全部 标题 作者
关键词 摘要

OALib Journal期刊
ISSN: 2333-9721
费用:99美元

查看量下载量

相关文章

更多...

Constrained Peptides as Miniature Protein Structures

DOI: 10.5402/2012/692190

Full-Text   Cite this paper   Add to My Lib

Abstract:

This paper discusses the recent developments of protein engineering using both covalent and noncovalent bonds to constrain peptides, forcing them into designed protein secondary structures. These constrained peptides subsequently can be used as peptidomimetics for biological functions such as regulations of protein-protein interactions. 1. Stabilized and Destabilized -Helices α-Helices have been found to be the secondary structure about 40% of all residues in natural proteins adopt [1], and they are widely used as fundamental recognition elements in many naturally occurring protein complexes, such as Bcl-2/Bak, MDM2/p53, calmodulin/smooth-muscle-myosin-light-chain kinase, Vav/DH domain, and CREB/CBP [2–5]. A typical α-helix completes one rotation with 3.6 amino acid residues, in which each has backbone dihedral angles of and [6]. This results in the helix having a rise of 1.5??/residue or 5.4??/turn [7]. Therefore, the side chain of a certain residue at the position projects from the same face with the side chains at the and the positions in the sequence (Figure 1). The backbone of the α-helix is primarily stabilized by hydrogen bonds between the carbonyl of residues and the carboamide of residues , which all point in the same direction [6]. Because the hydrogen bonding sites on the first and last turns of an α-helix are unfulfilled, a macrodipole is produced [8, 9]. The positive end of the dipole is centered at the N-terminus and the negative at the C-terminus. The total dipole is augmented if the peptide existed in conditions where both the termini are ionized. Figure 1: The structure of an α-helix. (a) The helical wheel diagram. (b) Surface displacement of residues on an α-helix surface. It can be imagined that isolated helical peptides would be ideal inhibitors of macromolecular interactions [10]. However, because many peptides, especially those with less than ten residues, rarely contain sizable degrees of helicity in isolation, much work has been done toward the goal of helix induction and stabilization [11]. The goal of this paper is to highlight the chemical strategies employed to stabilize protein secondary structures and the applications of the constrained peptides in regulating protein-protein interactions. 1.1. Covalent Stabilization The formation of covalent linkages between adjacent residues in peptides has been shown to impart stabilization to the helical form of the peptide. Disulfide bonds, lactam linkages, hydrazones, and carbon-carbon bonds have all been used to link to or residues in a peptide and promote helicity (Figure 2). The

