Abstract

•Single-molecule force spectroscopy probes the mechanics of 1-nm-size synthetic helices•Outstanding elasticity and rewinding capabilities of aromatic amide helices•Ultrafast nondissipative winding along precise trajectories•Reducing molecular size does not compromise mechanical properties Chemistry excels in designing and controlling molecular structures. Beyond structure, programming molecular dynamics is one of the next grand challenges. The conformation dynamics of natural elastomeric proteins, for example, are responsible for their ability to behave either as pure elastic spring-like molecules that can be reversibly stretched or as shock absorbers that dissipate energy. Here, we report the outstanding mechanical performances of synthetic helical molecules much smaller than proteins. The molecules were designed for interfacing with AFM force spectroscopy to study them one at a time, a true performance for such tiny objects. We show that the elastic response of helices as small as 1 nm is among the fastest and the most robust ever described. This unprecedented elastic behavior suggests that various applications may arise from their use as building blocks in molecular machines or in new classes of artificial elastomeric materials with tailored mechanical properties. Because of proteins’ many degrees of conformational freedom, programming protein folding dynamics, overall elasticity, and motor functions remains an elusive objective. Instead, smaller and simpler objects, such as synthetic foldamers, may be amenable to design. However, little is known about their mechanical performance. Here, we show that reducing molecular size may not compromise mechanical properties. We report that helical aromatic oligoamides as small as 1 nm possess outstanding elasticity and outperform most natural helices. Using single-molecule force spectroscopy, we characterize their folding trajectories and intermediate states. We show that they cooperatively and reversibly unwind at high forces. They extend up to 3.8 times their original length and rewind against considerable forces on a timescale of 10 μs. Pulling and relaxing cycles follow the same trace up to a very high loading rate, indicating that the mechanical energy accumulated during the stretching does not dissipate and is immediately reusable. Because of proteins’ many degrees of conformational freedom, programming protein folding dynamics, overall elasticity, and motor functions remains an elusive objective. Instead, smaller and simpler objects, such as synthetic foldamers, may be amenable to design. However, little is known about their mechanical performance. Here, we show that reducing molecular size may not compromise mechanical properties. We report that helical aromatic oligoamides as small as 1 nm possess outstanding elasticity and outperform most natural helices. Using single-molecule force spectroscopy, we characterize their folding trajectories and intermediate states. We show that they cooperatively and reversibly unwind at high forces. They extend up to 3.8 times their original length and rewind against considerable forces on a timescale of 10 μs. Pulling and relaxing cycles follow the same trace up to a very high loading rate, indicating that the mechanical energy accumulated during the stretching does not dissipate and is immediately reusable. Folding is the process by which the three-dimensional shapes and properties of proteins are generated from linear peptide chains. Folding is reversible and allows for some important dynamics to take place: molecular shape changes determine mechanical and motor functions in living systems. For example, reversible conformational changes in titin and elastin have been associated with tissue elasticity.1Labeit D. Watanabe K. Witt C. Fujita H. Wu Y. Lahmers S. Funck T. Labeit S. Granzier H. Calcium-dependent molecular spring elements in the giant protein titin.Proc. Natl. Acad. Sci. USA. 2003; 100: 13716-13721Crossref PubMed Scopus (287) Google Scholar,2Javadi Y. Fernandez J.M. Perez-Jimenez R. Protein folding under mechanical forces: a physiological view.Physiology. 