Generic placeholder image

Current Protein & Peptide Science

Editor-in-Chief

ISSN (Print): 1389-2037
ISSN (Online): 1875-5550

Review Article

About TFE: Old and New Findings

Author(s): Marian Vincenzi, Flavia A. Mercurio and Marilisa Leone*

Volume 20, Issue 5, 2019

Page: [425 - 451] Pages: 27

DOI: 10.2174/1389203720666190214152439

Price: $65

Abstract

The fluorinated alcohol 2,2,2-Trifluoroethanol (TFE) has been implemented for many decades now in conformational studies of proteins and peptides. In peptides, which are often disordered in aqueous solutions, TFE acts as secondary structure stabilizer and primarily induces an α -helical conformation. The exact mechanism through which TFE plays its stabilizing roles is still debated and direct and indirect routes, relying either on straight interaction between TFE and molecules or indirect pathways based on perturbation of solvation sphere, have been proposed. Another still unanswered question is the capacity of TFE to favor in peptides a bioactive or a native-like conformation rather than simply stimulate the raise of secondary structure elements that reflect only the inherent propensity of a specific amino-acid sequence. In protein studies, TFE destroys unique protein tertiary structure and often leads to the formation of non-native secondary structure elements, but, interestingly, gives some hints about early folding intermediates. In this review, we will summarize proposed mechanisms of TFE actions. We will also describe several examples, in which TFE has been successfully used to reveal structural properties of different molecular systems, including antimicrobial and aggregation-prone peptides, as well as globular folded and intrinsically disordered proteins.

Keywords: TFE, α-helix, conformational transitions, secondary structure, folding, amyloids, IDPs.