References

[1]  F. Ruan, Y. Chen, and P. B. Hopkins, “Metal ion enhanced helicity in synthetic peptides containing unnatural, metal-ligating residues,” Journal of the American Chemical Society, vol. 112, no. 25, pp. 9403–9404, 1990.
[2]  J. M. Adams and S. Cory, “The Bcl-2 protein family: arbiters of cell survival,” Science, vol. 281, no. 5381, pp. 1322–1326, 1998.
[3]  P. H. Kussie, S. Gorina, V. Marechal et al., “Structure of the MDM2 oncoprotein bound to the p53 tumor suppressor transactivation domain,” Science, vol. 274, no. 5289, pp. 948–953, 1996.
[4]  B. P. Orner, J. T. Ernst, and A. D. Hamilton, “Toward proteomimetics: terphenyl derivatives as structural and functional mimics of extended regions of an α-helix,” Journal of the American Chemical Society, vol. 123, no. 22, pp. 5382–5383, 2001.
[5]  B. Aghazadeh, W. E. Lowry, X. Y. Huang, and M. K. Rosen, “Structural basis for relief of autoinhibition of the Dbl homology domain of proto-oncogene Vav by tyrosine phosphorylation,” Cell, vol. 102, no. 5, pp. 625–633, 2000.
[6]  L. Pauling and R. B. Corey, “The configuration of polypeptide chains in proteins,” Fortschritte Der Chemie Organischer Naturstoffe, vol. 11, pp. 180–239, 1954.
[7]  L. Regan, “What determines where α-helices begin and end?” Proceedings of the National Academy of Sciences of the United States of America, vol. 90, no. 23, pp. 10907–10908, 1993.
[8]  E. Galoppini and M. A. Fox, “Effect of the electric field generated by the helix dipole on photoinduced intramolecular electron transfer in dichromophoric α-helical peptides,” Journal of the American Chemical Society, vol. 118, no. 9, pp. 2299–2300, 1996.
[9]  J. Martin Scholtz, E. J. York, J. M. Stewart, and R. L. Baldwin, “A neutral, water-soluble, α-helical peptide: the effect of ionic strength on the helix-coil equilibrium,” Journal of the American Chemical Society, vol. 113, no. 13, pp. 5102–5104, 1991.
[10]  D. P. Fairlie, M. L. West, and A. K. Wong, “Towards protein surface mimetics,” Current Medicinal Chemistry, vol. 5, no. 1, pp. 29–62, 1998.
[11]  M. J. I. Andrews and A. B. Tabor, “Forming stable helical peptides using natural and artificial amino acids,” Tetrahedron, vol. 55, no. 40, pp. 11711–11743, 1999.
[12]  D. Y. Jackson, D. S. King, J. Chmielewski, S. Singh, and P. G. Schultz, “General approach to the synthesis of short α-helical peptides,” Journal of the American Chemical Society, vol. 113, no. 24, pp. 9391–9392, 1991.
[13]  C. Yu and J. W. Taylor, “Synthesis and study of peptides with semirigid i and i+7 side-chain bridges designed for α-helix stabilization,” Bioorganic and Medicinal Chemistry, vol. 7, no. 1, pp. 161–175, 1999.
[14]  G. ?sapay and J. W. Taylor, “Multicyclic polypeptide model compounds. 2. Synthesis and conformational properties of a highly α-helical uncosapeptide constrained by three side-chain to side-chain lactam bridges,” Journal of the American Chemical Society, vol. 114, no. 18, pp. 6966–6973, 1992.
[15]  A. M. Leduc, J. O. Trent, J. L. Wittliff et al., “Helix-stabilized cyclic peptides as selective inhibitors of steroid receptor—coactivator interactions,” Proceedings of the National Academy of Sciences of the United States of America, vol. 100, no. 20, pp. 11273–11278, 2003.
[16]  T. R. Geistlinger and R. K. Guy, “Novel selective inhibitors of the interaction of individual nuclear hormone receptors with a mutually shared steroid receptor coactivator 2,” Journal of the American Chemical Society, vol. 125, no. 23, pp. 6852–6853, 2003.
[17]  N. E. Shepherd, H. N. Hoang, V. S. Desai, E. Letouze, P. R. Young, and D. P. Fairlie, “Modular α-helical mimetics with antiviral activity against respiratory syncitial virus,” Journal of the American Chemical Society, vol. 128, no. 40, pp. 13284–13289, 2006.
[18]  R. S. Harrison, N. E. Shepherd, H. N. Hoang et al., “Downsizing human, bacterial, and viral proteins to short water-stable alpha helices that maintain biological potency,” Proceedings of the National Academy of Sciences of the United States of America, vol. 107, no. 26, pp. 11686–11691, 2010.
[19]  D. G. Flint, J. R. Kumita, O. S. Smart, and G. A. Woolley, “Using an azobenzene cross-linker to either increase or decrease peptide helix content upon Trans-to-cis photoisomerization,” Chemistry and Biology, vol. 9, no. 3, pp. 391–397, 2002.
[20]  H. E. Blackwell and R. H. Grubbs, “Highly efficient synthesis of covalently cross-linked peptide helices by ring-closing metathesis,” Angewandte Chemie, vol. 37, no. 23, pp. 3281–3284, 1998.