2013; 28: 9-17Crossref PubMed Scopus (26) Google Scholar Conversely, the stiffness of collagen is responsible for tissue resistance.3Bao G. Protein mechanics: a new frontier in biomechanics.Exp. Mech. 2009; 49: 153-164Crossref PubMed Scopus (17) Google Scholar While there has been a considerable progress in the design of new protein folds,4Kuhlman B. Bradley P. Advances in protein structure prediction and design.Nat. Rev. Mol. Cell Biol. 2019; 20: 681-697Crossref PubMed Scopus (180) Google Scholar programming protein folding dynamics has remained an elusive objective because of the co-existence of multiple folding trajectories and of local energy minima, which slow down the process and compromise refolding on a short time scale. Smaller folded objects than proteins may provide entries into simpler folding dynamics and thus amenability to design. For example, the peptidic α-helix and nucleic acid helices have both been investigated in detail and show different behaviors intimately associated with their distinct chemical nature.5Hoffmann T. Dougan L. Single molecule force spectroscopy using polyproteins.Chem. Soc. Rev. 2012; 41: 4781-4796Crossref PubMed Scopus (117) Google Scholar, 6Camunas-Soler J. Ribezzi-Crivellari M. Ritort F. Elastic properties of nucleic acids by single-molecule force spectroscopy.Annu. Rev. Biophys. 2016; 45: 65-84Crossref PubMed Scopus (37) Google Scholar, 7Janshoff A. Neitzert M. Oberdörfer Y. Fuchs H. Force spectroscopy of molecular systems-single moleculespectroscopy of polymers and biomolecules.Angew. Chem. Int. Ed. Engl. 2000; 39: 3212-3237Crossref PubMed Google Scholar For these studies, AFM-based single-molecule force spectroscopy has emerged as a powerful method to probe molecular-level processes and mechanical forces with sub-nanometer resolution in (bio)-macromolecules.8Fisher T.E. Marszalek P.E. Fernandez J.M. Stretching single molecules into novel conformations using the atomic force microscope.Nat. Struct. Biol. 2000; 7: 719-724Crossref PubMed Scopus (280) Google Scholar, 9Bustamante C. Chemla Y.R. Forde N.R. Izhaky D. Mechanical processes in biochemistry.Annu. Rev. Biochem. 2004; 73: 705-748Crossref PubMed Scopus (601) Google Scholar, 10Neuman K.C. Nagy A. Single-molecule force spectroscopy: optical tweezers, magnetic tweezers and atomic force microscopy.Nat. Methods. 2008; 5: 491-505Crossref PubMed Scopus (1616) Google Scholar, 11Puchner E.M. Gaub H.E. Force and function: probing proteins with AFM-based force spectroscopy.Curr. Opin. Struct. Biol. 2009; 19: 605-614Crossref PubMed Scopus (208) Google Scholar, 12Liang J. Fernández J.M. Mechanochemistry: one bond at a time.ACS Nano. 2009; 3: 1628-1645Crossref PubMed Scopus (114) Google Scholar, 13Duwez A.-S. Willet N. Molecular Manipulation with Atomic Force Microscopy. CRC Press Press, 2012Google Scholar, 14Hughes M.L. Dougan L. The physics of pulling polyproteins: a review of single molecule force spectroscopy using the AFM to study protein unfolding.Rep. Prog. Phys. 2016; 79: 076601Crossref PubMed Scopus (70) Google Scholar In principle, all kinds of conformational transitions may give access to energy-dependent structural changes15Giannotti M.I. Vancso G.J. Interrogation of single synthetic polymer chains and polysaccharides by AFM-based force spectroscopy.ChemPhysChem. 2007; 8: 2290-2307Crossref PubMed Scopus (104) Google Scholar and to some elasticity, e.g., gauche-anti conformational transitions in poly(ethylene oxide) (PEO)16Oesterhelt F. Rief M. Gaub H.E. Single molecule force spectroscopy by AFM indicates helical structure of poly(ethylene-glycol) in water.New J. Phys. 1999; 1: 6Crossref Scopus (389) Google Scholar or bond bending, rotation, and boat-chair conformations in polysaccharides.17Marszalek P.E. Oberhauser A.F. Pang Y.P. Fernandez J.M. Polysaccharide elasticity governed by chair-boat transitions of the glucopyranose ring.Nature. 1998; 396: 661-664Crossref PubMed Scopus (389) Google Scholar,18Rief M. Oesterhelt F. Heymann B. Gaub H.E. Single molecule force spectroscopy on polysaccharides by atomic force microscopy.Science. 1997; 275: 1295-1297Crossref PubMed Scopus (980) Google Scholar But systems that would both adopt stable folded conformations that sustain high mechanical forces and undergo very fast reversible transitions between unfolded and folded states following well-defined pathways are yet to be reported. For this purpose, molecules having fewer degrees of freedom and steeper conformational energy landscapes than biopolymers would stand as valid candidates. In this respect, it is striking that the molecular mechanical properties of the numerous oligomeric and polymeric foldamers developed by chemists over the years, have not been investigated.19Guichard G. Huc I. Synthetic foldamers.Chem. Commun. 