Graphical Abstract
[1]
Buck, M. Trifluoroethanol and colleagues: Cosolvents come of age. Recent studies with peptides and proteins. Q. Rev. Biophys., 1998, 31(3), 297-355.
[2]
Searle, M.S.; Zerella, R.; Williams, D.H.; Packman, L.C. Native-like beta-hairpin structure in an isolated fragment from ferredoxin: NMR and CD studies of solvent effects on the N-terminal 20 residues. Protein Eng., 1996, 9(7), 559-565.
[3]
Hamada, D.; Goto, Y. The equilibrium intermediate of beta-lactoglobulin with non-native alpha-helical structure. J. Mol. Biol., 1997, 269(4), 479-487.
[4]
Kumar, S.; Modig, K.; Halle, B. Trifluoroethanol-induced beta -- alpha transition in beta-lactoglobulin: hydration and cosolvent binding studied by 2H, 17O, and 19F magnetic relaxation dispersion. Biochemistry, 2003, 42(46), 13708-13716.
[5]
Yang, Y.; Barker, S.; Chen, M.J.; Mayo, K.H. Effect of low molecular weight aliphatic alcohols and related compounds on platelet factor 4 subunit association. J. Biol. Chem., 1993, 268(13), 9223-9229.
[6]
Othon, C.M.; Kwon, O.H.; Lin, M.M.; Zewail, A.H. Solvation in protein (un)folding of melittin tetramer-monomer transition. Proc. Natl. Acad. Sci. USA, 2009, 106(31), 12593-12598.
[7]
Anderson, V.L.; Webb, W.W. A desolvation model for trifluoroethanol-induced aggregation of enhanced green fluorescent protein. Biophys. J., 2012, 102(4), 897-906.
[8]
Fioroni, M.; Diaz, M.D.; Burger, K.; Berger, S. Solvation phenomena of a tetrapeptide in water/trifluoroethanol and water/ethanol mixtures: A diffusion NMR, intermolecular NOE, and molecular dynamics study. J. Am. Chem. Soc., 2002, 124(26), 7737-7744.
[9]
Hong, D.P.; Hoshino, M.; Kuboi, R.; Goto, Y. Clustering of fluorine-substituted alcohols as a factor responsible for their marked effects on proteins and peptides. J. Am. Chem. Soc., 1999, 121(37), 8427-8433.
[10]
Roccatano, D.; Colombo, G.; Fioroni, M.; Mark, A.E. Mechanism by which 2,2,2-trifluoroethanol/water mixtures stabilize secondary-structure formation in peptides: A molecular dynamics study. Proc. Natl. Acad. Sci. USA, 2002, 99(19), 12179-12184.
[11]
Diaz, M.D.; Fioroni, M.; Burger, K.; Berger, S. Evidence of complete hydrophobic coating of bombesin by trifluoroethanol in aqueous solution: An NMR spectroscopic and molecular dynamics study. Chem. Eur. J., 2002, 8(7), 1663-1669.
[12]
Cantisani, M.; Finamore, E.; Mignogna, E.; Falanga, A.; Nicoletti, G.F.; Pedone, C.; Morelli, G.; Leone, M.; Galdiero, M.; Galdiero, S. Structural insights into and activity analysis of the antimicrobial peptide myxinidin. Antimicrob. Agents Chemother., 2014, 58(9), 5280-5290.
[13]
Cantisani, M.; Leone, M.; Mignogna, E.; Kampanaraki, K.; Falanga, A.; Morelli, G.; Galdiero, M.; Galdiero, S. Structure-activity Rrelations of myxinidin, an antibacterial peptide derived from the epidermal mucus of hagfish. Antimicrob. Agents Chemother., 2013, 57(11), 5665-5673.
[14]
Scudiero, O.; Nigro, E.; Cantisani, M.; Colavita, I.; Leone, M.; Mercurio, F.A.; Galdiero, M.; Pessi, A.; Daniele, A.; Salvatore, F.; Galdiero, S. Design and activity of a cyclic mini-beta-defensin analog: A novel antimicrobial tool. Int. J. Nanomedicine, 2015, 10, 6523-6539.
[15]
Buonanno, M.; Coppola, M.; Di Lelio, I.; Molisso, D.; Leone, M.; Pennacchio, F.; Langella, E.; Rao, R.; Monti, S.M. Prosystemin, a prohormone that modulates plant defense barriers, is an intrinsically disordered protein. Protein Sci., 2018, 27(3), 620-632.
[16]
Pirone, L.; Ripoll-Rozada, J.; Leone, M.; Ronca, R.; Lombardo, F.; Fiorentino, G.; Andersen, J.F.; Pereira, P.J.B.; Arca, B.; Pedone, E. Functional analyses yield detailed insight into the mechanism of thrombin inhibition by the antihemostatic salivary protein cE5 from Anopheles gambiae. J. Biol. Chem., 2017, 292(30), 12632-12642.
[17]
Mercurio, F.A.; Di Natale, C.; Pirone, L.; Scognamiglio, P.L.; Marasco, D.; Pedone, E.M.; Saviano, M.; Leone, M. Peptide fragments of Odin-Sam1: conformational analysis and interaction studies with EphA2-Sam. ChemBioChem, 2015, 16(11), 1629-1636.
[18]
Mercurio, F.A.; Scognamiglio, P.L.; Di Natale, C.; Marasco, D.; Pellecchia, M.; Leone, M. CD and NMR conformational studies of a peptide encompassing the Mid Loop interface of Ship2-Sam. Biopolymers, 2014, 101(11), 1088-1098.
[19]
Mercurio, F.A.; Costantini, S.; Di Natale, C.; Pirone, L.; Guariniello, S.; Scognamiglio, P.L.; Marasco, D.; Pedone, E.M.; Leone, M. Structural investigation of a C-terminal EphA2 receptor mutant: Does mutation affect the structure and interaction properties of the Sam domain? BBA-Proteins Proteom., 2017, 1865(9), 1095-1104.
[20]
Akitake, B.; Spelbrink, R.E.J.; Anishkin, A.; Killian, J.A.; de Kruijff, B.; Sukharev, S. 2,2,2-Trifluoroethanol changes the transition kinetics and subunit interactions in the small bacterial mechanosensitive channel MscS. Biophys. J., 2007, 92(8), 2771-2784.
[21]
Walgers, R.; Lee, T.C.; Cammers-Goodwin, A. An indirect chaotropic mechanism for the stabilization of helix conformation of peptides in aqueous trifluoroethanol and hexafluoro-2-propanol. J. Am. Chem. Soc., 1998, 120(20), 5073-5079.
[22]
Vijayalakshmi, L.; Krishna, R.; Sankaranarayanan, R.; Vijayan, M. An asymmetric dimer of beta-lactoglobulin in a low humidity crystal form - Structural changes that accompany partial dehydration and protein action. Proteins, 2008, 71(1), 241-249.
[23]
Mercurio, F.A.; Scaloni, A.; Caira, S.; Leone, M. The antimicrobial peptides casocidins I and II: Solution structural studies in water and different membrane-mimetic environments. Peptides, 2018.
[http://dx.doi.org/10.1016/j.peptides.2018.09.004]
[24]
Saxena, V.K.; Kumar, S.; Jha, B.K.; Kumar, A.; Kumar, D.; Naqvi, S.M.K. Study of conformational properties of solid phase synthesized ovine kisspeptin-14 using Circular Dichroism spectroscopy. Indian J. Exp. Biol., 2015, 53(10), 676-680.
[25]
Schonbrunner, N.; Wey, J.; Engels, J.; Georg, H.; Kiefhaber, T. Native-like beta-structure in a trifluoroethanol-induced partially folded state of the all-beta-sheet protein tendamistat. J. Mol. Biol., 1996, 260(3), 432-445.
[26]
Rajan, R.; Balaram, P. A model for the interaction of trifluoroethanol with peptides and proteins. Int. J. Pept. Protein Res., 1996, 48(4), 328-336.