[21]  C. E. Schafmeister, J. Po, and G. L. Verdine, “An all-hydrocarbon cross-linking system for enhancing the helicity and metabolic stability of peptides,” Journal of the American Chemical Society, vol. 122, no. 24, pp. 5891–5892, 2000.
[22]  L. D. Walensky, A. L. Kung, I. Escher et al., “Activation of apoptosis in vivo by a hydrocarbon-stapled BH3 helix,” Science, vol. 305, no. 5689, pp. 1466–1470, 2004.
[23]  F. Bernal, A. F. Tyler, S. J. Korsmeyer, L. D. Walensky, and G. L. Verdine, “Reactivation of the p53 tumor suppressor pathway by a stapled p53 peptide,” Journal of the American Chemical Society, vol. 129, no. 9, pp. 2456–2457, 2007.
[24]  R. E. Moellering, M. Cornejo, T. N. Davis et al., “Direct inhibition of the NOTCH transcription factor complex,” Nature, vol. 462, no. 7270, pp. 182–188, 2009.
[25]  A. Patgiri, A. L. Jochim, and P. S. Arora, “A hydrogen bond surrogate approach for stabilization of short peptide sequences in α-helical conformation,” Accounts of Chemical Research, vol. 41, no. 10, pp. 1289–1300, 2008.
[26]  B. H. Zimm and J. K. Bragg, “Theory of the phase transition between helix and random coil in polypeptide chains,” The Journal of Chemical Physics, vol. 31, no. 2, pp. 526–535, 1959.
[27]  S. Lifson and A. Roig, “On the theory of helix-coil transition in polypeptides,” The Journal of Chemical Physics, vol. 34, no. 6, pp. 1963–1974, 1961.
[28]  A. Patgiri, M. R. Witten, and P. S. Arora, “Solid phase synthesis of hydrogen bond surrogate derived α-helices: resolving the case of a difficult amide coupling,” Organic and Biomolecular Chemistry, vol. 8, no. 8, pp. 1773–1776, 2010.
[29]  D. Wang, K. Chen, J. L. Kulp, and P. S. Arora, “Evaluation of biologically relevant short α-helices stabilized by a main-chain hydrogen-bond surrogate,” Journal of the American Chemical Society, vol. 128, no. 28, pp. 9248–9256, 2006.
[30]  L. K. Henchey, S. Kushal, R. Dubey, R. N. Chapman, B. Z. Olenyuk, and P. S. Arora, “Inhibition of hypoxia inducible factor 1-transcription coactivator interaction by a hydrogen bond surrogate α-helix,” Journal of the American Chemical Society, vol. 132, no. 3, pp. 941–943, 2010.
[31]  D. Wang, M. Lu, and P. S. Arora, “Inhibition of HIV-1 fusion by hydrogen-bond-surrogate-based α helices,” Angewandte Chemie, vol. 47, no. 10, pp. 1879–1882, 2008.
[32]  M. J. Kelso, H. N. Hoang, T. G. Appleton, and D. P. Fairlie, “The first solution stucture of a single α-helical turn. A pentapeptide α-helix stabilized by a metal clip,” Journal of the American Chemical Society, vol. 122, no. 42, pp. 10488–10489, 2000.
[33]  M. J. Kelso, H. N. Hoang, W. Oliver et al., “A cyclic metallopeptide induces α helicity in short peptide fragments of thermolysin,” Angewandte Chemie, vol. 42, no. 4, pp. 421–424, 2003.
[34]  M. J. Kelso, R. L. Beyer, H. N. Hoang et al., “α-turn mimetics: short peptide α-helices composed of cyclic metallopentapeptide modules,” Journal of the American Chemical Society, vol. 126, no. 15, pp. 4828–4842, 2004.
[35]  I. Hamachi, R. Eboshi, J. I. Watanabe, and S. Shinkai, “Guest-induced umpolung on a protein surface: a strategy for regulation of enzymatic activity,” Journal of the American Chemical Society, vol. 122, no. 18, pp. 4530–4531, 2000.
[36]  I. Hamachi, Y. Yamada, T. Matsugi, and S. Shinkai, “Single-or dual-mode switching of semisynthetic ribonuclease S′ with an iminodiacetic acid moiety in response to the copper(II) concentration,” Chemistry, vol. 5, no. 5, pp. 1503–1511, 1999.
[37]  M. R. Ghadiri, C. Soares, and C. Choi, “A convergent approach to protein design. Metal ion-assisted spontaneous self-assembly of a polypeptide into a triple-helix bundle protein,” Journal of the American Chemical Society, vol. 114, no. 3, pp. 825–831, 1992.
[38]  M. R. Ghadiri, C. Soares, and C. Choi, “Design of an artificial four-helix bundle metalloprotein via a novel ruthenium(II)-assisted self-assembly process,” Journal of the American Chemical Society, vol. 114, pp. 4000–4002, 1992.
[39]  M. Lieberman and T. Sasaki, “Iron(II) organizes a synthetic peptide into three-helix bundles,” Journal of the American Chemical Society, vol. 113, pp. 1470–1471, 1991.
[40]  C. Y. Huang, S. He, W. F. DeGrado, D. G. McCafferty, and F. Gai, “Light-induced helix formation,” Journal of the American Chemical Society, vol. 124, no. 43, pp. 12674–12675, 2002.
[41]  D. J. Cline, C. Thorpe, and J. P. Schneider, “Effects of As(III) binding on α-helical structure,” Journal of the American Chemical Society, vol. 125, no. 10, pp. 2923–2929, 2003.