2011; 47: 5933-5941Crossref PubMed Scopus (582) Google Scholar, 20Cheng R.P. Gellman S.H. DeGrado W.F. β-peptides: From structure to function.Chem. Rev. 2001; 101: 3219-3232Crossref PubMed Scopus (1686) Google Scholar, 21Yashima E. Maeda K. Iida H. Furusho Y. Nagai K. Helical polymers: synthesis, structures, and functions.Chem. Rev. 2009; 109: 6102-6211Crossref PubMed Scopus (1213) Google Scholar, 22Yashima E. Ousaka N. Taura D. Shimomura K. Ikai T. Maeda K. Supramolecular helical systems: helical assemblies of small molecules, foldamers, and polymers with chiral amplification and their functions.Chem. Rev. 2016; 116: 13752-13990Crossref PubMed Scopus (942) Google Scholar A reason for this is that implementing single-molecule force spectroscopy on synthetic molecules much smaller than biopolymers remains very challenging: interfacing a small object with the force spectroscopy device and measuring transitions of minute amplitudes are both difficult. A few unfolding experiments have been carried out on relatively small (bio)molecules but they did not probe molecular elasticity and mechanical properties.23Kim J.S. Jung Y.J. Park J.W. Shaller A.D. Wan W. Li A.D.Q. Mechanically stretching folded nano- π-b; -stacks reveals pico-newton attractive forces.Adv. Mater. 2009; 21: 786-789Crossref Scopus (22) Google Scholar, 24Sluysmans D. Zhang L. Li X. Garci A. Stoddart J.F. Duwez A.S. Viologen tweezers to probe the force of individual donor−acceptor π-interactions.J. Am. Chem. Soc. 2020; 142: 21153-21159Crossref PubMed Scopus (7) Google Scholar, 25Sluysmans D. Willet N. Thevenot J. Lecommandoux S. Duwez A.S. Single- molecule mechanical unfolding experiments reveal a critical length for the formation of α-helices in peptides.Nanoscale Horiz. 2020; 5: 671-678Crossref PubMed Google Scholar To this date, very few single-molecule mechanics experiments have been successfully conducted on small molecules.26Lussis P. Svaldo-Lanero T. Bertocco A. Fustin C.A. Leigh D.A. Duwez A.S. A single synthetic small molecule that generates force against a load.Nat. Nanotechnol. 2011; 6: 553-557Crossref PubMed Scopus (81) Google Scholar, 27Van Quaethem A. Lussis P. Leigh D.A. Duwez A.-S. Fustin C.-A. Probing the mobility of catenane rings in single molecules.Chem. Sci. 2014; 5: 1449-1452Crossref Scopus (38) Google Scholar, 28Sluysmans D. Hubert S. Bruns C.J. Zhu Z. Stoddart J.F. Duwez A.S. Synthetic oligorotaxanes exert high forces when folding under mechanical load.Nat. Nanotechnol. 2018; 13: 209-213Crossref PubMed Scopus (32) Google Scholar, 29Sluysmans D. Devaux F. Bruns C.J. Stoddart J.F. Duwez A.S. Dynamic force spectroscopy of synthetic oligorotaxane foldamers.Proc. Natl. Acad. Sci. USA. 2018; 115: 9362-9366Crossref PubMed Scopus (30) Google Scholar Here, we report on the high-performance single-molecule mechanics of helical aromatic oligoamides30Huc I. Aromatic oligoamide foldamers.Eur. J. Org. Chem. 2004; 2004: 17-29Crossref Scopus (583) Google Scholar, 31Zhang D.W. Zhao X. Hou J.L. Li Z.T. Aromatic amide foldamers: structures, properties, and functions.Chem. Rev. 2012; 112: 5271-5316Crossref PubMed Scopus (469) Google Scholar, 32Qi T. Deschrijver T. Huc I. Large-scale and chromatography-free synthesis of an octameric quinoline-based aromatic amide helical foldamer.Nat. Protoc. 2013; 8: 693-708Crossref PubMed Scopus (34) Google Scholar as a representative of the broad family of foldamer backbones that possess aryl rings in their main chain. This emerging class of folded molecules are characterized by remarkably stable and predictable conformations. Many of them wind into helices in which aromatic rings stack face to face.33Jiang H. Léger J.M. Huc I. Aromatic δ-peptides.J. Am. Chem. Soc. 2003; 125: 3448-3449Crossref PubMed Scopus (252) Google Scholar, 34Nelson J.C. Saven J.G. Moore J.S. Wolynes P.G. Solvophobically driven folding of nonbiological oligomers.Science. 