New Compstatin Peptides Containing N-Terminal Extensions and Non-Natural Amino Acids Exhibit Potent Complement Inhibition and Improved Solubility CharacteristicsClick to copy article linkArticle link copied!
- Ronald D. Gorham, Jr.
- David L. Forest
- George A. Khoury
- James Smadbeck
- Consuelo N. Beecher
- Evangeline D. Healy
- Phanourios Tamamis
- Georgios Archontis
- Cynthia K. Larive
- Christodoulos A. Floudas
- Monte J. Radeke
- Lincoln V. Johnson
- Dimitrios Morikis
Abstract
Compstatin peptides are complement inhibitors that bind and inhibit cleavage of complement C3. Peptide binding is enhanced by hydrophobic interactions; however, poor solubility promotes aggregation in aqueous environments. We have designed new compstatin peptides derived from the W4A9 sequence (Ac-ICVWQDWGAHRCT-NH2, cyclized between C2 and C12), based on structural, computational, and experimental studies. Furthermore, we developed and utilized a computational framework for the design of peptides containing non-natural amino acids. These new compstatin peptides contain polar N-terminal extensions and non-natural amino acid substitutions at positions 4 and 9. Peptides with α-modified non-natural alanine analogs at position 9, as well as peptides containing only N-terminal polar extensions, exhibited similar activity compared to W4A9, as quantified via ELISA, hemolytic, and cell-based assays, and showed improved solubility, as measured by UV absorbance and reverse-phase HPLC experiments. Because of their potency and solubility, these peptides are promising candidates for therapeutic development in numerous complement-mediated diseases.
Introduction

Position refers to residue number within each compstatin sequence. For reference, the Cys residues are always at positions 2 and 12.
Non-natural amino acid abbreviations: meW = l-1-methyltryptophan; Nal = l-1-naphthylalanine; Rea = R-α-ethylalanine; Aal = R-α-allylalanine; Sea = S-α-ethylalanine; 2Nl = l-2-naphthylalanine. All peptides (except linear) are cyclized by a disulfide bond between C2 and C12.
Figure 1
Figure 1. Chemical structures of non-natural amino acids in compstatin peptide sequences. The abbreviations used in sequences (Table 1) are shown in parentheses.
Results
Complement Inhibition in ELISA and Hemolytic Assays
Figure 2
Figure 2. Concentration-dependent inhibition curves of compstatin peptides in C3b and C5b-9 ELISAs and hemolytic assays. Data show C3b, C5b-9, and hemolysis inhibition (from left to right) for set 1 (A–C), set 2 (D–F), set 3 (G–I), set 4 (J–L), set 5 (M–O), and control (P–R) peptides. Data points indicate mean percent inhibition ± SEM (standard error of the mean). The intersection of dashed lines shows the IC50 value for meW4A9 in all plots. Curves to the right and left of the intersection point represent peptides with higher IC50 values and lower IC50 values compared to meW4A9, respectively.
Figure 3
Figure 3. IC50 values for compstatin peptides. Bar plots show IC50 values for compstatin peptides 1–20 and positive control peptides W4A9, meW4A9, and Parent in C3b ELISA (A), C5b-9 ELISA (B), and hemolytic assay (C). Bars show mean IC50 from three independent runs of each experiment (±95% confidence interval). The dashed horizontal line shows the IC50 value for the meW4A9 control peptide, for ease of comparison.
Complement Inhibition in Retinal Pigmented Epithelial Cell Model
Figure 4
Figure 4. Effects of compstatin peptides on complement activation in the RPE cell in vitro assay. The ratio of C5b-9/ApoE fluorescence (±SEM, n = 10) is plotted as a percentage of the positive control (POS) for two hfRPE cell lines, 072810 (gray) and 081309 (black). Untreated cells that were not incubated with complement-competent human serum served as negative control (NEG). At 1 μM, the parent compound is not significantly different from the positive or linear peptide controls. All test peptides (W4A9, PEP 5, PEP 8, PEP 12, PEP 18, and PEP 19) displayed significant complement inhibition relative to their corresponding positive control (see Tables S5 and S6).
Figure 5
Figure 5. Effects of varying concentrations of Parent on complement activation in the RPE cell in vitro assay. The ratio of C5b-9/ApoE fluorescence (±SEM, n = 10) is plotted as a percentage of the positive control. Parent was tested at concentrations of 1, 10, and 50 μM (PAR1, PAR10, and PAR50). The concentration of W4A9 was 1 μM. All values are expressed relative to the positive control. Parent shows no significant difference from the positive control at 1 μM or 10 μM concentrations. At 50 μM the effect of Parent is equivalent to that of 1 μM W4A9. Both Parent at 50 μM and W4A9 at 1 μM are significantly different than the positive control (p-value of <0.001, two-tailed Mann–Whitney U test).
Solubility of Compstatin Peptides
Figure 6
Figure 6. Relation between activity and solubility of compstatin peptides. IC50 values for compstatin peptides in C3b ELISA (A), C5b-9 ELISA (B), and hemolytic assay (C) are plotted against solubility. Horizontal and vertical error bars represent the standard deviation and 95% confidence intervals of solubility and IC50, respectively. Points inside the ellipse (lower right corner) have a favorable balance between activity and solubility. Note the peptide 8 solubility was measured using less starting material compared to the rest, and thus, its solubility is not directly comparable to that of other peptides.
Relative Lipophilicity of Compstatin Peptides
Figure 7
Figure 7. RP-HPLC retention factors for compstatin peptides. Bar plots show log(k) values for compstatin peptides 1–20 and positive control peptides W4A9, meW4A9, and Parent in C3b ELISA. The dashed horizontal line shows the log(k) value for the meW4A9 control peptide, for ease of comparison.
Discussion
Experimental Section
De Novo Peptide Design with Natural Amino Acids
Sequence Selection






Fold Specificity

Approximate Binding Affinity
De Novo Peptide Design with Non-Natural Amino Acids









Constraints That Must be Met for Inclusion in Set Mp


Binary Decision Variables
Peptide Synthesis
Solubility Measurements
C3b and C5b-9 ELISA
Hemolytic Assays
RPE Cell Culture
Immunohistochemistry
Confocal Imaging and Analysis
RP-HPLC Study
Supporting Information
Tables and figures of peptide design rankings, compstatin peptide inhibitory activity (from ELISAs, hemolytic assays, and cell based RPE assays), and solubility/lipophilicity data. This material is available free of charge via the Internet at http://pubs.acs.org.
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.
Acknowledgment
D.M. and M.J.R. acknowledge donors to Macular Degeneration Research, a program of the BrightFocus Foundation, for support of this research (Grant M2013106). D.M. is the recipient of the Carolyn K. McGillvray Memorial Award for Macular Degeneration Research, administered by the BrightFocus Foundation. D.L.F. was supported by the Garland Initiative for Vision at University of California, Santa Barbara, and the BrightFocus Foundation (Grant M2013106). C.A.F. acknowledges support from the National Institutes of Health (Grant R01GM052032). G.A.K. is grateful for support by a National Science Foundation Graduate Research Fellowship under Grant DGE-1148900. The authors gratefully acknowledge that the design calculations were performed at the TIGRESS high performance computing center at Princeton University, NJ, which is supported by the Princeton Institute for Computational Science and Engineering (PICSciE) and the Princeton University Office of Information Technology.
C3 | complement component 3 |
Ac | acetyl group |
C3(H2O) | hydrolyzed C3 |
C3a | complement component 3a |
C3b | complement component 3b |
C3c | complement component 3c |
C5b-9 | membrane attack complex comprising complement proteins C5b, C6, C7, C8, and polymeric C9 |
meW | l-1-methyltryptophan |
Nal | l-1-naphthylalanine |
Rea | R-α-ethylalanine |
Aal | R-α-allylalanine |
Sea | S-α-ethylalanine |
2Nl | l-2-naphthylalanine |
Nmw | N-methyltryptophan |
Nma | N-methylalanine |
Pal | pyrenylalanine |
ELISA | enzyme-linked immunosorbent assay |
IC50 | half maximal inhibitory concentration |
UV | ultraviolet |
HPLC | high-performance liquid chromatography |
MD | molecular dynamics |
RP-HPLC | reverse phase high-performance liquid chromatography |
RPE | retinal pigmented epithelium |
AMD | age-related macular degeneration |
References
This article references 74 other publications.
- 1Walport, M. J. Complement. Second of two parts N. Engl. J. Med. 2001, 344, 1140– 1144Google Scholar1Complement in immune responsesWalport, Mark J.New England Journal of Medicine (2001), 344 (15), 1140-1144CODEN: NEJMAG; ISSN:0028-4793. (Massachusetts Medical Society)A review with 96 refs. focusing on the complement system. Topics discussed include the complement at the interface between innate and adaptive immunity; the link between the complement and inflammation; the assocn. between the complement and necrosis; the relationship between the complement and immune complexes; the role of complement deficiency in systemic lupus erythematosus; and the hypothesis that the failure of the complement to dispose of the remnants of apoptotic cells may evoke an autoimmune response.
- 2Ricklin, D.; Lambris, J. D. Complement-targeted therapeutics Nat. Biotechnol. 2007, 25, 1265– 1275Google ScholarThere is no corresponding record for this reference.
- 3Ricklin, D.; Lambris, J. D. Complement in immune and inflammatory disorders: therapeutic interventions J. Immunol. 2013, 190, 3839– 3847Google ScholarThere is no corresponding record for this reference.
- 4Qu, H.; Ricklin, D.; Lambris, J. D. Recent developments in low molecular weight complement inhibitors Mol. Immunol. 2009, 47, 185– 195Google Scholar4Recent developments in low molecular weight complement inhibitorsQu, Hongchang; Ricklin, Daniel; Lambris, John D.Molecular Immunology (2009), 47 (2-3), 185-195CODEN: MOIMD5; ISSN:0161-5890. (Elsevier Ltd.)A review. As a key part of the innate immune system, complement plays an important role not only in defending against invading pathogens but also in many other biol. processes. Inappropriate or excessive activation of complement has been linked to many autoimmune, inflammatory, and neurodegenerative diseases, as well as ischemia-reperfusion injury and cancer. A wide array of low mol. wt. complement inhibitors has been developed to target various components of the complement cascade. Their efficacy has been demonstrated in numerous in vitro and in vivo expts. Though none of these inhibitors has reached the market so far, some of them have entered clin. trials and displayed promising results. This review provides a brief overview of the currently developed low mol. wt. complement inhibitors, including short peptides and synthetic small mols., with an emphasis on those targeting components C1 and C3, and the anaphylatoxin receptors.
- 5Sahu, A.; Kay, B. Inhibition of human complement by a C3-binding peptide isolated from a phage-displayed random peptide library J. Immunol. 1996, 157, 884– 891Google ScholarThere is no corresponding record for this reference.
- 6Morikis, D.; Sahu, A.; Moore, W. T.; Lambris, J. D. Design, structure, function and application of compstatin. In Bioactive Peptides in Drug Discovery and Design: Medical Aspects; Matsoukas, J.; Mavromoustakos, T., Eds.; IOS Press: Amsterdam, The Netherlands, 1999; pp 235– 246.Google ScholarThere is no corresponding record for this reference.
- 7Sahu, A.; Morikis, D.; Lambris, J. D. Complement inhibitors targeting C3, C4, and C5. In Therapeutic Interventions in the Complement System; Lambris, J. D.; Holers, V. M., Eds.; Humana Press: Totowa, NJ, U.S., 2000; pp 75– 112.Google ScholarThere is no corresponding record for this reference.
- 8Morikis, D.; Lambris, J. D. Structural aspects and design of low-molecular-mass complement inhibitors Biochem. Soc. Trans. 2002, 30, 1026– 1036Google Scholar8Structural aspects and design of low-molecular-mass complement inhibitorsMorikis, D.; Lambris, J. D.Biochemical Society Transactions (2002), 30 (6), 1026-1036CODEN: BCSTB5; ISSN:0300-5127. (Portland Press Ltd.)A review. We present a mini-review on the structure-based design of three promising complement inhibitors. Firstly, we review compstatin, a 13-residue cyclic peptide that binds to C3 and inhibits the cleavage of C3 to C3a and C3b. Secondly, we review a six-residue cyclic peptide that binds to C5aR and antagonizes the binding of C5a to its receptor C5aR. Finally, we review three small mols. that bind to Factor D and inhibit the enzymic action of Factor D, during which Factor D proteolytically cleaves Factor B in complex with C3 or C3b.