[42]  S. Matsumura, S. Sakamoto, A. Ueno, and H. Mihara, “Construction of α-helix peptides with β-cyclodextrin and dansyl units and their conformational and molecular sensing properties,” Chemistry, vol. 6, no. 10, pp. 1781–1788, 2000.
[43]  N. Voyer and B. Guérin, “Design and synthesis of novel peptides bearing a host and a guest side chains,” Tetrahedron, vol. 50, no. 4, pp. 989–1010, 1994.
[44]  D. Wilson, L. Perlson, and R. Breslow, “Helical templating of oligopeptides by cyclodextrin dimers,” Bioorganic and Medicinal Chemistry, vol. 11, no. 12, pp. 2649–2653, 2003.
[45]  T. J. Shepodd, M. A. Petti, and D. A. Dougherty, “Molecular recognition in aqueous media: donor-acceptor and ion-dipole interactions produce tight binding for highly soluble guests,” Journal of the American Chemical Society, vol. 110, no. 6, pp. 1983–1985, 1988.
[46]  Z. Shi, C. A. Olson, and N. R. Kallenbach, “Cation-π interaction in model α-helical peptides,” Journal of the American Chemical Society, vol. 124, no. 13, pp. 3284–3291, 2002.
[47]  C. A. Olson, Z. Shi, and N. R. Kallenbach, “Polar interactions with aromatic side chains in α-helical peptides: Ch···O H-bonding and cation-π interactions,” Journal of the American Chemical Society, vol. 123, no. 26, pp. 6451–6452, 2001.
[48]  J. P. Gallivan and D. A. Dougherty, “Cation-π interactions in structural biology,” Proceedings of the National Academy of Sciences of the United States of America, vol. 96, no. 17, pp. 9459–9464, 1999.
[49]  L. K. Tsou, C. D. Tatko, and M. L. Waters, “Simple cation-π interaction between a phenyl ring and a protonated amine stabilizes an α-helix in water,” Journal of the American Chemical Society, vol. 124, no. 50, pp. 14917–14921, 2002.
[50]  J. S. Albert, M. S. Goodman, and A. D. Hamilton, “Molecular recognition of proteins: sequence-selective binding of aspartate pairs in helical peptides,” Journal of the American Chemical Society, vol. 117, no. 3, pp. 1143–1144, 1995.
[51]  J. S. Albert and A. D. Hamilton, “Stabilization of helical domains in short peptides using hydrophobic interactions,” Biochemistry, vol. 34, no. 3, pp. 984–990, 1995.
[52]  M. Tabet, V. Labroo, P. Sheppard, and T. Sasaki, “Spermine-induced conformational changes of a synthetic peptide,” Journal of the American Chemical Society, vol. 115, no. 10, pp. 3866–3868, 1993.
[53]  I. Hamachi, Y. Yamada, R. Eboshi, T. Hiraoka, and S. Shinkai, “Design and semisynthesis of spermine-sensitive Ribonuclease S',” Bioorganic and Medicinal Chemistry Letters, vol. 9, no. 9, pp. 1215–1218, 1999.
[54]  T. Haack, M. W. Peczuh, X. Salvatella et al., “Surface recognition and helix stabilization of a tetraaspartate peptide by shape and electrostatic complementarity of an artificial receptor,” Journal of the American Chemical Society, vol. 121, no. 50, pp. 11813–11820, 1999.
[55]  M. W. Peczuh, A. D. Hamilton, J. Sanchez-Quesada, J. De Mendoza, T. Haack, and E. Giralt, “Recognition and stabilization of an α-helical peptide by a synthetic receptor,” Journal of the American Chemical Society, vol. 119, no. 39, pp. 9327–9328, 1997.
[56]  B. P. Orner, X. Salvatella, J. Sánchez Quesada, J. De Mendoza, E. Giralt, and A. D. Hamilton, “De novo protein surface design: use of cation-π interactions to enhance binding between an α-helical peptide and a cationic molecule in 50% aqueous solution,” Angewandte Chemie, vol. 41, no. 1, pp. 117–119, 2002.
[57]  E. Schievano, A. Bisello, M. Chorev, A. Bisol, S. Mammi, and E. Peggion, “Aib-rich peptides containing lactam-bridged side chains as models of the 310-helix,” Journal of the American Chemical Society, vol. 123, no. 12, pp. 2743–2751, 2001.
[58]  C. Toniolo and E. Benedetti, “The polypeptide 310-helix,” Trends in Biochemical Sciences, vol. 16, no. 9, pp. 350–353, 1991.
[59]  C. Peggion, F. Formaggio, M. Crisma et al., “Folding of peptides characterized by c3Val, a highly constrained analogue of valine,” Biopolymers, vol. 68, no. 2, pp. 178–191, 2003.
[60]  R. P. Cheng, S. H. Gellman, and W. F. DeGrado, “β-peptides: from structure to function,” Chemical Reviews, vol. 101, no. 10, pp. 3219–3232, 2001.
[61]  R. P. Cheng and W. F. DeGrado, “De novo design of a monomeric helical β-peptide stabilized by electrostatic interactions,” Journal of the American Chemical Society, vol. 123, no. 21, pp. 5162–5163, 2001.
[62]  S. H. Gellman, “Foldamers: a Manifesto,” Accounts of Chemical Research, vol. 31, no. 4, pp. 173–180, 1998.