1997; 277: 1793-1796Crossref PubMed Scopus (754) Google Scholar, 35Gong B. Zeng H. Zhu J. Yuan L. Han Y. Cheng S. Furukawa M. Parra R.D. Kovalevsky A.Y. Mills J.L. et al.Creating nanocavities of tunable sizes: hollow helices.Proc. Natl. Acad. Sci. USA. 2002; 99: 11583-11588Crossref PubMed Scopus (137) Google Scholar, 36Ohkita M. Lehn J.M. Baum G. Fenske D. Helicity coding: programmed molecular self-organization of achiral nonbiological strands into multiturn helical superstructures: synthesis and characterization of alternating pyridine-pyrimidine oligomers.Chem. Eur. J. 1999; 5: 3471-3481Crossref Scopus (155) Google Scholar We now show that their folding trajectories are simple and may be amenable to design. Using AFM-based single-molecule force spectroscopy, we pulled on helical quinoline-based oligoamides of different lengths to investigate their mechanical properties. We found that their winding is extremely robust and fully reversible on a very short time scale, in contrast with typical biopolymer behavior (Figure 1A). The molecular helices behaved as springs that are energy loaded in their extended states. A detailed analysis, based, among others, on elements of stochastic calculus, allowed us to decipher precise winding/unwinding trajectories at the sub-molecular level and cooperative effects responsible for the observed elasticity. The analysis shows that at short times both winding and unwinding processes are driven by the same stochastic mechanisms. It also shows that at longer times a differentiation of mechanisms occurs. Such well-defined conformational trajectories open up new capabilities to orchestrate motion at the molecular scale. Oligoamides of 8-amino-2-quinolinecarboxylic acid Q (Figure 1C) were selected because their structures are well defined, and because their rigidity hinted at simple conformational dynamics. Qn oligomers adopt defectless helical conformations comprised of 2.5 quinoline units per turn that show high stability in a wide range of solvents.33Jiang H. Léger J.M. Huc I. Aromatic δ-peptides.J. Am. Chem. Soc. 2003; 125: 3448-3449Crossref PubMed Scopus (252) Google Scholar,37Qi T. Maurizot V. Noguchi H. Charoenraks T. Kauffmann B. Takafuji M. Ihara H. Huc I. Solvent dependence of helix stability in aromatic oligoamide foldamers.Chem. Commun. 2012; 48: 6337Crossref Scopus (65) Google Scholar Electrostatic repulsions, bifurcated hydrogen bonds, aryl-amide conjugation, and interactions between stacked aromatic rings synergistically concur to this high stability. For example, a simple octamer spanning about three helix turns shows no sign of denaturation at 120°C in DMSO.33Jiang H. Léger J.M. Huc I. Aromatic δ-peptides.J. Am. Chem. Soc. 2003; 125: 3448-3449Crossref PubMed Scopus (252) Google Scholar Furthermore, helix stability rapidly increases with oligomer length.38Delsuc N. Kawanami T. Lefeuvre J. Shundo A. Ihara H. Takafuji M. Huc I. Kinetics of helix-handedness inversion: folding and unfolding in aromatic amide oligomers.ChemPhysChem. 2008; 9: 1882-1890Crossref PubMed Scopus (68) Google Scholar,39Abramyan A.M. Liu Z. Pophristic V. Helix handedness inversion in arylamide foldamers: elucidation and free energy profile of a hopping mechanism.Chem. Commun. 2016; 52: 669-672Crossref PubMed Google Scholar The small number of units per turn, i.e., the high helix curvature, endows relatively short sequences with an aspect ratio larger than that of less curved helices,34Nelson J.C. Saven J.G. Moore J.S. Wolynes P.G. Solvophobically driven folding of nonbiological oligomers.Science. 1997; 277: 1793-1796Crossref PubMed Scopus (754) Google Scholar,35Gong B. Zeng H. Zhu J. Yuan L. Han Y. Cheng S. Furukawa M. Parra R.D. Kovalevsky A.Y. Mills J.L. et al.Creating nanocavities of tunable sizes: hollow helices.Proc. Natl. Acad. Sci. USA. 2002; 99: 11583-11588Crossref PubMed Scopus (137) Google Scholar and reduces the number of rotatable bonds per turn. Furthermore, stepwise segment-doubling synthesis of these oligomers gives access to long sequences with a perfectly uniform constitution.40Li X. Qi T. Srinivas K. Massip S. Maurizot V. Huc I. Synthesis and multibromination of nanosized helical aromatic amide foldamers via segment-doubling condensation.Org. Lett. 2016; 18: 1044-1047Crossref PubMed Scopus (16) Google Scholar Variants of these helices bearing specific terminal functions and side-chain or main-chain features display potentially useful properties in contexts as diverse as DNA mimicry and enzyme inhibition,41Ziach K. Chollet C. Parissi V. Prabhakaran P. Marchivie M. Corvaglia V. Bose P.P. Laxmi-Reddy K. Godde F. Schmitter J.