- 9Morikis, D.; Soulika, A. M.; Mallik, B.; Klepeis, J. L.; Floudas, C. A.; Lambris, J. D. Improvement of the anti-C3 activity of compstatin using rational and combinatorial approaches Biochem. Soc. Trans. 2004, 32, 28– 32Google Scholar9Improvement of the anti-C3 activity of compstatin using rational and combinatorial approachesMorikis, D.; Soulika, A. M.; Mallik, B.; Klepeis, J. L.; Floudas, C. A.; Lambris, J. D.Biochemical Society Transactions (2004), 32 (1), 28-32CODEN: BCSTB5; ISSN:0300-5127. (Portland Press Ltd.)A review. Compstatin is a 13-residue cyclic peptide that has the ability to inhibit the cleavage of C3 to C3a and C3b. The effects of targeting C3 cleavage are threefold, and result in hindrance of: (i) the generation of the pro-inflammatory peptide C3a, (ii) the generation of opsonin C3b (or its fragment C3d), and (iii) further complement activation of the common pathway (beyond C3) with the end result of the generation of the membrane attack complex. We will report on our progress on: (i) rational design of more active compstatin analogs based on the three-dimensional structure of compstatin, (ii) exptl. combinatorial design based on the generation of a phage-displayed peptide library partially randomized with the implementation of structure-induced restraints, and (iii) theor. combinatorial design based on a novel computational optimization method, structure-induced restraints and flexible structural templates. All three approaches have resulted in analogs with improved activities. Currently, the lead analog has the sequence acetyl-I[CVYQDWGAHRC]T-NH2 (where the brackets denote cyclization), and is 16-fold more active than the parent peptide. We will also report on our progress towards understanding the dynamic character of compstatin using mol. dynamics simulations. The identification of an ensemble of interconverting conformers of compstatin with variable populations is a first step towards the incorporation of dynamic elements in the design of new analogs using dynamics-activity relationships in addn. to structure-activity relationships.
- 10Holland, M. C. H.; Morikis, D.; Lambris, J. D. Synthetic small-molecule complement inhibitors Curr. Opin. Invest. Drugs 2004, 5, 1164– 1173Google ScholarThere is no corresponding record for this reference.
- 11Floudas, C.; Klepeis, J.; Lambris, J. D.; Morikis, D. De novo protein design: an interplay of global optimization, mixed-integer optimization and experiments. In FOCAPD 2004, Sixth International Conference on Foundations of Computer-Aided Process Design, Discovery through Product and Process Design; Floudas, C. A.; Agrawal, R., Eds.; CACHE Corporation: Austin, TX, U.S., 2004; pp 133– 146.Google ScholarThere is no corresponding record for this reference.
- 12Morikis, D.; Floudas, C.; Lambris, J. D. Structure-based integrative computational and experimental approach for the optimization of drug design. In ICCS 2005, Lecture Notes in Computer Science: Computational Science; Sunderam, V. S.; van Albada, G. D.; Sloot, P. M.; Dongarra, J., Eds.; Springer-Verlag: Berlin, Germany, 2005; pp 680– 688.Google ScholarThere is no corresponding record for this reference.
- 13Morikis, D.; Lambris, J. D. Structure, dynamics, activity, and function of compstatin and design of more potent analogs. In Structural Biology of the Complement System; Morikis, D.; Lambris, J. D., Eds.; CRC Press: Boca Raton, FL, U.S., 2005; pp 317– 340.Google ScholarThere is no corresponding record for this reference.
- 14Ricklin, D.; Lambris, J. D. Compstatin: a complement inhibitor on its way to clinical application. In Current Topics in Complement II: Advances in Experimental Medicine and Biology; Springer US: New York, NY, U.S., 2008; pp 262– 281.Google ScholarThere is no corresponding record for this reference.
- 15Morikis, D.; Assa-Munt, N.; Sahu, A.; Lambris, J. D. Solution structure of compstatin, a potent complement inhibitor Protein Sci. 1998, 7, 619– 627Google Scholar15Solution structure of Compstatin, a potent complement inhibitorMorikis, Dimitrios; Assa-Munt, Nuria; Sahu, Arvind; Lambris, John D.Protein Science (1998), 7 (3), 619-627CODEN: PRCIEI; ISSN:0961-8368. (Cambridge University Press)The third component of complement, C3, plays a central role in activation of the classical, alternative, and lectin pathways of complement activation. Recently, we have identified a 13-residue cyclic peptide (named Compstatin) that specifically binds to C3 and inhibits complement activation. To investigate the topol. and the contribution of each crit. residue to the binding of Compstatin to C3, we have now detd. the soln. structure using 2D NMR techniques; we have also synthesized substitution analogs and used these to study the structure-function relationships involved. Finally, we have generated an ensemble of a family of soln. structures of the peptide with a hybrid distance geometry-restrained simulated-annealing methodol., using distance, dihedral angle, and 3JNH-Hα-coupling const. restraints. The Compstatin structure contained a type I β-turn comprising the segment Gln5-Asp6-Trp7-Gly8. Preference for packing of the hydrophobic side chains of Val3, Val4, and Trp7 was obsd. The generated structure was also analyzed for consistency using NMR parameters such as NOE connectivity patterns, 3JNH-Hα-coupling consts., and chem. shifts. Anal. of Ala substitution analogs suggested that Val3, Gln5, Asp6, Trp7, and Gly8 contribute significantly to the inhibitory activity of the peptide. Substitution of Gly8 caused a 100-fold decrease in inhibitory potency. In contrast, substitution of Val4, His9, His10, and Arg11 resulted in minimal change in the activity. These findings indicate that specific side-chain interactions and the β-turn are crit. for preservation of the conformational stability of Compstatin and they might be significant for maintaining the functional activity of Compstatin.
- 16Sahu, A.; Soulika, A. M.; Morikis, D.; Spruce, L.; Moore, W. T.; Lambris, J. D. Binding kinetics, structure-activity relationship, and biotransformation of the complement inhibitor compstatin J. Immunol. 2000, 165, 2491– 2499Google ScholarThere is no corresponding record for this reference.
- 17Morikis, D.; Roy, M.; Sahu, A.; Troganis, A.; Jennings, P. A.; Tsokos, G. C.; Lambris, J. D. The structural basis of compstatin activity examined by structure-function-based design of peptide analogs and NMR J. Biol. Chem. 2002, 277, 14942– 14953Google ScholarThere is no corresponding record for this reference.
- 18Klepeis, J. L.; Floudas, C. A.; Morikis, D.; Tsokos, C. G.; Argyropoulos, E.; Spruce, L.; Lambris, J. D. Integrated computational and experimental approach for lead optimization and design of compstatin variants with improved activity J. Am. Chem. Soc. 2003, 125, 8422– 8423Google Scholar18Integrated Computational and Experimental Approach for Lead Optimization and Design of Compstatin Variants with Improved ActivityKlepeis, John L.; Floudas, Christodoulos A.; Morikis, Dimitrios; Tsokos, C. G.; Argyropoulos, E.; Spruce, L.; Lambris, John D.Journal of the American Chemical Society (2003), 125 (28), 8422-8423CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)A novel structure-activity-based combinatorial computational optimization methodol. for the design of peptides that are candidates to become therapeutics is presented. This methodol. has been successfully applied in the design of a 7-fold more active analog, among other active analogs, in the case of the complement inhibitor compstatin. The main steps of the approach involve the availability of NMR-derived structural templates, combinatorial selection of sequences based on optimization of parametrized pairwise residue interaction potentials, prediction of fold stabilities using deterministic global optimization, and exptl. validation with immunol. activity measurements. This work is direct evidence that an integrated exptl. and theor. approach can make the engineering of compds. with enhanced immunol. properties possible.
- 19Soulika, A. M.; Morikis, D.; Sarrias, M.-R.; Roy, M.; Spruce, L. A.; Sahu, A.; Lambris, J. D. Studies of structure-activity relations of complement inhibitor compstatin J. Immunol. 2003, 171, 1881– 1890Google ScholarThere is no corresponding record for this reference.
- 20Mallik, B.; Katragadda, M.; Spruce, L. A.; Carafides, C.; Tsokos, C. G.; Morikis, D.; Lambris, J. D. Design and NMR characterization of active analogues of compstatin containing non-natural amino acids J. Med. Chem. 2005, 48, 274– 286Google ScholarThere is no corresponding record for this reference.
- 21Katragadda, M.; Lambris, J. D. Expression of compstatin in Escherichia coli: incorporation of unnatural amino acids enhances its activity Protein Expression Purif. 2006, 47, 289– 295Google Scholar21Expression of compstatin in Escherichia coli: Incorporation of unnatural amino acids enhances its activityKatragadda, Madan; Lambris, John D.Protein Expression and Purification (2006), 47 (1), 289-295CODEN: PEXPEJ; ISSN:1046-5928. (Elsevier)Compstatin, a 13-residue cyclic peptide, is a complement inhibitor that shows therapeutic potential. Several previous approaches have improved the activity of this peptide several-fold. In the present study, we have expressed and purified compstatin from Escherichia coli in an effort to increase its potency and to generate it in high yield in a more economical fashion. An intein-based expression system was used to express compstatin in fusion with chitin-binding domain and Ssp DnaB intein, which were later cleaved from the expressed mol. at room temp. and pH 7.0 to yield pure compstatin in one step. The expressed compstatin showed activity similar to the synthetic compstatin in an ELISA-based assay. The same expression system and purifn. strategy were used to incorporate three tryptophan analogs, 6-fluoro-tryptophan, 5-hydroxy-tryptophan, and 7-aza-tryptophan, into compstatin. Interestingly, incorporation of 6-fluoro-tryptophan increased the activity three-fold relative to wild-type compstatin; in contrast, incorporation of 5-hydroxy- or 7-aza-tryptophan rendered compstatin less active than the wild-type form.
- 22Katragadda, M.; Magotti, P.; Sfyroera, G.; Lambris, J. D. Hydrophobic effect and hydrogen bonds account for the improved activity of a complement inhibitor, compstatin J. Med. Chem. 2006, 49, 4616– 4622Google ScholarThere is no corresponding record for this reference.
- 23Magotti, P.; Ricklin, D.; Qu, H.; Wu, Y.-Q.; Kaznessis, Y. N.; Lambris, J. D. Structure-kinetic relationship analysis of the therapeutic complement inhibitor compstatin J. Mol. Recognit. 2009, 22, 495– 505Google Scholar23Structure-kinetic relationship analysis of the therapeutic complement inhibitor compstatinMagotti, Paola; Ricklin, Daniel; Qu, Hongchang; Wu, You-Qiang; Kaznessis, Yiannis N.; Lambris, John D.Journal of Molecular Recognition (2009), 22 (6), 495-505CODEN: JMORE4; ISSN:0952-3499. (John Wiley & Sons Ltd.)Compstatin is a 13-residue peptide that inhibits activation of the complement system by binding to the central component C3 and its fragments C3b and C3c. A combination of theor. and exptl. approaches has previously allowed us to develop analogs of the original compstatin peptide with up to 264-fold higher activity; one of these analogs is now in clin. trials for the treatment of age-related macular degeneration (AMD). Here we used functional assays, surface plasmon resonance (SPR), and isothermal titrn. calorimetry (ITC) to assess the effect of modifications at three key residues (Trp-4, Asp-6, Ala-9) on the affinity and activity of compstatin and its analogs, and we correlated our findings to the recently reported co-crystal structure of compstatin and C3c. The KD values for the panel of tested analogs ranged from 10-6 to 10-8 M. These differences in binding affinity could be attributed mainly to differences in dissocn. rather than assocn. rates, with a >4-fold range in kon values (2-10 × 105 M-1 s-1) and a koff variation of >35-fold (1-37 × 10-2 s-1) being obsd. The stability of the C3b-compstatin complex seemed to be highly dependent on hydrophobic effects at position 4, and even small changes at position 6 resulted in a loss of complex formation. Induction of a β-turn shift by an A9P modification resulted in a more favorable entropy but a loss of binding specificity and stability. The results obtained by the three methods utilized here were highly correlated with regard to the activity/affinity of the analogs. Thus, our analyses have identified essential structural features of compstatin and provided important information to support the development of analogs with improved efficacy. Copyright © 2009 John Wiley & Sons, Ltd.
- 24Bellows, M. L.; Fung, H. K.; Taylor, M. S.; Floudas, C. A.; de Victoria, A. L.; Morikis, D. New compstatin variants through two de novo protein design frameworks Biophys. J. 2010, 98, 2337– 2346Google ScholarThere is no corresponding record for this reference.
- 25López de Victoria, A.; Gorham, R. D.; Bellows-Peterson, M. L.; Ling, J.; Lo, D. D.; Floudas, C. A.; Morikis, D. A new generation of potent complement inhibitors of the compstatin family Chem. Biol. Drug Des. 2011, 77, 431– 440Google Scholar25A new generation of potent complement inhibitors of the Compstatin familyLopez de Victoria, Aliana; Gorham, Ronald D., Jr.; Bellows-Peterson, Meghan L.; Ling, Jun; Lo, David D.; Floudas, Christodoulos A.; Morikis, DimitriosChemical Biology & Drug Design (2011), 77 (6), 431-440CODEN: CBDDAL; ISSN:1747-0277. (Wiley-Blackwell)Compstatin family peptides are potent inhibitors of the complement system and promising drug candidates against diseases involving under-regulated complement activation. Compstatin is a 13-residue cyclized peptide that inhibits cleavage of complement protein C3, preventing downstream complement activation. We present three new compstatin variants, characterized by tryptophan replacement at positions 1 and/or 13. Peptide design was based on physicochem. reasoning and was inspired by earlier work, which identified tryptophan substitutions at positions 1 and 13 in peptides with predicted C3c binding abilities. The new variants preserve distinct polar and nonpolar surfaces of compstatin, but have altered local interaction capabilities with C3. All three peptides exhibited potent C3 binding by surface plasmon resonance and potent complement inhibition by ELISA. We also present ELISA data and detailed surface plasmon resonance kinetic data of three peptides from previous computational design.