[63]  C. M. Goodman, S. Choi, S. Shandler, and W. F. DeGrado, “Foldamers as versatile frameworks for the design and evolution of function,” Nature Chemical Biology, vol. 3, no. 5, pp. 252–262, 2007.
[64]  D. Seebach, D. F. Hook, and A. Gl?ttli, “Helices and other secondary structures of β- and γ-peptides,” Biopolymers, vol. 84, no. 1, pp. 23–37, 2006.
[65]  D. Seebach and J. Gardiner, “β-Peptidic peptidomimetics,” Accounts of Chemical Research, vol. 41, no. 10, pp. 1366–1375, 2008.
[66]  W. S. Horne and S. H. Gellman, “Foldamers with heterogeneous backbones,” Accounts of Chemical Research, vol. 41, no. 10, pp. 1399–1408, 2008.
[67]  J. A. Kritzer, O. M. Stephens, D. A. Guarracino, S. K. Reznik, and A. Schepartz, “β-Peptides as inhibitors of protein-protein interactions,” Bioorganic and Medicinal Chemistry, vol. 13, no. 1, pp. 11–16, 2005.
[68]  W. S. Horne, L. M. Johnson, T. J. Ketas et al., “Structural and biological mimicry of protein surface recognition by α/β-peptide foldamers,” Proceedings of the National Academy of Sciences of the United States of America, vol. 106, no. 35, pp. 14751–14756, 2009.
[69]  D. S. Kemp, J. G. Boyd, and C. C. Muendel, “The helical s constant for alanine in water derived from template-nucleated helices,” Nature, vol. 352, no. 6334, pp. 451–454, 1991.
[70]  D. S. Kemp, “Efficient helix nucleation by a macrocyclic triproline-derived template,” Tetrahedron Letters, vol. 36, no. 23, pp. 4023–4026, 1995.
[71]  R. E. Austin, R. A. Maplestone, A. M. Sefler et al., “A template for stabilizatgion of a peptide α-helix: synthesis and evaluation of conformational effects by circular dichroism and NMR,” Journal of the American Chemical Society, vol. 119, no. 28, pp. 6461–6472, 1997.
[72]  W. M. Kazmierski, R. J. Hazen, A. Aulabaugh, and M. H. StClair, “Inhibitors of human immunodeficiency virus type 1 derived from gp41 transmembrane protein: structure-activity studies,” Journal of Medicinal Chemistry, vol. 39, no. 14, pp. 2681–2689, 1996.
[73]  A. A. Virgilio and J. A. Ellman, “Simultaneous solid-phase synthesis of β-turn mimetics incorporating side-chain functionality,” Journal of the American Chemical Society, vol. 116, no. 25, pp. 11580–11581, 1994.
[74]  A. A. Virgilio, S. C. Schürer, and J. A. Ellman, “Expedient solid-phase synthesis of putative β-turn mimetics incorporating the i + 1, i + 2, and i + 3 sidechains,” Tetrahedron Letters, vol. 37, no. 39, pp. 6961–6964, 1996.
[75]  K. Burgess, “Solid-phase syntheses of β-turn analogues to mimic or disrupt protein—protein interactions,” Accounts of Chemical Research, vol. 34, no. 10, pp. 826–835, 2001.
[76]  A. G. Cochran, N. J. Skelton, and M. A. Starovasnik, “Tryptophan zippers: stable, monomeric β-hairpins,” Proceedings of the National Academy of Sciences of the United States of America, vol. 98, no. 10, pp. 5578–5583, 2001.
[77]  A. G. Cochran, R. T. Tong, M. A. Starovasnik et al., “A minimal peptide scaffold for β-turn display: optimizing a strand position in disulfide-cyclized β-hairpins,” Journal of the American Chemical Society, vol. 123, no. 4, pp. 625–632, 2001.
[78]  K. Y. Tsang, H. Diaz, N. Graciani, and J. W. Kelly, “Hydrophobic cluster formation is necessary for dibenzofuran-based amino acids to function as β-sheet nucleators,” Journal of the American Chemical Society, vol. 116, no. 9, pp. 3988–4005, 1994.
[79]  H. Díaz, K. Y. Tsang, D. Choo, J. R. Espina, and J. W. Kelly, “Design, synthesis, and partial characterization of water-soluble β-sheets stabilized by a dibenzofuran-based amino acid,” Journal of the American Chemical Society, vol. 115, no. 9, pp. 3790–3791, 1993.
[80]  H. Diaz, J. R. Espina, and J. W. Kelly, “A dibenzofuran based amino acid designed to nucleate antiparallel β-sheet structure: evidence for intramolecular hydrogen bond formation,” Journal of the American Chemical Society, vol. 114, pp. 8316–8318, 1992.
[81]  H. Diaz and J. W. Kelly, “The synthesis of dibenzofuran based diacids and amino acids designed to nucleate parallel and antiparallel β-sheet formation,” Tetrahedron Letters, vol. 32, no. 41, pp. 5725–5728, 1991.
[82]  R. Kaul, A. R. Angeles, M. J?ger, E. T. Powers, and J. W. Kelly, “Incorporating β-turns and a turn mimetic out of context in loop 1 of the WW domain affords cooperatively folded β-sheets,” Journal of the American Chemical Society, vol. 123, no. 22, pp. 5206–5212, 2001.
[83]  R. Kaul, S. Deechongkit, and J. W. Kelly, “Synthesis of a negatively charged dibenzofuran-based β-turn mimetic and its incorporation into the WW miniprotein-enhanced solubility without a loss of thermodynamic stability,” Journal of the American Chemical Society, vol. 124, no. 40, pp. 11900–11907, 2002.