-M. et al.Single helically folded aromatic oligoamides that mimic the charge surface of double-stranded B-DNA.Nat. Chem. 2018; 10: 511-518Crossref PubMed Scopus (28) Google Scholar endomolecular recognition,42Chandramouli N. Ferrand Y. Lautrette G. Kauffmann B. MacKereth C.D. Laguerre M. Dubreuil D. Huc I. Iterative design of a helically folded aromatic oligoamide sequence for the selective encapsulation of fructose.Nat. Chem. 2015; 7: 334-341Crossref PubMed Scopus (163) Google Scholar and fast charge transport.43Méndez-Ardoy A. Markandeya N. Li X. Tsai Y.T. Pecastaings G. Buffeteau T. Maurizot V. Muccioli L. Castet F. Huc I. Bassani D.M. Multi-dimensional charge transport in supramolecular helical foldamer assemblies.Chem. Sci. 2017; 8: 7251-7257Crossref PubMed Google Scholar,44Li X. Markandeya N. Jonusauskas G. McClenaghan N.D. Maurizot V. Denisov S.A. Huc I. Photoinduced electron transfer and hole migration in nanosized helical aromatic oligoamide foldamers.J. Am. Chem. Soc. 2016; 138: 13568-13578Crossref PubMed Scopus (47) Google Scholar For this study, four oligomers comprised of 5, 9, 17, and 33 units were prepared (Figures 1C, S1, and S17–S32 and Supplemental experimental procedures). Monomers are δ- or ε-amino acids and equivalent in size to a dipeptide. Thus, even the longest sequence, Q33, is much smaller than proteins investigated so far by single-molecule force spectroscopy. All sequences bear a thiol group at their N terminus to ensure stable attachment to gold substrates, and polyethylene oxide (PEO) at the C terminus to be linked to the AFM tip for pulling. The molecules were grafted onto gold/silicon substrates at low grafting density to ensure large interspaces between them (Figure 1B, See Supplemental information for details). The AFM tip was brought into contact with the oligoamide-PEO substrate in N,N-dimethylformamide (DMF) to allow the linker to adsorb onto the tip. The caught molecule was then stretched in a controlled manner by moving the tip away from the substrate at a fixed pulling rate, and the force-extension profiles were measured. PEO behaves as a purely random coil in DMF and does not show specific features in the force curves.26Lussis P. Svaldo-Lanero T. Bertocco A. Fustin C.A. Leigh D.A. Duwez A.S. A single synthetic small molecule that generates force against a load.Nat. Nanotechnol. 2011; 6: 553-557Crossref PubMed Scopus (81) Google Scholar,27Van Quaethem A. Lussis P. Leigh D.A. Duwez A.-S. Fustin C.-A. Probing the mobility of catenane rings in single molecules.Chem. Sci. 2014; 5: 1449-1452Crossref Scopus (38) Google Scholar Any deviation from this behavior can thus be safely attributed to the oligoamide helix. Force-extension profiles revealed a consistent behavior for all four foldamers. A typical profile can be divided into different steps (Figure 2A). We attributed the initial part of the force-extension curve (up to about 100 pN) to the unraveling of the PEO random coil and the subsequent plateau to the unwinding of the helical structure. The occurrence of a plateau is characteristic of the breaking of intramolecular interactions in series and is thus usually associated to a non-cooperative unfolding. However, here, the plateau is slightly tilted toward lower forces when the distance increases (Figures 2A, 2C, and S2). It means that it gets easier and easier to open the helix in the course of the unwinding, a signature of a cooperative effect. Once an interaction is broken, the subsequent one is easier to break. At the end of the plateau, the force decreases below plateau level and fluctuates (see next section below). The length of the plateau is in a very good agreement with the theoretical difference in length between the wound and fully extended structures (Figures 1C and S3; Table S1. See Supplemental information for the details of calculation): the extended conformation is 3.8 times longer than the helical state. At the end of the plateau, the helix is completely unwound, and the force rises again as the PEO stretching