- 26Qu, H.; Magotti, P.; Ricklin, D.; Wu, E. L.; Kourtzelis, I.; Wu, Y.-Q.; Kaznessis, Y. N.; Lambris, J. D. Novel analogues of the therapeutic complement inhibitor compstatin with significantly improved affinity and potency Mol. Immunol. 2011, 48, 481– 489Google Scholar26Novel analogues of the therapeutic complement inhibitor compstatin with significantly improved affinity and potencyQu Hongchang; Magotti Paola; Ricklin Daniel; Wu Emilia L; Kourtzelis Ioannis; Wu You-Qiang; Kaznessis Yiannis N; Lambris John DMolecular immunology (2011), 48 (4), 481-9 ISSN:.Compstatin is a 13-residue disulfide-bridged peptide that inhibits a key step in the activation of the human complement system. Compstatin and its derivatives have shown great promise for the treatment of many clinical disorders associated with unbalanced complement activity. To obtain more potent compstatin analogues, we have now performed an N-methylation scan of the peptide backbone and amino acid substitutions at position 13. One analogue (Ac-I[CVW(Me)QDW-Sar-AHRC](NMe)I-NH(2)) displayed a 1000-fold increase in both potency (IC(50) = 62 nM) and binding affinity for C3b (K(D) = 2.3 nM) over that of the original compstatin. Biophysical analysis using surface plasmon resonance and isothermal titration calorimetry suggests that the improved binding originates from more favorable free conformation and stronger hydrophobic interactions. This study provides a series of significantly improved drug leads for therapeutic applications in complement-related diseases, and offers new insights into the structure-activity relationships of compstatin analogues.
- 27Tamamis, P.; López de Victoria, A.; Gorham, R. D.; Bellows-Peterson, M. L.; Pierou, P.; Floudas, C. A.; Morikis, D.; Archontis, G. Molecular dynamics in drug design: new generations of compstatin analogs Chem. Biol. Drug Des. 2012, 79, 703– 718Google Scholar27Molecular dynamics in drug design: new generations of compstatin analogsTamamis, Phanourios; Lopez de Victoria, Aliana; Gorham, Ronald D., Jr.; Bellows-Peterson, Meghan L.; Pierou, Panayiota; Floudas, Christodoulos A.; Morikis, Dimitrios; Archontis, GeorgiosChemical Biology & Drug Design (2012), 79 (5), 703-718CODEN: CBDDAL; ISSN:1747-0277. (Wiley-Blackwell)We report the computational and rational design of new generations of potential peptide-based inhibitors of the complement protein C3 from the compstatin family. The binding efficacy of the peptides is tested by extensive mol. dynamics-based structural and physicochem. anal., using 32 at. detail trajectories in explicit water for 22 peptides bound to human, rat or mouse target protein C3, with a total of 257 ns. The criteria for the new design are: (i) optimization for C3 affinity and for the balance between hydrophobicity and polarity to improve soly. compared to known compstatin analogs; and (ii) development of dual specificity, human-rat/mouse C3 inhibitors, which could be used in animal disease models. Three of the new analogs are analyzed in more detail as they possess strong and novel binding characteristics and are promising candidates for further optimization. This work paves the way for the development of an improved therapeutic for age-related macular degeneration, and other complement system-mediated diseases, compared to known compstatin variants.
- 28Gorham, R. D.; Forest, D. L.; Tamamis, P.; López de Victoria, A.; Kraszni, M.; Kieslich, C. A.; Banna, C. D.; Bellows-Peterson, M. L.; Larive, C. K.; Floudas, C. A.; Archontis, G.; Johnson, L. V.; Morikis, D. Novel compstatin family peptides inhibit complement activation by drusen-like deposits in human retinal pigmented epithelial cell cultures Exp. Eye Res. 2013, 116C, 96– 108Google ScholarThere is no corresponding record for this reference.
- 29Qu, H.; Ricklin, D.; Bai, H.; Chen, H.; Reis, E. S.; Maciejewski, M.; Tzekou, A.; DeAngelis, R. A.; Resuello, R. R. G.; Lupu, F.; Barlow, P. N.; Lambris, J. D. New analogs of the clinical complement inhibitor compstatin with subnanomolar affinity and enhanced pharmacokinetic properties Immunobiology 2013, 218, 496– 505Google ScholarThere is no corresponding record for this reference.
- 30Risitano, A. M.; Ricklin, D.; Huang, Y.; Reis, E. S.; Chen, H.; Ricci, P.; Lin, Z.; Pascariello, C.; Raia, M.; Sica, M.; del Vecchio, L.; Pane, F.; Lupu, F.; Notaro, R.; Resuello, R. R. G.; Deangelis, R. A.; Lambris, J. D. Peptide inhibitors of C3 activation as a novel strategy of complement inhibition for the treatment of paroxysmal nocturnal hemoglobinuria Blood 2014, 123, 2094– 2101Google ScholarThere is no corresponding record for this reference.
- 31Janssen, B. J. C.; Halff, E. F.; Lambris, J. D.; Gros, P. Structure of compstatin in complex with complement component C3c reveals a new mechanism of complement inhibition J. Biol. Chem. 2007, 282, 29241– 29247Google ScholarThere is no corresponding record for this reference.
- 32Klepeis, J. L.; Floudas, C. A.; Morikis, D.; Tsokos, C. G.; Lambris, J. D. Design of peptide analogues with improved activity using a novel de novo protein design approach Ind. Eng. Chem. Res. 2004, 43, 3817– 3826Google ScholarThere is no corresponding record for this reference.
- 33Walensky, L. D.; Bird, G. H. Hydrocarbon-stapled peptides: principles, practice, and progress J. Med. Chem. 2014, 57, 6275– 6288Google Scholar33Hydrocarbon-Stapled Peptides: Principles, Practice, and ProgressWalensky, Loren D.; Bird, Gregory H.Journal of Medicinal Chemistry (2014), 57 (15), 6275-6288CODEN: JMCMAR; ISSN:0022-2623. (American Chemical Society)A review. Protein structure underlies essential biol. processes and provides a blueprint for mol. mimicry that drives drug discovery. Although small mols. represent the lion's share of agents that target proteins for therapeutic benefit, there remains no substitute for the natural properties of proteins and their peptide subunits in the majority of biol. contexts. The peptide α-helix represents a common structural motif that mediates communication between signaling proteins. Because peptides can lose their shape when taken out of context, developing chem. interventions to stabilize their bioactive structure remains an active area of research. The all-hydrocarbon staple has emerged as one such soln., conferring α-helical structure, protease resistance, cellular penetrance, and biol. activity upon successful incorporation of a series of design and application principles. Here, we describe our more than decade-long experience in developing stapled peptides as biomedical research tools and prototype therapeutics, highlighting lessons learned, pitfalls to avoid, and keys to success.
- 34Chi, Z.-L.; Yoshida, T.; Lambris, J. D.; Iwata, T. Suppression of drusen formation by compstatin, a peptide inhibitor of complement C3 activation, on cynomolgus monkey with early-onset macular degeneration Adv. Exp. Med. Biol. 2010, 703, 127– 135Google ScholarThere is no corresponding record for this reference.
- 35Yehoshua, Z.; Rosenfeld, P. J.; Albini, T. A. Current clinical trials in dry AMD and the definition of appropriate clinical outcome measures Semin. Ophthalmol. 2011, 26, 167– 180Google Scholar35Current Clinical Trials in Dry AMD and the Definition of Appropriate Clinical Outcome MeasuresYehoshua Zohar; Rosenfeld Philip J; Albini Thomas ASeminars in ophthalmology (2011), 26 (3), 167-80 ISSN:.Currently, there is no proven drug treatment for dry age-related macular degeneration (AMD). Several different treatment strategies are being investigated, including complement inhibition, neuroprotection, and visual cycle inhibitors, and novel clinical trial endpoints are being explored. Studies have identified genetic predispositions for dry AMD associated with complement dysfunction. Consequently, complement-based therapeutic treatment modalities are promising.
- 36Hilbich, C.; Kisters-Woike, B.; Reed, J.; Masters, C. L.; Beyreuther, K. Substitutions of hydrophobic amino acids reduce the amyloidogenicity of Alzheimer’s disease beta A4 peptides J. Mol. Biol. 1992, 228, 460– 473Google ScholarThere is no corresponding record for this reference.
- 37Knerr, P. J.; Tzekou, A.; Ricklin, D.; Qu, H.; Chen, H.; van der Donk, W. A.; Lambris, J. D. Synthesis and activity of thioether-containing analogues of the complement inhibitor compstatin ACS Chem. Biol. 2011, 6, 753– 760Google ScholarThere is no corresponding record for this reference.
- 38Smadbeck, J.; Peterson, M. B.; Khoury, G. A.; Taylor, M. S.; Floudas, C. A. Protein WISDOM: a workbench for in silico de novo design of biomolecules J. Visualized Exp. 2013, e50476– e50476Google ScholarThere is no corresponding record for this reference.
- 39Rajgaria, R.; McAllister, S. R.; Floudas, C. A. Distance dependent centroid to centroid force fields using high resolution decoys Proteins 2008, 70, 950– 970Google ScholarThere is no corresponding record for this reference.
- 40Güntert, P. Automated NMR structure calculation with CYANA Methods Mol. Biol. 2004, 278, 353– 378Google Scholar40Automated NMR structure calculation with CYANAGuntert, PeterMethods in Molecular Biology (Totowa, NJ, United States) (2004), 278 (Protein NMR Techniques), 353-378CODEN: MMBIED; ISSN:1064-3745. (Humana Press Inc.)This chapter gives an introduction to automated NMR structure calcn. with the program CYANA. Given a sufficiently complete list of assigned chem. shifts and one or several lists of cross-peak positions and columns from two-, three-, or four-dimensional nuclear Overhauser effect spectroscopy (NOESY) spectra, the assignment of the NOESY cross-peaks and the three-dimensional structure of the protein in soln. can be calcd. automatically with CYANA.
- 41Güntert, P.; Mumenthaler, C.; Wüthrich, K. Torsion angle dynamics for NMR structure calculation with the new program DYANA J. Mol. Biol. 1997, 273, 283– 298Google Scholar41Torsion angle dynamics for NMR structure calculation with the new program DYANAGuntert, P.; Mumenthaler, C.; Wuthrich, K.Journal of Molecular Biology (1997), 273 (1), 283-298CODEN: JMOBAK; ISSN:0022-2836. (Academic)The new program DYANA (DYnamics Algorithm for Nmr Applications) for efficient calcn. of three-dimensional protein and nucleic acid structures from distance constraints and torsion angle constraints collected by NMR expts. performs simulated annealing by mol. dynamics in torsion angle space and uses a fast recursive algorithm to integrate the equations of motions. Torsion angle dynamics can be more efficient than mol. dynamics in Cartesian coordinate space because of the reduced no. of degrees of freedom and the concomitant absence of high-frequency bond and angle vibrations, which allows for the use of longer time-steps and/or higher temps. in the structure calcn. It also represents a significant advance over the variable target function method in torsion angle space with the REDAC strategy used by the predecessor program DIANA. DYANA computation times per accepted conformer in the "bundle" used to represent the NMR structure compare favorably with those of other presently available structure calcn. algorithms, and are of the order of 160 s for a protein of 165 amino acid residues when using a DEC Alpha 8400 5/300 computer. Test calcns. starting from conformers with random torsion angle values further showed that DYANA is capable of efficient calcn. of high-quality protein structures with up to 400 amino acid residues, and of nucleic acid structures.
- 42Ponder, J. W. TINKER, Software Tools for Molecular Design. Jay Ponder Lab, Washington Unversity: St. Louis, MO, 2004.Google ScholarThere is no corresponding record for this reference.
- 43Cornell, W. D.; Cieplak, P.; Bayly, C. I.; Gould, I. R.; Merz, K. M.; Ferguson, D. M.; Spellmeyer, D. C.; Fox, T.; Caldwell, J. W.; Kollman, P. A. A second generation force field for the simulation of proteins, nucleic acids, and organic molecules J. Am. Chem. Soc. 1995, 117, 5179– 5197Google Scholar43A Second Generation Force Field for the Simulation of Proteins, Nucleic Acids, and Organic MoleculesCornell, Wendy D.; Cieplak, Piotr; Bayly, Christopher I.; Gould, Ian R.; Merz, Kenneth M., Jr.; Ferguson, David M.; Spellmeyer, David C.; Fox, Thomas; Caldwell, James W.; Kollman, Peter A.Journal of the American Chemical Society (1995), 117 (19), 5179-97CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The authors present the derivation of a new mol. mech. force field for simulating the structures, conformational energies, and interaction energies of proteins, nucleic acids, and many related org. mols. in condensed phases. This effective two-body force field is the successor to the Weiner et al. force field and was developed with some of the same philosophies, such as the use of a simple diagonal potential function and electrostatic potential fit atom centered charges. The need for a 10-12 function for representing hydrogen bonds is no longer necessary due to the improved performance of the new charge model and new van der Waals parameters. These new charges are detd. using a 6-31G* basis set and restrained electrostatic potential (RESP) fitting and have been shown to reproduce interaction energies, free energies of solvation, and conformational energies of simple small mols. to a good degree of accuracy. Furthermore, the new RESP charges exhibit less variability as a function of the mol. conformation used in the charge detn. The new van der Waals parameters have been derived from liq. simulations and include hydrogen parameters which take into account the effects of any geminal electroneg. atoms. The bonded parameters developed by Weiner et al. were modified as necessary to reproduce exptl. vibrational frequencies and structures. Most of the simple dihedral parameters have been retained from Weiner et al., but a complex set of .vphi. and ψ parameters which do a good job of reproducing the energies of the low-energy conformations of glycyl and alanyl dipeptides was developed for the peptide backbone.