[84]  U. Arnold, M. P. Hinderaker, B. L. Nilsson, B. R. Huck, S. H. Gellman, and R. T. Raines, “Protein prosthesis: a semisynthetic enzyme with a β-peptide reverse turn,” Journal of the American Chemical Society, vol. 124, no. 29, pp. 8522–8523, 2002.
[85]  G. Platt, C. W. Chung, and M. S. Searle, “Design of histidine-Zn2+ binding sites within a β-hairpin peptide: enhancement of β-sheet stability through metal complexation,” Chemical Communications, no. 13, pp. 1162–1163, 2001.
[86]  S. R. Griffiths-Jones and M. S. Searle, “Structure, folding, and energetics of cooperative interactions between the ?-strands of a de Novo designed three-stranded antiparallel ?-sheet peptide,” Journal of the American Chemical Society, vol. 122, no. 35, pp. 8350–8356, 2000.
[87]  G. J. Sharman and M. S. Searle, “Cooperative interaction between the three strands of a designed antiparallel β-sheet,” Journal of the American Chemical Society, vol. 120, no. 21, pp. 5291–5300, 1998.
[88]  S. H. Gellman, “Minimal model systems for β sheet secondary structure in proteins,” Current Opinion in Chemical Biology, vol. 2, no. 6, pp. 717–725, 1998.
[89]  H. L. Schenck and S. H. Gellman, “Use of a designed triple-stranded antiparallel β-sheet to probe β-sheet cooperativity in aqueous solution,” Journal of the American Chemical Society, vol. 120, no. 19, pp. 4869–4870, 1998.
[90]  J. D. Fisk and S. H. Gellman, “A parallel β-sheet model system that folds in water,” Journal of the American Chemical Society, vol. 123, no. 2, pp. 343–344, 2001.
[91]  J. S. Nowick, E. M. Smith, J. W. Ziller, and A. J. Shaka, “Three-stranded mixed artificial β-sheets,” Tetrahedron, vol. 58, no. 4, pp. 727–739, 2002.
[92]  J. S. Nowick, D. M. Chung, K. Maitra, S. Maitra, K. D. Stigers, and Y. Sun, “An unnatural amino acid that mimics a tripeptide β-strand and forms β-sheetlike hydrogen-bonded dimers,” Journal of the American Chemical Society, vol. 122, no. 32, pp. 7654–7661, 2000.
[93]  E. M. Smith, D. L. Holmes, A. J. Shaka, and J. S. Nowick, “An artificial antiparallel β-sheet containing a new peptidomimetic template,” Journal of Organic Chemistry, vol. 62, no. 23, pp. 7906–7907, 1997.
[94]  J. S. Nowick, E. M. Smith, and M. Pairish, “Artificial β-sheets,” Chemical Society Reviews, vol. 25, no. 6, pp. 401–415, 1996.
[95]  J. S. Nowick, K. S. Lam, T. V. Khasanova et al., “An unnatural amino acid that induces β-sheet folding and interaction in peptides,” Journal of the American Chemical Society, vol. 124, no. 18, pp. 4972–4973, 2002.
[96]  J. P. Schneider and J. W. Kelly, “Synthesis and efficacy of square planar copper complexes designed to nucleate β-sheet structure,” Journal of the American Chemical Society, vol. 117, no. 9, pp. 2533–2546, 1995.
[97]  I. G. Jones and M. North, “The use of norbornene derivatives in the synthesis of conformationally constrained peptides and pseudo-peptides,” Letters in Peptide Science, vol. 5, no. 2-3, pp. 171–173, 1998.
[98]  D. Ranganathan, V. Haridas, S. Kurur et al., “Norbornene-constrained cyclic peptides with hairpin architecture: design, synthesis, conformation, and membrane ion transport,” Journal of Organic Chemistry, vol. 65, no. 2, pp. 365–374, 2000.
[99]  C. P. R. Hackenberger, I. Schiffers, J. Runsink, and C. Bolm, “General synthesis of unsymmetrical norbornane scaffolds as inducers for hydrogen bond interactions in peptides,” Journal of Organic Chemistry, vol. 69, no. 3, pp. 739–743, 2004.
[100]  H. Zeng, X. Yang, R. A. Flowers, and B. Gong, “A noncovalent approach to antiparallel β-sheet formation,” Journal of the American Chemical Society, vol. 124, no. 12, pp. 2903–2910, 2002.
[101]  A. M. Finch, A. K. Wong, N. J. Paczkowski et al., “Low-molecular-weight peptidic and cyclic antagonists of the receptor for the complement factor C5a,” Journal of Medicinal Chemistry, vol. 42, no. 11, pp. 1965–1974, 1999.
[102]  A. Short, A. K. Wong, A. M. Finch et al., “Effects of a new C5a receptor antagonist on C5a- and endotoxin-induced neutropenia in the rat,” British Journal of Pharmacology, vol. 126, no. 3, pp. 551–554, 1999.
[103]  C. Garcia-Echeverria, P. Chene, M. J. J. Blommers, and P. Furet, “Discovery of potent antagonists of the interaction between human double minute 2 and tumor suppressor p53,” Journal of Medicinal Chemistry, vol. 43, no. 17, pp. 3205–3208, 2000.
