fragmentation in neurodegeneration. Nat Rev Neurosci 2008, 9:505–518, 10.1038/nrn2417. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Mozdy AD, McCaffery JM, Shaw JM: Dnm1p GTPase-mediated mitochondrial fission is a multi-step process requiring the novel integral membrane component Fis1p. J Cell Biol 2000, 151:367–379, 10.1083/jcb.151.2.367. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Smirnova E, Griparic L, Shurland DL, van der Bliek AM: Dynamin-related protein Drp1 is required for mitochondrial division in mammalian cells. Mol Biol Cell 2001, 12:2245–2256, 10.1091/mbc.12.8.2245. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Ingerman E, Perkins EM, Marino M, Mears JA, McCaffery JM, Hinshaw JE, et al. : Dnm1 forms spirals that are structurally tailored to fit mitochondria. J Cell Biol 2005, 170:1021–1027, 10.1083/jcb.200506078. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Lackner LL, Horner JS, Nunnari J: Mechanistic analysis of a dynamin effector. Science 2009, 325:874–877, 10.1126/science.1176921. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Bhar D, Karren MA, Babst M, Shaw JM: Dimeric Dnm1-G385D interacts with Mdv1 on mitochondria and can be stimulated to assemble into fission complexes containing Mdv1 and Fis1. J Biol Chem 2006, 281:17312–17320, 10.1074/jbc.M513530200. [DOI] [PubMed] [Google Scholar]
- 102.Mears JA, Lackner LL, Fang S, Ingerman E, Nunnari J,Hinshaw JE: Conformational changes in Dnm1 support a contractile mechanism for mitochondrial fission. Nat Struct Mol Biol 2011, 18:20–26, 10.1038/nsmb.1949. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Daumke O, Praefcke GJK: Invited review: mechanisms of GTP hydrolysis and conformational transitions in the dynamin superfamily. Peptide Science 2016, 105:580–593, 10.1002/bip.22855. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Lee MW, Lee EY, Lai GH, Kennedy NW, Posey AE, Xian W,et al. : Molecular motor Dnm1 synergistically induces membrane curvature to facilitate mitochondrial fission. ACS Cent Sci 2017, 3:1156–1167, 10.1021/acscentsci.7b00338. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.Milletti F: Cell-penetrating peptides: classes, origin, and current landscape. Drug Discov Today 2012, 17:850–860, 10.1016/j.drudis.2012.03.002. [DOI] [PubMed] [Google Scholar]
- 106.Koren E, Torchilin VP: Cell-penetrating peptides: breaking through to the other side. Trends Mol Med 2012, 18:385–393, 10.1016/j.molmed.2012.04.012. [DOI] [PubMed] [Google Scholar]
- 107.Bechara C, Sagan S: Cell-penetrating peptides: 20 years later, where do we stand? FEBS Lett 2013, 587:1693–1702, 10.1016/j.febslet.2013.04.031. [DOI] [PubMed] [Google Scholar]
- 108.Mishra A, Lai GH, Schmidt NW, Sun VZ, Rodriguez AR, Tong R, et al. : Translocation of HIV TAT peptide and analogues induced by multiplexed membrane and cytoskeletal interactions. Proc Natl Acad Sci U S A 2011, 108:16883–16888, 10.1073/pnas.1108795108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Lam SJ, O’Brien-Simpson NM, Pantarat N, Sulistio1 A,Wong EHH, Chen Y-Y, et al. : Combating multidrug-resistant Gram-negative bacteria with structurally nanoengineered antimicrobial peptide polymers. Nature Microbiology 2016, 1: 1–11, 10.1038/nmicrobiol.2016.162. [DOI] [PubMed] [Google Scholar]
- 110.Zhao K, Choe U-J, Kamei DT, Wong GCL: Enhanced activity of cyclic transporter sequences driven by phase behavior of peptide-lipid complexes. Soft Matter 2012, 8:6430–6433, 10.1039/C2SM25405K. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Saleh AF, Arzumanov A, Abes R, Owen D, Lebleu B, Gait MJ: Synthesis and splice-redirecting activity of branched, arginine-rich peptide dendrimer conjugates of peptide nucleic acid oligonucleotides. Bioconjug Chem 2010, 21: 1902–1911, 10.1021/bc100275r. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Mandal D, Nasrolahi Shirazi A, Parang K: Cell-penetrating homochiral cyclic peptides as nuclear-targeting molecular transporters. Angew Chem Int Ed Engl 2011, 50:9633–9637, 10.1002/anie.201102572. [DOI] [PubMed] [Google Scholar]