[27]
Sundaralingam, M.; Sekharudu, Y.C. Water-inserted alpha-helical segments implicate reverse turns as folding intermediates. Science, 1989, 244(4910), 1333-1337.
[28]
Dicapua, F.M.; Swaminathan, S.; Beveridge, D.L. Theoretical evidence for destabilization of an alpha-helix by water insertion - molecular-dynamics of hydrated decaalanine. J. Am. Chem. Soc., 1990, 112(19), 6768-6771.
[29]
Deloof, H.; Nilsson, L.; Rigler, R. Molecular-dynamics simulation of galanin in aqueous and nonaqueous solution. J. Am. Chem. Soc., 1992, 114(11), 4028-4035.
[30]
Baker, E.N.; Hubbard, R.E. Hydrogen bonding in globular proteins. Prog. Biophys. Mol. Biol., 1984, 44(2), 97-179.
[31]
Shiraki, K.; Nishikawa, K.; Goto, Y. Trifluoroethanol-induced stabilization of the alpha-helical structure of beta-lactoglobulin - implication for non-hierarchical protein-folding. J. Mol. Biol., 1995, 245(2), 180-194.
[32]
Cruz, A.; Casals, C.; Perez-Gil, J. Conformational flexibility of pulmonary surfactant proteins SP-B and SP-C, studied in aqueous organic solvents. BBA-Lipid. Lipid Met., 1995, 1255(1), 68-76.
[33]
Liu, Z.P.; Rizo, J.; Gierasch, L.M. Equilibrium folding studies of cellular retinoic acid binding protein, a predominantly beta-sheet protein. Biochemistry, 1994, 33(1), 134-142.
[34]
Buck, M.; Radford, S.E.; Dobson, C.M. A partially folded state of hen egg white lysozyme in trifluoroethanol: structural characterization and implications for protein folding. Biochemistry, 1993, 32(2), 669-678.
[35]
Segawa, S.; Fukuno, T.; Fujiwara, K.; Noda, Y. Local structures in unfolded lysozyme and correlation with secondary structures in the native conformation - helix-forming or helix-breaking propensity of peptide segments. Biopolymers, 1991, 31(5), 497-509.
[36]
Fan, P.; Bracken, C.; Baum, J. Structural characterization of monellin in the alcohol-denatured state by NMR: evidence for beta-sheet to alpha-helix conversion. Biochemistry, 1993, 32(6), 1573-1582.
[37]
Vonstosch, A.G.; Kinzel, V.; Pipkorn, R.; Reed, J. Investigation of the structural components governing the polarity-dependent refolding of a Cd4-binding peptide from Gp120. J. Mol. Biol., 1995, 250(4), 507-513.
[38]
Chen, Y.; Liu, B.; Barkley, M.D. Trifluoroethanol quenches indole fluorescence by excited-state proton-transfer. J. Am. Chem. Soc., 1995, 117(20), 5608-5609.
[39]
Imai, T.; Kovalenko, A.; Hirata, F.; Kidera, A. Molecular thermodynamics of trifluoroethanol-induced helix formation: Analysis of the solvation structure and free energy by the 3D-RISM theory. Interdiscip. Sci., 2009, 1(2), 156-160.
[40]
Millhauser, G.L.; Stenland, C.J.; Hanson, P.; Bolin, K.A. vandeVen, F.J.M. Estimating the relative populations of 3(10)-helix and alpha-helix in Ala-rich peptides: A hydrogen exchange and high field NMR study. J. Mol. Biol., 1997, 267(4), 963-974.
[41]
Luo, P.; Baldwin, R.L. Mechanism of helix induction by trifluoroethanol: a framework for extrapolating the helix-forming properties of peptides from trifluoroethanol/water mixtures back to water. Biochemistry, 1997, 36(27), 8413-8421.
[42]
Kentsis, A.; Sosnick, T.R. Trifluoroethanol promotes helix formation by destabilizing backbone exposure: Desolvation rather than native hydrogen bonding defines the kinetic pathway of dimeric coiled coil folding. Biochemistry, 1998, 37(41), 14613-14622.
[43]
Xiong, K.; Asher, S.A. Circular dichroism and UV resonance raman study of the impact of alcohols on the Gibbs free energy landscape of an alpha-helical peptide. Biochemistry, 2010, 49(15), 3336-3342.
[44]
Vymetal, J.; Bednarova, L.; Vondrasek, J. Effect of TFE on the Helical Content of AK17 and HAL-1 Peptides: Theoretical Insights into the Mechanism of Helix Stabilization. J. Phys. Chem. B, 2016, 120(6), 1048-1059.
[45]
Carver, J.A.; Collins, J.G. Nmr identification of a partial helical conformation for bombesin in solution. Eur. J. Biochem., 1990, 187(3), 645-650.
[46]
Anastasi, A.; Erspamer, V.; Bucci, M. Isolation and amino acid sequences of alytesin and bombesin, two analogous active tetradecapeptides from the skin of European discoglossid frogs. Arch. Biochem. Biophys., 1972, 148(2), 443-446.
[47]
Diaz, M.D.; Berger, S. Preferential solvation of a tetrapeptide by trifluoroethanol as studied by intermolecular NOE. Magn. Reson. Chem., 2001, 39(7), 369-373.
[48]
Culik, R.M.; Abaskharon, R.M.; Pazos, I.M.; Gai, F. Experimental validation of the role of trifluoroethanol as a nanocrowder. J. Phys. Chem. B, 2014, 118(39), 11455-11461.
[49]
Gast, K.; Zirwer, D.; Muller-Frohne, M.; Damaschun, G. Trifluoroethanol-induced conformational transitions of proteins: insights gained from the differences between alpha-lactalbumin and ribonuclease A. Protein Sci., 1999, 8(3), 625-634.
[50]
Gast, K.; Siemer, A.; Zirwer, D.; Damaschun, G. Fluoroalcohol-induced structural changes of proteins: some aspects of cosolvent-protein interactions. Eur. Biophys. J., 2001, 30(4), 273-283.
[51]
Scharge, T.; Cezard, C.; Zielke, P.; Schutz, A.; Emmeluth, C.; Suhm, M.A. A peptide co-solvent under scrutiny: Self-aggregation of 2,2,2-trifluoroethanol. Phys. Chem. Chem. Phys., 2007, 9(32), 4472-4490.
[52]
Jalili, S.; Akhavan, M. Molecular dynamics simulation study of association in trifluoroethanol/water mixtures. J. Comput. Chem., 2010, 31(2), 286-294.
[53]
Gerig, J.T. Toward a molecular dynamics force field for simulations of 40% trifluoroethanol-water. J. Phys. Chem. B, 2014, 118(6), 1471-1480.
[54]
Zhou, H.X.; Rivas, G.; Minton, A.P. Macromolecular crowding and confinement: biochemical, biophysical, and potential physiological consequences. Annu. Rev. Biophys., 2008, 37, 375-397.
[55]
Wang, T.; Lau, W.L.; DeGrado, W.F.; Gai, F. T-jump infrared study of the folding mechanism of coiled-coil GCN4-p1. Biophys. J., 2005, 89(6), 4180-4187.
[56]
Reiersen, H.; Rees, A.R. Trifluoroethanol may form a solvent matrix for assisted hydrophobic interactions between peptide side chains. Protein Eng., 2000, 13(11), 739-743.
[57]
Munoz, V.; Serrano, L. Elucidating the folding problem of helical peptides using empirical parameters. Nat. Struct. Biol., 1994, 1(6), 399-409.
[58]