[104]  H. Yin, G. I. Lee, S. P. Hyung et al., “Terphenyl-based helical mimetics that disrupt the p53/HDM2 interaction,” Angewandte Chemie, vol. 44, no. 18, pp. 2704–2707, 2005.
[105]  R. Fasan, R. L. A. Dias, K. Moehle et al., “Using a β-hairpin to mimic an α-helix: cyclic peptidomimetic inhibitors of the p53-HDM2 protein-protein interaction,” Angewandte Chemie, vol. 43, no. 16, pp. 2109–2112, 2004.
[106]  J. A. Kritzer, J. D. Lear, M. E. Hodsdon, and A. Schepartz, “Helical β-peptide inhibitors of the p53-hDM2 interaction,” Journal of the American Chemical Society, vol. 126, no. 31, pp. 9468–9469, 2004.
[107]  J. K. Murray and S. H. Gellman, “Targeting protein-protein interactions: lessons from p53/MDM2,” Biopolymers, vol. 88, no. 5, pp. 657–686, 2007.
[108]  J. P. Plante, T. Burnley, B. Malkova et al., “Oligobenzamide proteomimetic inhibitors of the p53-hDM2 protein-protein interaction,” Chemical Communications, no. 34, pp. 5091–5093, 2009.
[109]  A. Shaginian, L. R. Whitby, S. Hong et al., “Design, synthesis, and evaluation of an a-helix mimetic library targeting protein-protein interactions,” Journal of the American Chemical Society, vol. 131, no. 15, pp. 5564–5572, 2009.
[110]  A. B?ttger, V. B?ttger, A. Sparks, W.-L. Liu, S. F. Howard, and D. P. Lane, “Design of a synthetic Mdm2-binding mini protein that activates the p53 response in vivo,” Current Biology, vol. 7, no. 11, pp. 860–869, 1997.
[111]  C. Li, M. Liu, J. Monbo et al., “Turning a scorpion toxin into an antitumor miniprotein,” Journal of the American Chemical Society, vol. 130, no. 41, pp. 13546–13548, 2008.
[112]  C. Li, M. Pazgier, M. Liu, W. Y. Lu, and W. Lu, “Apamin as a template for structure-based rational design of potent peptide activators of p53,” Angewandte Chemie, vol. 48, no. 46, pp. 8712–8715, 2009.
[113]  J. A. Kritzer, R. Zutshi, M. Cheah et al., “Miniature protein inhibitors of the p53-hDM2 interaction,” ChemBioChem, vol. 7, no. 1, pp. 29–31, 2006.
[114]  E. A. Harker and A. Schepartz, “Cell-permeable β-peptide inhibitors of p53/hDM2 complexation,” ChemBioChem, vol. 10, no. 6, pp. 990–993, 2009.
[115]  D. C. Chan, D. Fass, J. M. Berger, and P. S. Kim, “Core structure of gp41 from the HIV envelope glycoprotein,” Cell, vol. 89, no. 2, pp. 263–273, 1997.
[116]  K. Tan, J. H. Liu, J. H. Wang, S. Shen, and M. Lu, “Atomic structure of a thermostable subdomain of HIV-1 gp41,” Proceedings of the National Academy of Sciences of the United States of America, vol. 94, no. 23, pp. 12303–12308, 1997.
[117]  D. M. Eckert, V. N. Malashkevich, L. H. Hong, P. A. Carr, and P. S. Kim, “Inhibiting HIV-1 entry: discovery of D-peptide inhibitors that target the gp41 coiled-coil pocket,” Cell, vol. 99, no. 1, pp. 103–115, 1999.
[118]  M. Ferrer, T. M. Kapoor, T. Strassmaier et al., “Selection of gp41-mediated HIV-1 cell entry inhibitors from biased combinatorial libraries of non-natural binding elements,” Nature Structural Biology, vol. 6, no. 10, pp. 953–960, 1999.
[119]  D. C. Chan, C. T. Chutkowski, and P. S. Kim, “Evidence that a prominent cavity in the coiled coil of HIV type 1 gp41 is an attractive drug target,” Proceedings of the National Academy of Sciences of the United States of America, vol. 95, no. 26, pp. 15613–15617, 1998.
[120]  C. T. Wild, D. C. Shugars, T. K. Greenwell, C. B. McDanal, and T. J. Matthews, “Peptides corresponding to a predictive α-helical domain of human immunodeficiency virus type 1 gp41 are potent inhibitors of virus infection,” Proceedings of the National Academy of Sciences of the United States of America, vol. 91, no. 21, pp. 9770–9774, 1994.
[121]  S. Jiang, K. Lin, N. Strick, and A. R. Neurath, “HIV-1 inhibition by a peptide,” Nature, vol. 365, no. 6442, p. 113, 1993.
[122]  J. K. Judice, J. Y. K. Tom, W. Huang et al., “Inhibition of HIV type 1 infectivity by constrained α-helical peptides: implications for the viral fusion mechanism,” Proceedings of the National Academy of Sciences of the United States of America, vol. 94, no. 25, pp. 13426–13430, 1997.
[123]  S. K. Sia, P. A. Carr, A. G. Cochran, V. N. Malashkevich, and P. S. Kim, “Short constrained peptides that inhibit HIV-1 entry,” Proceedings of the National Academy of Sciences of the United States of America, vol. 99, no. 23, pp. 14664–14669, 2002.
[124]  J. R. Barbier, W. Neugebauer, P. Morley et al., “Bioactivities and secondary structures of constrained analogues of human parathyroid hormone: cyclic lactams of the receptor binding region,” Journal of Medicinal Chemistry, vol. 40, no. 9, pp. 1373–1380, 1997.