- 113.Angeles-Boza AM, Erazo-Oliveras A, Lee Y-J, Pellois J-P: Generation of endosomolytic reagents by branching of cellpenetrating peptides: tools for the delivery of bioactive compounds to live cells in cis or trans. Bioconjug Chem 2010, 21:2164–2167, 10.1021/bc100130r. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Lee MW, Han M, Bossa GV, Snell C, Song Z, Tang H, et al. : Interactions between membranes and “metaphilic” polypeptide architectures with diverse side-chain populations. ACS Nano 2017, 11:2858–2871, 10.1021/acsnano.6b07981. [DOI] [PubMed] [Google Scholar]
- 115.Bowdish DME, Davidson DJ, Hancock REW: Immunomodulatory properties of defensins and cathelicidins. Curr Top Microbiol Immunol 2006, 306:27–66, 10.1007/3540-29916-5_2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Schmidt NW, Jin F, Lande R, Curk T, Xian W, Lee C, et al. : Liquid-crystalline ordering of antimicrobial peptide-DNA complexes controls TLR9 activation. Nat Mater 2015, 14: 696–700, 10.1038/nmat4298. [DOI] [PubMed] [Google Scholar]
- 117.Lee EY, Takahashi T, Curk T, Dobnikar J, Gallo RL, Wong GCL: Crystallinity of double-stranded RNA-antimicrobial peptide complexes modulates toll-like receptor 3-mediated inflammation. ACS Nano 2017, 11:12145–12155, 10.1021/acsnano.7b05234. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Lee EY, Lee MW, Wong GCL: Modulation of Toll-like receptor signaling by antimicrobial peptides. Semin Cell Dev Biol 2018, 10.1016/j.semcdb.2018.02.002. [DOI] [PMC free article] [PubMed]
- 119.Ohto U, Shibata T, Tanji H, Ishida H, Krayukhina E, Uchiyama S, et al. : Structural basis of CpG and inhibitory DNA recognition by Toll-like receptor 9. Nature 2015, 520:702–705, 10.1038/nature14138. [DOI] [PubMed] [Google Scholar]
- 120.Lande R, Gregorio J, Facchinetti V, Chatterjee B, Wang Y-H,Homey B, et al. : Plasmacytoid dendritic cells sense self-DNA coupled with antimicrobial peptide. Nature 2007, 449: 564–569, 10.1038/nature06116. [DOI] [PubMed] [Google Scholar]
- 121.Gilliet M, Lande R: Antimicrobial peptides and self-DNA in autoimmune skin inflammation. Curr Opin Immunol 2008, 20: 401–407, 10.1016/j.coi.2008.06.008. [DOI] [PubMed] [Google Scholar]
- 122.Lande R, Ganguly D, Facchinetti V, Frasca L, Conrad C,Gregorio J, et al. : Neutrophils activate plasmacytoid dendritic cells by releasing self-DNA-peptide complexes in systemic lupus erythematosus. Sci Transl Med 2011, 3:73ra19, 10.1126/scitranslmed.3001180. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Lande R, Botti E, Jandus C, Dojcinovic D, Fanelli G, Conrad C, et al. : The antimicrobial peptide LL37 is a T-cell autoantigen in psoriasis. Nat Commun 2014, 5:5621, 10.1038/ncomms6621. [DOI] [PubMed] [Google Scholar]
- 124.Morizane S, Yamasaki K, Mühleisen B: Cathelicidin antimicrobial peptide LL-37 in psoriasis enables keratinocyte reactivity against TLR9 ligands. J Invest Dermatol 2012, 132:135–143, 10.1038/jid.2011.259. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Lande R, Chamilos G, Ganguly D, Demaria O, Frasca L, Durr S, et al. : Cationic antimicrobial peptides in psoriatic skin cooprate to break innate tolerance to self-DNA. Eur J Immunol 2015, 45:203–213, 10.1002/eji.201344277. [DOI] [PubMed] [Google Scholar]
- 126.Lee EY, Lee CK, Schmidt NW, Jin F, Lande R, Curk T, et al. : A review of immune amplification via ligand clustering by self-assembled liquid-crystalline DNA complexes. Adv Colloid Interface Sci 2016, 232:17–24, 10.1016/j.cis.2016.02.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127.Tursi SA, Lee EY, Medeiros NJ, Lee MH, Nicastro LK, Buttaro B, et al. : Bacterial amyloid curli acts as a Carrier for DNA to elicit an autoimmune response via TLR2 and TLR9. PLoS Pathog 2017, 13:e1006315, 10.1371/journal.ppat.1006315. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128.Alexopoulou L, Holt AC, Medzhitov R, Flavell RA: Recognition of double-stranded RNA and activation of NF-kappaB by Tolllike receptor 3. Nature 2001, 413:732–738, 10.1038/35099560. [DOI] [PubMed] [Google Scholar]