Najbar, L.V.; Craik, D.J.; Wade, J.D.; Salvatore, D.; McLeish, M.J. Conformational analysis of LYS(11-36), a peptide derived from the beta-sheet region of T4 lysozyme, in TFE and SDS. Biochemistry, 1997, 36(38), 11525-11533.
[59]
Wang, Y.J.; Henz, M.E.; Gallagher, N.L.F.; Chai, S.Y.; Gibbs, A.C.; Yan, L.Z.; Stiles, M.E.; Wishart, D.S.; Vederas, J.C. Solution structure of carnobacteriocin B2 and implications for structure-activity relationships among type IIa bacteriocins from lactic acid bacteria. Biochemistry, 1999, 38(47), 15438-15447.
[60]
Starzyk, A.; Barber-Armstrong, W.; Sridharan, M.; Decatur, S.M. Spectroscopic evidence for backbone desolvation of helical peptides by 2,2,2-trifluoroethanol: An isotope-edited FTIR study. Biochemistry, 2005, 44(1), 369-376.
[61]
Monincova, L.; Budesinsky, M.; Slaninova, J.; Hovorka, O.; Cvacka, J.; Voburka, Z.; Fucik, V.; Borovickova, L.; Bednarova, L.; Straka, J.; Cerovsky, V. Novel antimicrobial peptides from the venom of the eusocial bee Halictus sexcinctus (Hymenoptera: Halictidae) and their analogs. Amino Acids, 2010, 39(3), 763-775.
[62]
Sonnichsen, F.D.; Vaneyk, J.E.; Hodges, R.S.; Sykes, B.D. Effect of trifluoroethanol on protein secondary structure - an NMR and CD study using a synthetic actin peptide. Biochemistry, 1992, 31(37), 8790-8798.
[63]
Lehrman, S.R.; Tuls, J.L.; Lund, M. Peptide alpha-helicity in aqueous trifluoroethanol: correlations with predicted alpha-helicity and the secondary structure of the corresponding regions of bovine growth hormone. Biochemistry, 1990, 29(23), 5590-5596.
[64]
Mercurio, F.A.; Marasco, D.; Di Natale, C.; Pirone, L.; Costantini, S.; Pedone, E.M.; Leone, M. Targeting EphA2-Sam and its interactome: design and evaluation of helical peptides enriched in charged residues. ChemBioChem, 2016, 17(22), 2179-2188.
[65]
Zhou, P.; Zhao, H.; Chen, C.; Bai, J.; Wang, D. The stability of alpha-helix of the helical antimicrobial peptide in polar/apolar solvent. Int. J. Biosci. Biochem. Bioinform., 2015, 5(4), 249-255.
[66]
Maroun, R.G.; Krebs, D.; El Antri, S.; Deroussent, A.; Lescot, E.; Troalen, F.; Porumb, H.; Goldberg, M.E.; Fermandjian, S. Self-association and domains of interactions of an amphipathic helix peptide inhibitor of HIV-1 integrase assessed by analytical ultracentrifugation and NMR experiments in trifluoroethanol/H2O mixtures. J. Biol. Chem., 1999, 274(48), 34174-34185.
[67]
Choy, N.; Raussens, V.; Narayanaswami, V. Inter-molecular coiled-coil formation in human apolipoprotein E C-terminal domain. J. Mol. Biol., 2003, 334(3), 527-539.
[68]
Goetz, M.; Carlotti, C.; Bontems, F.; Dufourc, E.J. Evidence for an alpha-helix - pi-bulge helicity modulation for the neu/erbB-2 membrane-spanning segment. A 1H NMR and circular dichroism study. Biochemistry, 2001, 40(21), 6534-6540.
[69]
Roa, J.; Aguilar, E.; Dieguez, C.; Pinilla, L.; Tena-Sempere, M. New frontiers in kisspeptin/GPR54 physiology as fundamental gatekeepers of reproductive function. Front. Neuroendocrinol., 2008, 29(1), 48-69.
[70]
Cornish, V.W.; Kaplan, M.I.; Veenstra, D.L.; Kollman, P.A.; Schultz, P.G. Stabilizing and destabilizing effects of placing beta-branched amino-acids in protein alpha-helices. Biochemistry, 1994, 33(40), 12022-12031.
[71]
Lewandowska, A.; Oldziej, S.; Liwo, A.; Scheraga, H.A. Beta-hairpin-forming peptides; models of early stages of protein folding. Biophys. Chem., 2010, 151(1-2), 1-9.
[72]
Santiveri, C.M.; Pantoja-Uceda, D.; Rico, M.; Jimenez, M.A. Beta-hairpin formation in aqueous solution and in the presence of trifluoroethanol: A H-1 and C-13 nuclear magnetic resonance conformational study of designed peptides. Biopolymers, 2005, 79(3), 150-162.
[73]
Mirtic, A.; Grdadolnik, J. The structure of poly-L-lysine in different solvents. Biophys. Chem., 2013, 175, 47-53.
[74]
Arunkumar, A.I.; Kumar, T.K.S.; Yu, C. Specificity of helix-induction by 2,2,2-trifluoroethanol in polypeptides. Int. J. Biol. Macromol., 1997, 21(3), 223-230.
[75]
Drozdov, A.N.; Grossfield, A.; Pappu, R.V. Role of solvent in determining conformational preferences of alanine dipeptide in water. J. Am. Chem. Soc., 2004, 126(8), 2574-2581.
[76]
Tiffany, M.L.; Krimm, S. New chain conformations of poly(glutamic acid) and polylysine. Biopolymers, 1968, 6(9), 1379-1382.
[77]
Ataei, F.; Hosseinkhani, S. Impact of trifluoroethanol-induced structural changes on luciferase cleavage sites. J. Photochem. Photobiol. B, 2015, 144, 1-7.
[78]
Myers, J.K.; Pace, C.N.; Scholtz, J.M. Trifluoroethanol effects on helix propensity and electrostatic interactions in the helical peptide from ribonuclease T1. Protein Sci., 1998, 7(2), 383-388.
[79]
Rohl, C.A.; Chakrabartty, A.; Baldwin, R.L. Helix propagation and N-cap propensities of the amino acids measured in alanine-based peptides in 40 volume percent trifluoroethanol. Protein Sci., 1996, 5(12), 2623-2637.
[80]
Daoust, H.; St-Cyr, D. Effect of the cation size and of the solvent composition on the conformation of Poly(L-glutamic acid) alkaline metal salts. Polym. J., 1982, 14(11), 831-838.
[81]
Thennarasu, S.; Nagaraj, R. Effects of salt and denaturant on structure of the amino terminal alpha-helical segment of an antibacterial peptide dermaseptin and its binding to model membranes. Indian J. Biochem. Biophys., 2001, 38(3), 142-148.
[82]
Buck, M.; Boyd, J.; Redfield, C.; MacKenzie, D.A.; Jeenes, D.J.; Archer, D.B.; Dobson, C.M. Structural determinants of protein dynamics: analysis of 15N NMR relaxation measurements for main-chain and side-chain nuclei of hen egg white lysozyme. Biochemistry, 1995, 34(12), 4041-4055.
[83]
Povey, J.F.; Smales, C.M.; Hassard, S.J.; Howard, M.J. Comparison of the effects of 2,2,2-trifluoroethanol on peptide and protein structure and function. J. Struct. Biol., 2007, 157(2), 329-338.
[84]
Buck, M.; Schwalbe, H.; Dobson, C.M. Characterization of conformational preferences in a partly folded protein by heteronuclear NMR spectroscopy assignment and secondary structureanalysis of hen egg-white lysozyme in trifluoroethanol. Biochemistry, 1995, 34(40), 13219-13232.
[85]
Williams, M.A.; Thornton, J.M.; Goodfellow, J.M. Modelling protein unfolding: Hen egg-white lysozyme. Protein Eng., 1997, 10(8), 895-903.