continues. The force of unwinding, 101 pN for Q33 (Figure 2B), is much higher than for natural helices. For example, double-helical coiled-coil protein structures, triple-helical spectrin repeats, parallel-stacked α-helical ankyrin repeats, and myomesin α-helical linkers in muscles exhibit unwinding forces of about 20 pN at comparable loading rates.45Rief M. Pascual J. Saraste M. Gaub H.E. Single molecule force spectroscopy of spectrin repeats: low unfolding forces in helix bundles.J. Mol. Biol. 1999; 286: 553-561Crossref PubMed Scopus (469) Google Scholar, 46Berkemeier F. Bertz M. Xiao S. Pinotsis N. Wilmanns M. Gräter F. Rief M. Fast-folding alpha-helices as reversible strain absorbers in the muscle protein myomesin.Proc. Natl. Acad. Sci. USA. 2011; 108: 14139-14144Crossref PubMed Scopus (45) Google Scholar, 47Schwaiger I. Sattler C. Hostetter D.R. Rief M. The myosin coiled-coil is a truly elastic protein structure.Nat. Mater. 2002; 1: 232-235Crossref PubMed Scopus (196) Google Scholar, 48Li L. Wetzel S. Plückthun A. Fernandez J.M. Stepwise unfolding of ankyrin repeats in a single protein revealed by atomic force microscopy.Biophys. J. 2006; 90: L30-L32Abstract Full Text Full Text PDF PubMed Scopus (69) Google Scholar, 49Takahashi H. Rico F. Chipot C. Scheuring S. α-helix unwinding as force buffer in spectrins.ACS Nano. 2018; 12: 2719-2727Crossref PubMed Scopus (22) Google Scholar The unwinding of long B-DNA takes place at 65 pN50Cluzel P. Lebrun A. Heller C. Lavery R. Viovy J.L. Chatenay D. Caron F. DNA: an extensible molecule.Science. 1996; 271: 792-794Crossref PubMed Scopus (861) Google Scholar,51Smith S.B. Cui Y. Bustamante C. Overstretching B-DNA: the elastic response of individual double-stranded and single-stranded DNA molecules.Science. 1996; 271: 795-799Crossref PubMed Scopus (2282) Google Scholar and helical structures in polysaccharides unwind between 20 and 60 pN.52Zhang Q. Marszalek P.E. Identification of sugar isomers by single-molecule force spectroscopy.J. Am. Chem. Soc. 2006; 128: 5596-5597Crossref PubMed Scopus (29) Google Scholar,53Zhang L. Wang C. Cui S. Wang Z. Zhang X. Single-molecule force spectroscopy on curdlan: unwinding helical structures and random coils.Nano Lett. 2003; 3: 1119-1124Crossref Scopus (50) Google Scholar Force-extension profiles of the foldamers obtained in other solvents (toluene, ethanol, tetrachloroethane, dimethylsulfoxide, acetonitrile, and water) show similar features, evidencing the robustness of the winding in a wide range of media (Figure S4). The force profiles for the shorter helices are shown in Figures 2C–2E. For Q17, we observed the same characteristic force curves with the length of the plateau being proportional to the length of the molecule (Figures 2E and S2). Q9 displayed a small plateau followed by predominant large fluctuations (Figure S5). Q5 showed only fluctuations or a single rupture peak between the two states (Figure S6), which confirms the cooperativity. The histograms of the length of unwinding signal for each helix show Gaussian-like monomodal distributions centered on a value proportional to the number of units in the helix (Figure 2E). This observation implies that all the helices adopt a unique wound structure in DMF, in agreement with earlier solution and solid-state studies.32Qi T. Deschrijver T. Huc I. Large-scale and chromatography-free synthesis of an octameric quinoline-based aromatic amide helical foldamer.Nat. Protoc. 2013; 8: 693-708Crossref PubMed Scopus (34) Google Scholar,33Jiang H. Léger J.M. Huc I. Aromatic δ-peptides.J. Am. Chem. Soc. 2003; 125: 3448-3449Crossref PubMed Scopus (252) Google Scholar Indeed, narrow monomodal distributions exclude the possibility of populated partially unfolded or differently folded conformations in solution. The unwinding force ranges from 74 to 101 pN from the smallest to the longest helix (Figure 2D). The average unwinding force increases with the number of quinoline units until it reaches a maximum value between 9 and 17 units and plateaus between 17 and 33 units. This is probably an additional signature of cooperativity in winding. Such effects had never been identified in earlier experimental ensemble studies because completely unwound conformations were never reached, only partly unwound transition states were identified. For instance, helix handedness inversion was shown to proceed through a nucleation-propagation mechanism involving the unwinding of only two units.38Delsuc N. Kawanami T. Lefeuvre J. Shundo A. Ihara H. Takafuji M. Huc I. Kinetics of helix-handedness inversion: folding and unfolding in aromatic amide oligomers.ChemPhysChem. 