- 44Lilien, R. H.; Stevens, B. W.; Anderson, A. C.; Donald, B. R. A novel ensemble-based scoring and search algorithm for protein redesign and its application to modify the substrate specificity of the gramicidin synthetase a phenylalanine adenylation enzyme J. Comput. Biol. 2005, 12, 740– 761Google Scholar44A Novel Ensemble-Based Scoring and Search Algorithm for Protein Redesign and Its Application to Modify the Substrate Specificity of the Gramicidin Synthetase A Phenylalanine Adenylation EnzymeLilien, Ryan H.; Stevens, Brian W.; Anderson, Amy C.; Donald, Bruce R.Journal of Computational Biology (2005), 12 (6), 740-761CODEN: JCOBEM; ISSN:1066-5277. (Mary Ann Liebert, Inc.)Realization of novel mol. function requires the ability to alter mol. complex formation. Enzymic function can be altered by changing enzyme-substrate interactions via modification of an enzyme's active site. A redesigned enzyme may either perform a novel reaction on its native substrates or its native reaction on novel substrates. A no. of computational approaches have been developed to address the combinatorial nature of the protein redesign problem. These approaches typically search for the global min. energy conformation among an exponential no. of protein conformations. We present a novel algorithm for protein redesign, which combines a statistical mechanics-derived ensemble-based approach to computing the binding const. with the speed and completeness of a branch-and-bound pruning algorithm. In addn., we developed an efficient deterministic approxn. algorithm, capable of approximating our scoring function to arbitrary precision. In practice, the approxn. algorithm decreases the execution time of the mutation search by a factor of ten. To test our method, we examd. the Phe-specific adenylation domain of the nonribosomal peptide synthetase gramicidin synthetase A (GrsA-PheA). Ensemble scoring, using a rotameric approxn. to the partition functions of the bound and unbound states for GrsA-PheA, is first used to predict binding of the wild-type protein and a previously described mutant (selective for leucine), and second, to switch the enzyme specificity toward leucine, using two novel active site sequences computationally predicted by searching through the space of possible active site mutations. The top scoring in silico mutants were created in the wet-lab. and dissocn./binding consts. were detd. by fluorescence quenching. These tested mutations exhibit the desired change in specificity from Phe to Leu. Our ensemble-based algorithm, which flexibly models both protein and ligand using rotamer-based partition functions, has application in enzyme redesign, the prediction of protein-ligand binding, and computer-aided drug design.
- 45Lee, M. R.; Baker, D.; Kollman, P. A. 2.1 and 1.8 Å average Cα rmsd structure predictions on two small proteins, HP-36 and S15 J. Am. Chem. Soc. 2001, 123, 1040– 1046Google ScholarThere is no corresponding record for this reference.
- 46Rohl, C. A.; Baker, D. De novo determination of protein backbone structure from residual dipolar couplings using Rosetta J. Am. Chem. Soc. 2002, 124, 2723– 2729Google ScholarThere is no corresponding record for this reference.
- 47Rohl, C. A.; Strauss, C. E. M.; Misura, K. M. S.; Baker, D. Protein structure prediction using Rosetta. In Methods in Enzymology; Elsevier: San Diego, CA, 2004; Vol. 383, pp 66– 93.Google ScholarThere is no corresponding record for this reference.
- 48DiMaggio, P. A.; McAllister, S. R.; Floudas, C. A.; Feng, X.-J.; Rabinowitz, J. D.; Rabitz, H. A. Biclustering via optimal re-ordering of data matrices in systems biology: rigorous methods and comparative studies BMC Bioinf. 2008, 9, 458Google Scholar48Biclustering via optimal re-ordering of data matrices in systems biology: rigorous methods and comparative studiesDiMaggio Peter A Jr; McAllister Scott R; Floudas Christodoulos A; Feng Xiao-Jiang; Rabinowitz Joshua D; Rabitz Herschel ABMC bioinformatics (2008), 9 (), 458 ISSN:.BACKGROUND: The analysis of large-scale data sets via clustering techniques is utilized in a number of applications. Biclustering in particular has emerged as an important problem in the analysis of gene expression data since genes may only jointly respond over a subset of conditions. Biclustering algorithms also have important applications in sample classification where, for instance, tissue samples can be classified as cancerous or normal. Many of the methods for biclustering, and clustering algorithms in general, utilize simplified models or heuristic strategies for identifying the "best" grouping of elements according to some metric and cluster definition and thus result in suboptimal clusters. RESULTS: In this article, we present a rigorous approach to biclustering, OREO, which is based on the Optimal RE-Ordering of the rows and columns of a data matrix so as to globally minimize the dissimilarity metric. The physical permutations of the rows and columns of the data matrix can be modeled as either a network flow problem or a traveling salesman problem. Cluster boundaries in one dimension are used to partition and re-order the other dimensions of the corresponding submatrices to generate biclusters. The performance of OREO is tested on (a) metabolite concentration data, (b) an image reconstruction matrix, (c) synthetic data with implanted biclusters, and gene expression data for (d) colon cancer data, (e) breast cancer data, as well as (f) yeast segregant data to validate the ability of the proposed method and compare it to existing biclustering and clustering methods. CONCLUSION: We demonstrate that this rigorous global optimization method for biclustering produces clusters with more insightful groupings of similar entities, such as genes or metabolites sharing common functions, than other clustering and biclustering algorithms and can reconstruct underlying fundamental patterns in the data for several distinct sets of data matrices arising in important biological applications.
- 49DiMaggio, P. A., Jr; McAllister, S. R.; Floudas, C. A.; Feng, X.-J.; Rabinowitz, J. D.; Rabitz, H. A. A network flow model for biclustering via optimal re-ordering of data matrices J. Global Optim. 2010, 47, 343– 354Google ScholarThere is no corresponding record for this reference.
- 50Gray, J. J.; Moughon, S.; Wang, C.; Schueler-Furman, O.; Kuhlman, B.; Rohl, C. A.; Baker, D. Protein-protein docking with simultaneous optimization of rigid-body displacement and side-chain conformations J. Mol. Biol. 2003, 331, 281– 299Google Scholar50Protein-protein docking with simultaneous optimization of rigid-body displacement and side-chain conformationsGray, Jeffrey J.; Moughon, Stewart; Wang, Chu; Schueler-Furman, Ora; Kuhlman, Brian; Rohl, Carol A.; Baker, DavidJournal of Molecular Biology (2003), 331 (1), 281-299CODEN: JMOBAK; ISSN:0022-2836. (Elsevier Science Ltd.)Protein-protein docking algorithms provide a means to elucidate structural details for presently unknown complexes. Here, we present and evaluate a new method to predict protein-protein complexes from the coordinates of the unbound monomer components. The method employs a low-resoln., rigid-body, Monte Carlo search followed by simultaneous optimization of backbone displacement and side-chain conformations using Monte Carlo minimization. Up to 105 independent simulations are carried out, and the resulting "decoys" are ranked using an energy function dominated by van der Waals interactions, an implicit solvation model, and an orientation-dependent hydrogen bonding potential. Top-ranking decoys are clustered to select the final predictions. Small-perturbation studies reveal the formation of binding funnels in 42 of 54 cases using coordinates derived from the bound complexes and in 32 of 54 cases using independently detd. coordinates of one or both monomers. Exptl. binding affinities correlate with the calcd. score function and explain the predictive success or failure of many targets. Global searches using one or both unbound components predict at least 25% of the native residue-residue contacts in 28 of the 32 cases where binding funnels exist. The results suggest that the method may soon be useful for generating models of biol. important complexes from the structures of the isolated components, but they also highlight the challenges that must be met to achieve consistent and accurate prediction of protein-protein interactions.
- 51Kuhlman, B.; Baker, D. Native protein sequences are close to optimal for their structures Proc. Natl. Acad. Sci. U.S.A. 2000, 97, 10383– 10388Google Scholar51Native protein sequences are close to optimal for their structuresKuhlman, Brian; Baker, DavidProceedings of the National Academy of Sciences of the United States of America (2000), 97 (19), 10383-10388CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)How large is the vol. of sequence space that is compatible with a given protein structure. Starting from random sequences, low free energy sequences were generated for 108 protein backbone structures by using a Monte Carlo optimization procedure and a free energy function based primarily on Lennard-Jones packing interactions and the Lazaridis-Karplus implicit solvation model. Remarkably, in the designed sequences 51% of the core residues and 27% of all residues were identical to the amino acids in the corresponding positions in the native sequences. The lowest free energy sequences obtained for ensembles of native-like backbone structures were also similar to the native sequence. Furthermore, both the individual residue frequencies and the covariances between pairs of positions obsd. in the very large SH3 domain family were recapitulated in core sequences designed for SH3 domain structures. Taken together, these results suggest that the vol. of sequence space optimal for a protein structure is surprisingly restricted to a region around the native sequence.
- 52Drew, K.; Renfrew, P. D.; Craven, T. W.; Butterfoss, G. L.; Chou, F.-C.; Lyskov, S.; Bullock, B. N.; Watkins, A.; Labonte, J. W.; Pacella, M.; Kilambi, K. P.; Leaver-Fay, A.; Kuhlman, B.; Gray, J. J.; Bradley, P.; Kirshenbaum, K.; Arora, P. S.; Das, R.; Bonneau, R. Adding diverse noncanonical backbones to Rosetta: enabling peptidomimetic design PLoS One 2013, 8, e67051Google ScholarThere is no corresponding record for this reference.
- 53Renfrew, P. D.; Choi, E. J.; Bonneau, R.; Kuhlman, B. Incorporation of noncanonical amino acids into Rosetta and use in computational protein-peptide interface design PLoS One 2012, 7, e32637Google Scholar53Incorporation of noncanonical amino acids into Rosetta and use in computational protein-peptide interface designRenfrew, P. Douglas; Choi, Eun Jung; Bonneau, Richard; Kuhlman, BrianPLoS One (2012), 7 (3), e32637CODEN: POLNCL; ISSN:1932-6203. (Public Library of Science)Noncanonical amino acids (NCAAs) can be used in a variety of protein design contexts. For example, they can be used in place of the canonical amino acids (CAAs) to improve the biophys. properties of peptides that target protein interfaces. We describe the incorporation of 114 NCAAs into the protein-modeling suite Rosetta. We describe our methods for building backbone dependent rotamer libraries and the parameterization and construction of a scoring function that can be used to score NCAA contg. peptides and proteins. We validate these addns. to Rosetta and our NCAA-rotamer libraries by showing that we can improve the binding of a calpastatin derived peptides to calpain-1 by substituting NCAAs for native amino acids using Rosetta. Rosetta (executables and source), auxiliary scripts and code, and documentation can be found at online.
- 54Mills, J. H.; Khare, S. D.; Bolduc, J. M.; Forouhar, F.; Mulligan, V. K.; Lew, S.; Seetharaman, J.; Tong, L.; Stoddard, B. L.; Baker, D. Computational design of an unnatural amino acid dependent metalloprotein with atomic level accuracy J. Am. Chem. Soc. 2013, 135, 13393– 13399Google Scholar54Computational Design of an Unnatural Amino Acid Dependent Metalloprotein with Atomic Level AccuracyMills, Jeremy H.; Khare, Sagar D.; Bolduc, Jill M.; Forouhar, Farhad; Mulligan, Vikram Khipple; Lew, Scott; Seetharaman, Jayaraman; Tong, Liang; Stoddard, Barry L.; Baker, DavidJournal of the American Chemical Society (2013), 135 (36), 13393-13399CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Genetically encoded unnatural amino acids could facilitate the design of proteins and enzymes of novel function, but correctly specifying sites of incorporation and the identities and orientations of surrounding residues represents a formidable challenge. Computational design methods have been used to identify optimal locations for functional sites in proteins and design the surrounding residues but have not incorporated unnatural amino acids in this process. The authors extended the Rosetta design methodol. to design metalloproteins in which the amino acid (2,2'-bipyridin-5-yl)-alanine (Bpy-Ala) is a primary ligand of a bound metal ion. Following initial results that indicated the importance of buttressing the Bpy-Ala amino acid, the authors designed a buried metal binding site with octahedral coordination geometry consisting of Bpy-Ala, two protein-based metal ligands, and two metal-bound water mols. Exptl. characterization revealed a Bpy-Ala-mediated metalloprotein with the ability to bind divalent cations including Co2+, Zn2+, Fe2+, and Ni2+, with a Kd for Zn2+ of ∼40 pM. X-ray crystal structures of the designed protein bound to Co2+ and Ni2+ have RMSDs to the design model of 0.9 and 1.0 Å resp. over all atoms in the binding site.