[86]
Eyles, S.J.; Radford, S.E.; Robinson, C.V.; Dobson, C.M. Kinetic consequences of the removal of a disulfide bridge on the folding of hen lysozyme. Biochemistry, 1994, 33(44), 13038-13048.
[87]
Radford, S.E.; Dobson, C.M.; Evans, P.A. The folding of hen lysozyme involves partially structured intermediates and multiple pathways. Nature, 1992, 358(6384), 302-307.
[88]
Liu, F.T.; Patterson, R.J.; Wang, J.L. Intracellular functions of galectins. BBA-Gen. Subjects, 2002, 1572(2-3), 263-273.
[89]
Mandal, P.; Molla, A.R.; Mandal, D.K. Denaturation of bovine spleen galectin-1 in guanidine hydrochloride and fluoroalcohols: Structural characterization and implications for protein folding. J. Biochem., 2013, 154(6), 531-540.
[90]
Wright, P.E.; Dyson, H.J. Intrinsically unstructured proteins: Re-assessing the protein structure-function paradigm. J. Mol. Biol., 1999, 293(2), 321-331.
[91]
Iakoucheva, L.M.; Brown, C.J.; Lawson, J.D.; Obradovic, Z.; Dunker, A.K. Intrinsic disorder in cell-signaling and cancer-associated proteins. J. Mol. Biol., 2002, 323(3), 573-584.
[92]
Liu, J.G.; Perumal, N.B.; Oldfield, C.J.; Su, E.W.; Uversky, V.N.; Dunker, A.K. Intrinsic disorder in transcription factors. Biochemistry, 2006, 45(22), 6873-6888.
[93]
Galea, C.A.; Wang, Y.; Sivakolundu, S.G.; Kriwacki, R.W. Regulation of cell division by intrinsically unstructured proteins: Intrinsic flexibility, modularity, and signaling conduits. Biochemistry, 2008, 47(29), 7598-7609.
[94]
Uversky, V.N.; Oldfield, C.J.; Dunker, A.K. Intrinsically disordered proteins in human diseases: Introducing the D-2 concept. Annu. Rev. Biophys., 2008, 37, 215-246.
[95]
Uversky, V.N. Intrinsically disordered proteins from A to Z. Int. J. Biochem. Cell Biol., 2011, 43(8), 1090-1103.
[96]
Singh, G.P.; Dash, D. Intrinsic disorder in yeast transcriptional regulatory network. Proteins, 2007, 68(3), 602-605.
[97]
Habchi, J.; Tompa, P.; Longhi, S.; Uversky, V.N. Introducing protein intrinsic disorder. Chem. Rev., 2014, 114(13), 6561-6588.
[98]
Hamdi, K.; Salladini, E.; O’Brien, D.P.; Brier, S.; Chenal, A.; Yacoubi, I.; Longhi, S. Structural disorder and induced folding within two cereal, ABA stress and ripening (ASR) proteins. Sci. Rep., 2017, 7(1), 15544.
[99]
Sun, X.L.; Rikkerink, E.H.A.; Jones, W.T.; Uversky, V.N. Multifarious roles of intrinsic disorder in proteins illustrate its broad impact on plant biology. Plant Cell, 2013, 25(1), 38-55.
[100]
Fontana, A.; de Laureto, P.P.; Spolaore, B.; Frare, E.; Picotti, P.; Zambonin, M. Probing protein structure by limited proteolysis. Acta Biochim. Pol., 2004, 51(2), 299-321.
[101]
Tompa, P. Intrinsically unstructured proteins. Trends Biochem. Sci., 2002, 27(10), 527-533.
[102]
Receveur-Brechot, V.; Bourhis, J.M.; Uversky, V.N.; Canard, B.; Longhi, S. Assessing protein disorder and induced folding. Proteins, 2006, 62(1), 24-45.
[103]
Kaczka, P.; Winiewska, M.; Zhukov, I.; Rempola, B.; Bolewska, K.; Lozinski, T.; Ejchart, A.; Poznanska, A.; Wierzchowski, K.L.; Poznanski, J. The TFE-induced transient native-like structure of the intrinsically disordered domain of Escherichia coli RNA polymerase. Eur. Biophys. J., 2014, 43(12), 581-594.
[104]
Duvignaud, J.B.; Savard, C.; Fromentin, R.; Majeau, N.; Leclerc, D.; Gagne, S.M. Structure and dynamics of the N-terminal half of hepatitis C virus core protein: An intrinsically unstructured protein. Biochem. Biophys. Res. Commun., 2009, 378(1), 27-31.
[105]
Majeau, N.; Gagne, V.; Boivin, A.; Bolduc, M.; Majeau, J.A.; Ouellet, D.; Leclerc, D. The N-terminal half of the core protein of hepatitis C virus is sufficient for nucleocapsid formation. J. Gen. Virol., 2004, 85(Pt. 4), 971-981.
[106]
Chen, C.M.; You, L.R.; Hwang, L.H.; Lee, Y.H.W. Direct interaction of hepatitis C virus core protein with the cellular lymphotoxin-beta receptor modulates the signal pathway of the lymphotoxin-beta receptor. J. Virol., 1997, 71(12), 9417-9426.
[107]
Mamiya, N.; Worman, H.J. Hepatitis C virus core protein binds to a DEAD box RNA helicase. J. Biol. Chem., 1999, 274(22), 15751-15756.
[108]
Wang, F.; Yoshida, I.; Takamatsu, M.; Ishido, S.; Fujita, T.; Oka, K.; Hotta, H. Complex formation between hepatitis C virus core protein and p21Waf1/Cip1/Sdi1. Biochem. Biophys. Res. Commun., 2000, 273(2), 479-484.
[109]
Duvignaud, J.B.; Leclerc, D.; Gagne, S.M. Structure and dynamics changes induced by 2,2,2-trifluoro-ethanol (TFE) on the N-terminal half of hepatitis C virus core protein. Biochem. Cell Biol., 2010, 88(2), 315-323.
[110]
Muller, I.; Sarramegna, V.; Milon, A.; Talmont, F.J. The N-terminal end truncated mu-opioid receptor: from expression to circular dichroism analysis. Appl. Biochem. Biotechnol., 2010, 160(7), 2175-2186.
[111]
Moncoq, K.; Broutin, I.; Craescu, C.T.; Vachette, P.; Ducruix, A.; Durand, D. SAXS study of the PIR domain from the Grb14 molecular adaptor: a natively unfolded protein with a transient structure primer? Biophys. J., 2004, 87(6), 4056-4064.
[112]
Boulant, S.; Vanbelle, C.; Ebel, C.; Penin, F.; Lavergne, J.P. Hepatitis C virus core protein is a dimeric alpha-helical protein exhibiting membrane protein features. J. Virol., 2005, 79(17), 11353-11365.
[113]
Rodriguez-Casado, A.; Molina, M.; Carmona, P. Core protein-nucleic acid interactions in hepatitis C virus as revealed by Raman and circular dichroism spectroscopy. Appl. Spectrosc., 2007, 61(11), 1219-1224.
[114]
Tantos, A.; Szrnka, K.; Szabo, B.; Bokor, M.; Kamasa, P.; Matus, P.; Bekesi, A.; Tompa, K.; Han, K.H.; Tompa, P. Structural disorder and local order of hNopp140. BBA-Proteins Proteom., 2013, 1834(1), 342-350.
[115]
Luo, P.Z.; Baldwin, R.L. Mechanism of helix induction by trifluoroethanol: A framework for extrapolating the helix-forming properties of peptides from trifluoroethanol/water mixtures back to water. Biochemistry, 1997, 36(27), 8413-8421.
[116]
Chiti, F.; Dobson, C.M. Protein misfolding, functional amyloid, and human disease. Annu. Rev. Biochem., 2006, 75, 333-366.
[117]
Nelson, R.; Sawaya, M.R.; Balbirnie, M.; Madsen, A.O.; Riekel, C.; Grothe, R.; Eisenberg, D. Structure of the cross-beta spine of amyloid-like fibrils. Nature, 2005, 435(7043), 773-778.