- 129.Nelson AM, Reddy SK, Ratliff TS, Hossain MZ, Katseff AS,Zhu AS, et al. : dsRNA released by tissue damage activates TLR3 to drive skin regeneration. Cell Stem Cell 2015, 17: 139–151, 10.1016/j.stem.2015.07.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130.Adase CA, Borkowski AW, Zhang L-J, Williams M, Sato E, Sanford JA, et al. : Non-coding double-stranded RNA and LL37 induce growth factor expression from keratinocytes and endothelial cells. J Biol Chem 2016, 291:11635–11646, 10.1074/jbc.M116.725317. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131.Zhang L-J, Sen GL, Ward NL, Johnston A, Chun K, Chen Y, et al. : Antimicrobial peptide LL37 and MAVS signaling drive interferon-b production by epidermal keratinocytes during skin injury. Immunity 2016, 45:119–130, 10.1016/j.immuni.2016.06.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.Lai Y, Adhikarakunnathu S, Bhardwaj K, Ranjith-Kumar CT, Wen Y, Jordan JL, et al. : LL37 and Cationic Peptides Enhance TLR3 Signaling by Viral Double-stranded RNAs. PloS One 2011, 6:e26632, 10.1371/journal.pone.0026632. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133.Chen X, Takai T, Xie Y, Niyonsaba F, Okumura K, Ogawa H: Human antimicrobial peptide LL-37 modulates proinflammatory responses induced by cytokine milieus and double-stranded RNA in human keratinocytes. Biochem Biophys Res Commun 2013, 433:532–537, 10.1016/j.bbrc.2013.03.024. [DOI] [PubMed] [Google Scholar]
- 134.Chen X, Takai T, Xie Y, Okumura K, Ikeda S, Ogawa H: Modulation of double-stranded RNAand cytokineinduced responses of human keratinocytes by LL-37. J Dermatol Sci 2013, 69:e14, 10.1016/j.jdermsci.2012.11.339. [DOI] [Google Scholar]
- 135.Takahashi T, Kulkarni NN, Lee EY, Zhang L, Wong GCL, Aiba S, et al. : 886 Discovery of a receptor-dependent step in cathelicidin activation of inflammation identifies a novel therapeutic target for psoriasis and rosacea. J Invest Dermatol 2018, 138:S151, 10.1016/j.jid.2018.03.898. [DOI] [Google Scholar]
- 136.Takahashi T, Kulkarni NN, Lee EY, Zhang L-J, Wong GCL, Gallo RL: Cathelicidin promotes inflammation by enabling binding of self-RNA to cell surface scavenger receptors. Sci Rep 2018, 8:4032, 10.1038/s41598-018-22409-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137.Lee EY, Takahashi T, Curk T, Dobnikar J, Gallo RL, Wong GCL: 070 Liquid crystalline ordering of antimicrobial peptide-RNA complexes controls TLR3 activation. J Invest Dermatol 2017, 137:S12, 10.1016/j.jid.2017.02.083. [DOI] [Google Scholar]
- 138.Glotzer SC, Solomon MJ: Anisotropy of building blocks and their assembly into complex structures. Nat Mater 2007, 6: 557–562, 10.1038/nmat1949. [DOI] [PubMed] [Google Scholar]
- 139.Wang Y, Wang Y, Breed DR, Manoharan VN, Feng L, Hollingsworth AD, et al. : Colloids with valence and specific directional bonding. Nature 2012, 491:51–55, 10.1038/nature11564. [DOI] [PubMed] [Google Scholar]
- 140.Chen Q, Bae SC, Granick S: Directed self-assembly of a colloidal kagome lattice. Nature 2011, 469:381–384, 10.1038/nature09713. [DOI] [PubMed] [Google Scholar]
- 141.Ye X, Chen J, Engel M, Millan JA, Li W, Qi L, et al. : Competition of shape and interaction patchiness for self-assembling nanoplates. Nat Chem 2013, 5:466–473, 10.1038/nchem.1651. [DOI] [PubMed] [Google Scholar]
- 142.Schnaider L, Brahmachari S, Schmidt NW, Mensa B, ShahamNiv S, Bychenko D, et al. : Self-assembling dipeptide antibacterial nanostructures with membrane disrupting activity. Nat Commun 2017, 8:1365, 10.1038/s41467-017-01447-x.s [DOI] [PMC free article] [PubMed] [Google Scholar]