- 55Yu, H.; Daura, X.; van Gunsteren, W. F. Molecular dynamics simulations of peptides containing an unnatural amino acid: dimerization, folding, and protein binding Proteins 2004, 54, 116– 127Google ScholarThere is no corresponding record for this reference.
- 56Daura, X.; van Gunsteren, W. F.; Mark, A. E. Folding-unfolding thermodynamics of a beta-heptapeptide from equilibrium simulations Proteins 1999, 34, 269– 280Google ScholarThere is no corresponding record for this reference.
- 57Daura, X.; Gademann, K.; Schäfer, H.; Jaun, B.; Seebach, D.; van Gunsteren, W. F. The β-peptide hairpin in solution: conformational study of a β-hexapeptide in methanol by NMR spectroscopy and MD simulation J. Am. Chem. Soc. 2001, 123, 2393– 2404Google Scholar57The β-Peptide Hairpin in Solution: Conformational Study of a β-Hexapeptide in Methanol by NMR Spectroscopy and MD SimulationDaura, Xavier; Gademann, Karl; Schaefer, Heiko; Jaun, Bernhard; Seebach, Dieter; van Gunsteren, Wilfred F.Journal of the American Chemical Society (2001), 123 (10), 2393-2404CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The structural and thermodn. properties of a 6-residue β-peptide I that was designed to form a hairpin conformation have been studied by NMR spectroscopy and mol. dynamics simulation in methanol soln. The predicted hairpin would be characterized by a 10-membered hydrogen-bonded turn involving residues 3 and 4, and two extended antiparallel strands. The interproton distances and backbone torsional dihedral angles derived from the NMR expts. at room temp. are in general terms compatible with the hairpin conformation. Two trajectories of system configurations from 100-ns mol.-dynamics simulations of the peptide in soln. at 298 and 340 K have been analyzed. In both simulations, reversible folding to the hairpin conformation is obsd. Interestingly, there is a significant conformational overlap between the unfolded state of the peptide at each of the temps. As already obsd. in previous studies of peptide folding, the unfolded state is composed of a (relatively) small no. of predominant conformers and in this case lacks any type of secondary-structure element. The trajectories provide an excellent ground for the interpretation of the NMR-derived data in terms of ensemble avs. and distributions as opposed to single-conformation interpretations. From this perspective, a relative population of the hairpin conformation of 20% to 30% would suffice to explain the NMR-derived data. Surprisingly, however, the ensemble of structures from the simulation at 340 K reproduces more accurately the NMR-derived data than the ensemble from the simulation at 298 K, and this point needs further investigation.
- 58Schäfer, H.; Daura, X.; Mark, A. E.; van Gunsteren, W. F. Entropy calculations on a reversibly folding peptide: changes in solute free energy cannot explain folding behavior Proteins 2001, 43, 45– 56Google ScholarThere is no corresponding record for this reference.
- 59Rathore, N.; Gellman, S. H.; de Pablo, J. J. Thermodynamic stability of β-peptide helices and the role of cyclic residues Biophys. J. 2006, 91, 3425– 3435Google ScholarThere is no corresponding record for this reference.
- 60McGovern, M.; Abbott, N.; de Pablo, J. J. Dimerization of helical β-peptides in solution Biophys. J. 2012, 102, 1435– 1442Google ScholarThere is no corresponding record for this reference.
- 61Khoury, G. A.; Thompson, J. P.; Smadbeck, J.; Kieslich, C. A.; Floudas, C. A. Forcefield_PTM: ab initio charge and AMBER forcefield parameters for frequently occurring post-translational modifications J. Chem. Theory Comput. 2013, 9, 5653– 5674Google Scholar61Forcefield_PTM: Ab Initio Charge and AMBER Forcefield Parameters for Frequently Occurring Post-Translational ModificationsKhoury, George A.; Thompson, Jeff P.; Smadbeck, James; Kieslich, Chris A.; Floudas, Christodoulos A.Journal of Chemical Theory and Computation (2013), 9 (12), 5653-5674CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)The authors introduce Forcefield_PTM, a set of AMBER forcefield parameters consistent with ff03 for 32 common post-translational modifications. Partial charges were calcd. through ab initio calcns. and a two-stage RESP-fitting procedure in an ether-like implicit solvent environment. The charges are generally consistent with others previously reported for phosphorylated amino acids, and trimethyllysine, using different parametrization methods. Pairs of modified structures and their corresponding unmodified structures were curated from the PDB for both single and multiple modifications. Background structural similarity was assessed in the context of secondary and tertiary structures from the global data set. Next, the charges derived for Forcefield_PTM were tested on a macroscopic scale using unrestrained all-atom Langevin mol. dynamics simulations in AMBER for 34 (17 pairs of modified/unmodified) systems in implicit solvent. Assessment was performed in the context of secondary structure preservation, stability in energies, and correlations between the modified and unmodified structure trajectories on the aggregate. As an illustration of their utility, the parameters were used to compare the structural stability of the phosphorylated and dephosphorylated forms of OdhI. Microscopic comparisons between quantum and AMBER single point energies along key χ torsions on several PTMs were performed, and corrections to improve their agreement in terms of mean-squared errors and squared correlation coeffs. were parametrized. This forcefield for post-translational modifications in condensed-phase simulations can be applied to a no. of biol. relevant and timely applications including protein structure prediction, protein and peptide design, and docking and to study the effect of PTMs on folding and dynamics. The authors make the derived parameters and an assocd. interactive webtool capable of performing post-translational modifications on proteins using Forcefield_PTM available at http://selene.princeton.edu/FFPTM.
- 62Khoury, G. A.; Smadbeck, J.; Tamamis, P.; Vandris, A. C.; Kieslich, C. A.; Floudas, C. A. Forcefield_NCAA: ab initio charge parameters to aid in the discovery and design of therapeutic proteins and peptides with unnatural amino acids and their application to complement inhibitors of the compstatin family. ACS Synth. Biol. [Online early access]. DOI: DOI: 10.1021/sb400168u. Published Online: January 6, 2014.Google ScholarThere is no corresponding record for this reference.
- 63Mills, J. E.; Dean, P. M. Three-dimensional hydrogen-bond geometry and probability information from a crystal survey J. Comput.-Aided Mol. Des. 1996, 10, 607– 622Google Scholar63Three-dimensional hydrogen-bond geometry and probability information from a crystal surveyMills, J.E.J.; Dean, P.M.Journal of Computer-Aided Molecular Design (1996), 10 (6), 607-622CODEN: JCADEQ; ISSN:0920-654X. (ESCOM)An extensive crystal survey of the Cambridge Structural Database has been carried out to provide hydrogen-bond data for use in drug-design strategies. Previous crystal surveys have generated 1D frequency distributions of hydrogen-bond distances and angles, which are not sufficient to model the hydrogen bond as a ligand-receptor interaction. For each hydrogen-bonding group of interest to the drug designer, geometric hydrogen-bond criteria have been derived. The 3D distribution of complementary atoms about each hydrogen-bonding group has been ascertained by dividing the space about each group into bins of equal vol. and continuing the no. of obsd. hydrogen-bonding contacts in each bin. Finally, the propensity of each group to form a hydrogen bond has been calcd. Together, these data can be used to predict the potential site points with which a ligand could interact and therefore could be used in mol.-similarity studies, pharmacophore query searching of databases, or de novo design algorithms.
- 64Pettersen, E. F.; Goddard, T. D.; Huang, C. C.; Couch, G. S.; Greenblatt, D. M.; Meng, E. C.; Ferrin, T. E. UCSF Chimera: a visualization system for exploratory research and analysis J. Comput. Chem. 2004, 25, 1605– 1612Google Scholar64UCSF Chimera-A visualization system for exploratory research and analysisPettersen, Eric F.; Goddard, Thomas D.; Huang, Conrad C.; Couch, Gregory S.; Greenblatt, Daniel M.; Meng, Elaine C.; Ferrin, Thomas E.Journal of Computational Chemistry (2004), 25 (13), 1605-1612CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)The design, implementation, and capabilities of an extensible visualization system, UCSF Chimera, are discussed. Chimera is segmented into a core that provides basic services and visualization, and extensions that provide most higher level functionality. This architecture ensures that the extension mechanism satisfies the demands of outside developers who wish to incorporate new features. Two unusual extensions are presented: Multiscale, which adds the ability to visualize large-scale mol. assemblies such as viral coats, and Collab., which allows researchers to share a Chimera session interactively despite being at sep. locales. Other extensions include Multalign Viewer, for showing multiple sequence alignments and assocd. structures; ViewDock, for screening docked ligand orientations; Movie, for replaying mol. dynamics trajectories; and Vol. Viewer, for display and anal. of volumetric data. A discussion of the usage of Chimera in real-world situations is given, along with anticipated future directions. Chimera includes full user documentation, is free to academic and nonprofit users, and is available for Microsoft Windows, Linux, Apple Mac OS X, SGI IRIX, and HP Tru64 Unix from http://www.cgl.ucsf.edu/chimera/.
- 65Hubbart, S. J.; Thornton, J. M.NACCESS (computer program); Department of Biochemistry and Molecular Biology, University College London: London, 1993.Google ScholarThere is no corresponding record for this reference.
- 66Bellows, M. L.; Taylor, M. S.; Cole, P. A.; Shen, L.; Siliciano, R. F. Discovery of entry inhibitors for HIV-1 via a new de novo protein design framework Biophys. J. 2010, 99, 3445– 3453Google ScholarThere is no corresponding record for this reference.
- 67Bellows, M. L.; Floudas, C. A. Computational methods for de novo protein design and its applications to the human immunodeficiency virus 1, purine nucleoside phosphorylase, ubiquitin specific protease 7, and histone demethylases Curr. Drug Targets 2010, 11, 264– 278Google Scholar67Computational methods for de novo protein design and its applications to the human immunodeficiency virus 1, purine nucleoside phosphorylase, ubiquitin specific protease 7, and histone demethylasesBellows, M. L.; Floudas, C. A.Current Drug Targets (2010), 11 (3), 264-278CODEN: CDTUAU; ISSN:1389-4501. (Bentham Science Publishers Ltd.)A review. This paper provides an overview of computational de novo protein design methods, highlighting recent advances and successes. Four protein systems are described that are important targets for drug design: human immunodeficiency virus 1, purine nucleoside phosphorylase, ubiquitin specific protease 7, and histone demethylases. Target areas for drug design for each protein are described, along with known inhibitors, focusing on peptidic inhibitors, but also describing some small-mol. inhibitors. Computational design methods that have been employed in elucidating these inhibitors for each protein are outlined, along with steps that can be taken in order to apply computational protein design to a system that has mainly used exptl. methods to date.
- 68Bellows-Peterson, M. L.; Fung, H. K.; Floudas, C. A.; Kieslich, C. A.; Zhang, L.; Morikis, D.; Wareham, K. J.; Monk, P. N.; Hawksworth, O. A.; Woodruff, T. M. De novo peptide design with C3a receptor agonist and antagonist activities: theoretical predictions and experimental validation J. Med. Chem. 2012, 55, 4159– 4168Google ScholarThere is no corresponding record for this reference.
- 69Bellows, M. L.; Fung, H. K.; Floudas, C. A. Recent advances in de novo protein design. In Molecular Systems Engineering; Adjiman, C. S.; Galindo, A., Eds.; Process Systems Engineering, Vol. 6; Wiley-VCH: Weinheim, Germany, 2010; pp 207– 232.Google ScholarThere is no corresponding record for this reference.
- 70Case, D. A.; Darden, T. A.; Cheatham, T. E., III; Simmerling, C. L.; Wang, J.; Duke, R. E.; Luo, R.; Walker, R. C.; Zhang, W.; Merz, K. M.; Roberts, B.; Wang, B.; Hayik, S.; Roitberg, A.; Seabra, G.; Kolossváry, I.; Wong, K. F.; Paesani, F.; Vanicek, J.; Liu, J.; Wu, X.; Brozell, S. R.; Steinbrecher, T.; Gohlke, H.; Cai, Q.; Ye, X.; Wang, J.; Hsieh, M.-J.; Cui, G.; Roe, D. R.; Mathews, D. H.; Seetin, M. G.; Sagui, C.; Babin, V.; Luchko, T.; Gusarov, S.; Kovalenko, A.; Kollman, P. A.AMBER 11; University of California: San Francisco, CA, 2010.Google ScholarThere is no corresponding record for this reference.
- 71Onufriev, A.; Bashford, D.; Case, D. A. Modification of the generalized Born model suitable for macromolecules J. Phys. Chem. B 2000, 104, 3712– 3720Google Scholar71Modification of the Generalized Born Model Suitable for MacromoleculesOnufriev, Alexey; Bashford, Donald; Case, David A.Journal of Physical Chemistry B (2000), 104 (15), 3712-3720CODEN: JPCBFK; ISSN:1089-5647. (American Chemical Society)The analytic generalized Born approxn. is an efficient electrostatic model that describes mols. in soln. Here it is modified to permit a more accurate description of large macromols., while its established performance on small compds. is nearly unaffected. The modified model is also adapted to describe mols. with an interior dielec. const. not equal to unity. The model was tested by computations of pK shifts for a no. of titratable residues in lysozyme, myoglobin, and bacteriorhodopsin. In general, except for some deeply buried residues of bacteriorhodopsin, the results show reasonable agreement with both exptl. data and calcns. based on numerical soln. of the Poisson-Boltzmann equation. A very close agreement between the two models is obtained in prediction of the pK shifts assocd. with conformational change. The calcns. based on this version of the generalized Born approxn. are much faster than finite difference solns. of the Poisson-Boltzmann equation, which makes the present method useful for a variety of other applications where computational time is a crit. factor. The model may also be integrated into mol. dynamics programs to replace explicit solvent simulations which are particularly time-consuming for large mols.