[118]
Bhak, G.; Choe, Y.J.; Paik, S.R. Mechanism of amyloidogenesis: nucleation-dependent fibrillation versus double-concerted fibrillation. BMB Rep., 2009, 42(9), 541-551.
[119]
Abedini, A.; Raleigh, D.P. A role for helical intermediates in amyloid formation by natively unfolded polypeptides? Phys. Biol., 2009, 6(1), 015005.
[120]
Selkoe, D.J. Folding proteins in fatal ways. Nature, 2003, 426(6968), 900-904.
[121]
Abedini, A.; Raleigh, D.P. A critical assessment of the role of helical intermediates in amyloid formation by natively unfolded proteins and polypeptides. Protein Eng. Des. Sel., 2009, 22(8), 453-459.
[122]
Tycko, R. Progress towards a molecular-level structural understanding of amyloid fibrils. Curr. Opin. Struct. Biol., 2004, 14(1), 96-103.
[123]
Petkova, A.T.; Ishii, Y.; Balbach, J.J.; Antzutkin, O.N.; Leapman, R.D.; Delaglio, F.; Tycko, R. A structural model for Alzheimer’s beta-amyloid fibrils based on experimental constraints from solid state NMR. Proc. Natl. Acad. Sci. USA, 2002, 99(26), 16742-16747.
[124]
Santambrogio, C.; Ricagno, S.; Sobott, F.; Colombo, M.; Bolognesi, M.; Grandori, R. Characterization of beta 2-microglobulin conformational intermediates associated to different fibrillation conditions. J. Mass Spectrom., 2011, 46(8), 734-741.
[125]
Morris, A.M.; Watzky, M.A.; Finke, R.G. Protein aggregation kinetics, mechanism, and curve-fitting: A review of the literature. BBA-Proteins Proteom., 2009, 1794(3), 375-397.
[126]
Ban, T.; Yamaguchi, K.; Goto, Y. Direct observation of amyloid fibril growth, propagation, and adaptation. Acc. Chem. Res., 2006, 39(9), 663-670.
[127]
Xi, W.H.; Wei, G.H. Amyloid-beta peptide aggregation and the influence of carbon nanoparticles. Chin. Phys. B, 2016, 25(1), 18704-018704.
[128]
Dammers, C.; Gremer, L.; Reiss, K.; Klein, A.N.; Neudecker, P.; Hartmann, R.; Sun, N.; Demuth, H.U.; Schwarten, M.; Willbold, D. Structural analysis and aggregation propensity of pyroglutamate A beta(3-40) in aqueous trifluoroethanol. PLoS One, 2015, 10(11), e0143647.
[129]
O’Brien, R.J.; Wong, P.C. Amyloid precursor protein processing and Alzheimer’s disease. Annu. Rev. Neurosci., 2011, 34, 185-204.
[130]
Sticht, H.; Bayer, P.; Willbold, D.; Dames, S.; Hilbich, C.; Beyreuther, K.; Frank, R.W.; Rosch, P. Structure of amyloid A4-(1-40)-peptide of Alzheimer’s disease. Eur. J. Biochem., 1995, 233(1), 293-298.
[131]
Sun, N.; Hartmann, R.; Lecher, J.; Stoldt, M.; Funke, S.A.; Gremer, L.; Ludwig, H.H.; Demuth, H.U.; Kleinschmidt, M.; Willbold, D. Structural analysis of the pyroglutamate-modified isoform of the Alzheimer’s disease-related amyloid-beta using NMR spectroscopy. J. Pept. Sci., 2012, 18(11), 691-695.
[132]
Anderson, V.L.; Ramlall, T.F.; Rospigliosi, C.C.; Webb, W.W.; Eliezer, D. Identification of a helical intermediate in trifluoroethanol-induced alpha-synuclein aggregation. Proc. Natl. Acad. Sci. USA, 2010, 107(44), 18850-18855.
[133]
Spillantini, M.G. Parkinson’s disease, dementia with Lewy bodies and multiple system atrophy are alpha-synucleinopathies. Parkinsonism Relat. Disord., 1999, 5(4), 157-162.
[134]
Eliezer, D.; Kutluay, E.; Bussell, R.; Browne, G. Conformational properties of alpha-synuclein in its free and lipid-associated states. J. Mol. Biol., 2001, 307(4), 1061-1073.
[135]
Khan, M.S.; Tabrez, S.; Bhat, S.A.; Rabbani, N.; Al-Senaidy, A.M.; Bano, B. Effect of trifluoroethanol on alpha-crystallin: folding, aggregation, amyloid, and cytotoxicity analysis. J. Mol. Recognit., 2016, 29(1), 33-40.
[136]
Ehrnsperger, M.; Graber, S.; Gaestel, M.; Buchner, J. Binding of non-native protein to Hsp25 during heat shock creates a reservoir of folding intermediates for reactivation. EMBO J., 1997, 16(2), 221-229.
[137]
Lee, G.J.; Roseman, A.M.; Saibil, H.R.; Vierling, E. A small heat shock protein stably binds heat-denatured model substrates and can maintain a substrate in a folding-competent state. EMBO J., 1997, 16(3), 659-671.
[138]
Kumar, M.S.; Reddy, P.Y.; Kumar, P.A.; Surolia, I.; Reddy, G.B. Effect of dicarbonyl-induced browning on alpha-crystallin chaperone-like activity: physiological significance and caveats of in vitro aggregation assays. Biochem. J., 2004, 379(Pt 2), 273-282.
[139]
Ehrnsperger, M.; Hergersberg, C.; Wienhues, U.; Nichtl, A.; Buchner, J. Stabilization of proteins and peptides in diagnostic immunological assays by the molecular chaperone Hsp25. Anal. Biochem., 1998, 259(2), 218-225.
[140]
Hollmann, A.; Martinez, M.; Maturana, P.; Semorile, L.C.; Maffia, P.C. Antimicrobial peptides: interaction with model and biological membranes and synergism with chemical antibiotics. Front Chem., 2018, 6, 204.
[141]
da Costa, J.P.; Cova, M.; Ferreira, R.; Vitorino, R. Antimicrobial peptides: An alternative for innovative medicines? Appl. Microbiol. Biotechnol., 2015, 99(5), 2023-2040.
[142]
Brogden, K.A. Antimicrobial peptides: Pore formers or metabolic inhibitors in bacteria? Nat. Rev. Microbiol., 2005, 3(3), 238-250.
[143]
Wimley, W.C. Describing the mechanism of antimicrobial peptide action with the interfacial activity model. ACS Chem. Biol., 2010, 5(10), 905-917.
[144]
Haney, E.F.; Vogel, H.J. NMR of antimicrobial peptides. Annu. Rep. NMR Spectrosc., 2009, 65, 1-51.
[145]
Tack, B.F.; Sawai, M.V.; Kearney, W.R.; Robertson, A.D.; Sherman, M.A.; Wang, W.; Hong, T.; Lee, M.B.; Wu, H.; Waring, A.J.; Lehrer, R.I. SMAP-29 has two LPS-binding sites and a central hinge. Eur. J. Biochem., 2002, 269(4), 1181-1189.
[146]
Sawai, M.V.; Waring, A.J.; Kearney, W.R.; McCray, P.B.; Forsyth, W.R.; Lehrer, R.I.; Tack, B.F. Impact of single-residue mutations on the structure and function of ovispirin/novispirin antimicrobial peptides. Protein Eng., 2002, 15(3), 225-232.
[147]
Chen, C.P.; Brock, R.; Luh, F.; Chou, P.J.; Larrick, J.W.; Huang, R.F.; Huang, T.H. The solution structure of the active domain of Cap18 - a lipopolysaccharide-binding protein from rabbit leukocytes. FEBS Lett., 1995, 370(1-2), 46-52.
[148]
Uteng, M.; Hauge, H.H.; Markwick, P.R.; Fimland, G.; Mantzilas, D.; Nissen-Meyer, J.; Muhle-Goll, C. Three-dimensional structure in lipid micelles of the pediocin-like antimicrobial peptide sakacin P and a sakacin P variant that is structurally stabilized by an inserted C-terminal disulfide bridge. Biochemistry, 2003, 42(39), 11417-11426.