[125]  S. M. Condon, I. Morize, S. Darnbrough et al., “The bioactive conformation of human parathyroid hormone. Structural evidence for the extended helix postulate,” Journal of the American Chemical Society, vol. 122, no. 13, pp. 3007–3014, 2000.
[126]  E. Peggion, S. Mammi, E. Schievano et al., “Structure-function relationship studies of bovine parathyroid hormone [bPTH(1-34)] analogues containing α-amino-iso-butyric acid (Aib) residues,” Biopolymers, vol. 68, no. 3, pp. 437–457, 2003.
[127]  C. Das, O. Berezovska, T. S. Diehl et al., “Designed helical peptides inhibit an intramembrane protease,” Journal of the American Chemical Society, vol. 125, no. 39, pp. 11794–11795, 2003.
[128]  I. L. Karle and P. Balaram, “Structural characteristics of α-helical peptide molecules containing Aib residues,” Biochemistry, vol. 29, no. 29, pp. 6747–6756, 1990.
[129]  J. W. Chin and A. Schepartz, “Design and evolution of a miniature Bcl-2 binding protein,” Angewandte Chemie, vol. 40, no. 20, pp. 3806–3809, 2001.
[130]  J. W. Chin, R. M. Grotzfeld, M. A. Fabian, and A. Schepartz, “Methodology for optimizing functional miniature proteins based on avian pancreatic polypeptide using phage display,” Bioorganic and Medicinal Chemistry Letters, vol. 11, no. 12, pp. 1501–1505, 2001.
[131]  J. W. Chin and A. Schepartz, “Concerted evolution of structure and function in a miniature protein,” Journal of the American Chemical Society, vol. 123, no. 12, pp. 2929–2930, 2001.
[132]  N. J. Zondlo and A. Schepartz, “Highly specific DNA recognition by a designed miniature protein,” Journal of the American Chemical Society, vol. 121, no. 29, pp. 6938–6939, 1999.
[133]  J. K. Montclare and A. Schepartz, “Miniature homeodomains: high specificity without an N-terminal arm,” Journal of the American Chemical Society, vol. 125, no. 12, pp. 3416–3417, 2003.
[134]  S. E. Rutledge, H. M. Volkman, and A. Schepartz, “Molecular recognition of protein surfaces: high affinity ligands for the CBP KIX domain,” Journal of the American Chemical Society, vol. 125, no. 47, pp. 14336–14347, 2003.
[135]  I. Radhakrishnan, G. C. Pérez-Alvarado, D. Parker, H. J. Dyson, M. R. Montminy, and P. E. Wright, “Solution structure of the KIX domain of CBP bound to the transactivation domain of CREB: a model for activator:coactivator interactions,” Cell, vol. 91, no. 6, pp. 741–752, 1997.
[136]  I. Glover, I. Haneef, and J. Pitts, “Conformational flexibility in a small globular hormone: X-ray analysis of avian pancreatic polypeptide at 0.98-? resolution,” Biopolymers, vol. 22, no. 1, pp. 293–304, 1983.
[137]  H. M. Volkman, S. E. Rutledge, and A. Schepartz, “Binding mode and transcriptional activation potential of high affinity ligands for the CBP KIX domain,” Journal of the American Chemical Society, vol. 127, no. 13, pp. 4649–4658, 2005.
[138]  M. Abul Fazal, B. C. Roy, S. Sun, S. Mallik, and K. R. Rodgers, “Surface recognition of a protein using designed transition metal complexes,” Journal of the American Chemical Society, vol. 123, no. 26, pp. 6283–6290, 2001.
[139]  W. E. Meador, A. R. Means, and F. A. Quiocho, “Target enzyme recognition by calmodulin: 2.4 ? structure of a calmodulin-peptide complex,” Science, vol. 257, no. 5074, pp. 1251–1255, 1992.
[140]  A. J. Souers, A. A. Virgilio, S. S. Schürer et al., “Novel inhibitors of α4β1 integrin receptor interactions through library synthesis and screening,” Bioorganic and Medicinal Chemistry Letters, vol. 8, no. 17, pp. 2297–2302, 1998.
[141]  A. J. Souers, A. A. Virgilio, ?. Rosenquist, W. Fenuik, and J. A. Ellman, “Identification of a potent heterocyclic ligand to somatostatin receptor subtype 5 by the synthesis and screening of β-turn mimetic libraries,” Journal of the American Chemical Society, vol. 121, no. 9, pp. 1817–1825, 1999.
[142]  G. Lauri and P. A. Bartlett, “CAVEAT: a program to facilitate the design of organic molecules,” Journal of Computer-Aided Molecular Design, vol. 8, no. 1, pp. 51–66, 1994.
[143]  F. A. Etzkorn, T. Guo, M. A. Lipton, S. D. Goldberg, and P. A. Bartlett, “Cyclic hexapeptides and chimeric peptides as mimics of tendamistat,” Journal of the American Chemical Society, vol. 116, no. 23, pp. 10412–10425, 1994.
[144]  A. N. Koehler, “A complex task? Direct modulation of transcription factors with small molecules,” Current Opinion in Chemical Biology, vol. 14, no. 3, pp. 331–340, 2010.

Full-Text

comments powered by Disqus

Contact Us

service@oalib.com

QQ:3279437679

WhatsApp +8615387084133