- 72Onufriev, A.; Bashford, D.; Case, D. A. Exploring protein native states and large-scale conformational changes with a modified generalized Born model Proteins 2004, 55, 383– 394Google Scholar72Exploring protein native states and large-scale conformational changes with a modified Generalized Born modelOnufriev, Alexey; Bashford, Donald; Case, David A.Proteins: Structure, Function, and Bioinformatics (2004), 55 (2), 383-394CODEN: PSFBAF ISSN:. (Wiley-Liss, Inc.)Implicit solvation models provide, for many applications, a reasonably accurate and computationally effective way to describe the electrostatics of aq. solvation. Here, a popular anal. Generalized Born (GB) solvation model is modified to improve its accuracy in calcg. the solvent polarization part of free energy changes in large-scale conformational transitions, such as protein folding. In contrast to an earlier GB model (implemented in the AMBER-6 program), the improved version does not overstabilize the native structures relative to the finite-difference Poisson-Boltzmann continuum treatment. In addn. to improving the energy balance between folded and unfolded conformers, the algorithm (available in the AMBER-7 and NAB mol. modeling packages) is shown to perform well in more than 50 ns of native-state mol. dynamics (MD) simulations of thioredoxin, protein-A, and ubiquitin, as well as in a simulation of Barnase/Barstar complex formation. For thioredoxin, various combinations of input parameters have been explored, such as the underlying gas-phase force fields and the at. radii. The best performance is achieved with a previously proposed modification to the torsional potential in the Amber ff99 force field, which yields stable native trajectories for all of the tested proteins, with back-bone root-mean-square deviations from the native structures being ∼ 1.5 Å after 6 ns of simulation time. The structure of Barnase/Barstar complex is regenerated, starting from an unbound state, to within 1.9 Å relative to the crystal structure of the complex.
- 73Johnson, L. V.; Forest, D. L.; Banna, C. D.; Radeke, C. M.; Maloney, M. A.; Hu, J.; Spencer, C. N.; Walker, A. M.; Tsie, M. S.; Bok, D.; Radeke, M. J.; Anderson, D. H. Cell culture model that mimics drusen formation and triggers complement activation associated with age-related macular degeneration Proc. Natl. Acad. Sci. U.S.A. 2011, 1, 18277– 18282Google ScholarThere is no corresponding record for this reference.
- 74Henchoz, Y.; Bard, B.; Guillarme, D.; Carrupt, P.-A.; Veuthey, J.-L.; Martel, S. Analytical tools for the physicochemical profiling of drug candidates to predict absorption/distribution Anal. Bioanal. Chem. 2009, 394, 707– 729Google Scholar74Analytical tools for the physicochemical profiling of drug candidates to predict absorption/distributionHenchoz, Yveline; Bard, Bruno; Guillarme, Davy; Carrupt, Pierre-Alain; Veuthey, Jean-Luc; Martel, SophieAnalytical and Bioanalytical Chemistry (2009), 394 (3), 707-729CODEN: ABCNBP; ISSN:1618-2642. (Springer)A review. The measurement of physicochem. properties at an early phase of drug discovery and development is crucial to reduce attrition rates due to poor biopharmaceutical properties. Among these properties, ionization, lipophilicity, soly. and permeability are mandatory to predict the pharmacokinetic behavior of NCEs (new chem. entities). Due to the high no. of NCEs, the anal. tools used to measure these properties are automated and progressively adapted to high-throughput technologies. The present review is dedicated to exptl. methods applied in the early drug discovery process for the detn. of soly., ionization consts., lipophilicity and permeability of small mols. The principles and exptl. conditions of the different methods are described, and important enhancements in terms of throughput are highlighted.
Cited By
Smart citations by scite.ai include citation statements extracted from the full text of the citing article. The number of the statements may be higher than the number of citations provided by ACS Publications if one paper cites another multiple times or lower if scite has not yet processed some of the citing articles.
This article is cited by 15 publications.
- Joseph M. Jakubowski, Asuka A. Orr, Doan A. Le, Phanourios Tamamis. Interactions between Curcumin Derivatives and Amyloid-β Fibrils: Insights from Molecular Dynamics Simulations. Journal of Chemical Information and Modeling 2020, 60
(1)
, 289-305. https://doi.org/10.1021/acs.jcim.9b00561
- Hugh D. Glossop, Gayan Heruka De Zoysa, Yacine Hemar, Priscila Cardoso, Kelvin Wang, Jun Lu, Céline Valéry, Vijayalekshmi Sarojini. Battacin-Inspired Ultrashort Peptides: Nanostructure Analysis and Antimicrobial Activity. Biomacromolecules 2019, 20
(7)
, 2515-2529. https://doi.org/10.1021/acs.biomac.9b00291
- Kuen-Da Wu, Grace Shiahuy Chen, Jia-Rong Liu, Chen-En Hsieh, Ji-Wang Chern. Acrylamide Functional Group Incorporation Improves Drug-like Properties: An Example with EGFR Inhibitors. ACS Medicinal Chemistry Letters 2019, 10
(1)
, 22-26. https://doi.org/10.1021/acsmedchemlett.8b00270
- Nadja Berger, Tchilabalo Dilezitoko Alayi, Ranillo R. G. Resuello, Joel V. Tuplano, Edimara S. Reis, John D. Lambris. New Analogs of the Complement C3 Inhibitor Compstatin with Increased Solubility and Improved Pharmacokinetic Profile. Journal of Medicinal Chemistry 2018, 61
(14)
, 6153-6162. https://doi.org/10.1021/acs.jmedchem.8b00560
- Rohith
R. Mohan, Mark Wilson, Ronald D. Gorham, Jr., Reed E. S. Harrison, Vasilios A. Morikis, Chris A. Kieslich, Asuka A. Orr, Alexis V. Coley, Phanourios Tamamis, Dimitrios Morikis. Virtual Screening of Chemical Compounds for Discovery of Complement C3 Ligands. ACS Omega 2018, 3
(6)
, 6427-6438. https://doi.org/10.1021/acsomega.8b00606
- Tulip Lifa, William Tieu, Rosalie K. Hocking, and Rachel Codd . Forward and Reverse (Retro) Iron(III) or Gallium(III) Desferrioxamine E and Ring-Expanded Analogues Prepared Using Metal-Templated Synthesis from endo-Hydroxamic Acid Monomers. Inorganic Chemistry 2015, 54
(7)
, 3573-3583. https://doi.org/10.1021/acs.inorgchem.5b00141
- Selvi Putri, Rani Maharani, Iman Maksum, Teruna Siahaan. Peptide Design for Enhanced Anti-Melanogenesis: Optimizing Molecular Weight, Polarity, and Cyclization. Drug Design, Development and Therapy 2025, Volume 19 , 645-670. https://doi.org/10.2147/DDDT.S500004
- Asuka A. Orr, Sarah K. Kuhlmann, Phanourios Tamamis. Computational design of a β-wrapin's N-terminal domain with canonical and non-canonical amino acid modifications mimicking curcumin's proposed inhibitory function. Biophysical Chemistry 2022, 286 , 106805. https://doi.org/10.1016/j.bpc.2022.106805
- Trina Ekawati Tallei, Fatimawali, Ahmad Akroman Adam, Mona M. Elseehy, Ahmed M. El-Shehawi, Eman A. Mahmoud, Adinda Dwi Tania, Nurdjannah Jane Niode, Diah Kusumawaty, Souvia Rahimah, Yunus Effendi, Rinaldi Idroes, Ismail Celik, Md. Jamal Hossain, Talha Bin Emran. Fruit Bromelain-Derived Peptide Potentially Restrains the Attachment of SARS-CoV-2 Variants to hACE2: A Pharmacoinformatics Approach. Molecules 2022, 27
(1)
, 260. https://doi.org/10.3390/molecules27010260
- Didier Devaurs, Dinler A. Antunes, Lydia E. Kavraki. Computational analysis of complement inhibitor compstatin using molecular dynamics. Journal of Molecular Modeling 2020, 26
(9)
https://doi.org/10.1007/s00894-020-04472-8
- David J. Diller, Jon Swanson, Alexander S. Bayden, Chris J. Brown, Dawn Thean, David P. Lane, Anthony W. Partridge, Tomi K. Sawyer, Joseph Audie. Rigorous Computational and Experimental Investigations on MDM2/MDMX-Targeted Linear and Macrocyclic Peptides. Molecules 2019, 24
(24)
, 4586. https://doi.org/10.3390/molecules24244586
- Kyriakos Gabriel Varnava, Jóhannes Reynisson, Srinivasrao Raghothama, Vijayalekshmi Sarojini. Synthesis, antibacterial, and antibiofilm potential of human autophagy 16 polypeptide and analogues. Peptide Science 2018, 110
(4)
https://doi.org/10.1002/pep2.24076
- Nehemiah Zewde, Dimitrios Morikis, . A computational model for the evaluation of complement system regulation under homeostasis, disease, and drug intervention. PLOS ONE 2018, 13
(6)
, e0198644. https://doi.org/10.1371/journal.pone.0198644
- Savvas Polydorides, Eleni Michael, David Mignon, Karen Druart, Georgios Archontis, Thomas Simonson. Proteus and the Design of Ligand Binding Sites. 2016, 77-97. https://doi.org/10.1007/978-1-4939-3569-7_6
- David J Diller, Jon Swanson, Alexander S Bayden, Mark Jarosinski, Joseph Audie. Rational, Computer-Enabled Peptide Drug Design: Principles, Methods, Applications and Future Directions. Future Medicinal Chemistry 2015, 7
(16)
, 2173-2193. https://doi.org/10.4155/fmc.15.142
Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.
Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.
The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.
Recommended Articles
Abstract
Figure 1
Figure 1. Chemical structures of non-natural amino acids in compstatin peptide sequences. The abbreviations used in sequences (Table 1) are shown in parentheses.
Figure 2
Figure 2. Concentration-dependent inhibition curves of compstatin peptides in C3b and C5b-9 ELISAs and hemolytic assays. Data show C3b, C5b-9, and hemolysis inhibition (from left to right) for set 1 (A–C), set 2 (D–F), set 3 (G–I), set 4 (J–L), set 5 (M–O), and control (P–R) peptides. Data points indicate mean percent inhibition ± SEM (standard error of the mean). The intersection of dashed lines shows the IC50 value for meW4A9 in all plots. Curves to the right and left of the intersection point represent peptides with higher IC50 values and lower IC50 values compared to meW4A9, respectively.
Figure 3
Figure 3. IC50 values for compstatin peptides. Bar plots show IC50 values for compstatin peptides 1–20 and positive control peptides W4A9, meW4A9, and Parent in C3b ELISA (A), C5b-9 ELISA (B), and hemolytic assay (C). Bars show mean IC50 from three independent runs of each experiment (±95% confidence interval). The dashed horizontal line shows the IC50 value for the meW4A9 control peptide, for ease of comparison.
Figure 4
Figure 4. Effects of compstatin peptides on complement activation in the RPE cell in vitro assay. The ratio of C5b-9/ApoE fluorescence (±SEM, n = 10) is plotted as a percentage of the positive control (POS) for two hfRPE cell lines, 072810 (gray) and 081309 (black). Untreated cells that were not incubated with complement-competent human serum served as negative control (NEG). At 1 μM, the parent compound is not significantly different from the positive or linear peptide controls. All test peptides (W4A9, PEP 5, PEP 8, PEP 12, PEP 18, and PEP 19) displayed significant complement inhibition relative to their corresponding positive control (see Tables S5 and S6).
Figure 5
Figure 5. Effects of varying concentrations of Parent on complement activation in the RPE cell in vitro assay. The ratio of C5b-9/ApoE fluorescence (±SEM, n = 10) is plotted as a percentage of the positive control. Parent was tested at concentrations of 1, 10, and 50 μM (PAR1, PAR10, and PAR50). The concentration of W4A9 was 1 μM. All values are expressed relative to the positive control. Parent shows no significant difference from the positive control at 1 μM or 10 μM concentrations. At 50 μM the effect of Parent is equivalent to that of 1 μM W4A9. Both Parent at 50 μM and W4A9 at 1 μM are significantly different than the positive control (p-value of <0.001, two-tailed Mann–Whitney U test).
Figure 6
Figure 6. Relation between activity and solubility of compstatin peptides. IC50 values for compstatin peptides in C3b ELISA (A), C5b-9 ELISA (B), and hemolytic assay (C) are plotted against solubility. Horizontal and vertical error bars represent the standard deviation and 95% confidence intervals of solubility and IC50, respectively. Points inside the ellipse (lower right corner) have a favorable balance between activity and solubility. Note the peptide 8 solubility was measured using less starting material compared to the rest, and thus, its solubility is not directly comparable to that of other peptides.