[149]
Landon, C.; Meudal, H.; Boulanger, N.; Bulet, P.; Vovelle, F. Solution structures of stomoxyn and spinigerin, two insect antimicrobial peptides with an alpha-helical conformation. Biopolymers, 2006, 81(2), 92-103.
[150]
Xiao, Y.; Dai, H.; Bommineni, Y.R.; Soulages, J.L.; Gong, Y.X.; Prakash, O.; Zhang, G. Structure-activity relationships of fowlicidin-1, a cathelicidin antimicrobial peptide in chicken. FEBS J., 2006, 273(12), 2581-2593.
[151]
Bommineni, Y.R.; Dai, H.E.; Gong, Y.X.; Soulages, J.L.; Fernando, S.C.; DeSilva, U.; Prakash, O.; Zhang, G.L. Fowlicidin-3 is an alpha-helical cationic host defense peptide with potent antibacterial and lipopolysaccharide-neutralizing activities. FEBS J., 2007, 274(2), 418-428.
[152]
Rogne, P.; Fimland, G.; Nissen-Meyer, J.; Kristiansen, P.E. Threedimensional structure of the two peptides that constitute the twopeptide bacteriocin lactococcin G. BBA-Proteins Proteom, 2008, 1784 (3), 543-554.
[153]
Verly, R.M.; de Moraes, C.M.; Resende, J.M.; Aisenbrey, C.; Bernquerer, M.P.; Pilo-Veloso, D.; Valente, A.P.; Almeida, F.C.L.; Bechinger, B. Structure and membrane interactions of the antibiotic peptide dermadistinctin K by multidimensional solution and oriented N-15 and P-31 solid-state NMR spectroscopy. Biophys. J., 2009, 96(6), 2194-2203.
[154]
Subasinghage, A.R.; Conlon, J.M.; Hewage, C.M. Conformational analysis of the broad-spectrum antibacterial peptide, ranatuerin-2CSa: Identification of a full length helix-turn-helix motif. BBA-Proteins Proteom., 2008, 1784(6), 924-929.
[155]
Rogne, P.; Haugen, C.; Fimland, G.; Nissen-Meyer, J.; Kristiansen, P.E. Three-dimensional structure of the two-peptide bacteriocin plantaricin JK. Peptides, 2009, 30(9), 1613-1621.
[156]
Gao, B.; Xu, J.; Rodriguez, M.D.; Lanz-Mendoza, H.; Hernandez-Rivas, R.; Du, W.H.; Zhu, S.Y. Characterization of two linear cationic antimalarial peptides in the scorpion Mesobuthus eupeus. Biochimie, 2010, 92(4), 350-359.
[157]
Fregeau Gallagher, N.L.; Sailer, M.; Niemczura, W.P.; Nakashima, T.T.; Stiles, M.E.; Vederas, J.C. Three-dimensional structure of leucocin A in trifluoroethanol and dodecylphosphocholine micelles: Spatial location of residues critical for biological activity in type IIa bacteriocins from lactic acid bacteria. Biochemistry, 1997, 36(49), 15062-15072.
[158]
Ovchinnikov, K.V.; Kristiansen, P.E.; Uzelac, G.; Topisirovic, L.; Kojic, M.; Nissen-Meyer, J.; Nes, I.F.; Diep, D.B. Defining the structure and receptor binding domain of the leaderless bacteriocin LsbB. J. Biol. Chem., 2014, 289(34), 23838-23845.
[159]
Godreuil, S.; Leban, N.; Padilla, A.; Hamel, R.; Luplertlop, N.; Chauffour, A.; Vittecoq, M.; Hoh, F.; Thomas, F.; Sougakoff, W.; Lionne, C.; Yssel, H.; Misse, D. Aedesin: structure and antimicrobial activity against multidrug resistant bacterial strains. PLoS One, 2014, 9(8), e105441.
[160]
Lopez-Abarrategui, C.; McBeth, C.; Mandal, S.M.; Sun, Z.Y.J.; Heffron, G.; Alba-Menendez, A.; Migliolo, L.; Reyes-Acosta, O.; Garcia-Villarino, M.; Nolasco, D.O.; Falcao, R.; Cherobim, M.D.; Dias, S.C.; Brandt, W.; Wessjohann, L.; Starnbach, M.; Franco, O.L.; Otero-Gonzalez, A.J. Cm-p5: An antifungal hydrophilic peptide derived from the coastal mollusk Cenchritis muricatus (Gastropoda: Littorinidae). FASEB J., 2015, 29(8), 3315-3325.
[161]
Arbulu, S.; Lohans, C.T.; van Belkum, M.J.; Cintas, L.M.; Herranz, C.; Vederas, J.C.; Hernandez, P.E. Solution structure of enterocin HF, an antilisterial bacteriocin produced by enterococcus faecium M3K31. J. Agric. Food Chem., 2015, 63(49), 10689-10695.
[162]
Acedo, J.Z.; van Belkum, M.J.; Lohans, C.T.; Towle, K.M.; Miskolzie, M.; Vederas, J.C. Nuclear magnetic resonance solution structures of lacticin Q and aureocin A53 reveal a structural motif conserved among leaderless bacteriocins with broad-spectrum activity. Biochemistry, 2016, 55(4), 733-742.
[163]
Ovchinnikov, K.V.; Kristiansen, P.E.; Straume, D.; Jensen, M.S.; Aleksandrzak-Piekarczyk, T.; Nes, I.F.; Diep, D.B. The leaderless bacteriocin enterocin K1 is highly potent against enterococcus faecium: A study on structure, target spectrum and receptor. Front. Microbiol., 2017, 8, 774.
[164]
Gomes, K.A.G.G.; dos Santos, D.M.; Santos, V.M.; Pilo-Veloso, D.; Mundim, H.M.; Rodrigues, L.V.; Liao, L.M.; Verly, R.M.; de Lima, M.E.; Resende, J.M. NMR structures in different membrane environments of three ocellatin peptides isolated from Leptodactylus labyrinthicus. Peptides, 2018, 103, 72-83.
[165]
Silva, O.N.; Alves, E.S.F.; de la Fuente-Nunez, C.; Ribeiro, S.M.; Mandal, S.M.; Gaspar, D.; Veiga, A.S.; Castanho, M.A.R.B.; Andrade, C.A.S.; Nascimento, J.M.; Fensterseifer, I.C.M.; Porto, W.F.; Correa, J.R.; Hancock, R.E.W.; Korpole, S.; Oliveira, A.L.; Liao, L.M.; Franco, O.L. Structural studies of a lipid-binding peptide from tunicate hemocytes with anti-biofilm activity. Sci. Rep., 2016, 6, 27128.
[166]
Scorciapino, M.A.; Rinaldi, A.C. Antimicrobial peptidomimetics: reinterpreting nature to deliver innovative therapeutics. Front. Immunol., 2012, 3, 171.
[167]
Gaglione, R.; Dell’Olmo, E.; Bosso, A.; Chino, M.; Pane, K.; Ascione, F.; Itri, F.; Caserta, S.; Amoresano, A.; Lombardi, A.; Haagsman, H.P.; Piccoli, R.; Pizzo, E.; Veldhuizen, E.J.A.; Notomista, E.; Arciello, A. Novel human bioactive peptides identified in Apolipoprotein B: Evaluation of their therapeutic potential. Biochem. Pharmacol., 2017, 130, 34-50.
[168]
Cui, P.F.; Dong, Y.; Li, Z.J.; Zhang, Y.B.; Zhang, S.C. Identification and functional characterization of an uncharacterized antimicrobial peptide from a ciliate Paramecium caudatum. Dev. Comp. Immunol., 2016, 60, 53-65.
[169]
Pizzo, E.; Pane, K.; Bosso, A.; Landi, N.; Ragucci, S.; Russo, R.; Gaghone, R.; Torres, M.D.T.; de la Fuente-Nunez, C.; Arciello, A.; Di Donato, A.; Notomista, E.; Di Maro, A. Novel bioactive peptides from PD-L1/2, a type 1 ribosome inactivating protein from Phytolacca dioica L. Evaluation of their antimicrobial properties and anti-biofilm activities. BBA-Biomembranes, 2018, 1860(7), 1425-1435.
[170]