2008; 9: 1882-1890Crossref PubMed Scopus (68) Google Scholar,39Abramyan A.M. Liu Z. Pophristic V. Helix handedness inversion in arylamide foldamers: elucidation and free energy profile of a hopping mechanism.Chem. Commun. 2016; 52: 669-672Crossref PubMed Google Scholar Before the plateau starts, we often observed fluctuations (Figures 2A and S7). These fluctuations can be explained by the reforming of an interaction a short time after being broken, and thus occur between wound and partially unwound states. It means that during pulling, the oligoamide pulls on the AFM cantilever in the opposite direction and rewinds against the mechanical load. The average force in the hopping region was about 10 pN higher than the plateau force. This difference in force can be associated to a barrier to overcome in order to trigger the unwinding of the helix. Such a barrier has been described by molecular dynamics simulations and was shown to correspond to the spring-like deformation of the helix.54Zegarra F.C. Peralta G.N. Coronado A.M. Gao Y.Q. Free energies and forces in helix–coil transition of homopolypeptides under stretching.Phys. Chem. Chem. Phys. 2009; 11: 4019-4024Crossref PubMed Scopus (9) Google Scholar Then, the force decreases, and the plateau starts. At the end of the plateau, the force largely decreases (about 30 pN), and we again detect significant hopping states. In this phase, unwinding requires a lower force and a segment of the foldamer dynamically fluctuates between a partially unwound and the fully unwound state. This is a clear signature of the cooperativity of mechanical unwinding: when the interactions are progressively broken, the subsequent opening is facilitated. The interactions weaken progressively as neighboring interactions disappear along the extension. Weaker forces are thus sufficient to unfold the last part of the sequences, and the drop in force is significant when the length reaches less than two turns. At the upper limit of sampling rate (4,200 nm·s−1), we measured a fluctuation rate between wound and unwound states of about 7,500 s–1 at forces of 100 pN (Figure 3A). From these fluctuations, we could estimate a minimum rewinding rate of about 25,000 s−1 under a force of 100 pN (see Supplemental information for details of the analysis). This estimation is a lower limit. We were indeed limited by the response time of the cantilever and the sampling rate (Figure S14). It means that the rewinding may be even faster than our estimation. The fastest folding kinetics for single α-helices that has been reported so far by direct time-resolved experimental measurements of fluctuations was 30,000 s−1 for villin, an actin-binding protein. This rate is thus comparable to the one obtained here (25,000 s−1), but it was measured at a much lower force of 10 pN.55Zoldák G. Stigler J. Pelz B. Li H. Rief M. Ultrafast folding kinetics and cooperativity of villin headpiece in single-molecule force spectroscopy.Proc. Natl. Acad. Sci. USA. 2013; 110: 18156-18161Crossref PubMed Scopus (58) Google Scholar As the folding rate is known to drastically decrease when the applied force increases,46Berkemeier F. Bertz M. Xiao S. Pinotsis N. Wilmanns M. Gräter F. Rief M. Fast-folding alpha-helices as reversible strain absorbers in the muscle protein myomesin.Proc. Natl. Acad. Sci. USA. 2011; 108: 14139-14144Crossref PubMed Scopus (45) Google Scholar we can conclude that the aromatic amide helices are faster to rewind than those natural single α-helices. It is worth mentioning that here we compare the aromatic amide helices, which are single isolated helices, with natural single α-helices only. Other protein structures described in the literature, such as β-sheets, trimeric α-helices in spectrin, and a series of α-helices in bacte

Full Text
Published version (Free)

Talk to us

Join us for a 30 min session where you can share your feedback and ask us any queries you have

Schedule a call