Figure 7
Figure 7. RP-HPLC retention factors for compstatin peptides. Bar plots show log(k) values for compstatin peptides 1–20 and positive control peptides W4A9, meW4A9, and Parent in C3b ELISA. The dashed horizontal line shows the log(k) value for the meW4A9 control peptide, for ease of comparison.
References
This article references 74 other publications.
- 1Walport, M. J. Complement. Second of two parts N. Engl. J. Med. 2001, 344, 1140– 11441Complement in immune responsesWalport, Mark J.New England Journal of Medicine (2001), 344 (15), 1140-1144CODEN: NEJMAG; ISSN:0028-4793. (Massachusetts Medical Society)A review with 96 refs. focusing on the complement system. Topics discussed include the complement at the interface between innate and adaptive immunity; the link between the complement and inflammation; the assocn. between the complement and necrosis; the relationship between the complement and immune complexes; the role of complement deficiency in systemic lupus erythematosus; and the hypothesis that the failure of the complement to dispose of the remnants of apoptotic cells may evoke an autoimmune response.
- 2Ricklin, D.; Lambris, J. D. Complement-targeted therapeutics Nat. Biotechnol. 2007, 25, 1265– 1275There is no corresponding record for this reference.
- 3Ricklin, D.; Lambris, J. D. Complement in immune and inflammatory disorders: therapeutic interventions J. Immunol. 2013, 190, 3839– 3847There is no corresponding record for this reference.
- 4Qu, H.; Ricklin, D.; Lambris, J. D. Recent developments in low molecular weight complement inhibitors Mol. Immunol. 2009, 47, 185– 1954Recent developments in low molecular weight complement inhibitorsQu, Hongchang; Ricklin, Daniel; Lambris, John D.Molecular Immunology (2009), 47 (2-3), 185-195CODEN: MOIMD5; ISSN:0161-5890. (Elsevier Ltd.)A review. As a key part of the innate immune system, complement plays an important role not only in defending against invading pathogens but also in many other biol. processes. Inappropriate or excessive activation of complement has been linked to many autoimmune, inflammatory, and neurodegenerative diseases, as well as ischemia-reperfusion injury and cancer. A wide array of low mol. wt. complement inhibitors has been developed to target various components of the complement cascade. Their efficacy has been demonstrated in numerous in vitro and in vivo expts. Though none of these inhibitors has reached the market so far, some of them have entered clin. trials and displayed promising results. This review provides a brief overview of the currently developed low mol. wt. complement inhibitors, including short peptides and synthetic small mols., with an emphasis on those targeting components C1 and C3, and the anaphylatoxin receptors.
- 5Sahu, A.; Kay, B. Inhibition of human complement by a C3-binding peptide isolated from a phage-displayed random peptide library J. Immunol. 1996, 157, 884– 891There is no corresponding record for this reference.
- 6Morikis, D.; Sahu, A.; Moore, W. T.; Lambris, J. D. Design, structure, function and application of compstatin. In Bioactive Peptides in Drug Discovery and Design: Medical Aspects; Matsoukas, J.; Mavromoustakos, T., Eds.; IOS Press: Amsterdam, The Netherlands, 1999; pp 235– 246.There is no corresponding record for this reference.
- 7Sahu, A.; Morikis, D.; Lambris, J. D. Complement inhibitors targeting C3, C4, and C5. In Therapeutic Interventions in the Complement System; Lambris, J. D.; Holers, V. M., Eds.; Humana Press: Totowa, NJ, U.S., 2000; pp 75– 112.There is no corresponding record for this reference.
- 8Morikis, D.; Lambris, J. D. Structural aspects and design of low-molecular-mass complement inhibitors Biochem. Soc. Trans. 2002, 30, 1026– 10368Structural aspects and design of low-molecular-mass complement inhibitorsMorikis, D.; Lambris, J. D.Biochemical Society Transactions (2002), 30 (6), 1026-1036CODEN: BCSTB5; ISSN:0300-5127. (Portland Press Ltd.)A review. We present a mini-review on the structure-based design of three promising complement inhibitors. Firstly, we review compstatin, a 13-residue cyclic peptide that binds to C3 and inhibits the cleavage of C3 to C3a and C3b. Secondly, we review a six-residue cyclic peptide that binds to C5aR and antagonizes the binding of C5a to its receptor C5aR. Finally, we review three small mols. that bind to Factor D and inhibit the enzymic action of Factor D, during which Factor D proteolytically cleaves Factor B in complex with C3 or C3b.
- 9Morikis, D.; Soulika, A. M.; Mallik, B.; Klepeis, J. L.; Floudas, C. A.; Lambris, J. D. Improvement of the anti-C3 activity of compstatin using rational and combinatorial approaches Biochem. Soc. Trans. 2004, 32, 28– 329Improvement of the anti-C3 activity of compstatin using rational and combinatorial approachesMorikis, D.; Soulika, A. M.; Mallik, B.; Klepeis, J. L.; Floudas, C. A.; Lambris, J. D.Biochemical Society Transactions (2004), 32 (1), 28-32CODEN: BCSTB5; ISSN:0300-5127. (Portland Press Ltd.)A review. Compstatin is a 13-residue cyclic peptide that has the ability to inhibit the cleavage of C3 to C3a and C3b. The effects of targeting C3 cleavage are threefold, and result in hindrance of: (i) the generation of the pro-inflammatory peptide C3a, (ii) the generation of opsonin C3b (or its fragment C3d), and (iii) further complement activation of the common pathway (beyond C3) with the end result of the generation of the membrane attack complex. We will report on our progress on: (i) rational design of more active compstatin analogs based on the three-dimensional structure of compstatin, (ii) exptl. combinatorial design based on the generation of a phage-displayed peptide library partially randomized with the implementation of structure-induced restraints, and (iii) theor. combinatorial design based on a novel computational optimization method, structure-induced restraints and flexible structural templates. All three approaches have resulted in analogs with improved activities. Currently, the lead analog has the sequence acetyl-I[CVYQDWGAHRC]T-NH2 (where the brackets denote cyclization), and is 16-fold more active than the parent peptide. We will also report on our progress towards understanding the dynamic character of compstatin using mol. dynamics simulations. The identification of an ensemble of interconverting conformers of compstatin with variable populations is a first step towards the incorporation of dynamic elements in the design of new analogs using dynamics-activity relationships in addn. to structure-activity relationships.
- 10Holland, M. C. H.; Morikis, D.; Lambris, J. D. Synthetic small-molecule complement inhibitors Curr. Opin. Invest. Drugs 2004, 5, 1164– 1173There is no corresponding record for this reference.
- 11Floudas, C.; Klepeis, J.; Lambris, J. D.; Morikis, D. De novo protein design: an interplay of global optimization, mixed-integer optimization and experiments. In FOCAPD 2004, Sixth International Conference on Foundations of Computer-Aided Process Design, Discovery through Product and Process Design; Floudas, C. A.; Agrawal, R., Eds.; CACHE Corporation: Austin, TX, U.S., 2004; pp 133– 146.There is no corresponding record for this reference.
- 12Morikis, D.; Floudas, C.; Lambris, J. D. Structure-based integrative computational and experimental approach for the optimization of drug design. In ICCS 2005, Lecture Notes in Computer Science: Computational Science; Sunderam, V. S.; van Albada, G. D.; Sloot, P. M.; Dongarra, J., Eds.; Springer-Verlag: Berlin, Germany, 2005; pp 680– 688.There is no corresponding record for this reference.
- 13Morikis, D.; Lambris, J. D. Structure, dynamics, activity, and function of compstatin and design of more potent analogs. In Structural Biology of the Complement System; Morikis, D.; Lambris, J. D., Eds.; CRC Press: Boca Raton, FL, U.S., 2005; pp 317– 340.There is no corresponding record for this reference.
- 14Ricklin, D.; Lambris, J. D. Compstatin: a complement inhibitor on its way to clinical application. In Current Topics in Complement II: Advances in Experimental Medicine and Biology; Springer US: New York, NY, U.S., 2008; pp 262– 281.There is no corresponding record for this reference.
- 15Morikis, D.; Assa-Munt, N.; Sahu, A.; Lambris, J. D. Solution structure of compstatin, a potent complement inhibitor Protein Sci. 1998, 7, 619– 62715Solution structure of Compstatin, a potent complement inhibitorMorikis, Dimitrios; Assa-Munt, Nuria; Sahu, Arvind; Lambris, John D.Protein Science (1998), 7 (3), 619-627CODEN: PRCIEI; ISSN:0961-8368. (Cambridge University Press)The third component of complement, C3, plays a central role in activation of the classical, alternative, and lectin pathways of complement activation. Recently, we have identified a 13-residue cyclic peptide (named Compstatin) that specifically binds to C3 and inhibits complement activation. To investigate the topol. and the contribution of each crit. residue to the binding of Compstatin to C3, we have now detd. the soln. structure using 2D NMR techniques; we have also synthesized substitution analogs and used these to study the structure-function relationships involved. Finally, we have generated an ensemble of a family of soln. structures of the peptide with a hybrid distance geometry-restrained simulated-annealing methodol., using distance, dihedral angle, and 3JNH-Hα-coupling const. restraints. The Compstatin structure contained a type I β-turn comprising the segment Gln5-Asp6-Trp7-Gly8. Preference for packing of the hydrophobic side chains of Val3, Val4, and Trp7 was obsd. The generated structure was also analyzed for consistency using NMR parameters such as NOE connectivity patterns, 3JNH-Hα-coupling consts., and chem. shifts. Anal. of Ala substitution analogs suggested that Val3, Gln5, Asp6, Trp7, and Gly8 contribute significantly to the inhibitory activity of the peptide. Substitution of Gly8 caused a 100-fold decrease in inhibitory potency. In contrast, substitution of Val4, His9, His10, and Arg11 resulted in minimal change in the activity. These findings indicate that specific side-chain interactions and the β-turn are crit. for preservation of the conformational stability of Compstatin and they might be significant for maintaining the functional activity of Compstatin.
- 16Sahu, A.; Soulika, A. M.; Morikis, D.; Spruce, L.; Moore, W. T.; Lambris, J. D. Binding kinetics, structure-activity relationship, and biotransformation of the complement inhibitor compstatin J. Immunol. 2000, 165, 2491– 2499There is no corresponding record for this reference.
- 17Morikis, D.; Roy, M.; Sahu, A.; Troganis, A.; Jennings, P. A.; Tsokos, G. C.; Lambris, J. D. The structural basis of compstatin activity examined by structure-function-based design of peptide analogs and NMR J. Biol. Chem. 2002, 277, 14942– 14953There is no corresponding record for this reference.
- 18Klepeis, J. L.; Floudas, C. A.; Morikis, D.; Tsokos, C. G.; Argyropoulos, E.; Spruce, L.; Lambris, J. D. Integrated computational and experimental approach for lead optimization and design of compstatin variants with improved activity J. Am. Chem. Soc. 2003, 125, 8422– 842318Integrated Computational and Experimental Approach for Lead Optimization and Design of Compstatin Variants with Improved ActivityKlepeis, John L.; Floudas, Christodoulos A.; Morikis, Dimitrios; Tsokos, C. G.; Argyropoulos, E.; Spruce, L.; Lambris, John D.Journal of the American Chemical Society (2003), 125 (28), 8422-8423CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)A novel structure-activity-based combinatorial computational optimization methodol. for the design of peptides that are candidates to become therapeutics is presented. This methodol. has been successfully applied in the design of a 7-fold more active analog, among other active analogs, in the case of the complement inhibitor compstatin. The main steps of the approach involve the availability of NMR-derived structural templates, combinatorial selection of sequences based on optimization of parametrized pairwise residue interaction potentials, prediction of fold stabilities using deterministic global optimization, and exptl. validation with immunol. activity measurements. This work is direct evidence that an integrated exptl. and theor. approach can make the engineering of compds. with enhanced immunol. properties possible.
- 19Soulika, A. M.; Morikis, D.; Sarrias, M.-R.; Roy, M.; Spruce, L. A.; Sahu, A.; Lambris, J. D. Studies of structure-activity relations of complement inhibitor compstatin J. Immunol. 2003, 171, 1881– 1890There is no corresponding record for this reference.
- 20Mallik, B.; Katragadda, M.; Spruce, L. A.; Carafides, C.; Tsokos, C. G.; Morikis, D.; Lambris, J. D. Design and NMR characterization of active analogues of compstatin containing non-natural amino acids J. Med. Chem. 2005, 48, 274– 286There is no corresponding record for this reference.