Liu, B.; Huang, H.; Yang, Z.; Liu, B.; Gou, S.; Zhong, C.; Han, X.; Zhang, Y.; Ni, J.; Wang, R. Design of novel antimicrobial peptide dimer analogues with enhanced antimicrobial activity in vitro and in vivo by intermolecular triazole bridge strategy. Peptides, 2017, 88, 115-125.
[171]
Salas, R.L.; Garcia, J.K.D.L.; Miranda, A.C.R.; Rivera, W.L.; Nellas, R.B.; Sabido, P.M.G. Effects of truncation of the peptide chain on the secondary structure and bioactivities of palmitoylated anoplin. Peptides, 2018, 104, 7-14.
[172]
Huang, Y.B.; He, L.Y.; Li, G.R.; Zhai, N.C.; Jiang, H.Y.; Chen, Y.X. Role of helicity of alpha-helical antimicrobial peptides to improve specificity. Protein Cell, 2014, 5(8), 631-642.
[173]
Harris, F.; Dennison, S.R.; Phoenix, D.A. Anionic antimicrobial peptides from eukaryotic organisms. Curr. Protein Pept. Sci., 2009, 10(6), 585-606.
[174]
Dashper, S.G.; Liu, S.W.; Reynolds, E.C. Antimicrobial peptides and their potential as oral therapeutic agents. Int. J. Pept. Res. Ther., 2007, 13(4), 505-516.
[175]
Malkoski, M.; Dashper, S.G.; O’Brien-Simpson, N.M.; Talbo, G.H.; Macris, M.; Cross, K.J.; Reynolds, E.C. Kappacin, a novel antibacterial peptide from bovine milk. Antimicrob. Agents Chemother., 2001, 45(8), 2309-2315.
[176]
Dashper, S.G.; O’Brien-Simpson, N.M.; Cross, K.J.; Paolini, R.A.; Hoffmann, B.; Catmull, D.V.; Malkoski, M.; Reynolds, E.C. Divalent metal cations increase the activity of the antimicrobial Peptide kappacin. Antimicrob. Agents Chemother., 2005, 49(6), 2322-2328.
[177]
Plowman, J.E.; Creamer, L.K.; Liddell, M.J.; Cross, J.J. Solution conformation of a peptide corresponding to bovine kappa-casein B residues 130-153 by circular dichroism spectroscopy and 1H-nuclear magnetic resonance spectroscopy. J. Dairy Res., 1997, 64(3), 377-397.
[178]
Boucher, L.E.; Lopez, D.D.C.; Miller, A.S.; Stamm, S.M.; Bosch, J. Targeting protein-protein-interactions for antimalarial drug discovery. Biophys. J., 2015, 108(2), 148.
[179]
Amartely, H.; Iosub-Amir, A.; Friedler, A. Identifying protein-protein interaction sites using peptide arrays. J. Vis. Exp., 2014, (93), 52097.
[180]
Reymond, M.T.; Merutka, G.; Dyson, H.J.; Wright, P.E. Folding propensities of peptide fragments of myoglobin. Protein Sci., 1997, 6(3), 706-716.
[181]
Mercurio, F.A.; Di Natale, C.; Pirone, L.; Iannitti, R.; Marasco, D.; Pedone, E.M.; Palumbo, R.; Leone, M. The Sam-Sam interaction between Ship2 and the EphA2 receptor: Design and analysis of peptide inhibitors. Sci. Rep., 2017, 7(1), 17474.
[182]
Kim, C.A.; Bowie, J.U. SAM domains: Uniform structure, diversity of function. Trends Biochemy. Sci., 2003, 28(12), 625-628.
[183]
Yang, N.Y.; Fernandez, C.; Richter, M.; Xiao, Z.; Valencia, F.; Tice, D.A.; Pasquale, E.B. Crosstalk of the EphA2 receptor with a serine/threonine phosphatase suppresses the Akt-mTORC1 pathway in cancer cells. Cell. Signal., 2011, 23(1), 201-212.
[184]
Miao, H.; Li, D.Q.; Mukherjee, A.; Guo, H.; Petty, A.; Cutter, J.; Basilion, J.P.; Sedor, J.; Wu, J.; Danielpour, D.; Sloan, A.E.; Cohen, M.L.; Wang, B. EphA2 mediates ligand-dependent inhibition and ligand-independent promotion of cell migration and invasion via a reciprocal regulatory loop with Akt. Cancer Cell, 2009, 16(1), 9-20.
[185]
Zhuang, G.L.; Hunter, S.; Hwang, Y.; Chen, J. Regulation of EphA2 receptor endocytosis by SHIP2 lipid phosphatase via phosphatidylinositol 3-kinase-dependent Rac1 activation. J. Biol. Chem., 2007, 282(4), 2683-2694.
[186]
Leone, M.; Cellitti, J.; Pellecchia, M. NMR studies of a heterotypic Sam-Sam domain association: the interaction between the lipid phosphatase Ship2 and the EphA2 receptor. Biochemistry, 2008, 47(48), 12721-12728.
[187]
Singh, D.R.; Ahmed, F.; Paul, M.D.; Gedam, M.; Pasquale, E.B.; Hristova, K. The SAM domain inhibits EphA2 interactions in the plasma membrane. BBA-Mol. Cell Res., 2017, 1864(1), 31-38.
[188]
Shi, X.; Hapiak, V.; Zheng, J.; Muller-Greven, J.; Bowman, D.; Lingerak, R.; Buck, M.; Wang, B.C.; Smith, A.W. A role of the SAM domain in EphA2 receptor activation. Sci. Rep., 2017, 7, 45084.
[189]
Lee, H.J.; Hota, P.K.; Chugha, P.; Guo, H.; Miao, H.; Zhang, L.Q.; Kim, S.J.; Stetzik, L.; Wang, B.C.; Buck, M. NMR structure of a heterodimeric SAM:SAM complex: Characterization and manipulation of EphA2 binding reveal new cellular functions of SHIP2. Structure, 2012, 20(1), 41-55.
[190]
Mercurio, F.A.; Marasco, D.; Pirone, L.; Pedone, E.M.; Pellecchia, M.; Leone, M. Solution structure of the first Sam domain of Odin and binding studies with the EphA2 receptor. Biochemistry, 2012, 51(10), 2136-2145.
[191]
Kim, J.; Lee, H.; Kim, Y.; Yoo, S.; Park, E.; Park, S. The SAM domains of Anks family proteins are critically involved in modulating the degradation of EphA receptors. Mol. Cell. Biol., 2010, 30(7), 1582-1592.
[192]
Wang, Y.; Shang, Y.; Li, J.C.; Chen, W.D.; Li, G.; Wan, J.; Liu, W.; Zhang, M.J. Specific Eph receptor-cytoplasmic effector signaling mediated by SAM-SAM domain interactions. eLife, 2018, 7, e35677.
[193]
Neira, J.L. Structural dissection of the C-terminal sterile alpha motif (SAM) of human p73. Arch. Biochem. Biophys., 2014, 558, 133-142.
[194]
Chi, S.W.; Ayed, A.; Arrowsmith, C.H. Solution structure of a conserved C-terminal domain of p73 with structural homology to the SAM domain. EMBO J., 1999, 18(16), 4438-4445.
[195]
Mercurio, F.A.; Pirone, L.; Di Natale, C.; Marasco, D.; Pedone, E.M.; Leone, M. Sam domain-based stapled peptides: Structural analysis and interaction studies with the Sam domains from the EphA2 receptor and the lipid phosphatase Ship2. Bioorg. Chem., 2018, 80, 602-610.
[196]
Mercurio, F.A.; Leone, M. The Sam domain of EphA2 receptor and its relevance to cancer: A novel challenge for drug discovery? Curr. Med. Chem., 2016, 23(42), 4718-4734.
[197]
Rothemund, S.; Weisshoff, H.; Beyermann, M.; Krause, E.; Bienert, M.; Mugge, C.; Sykes, B.D.; Sonnichsen, F.D. Temperature coefficients of amide proton NMR resonance frequencies in trifluoroethanol: A monitor of intramolecular hydrogen bonds in helical peptides. J. Biomol. NMR, 1996, 8(1), 93-97.

Rights & Permissions Print Cite
© 2024 Bentham Science Publishers | Privacy Policy