- 21Katragadda, M.; Lambris, J. D. Expression of compstatin in Escherichia coli: incorporation of unnatural amino acids enhances its activity Protein Expression Purif. 2006, 47, 289– 29521Expression of compstatin in Escherichia coli: Incorporation of unnatural amino acids enhances its activityKatragadda, Madan; Lambris, John D.Protein Expression and Purification (2006), 47 (1), 289-295CODEN: PEXPEJ; ISSN:1046-5928. (Elsevier)Compstatin, a 13-residue cyclic peptide, is a complement inhibitor that shows therapeutic potential. Several previous approaches have improved the activity of this peptide several-fold. In the present study, we have expressed and purified compstatin from Escherichia coli in an effort to increase its potency and to generate it in high yield in a more economical fashion. An intein-based expression system was used to express compstatin in fusion with chitin-binding domain and Ssp DnaB intein, which were later cleaved from the expressed mol. at room temp. and pH 7.0 to yield pure compstatin in one step. The expressed compstatin showed activity similar to the synthetic compstatin in an ELISA-based assay. The same expression system and purifn. strategy were used to incorporate three tryptophan analogs, 6-fluoro-tryptophan, 5-hydroxy-tryptophan, and 7-aza-tryptophan, into compstatin. Interestingly, incorporation of 6-fluoro-tryptophan increased the activity three-fold relative to wild-type compstatin; in contrast, incorporation of 5-hydroxy- or 7-aza-tryptophan rendered compstatin less active than the wild-type form.
- 22Katragadda, M.; Magotti, P.; Sfyroera, G.; Lambris, J. D. Hydrophobic effect and hydrogen bonds account for the improved activity of a complement inhibitor, compstatin J. Med. Chem. 2006, 49, 4616– 4622There is no corresponding record for this reference.
- 23Magotti, P.; Ricklin, D.; Qu, H.; Wu, Y.-Q.; Kaznessis, Y. N.; Lambris, J. D. Structure-kinetic relationship analysis of the therapeutic complement inhibitor compstatin J. Mol. Recognit. 2009, 22, 495– 50523Structure-kinetic relationship analysis of the therapeutic complement inhibitor compstatinMagotti, Paola; Ricklin, Daniel; Qu, Hongchang; Wu, You-Qiang; Kaznessis, Yiannis N.; Lambris, John D.Journal of Molecular Recognition (2009), 22 (6), 495-505CODEN: JMORE4; ISSN:0952-3499. (John Wiley & Sons Ltd.)Compstatin is a 13-residue peptide that inhibits activation of the complement system by binding to the central component C3 and its fragments C3b and C3c. A combination of theor. and exptl. approaches has previously allowed us to develop analogs of the original compstatin peptide with up to 264-fold higher activity; one of these analogs is now in clin. trials for the treatment of age-related macular degeneration (AMD). Here we used functional assays, surface plasmon resonance (SPR), and isothermal titrn. calorimetry (ITC) to assess the effect of modifications at three key residues (Trp-4, Asp-6, Ala-9) on the affinity and activity of compstatin and its analogs, and we correlated our findings to the recently reported co-crystal structure of compstatin and C3c. The KD values for the panel of tested analogs ranged from 10-6 to 10-8 M. These differences in binding affinity could be attributed mainly to differences in dissocn. rather than assocn. rates, with a >4-fold range in kon values (2-10 × 105 M-1 s-1) and a koff variation of >35-fold (1-37 × 10-2 s-1) being obsd. The stability of the C3b-compstatin complex seemed to be highly dependent on hydrophobic effects at position 4, and even small changes at position 6 resulted in a loss of complex formation. Induction of a β-turn shift by an A9P modification resulted in a more favorable entropy but a loss of binding specificity and stability. The results obtained by the three methods utilized here were highly correlated with regard to the activity/affinity of the analogs. Thus, our analyses have identified essential structural features of compstatin and provided important information to support the development of analogs with improved efficacy. Copyright © 2009 John Wiley & Sons, Ltd.
- 24Bellows, M. L.; Fung, H. K.; Taylor, M. S.; Floudas, C. A.; de Victoria, A. L.; Morikis, D. New compstatin variants through two de novo protein design frameworks Biophys. J. 2010, 98, 2337– 2346There is no corresponding record for this reference.
- 25López de Victoria, A.; Gorham, R. D.; Bellows-Peterson, M. L.; Ling, J.; Lo, D. D.; Floudas, C. A.; Morikis, D. A new generation of potent complement inhibitors of the compstatin family Chem. Biol. Drug Des. 2011, 77, 431– 44025A new generation of potent complement inhibitors of the Compstatin familyLopez de Victoria, Aliana; Gorham, Ronald D., Jr.; Bellows-Peterson, Meghan L.; Ling, Jun; Lo, David D.; Floudas, Christodoulos A.; Morikis, DimitriosChemical Biology & Drug Design (2011), 77 (6), 431-440CODEN: CBDDAL; ISSN:1747-0277. (Wiley-Blackwell)Compstatin family peptides are potent inhibitors of the complement system and promising drug candidates against diseases involving under-regulated complement activation. Compstatin is a 13-residue cyclized peptide that inhibits cleavage of complement protein C3, preventing downstream complement activation. We present three new compstatin variants, characterized by tryptophan replacement at positions 1 and/or 13. Peptide design was based on physicochem. reasoning and was inspired by earlier work, which identified tryptophan substitutions at positions 1 and 13 in peptides with predicted C3c binding abilities. The new variants preserve distinct polar and nonpolar surfaces of compstatin, but have altered local interaction capabilities with C3. All three peptides exhibited potent C3 binding by surface plasmon resonance and potent complement inhibition by ELISA. We also present ELISA data and detailed surface plasmon resonance kinetic data of three peptides from previous computational design.
- 26Qu, H.; Magotti, P.; Ricklin, D.; Wu, E. L.; Kourtzelis, I.; Wu, Y.-Q.; Kaznessis, Y. N.; Lambris, J. D. Novel analogues of the therapeutic complement inhibitor compstatin with significantly improved affinity and potency Mol. Immunol. 2011, 48, 481– 48926Novel analogues of the therapeutic complement inhibitor compstatin with significantly improved affinity and potencyQu Hongchang; Magotti Paola; Ricklin Daniel; Wu Emilia L; Kourtzelis Ioannis; Wu You-Qiang; Kaznessis Yiannis N; Lambris John DMolecular immunology (2011), 48 (4), 481-9 ISSN:.Compstatin is a 13-residue disulfide-bridged peptide that inhibits a key step in the activation of the human complement system. Compstatin and its derivatives have shown great promise for the treatment of many clinical disorders associated with unbalanced complement activity. To obtain more potent compstatin analogues, we have now performed an N-methylation scan of the peptide backbone and amino acid substitutions at position 13. One analogue (Ac-I[CVW(Me)QDW-Sar-AHRC](NMe)I-NH(2)) displayed a 1000-fold increase in both potency (IC(50) = 62 nM) and binding affinity for C3b (K(D) = 2.3 nM) over that of the original compstatin. Biophysical analysis using surface plasmon resonance and isothermal titration calorimetry suggests that the improved binding originates from more favorable free conformation and stronger hydrophobic interactions. This study provides a series of significantly improved drug leads for therapeutic applications in complement-related diseases, and offers new insights into the structure-activity relationships of compstatin analogues.
- 27Tamamis, P.; López de Victoria, A.; Gorham, R. D.; Bellows-Peterson, M. L.; Pierou, P.; Floudas, C. A.; Morikis, D.; Archontis, G. Molecular dynamics in drug design: new generations of compstatin analogs Chem. Biol. Drug Des. 2012, 79, 703– 71827Molecular dynamics in drug design: new generations of compstatin analogsTamamis, Phanourios; Lopez de Victoria, Aliana; Gorham, Ronald D., Jr.; Bellows-Peterson, Meghan L.; Pierou, Panayiota; Floudas, Christodoulos A.; Morikis, Dimitrios; Archontis, GeorgiosChemical Biology & Drug Design (2012), 79 (5), 703-718CODEN: CBDDAL; ISSN:1747-0277. (Wiley-Blackwell)We report the computational and rational design of new generations of potential peptide-based inhibitors of the complement protein C3 from the compstatin family. The binding efficacy of the peptides is tested by extensive mol. dynamics-based structural and physicochem. anal., using 32 at. detail trajectories in explicit water for 22 peptides bound to human, rat or mouse target protein C3, with a total of 257 ns. The criteria for the new design are: (i) optimization for C3 affinity and for the balance between hydrophobicity and polarity to improve soly. compared to known compstatin analogs; and (ii) development of dual specificity, human-rat/mouse C3 inhibitors, which could be used in animal disease models. Three of the new analogs are analyzed in more detail as they possess strong and novel binding characteristics and are promising candidates for further optimization. This work paves the way for the development of an improved therapeutic for age-related macular degeneration, and other complement system-mediated diseases, compared to known compstatin variants.
- 28Gorham, R. D.; Forest, D. L.; Tamamis, P.; López de Victoria, A.; Kraszni, M.; Kieslich, C. A.; Banna, C. D.; Bellows-Peterson, M. L.; Larive, C. K.; Floudas, C. A.; Archontis, G.; Johnson, L. V.; Morikis, D. Novel compstatin family peptides inhibit complement activation by drusen-like deposits in human retinal pigmented epithelial cell cultures Exp. Eye Res. 2013, 116C, 96– 108There is no corresponding record for this reference.
- 29Qu, H.; Ricklin, D.; Bai, H.; Chen, H.; Reis, E. S.; Maciejewski, M.; Tzekou, A.; DeAngelis, R. A.; Resuello, R. R. G.; Lupu, F.; Barlow, P. N.; Lambris, J. D. New analogs of the clinical complement inhibitor compstatin with subnanomolar affinity and enhanced pharmacokinetic properties Immunobiology 2013, 218, 496– 505There is no corresponding record for this reference.
- 30Risitano, A. M.; Ricklin, D.; Huang, Y.; Reis, E. S.; Chen, H.; Ricci, P.; Lin, Z.; Pascariello, C.; Raia, M.; Sica, M.; del Vecchio, L.; Pane, F.; Lupu, F.; Notaro, R.; Resuello, R. R. G.; Deangelis, R. A.; Lambris, J. D. Peptide inhibitors of C3 activation as a novel strategy of complement inhibition for the treatment of paroxysmal nocturnal hemoglobinuria Blood 2014, 123, 2094– 2101There is no corresponding record for this reference.
- 31Janssen, B. J. C.; Halff, E. F.; Lambris, J. D.; Gros, P. Structure of compstatin in complex with complement component C3c reveals a new mechanism of complement inhibition J. Biol. Chem. 2007, 282, 29241– 29247There is no corresponding record for this reference.
- 32Klepeis, J. L.; Floudas, C. A.; Morikis, D.; Tsokos, C. G.; Lambris, J. D. Design of peptide analogues with improved activity using a novel de novo protein design approach Ind. Eng. Chem. Res. 2004, 43, 3817– 3826There is no corresponding record for this reference.
- 33Walensky, L. D.; Bird, G. H. Hydrocarbon-stapled peptides: principles, practice, and progress J. Med. Chem. 2014, 57, 6275– 628833Hydrocarbon-Stapled Peptides: Principles, Practice, and ProgressWalensky, Loren D.; Bird, Gregory H.Journal of Medicinal Chemistry (2014), 57 (15), 6275-6288CODEN: JMCMAR; ISSN:0022-2623. (American Chemical Society)A review. Protein structure underlies essential biol. processes and provides a blueprint for mol. mimicry that drives drug discovery. Although small mols. represent the lion's share of agents that target proteins for therapeutic benefit, there remains no substitute for the natural properties of proteins and their peptide subunits in the majority of biol. contexts. The peptide α-helix represents a common structural motif that mediates communication between signaling proteins. Because peptides can lose their shape when taken out of context, developing chem. interventions to stabilize their bioactive structure remains an active area of research. The all-hydrocarbon staple has emerged as one such soln., conferring α-helical structure, protease resistance, cellular penetrance, and biol. activity upon successful incorporation of a series of design and application principles. Here, we describe our more than decade-long experience in developing stapled peptides as biomedical research tools and prototype therapeutics, highlighting lessons learned, pitfalls to avoid, and keys to success.
- 34Chi, Z.-L.; Yoshida, T.; Lambris, J. D.; Iwata, T. Suppression of drusen formation by compstatin, a peptide inhibitor of complement C3 activation, on cynomolgus monkey with early-onset macular degeneration Adv. Exp. Med. Biol. 2010, 703, 127– 135There is no corresponding record for this reference.
- 35Yehoshua, Z.; Rosenfeld, P. J.; Albini, T. A. Current clinical trials in dry AMD and the definition of appropriate clinical outcome measures Semin. Ophthalmol. 2011, 26, 167– 18035Current Clinical Trials in Dry AMD and the Definition of Appropriate Clinical Outcome MeasuresYehoshua Zohar; Rosenfeld Philip J; Albini Thomas ASeminars in ophthalmology (2011), 26 (3), 167-80 ISSN:.Currently, there is no proven drug treatment for dry age-related macular degeneration (AMD). Several different treatment strategies are being investigated, including complement inhibition, neuroprotection, and visual cycle inhibitors, and novel clinical trial endpoints are being explored. Studies have identified genetic predispositions for dry AMD associated with complement dysfunction. Consequently, complement-based therapeutic treatment modalities are promising.
- 36Hilbich, C.; Kisters-Woike, B.; Reed, J.; Masters, C. L.; Beyreuther, K. Substitutions of hydrophobic amino acids reduce the amyloidogenicity of Alzheimer’s disease beta A4 peptides J. Mol. Biol. 1992, 228, 460– 473There is no corresponding record for this reference.
- 37Knerr, P. J.; Tzekou, A.; Ricklin, D.; Qu, H.; Chen, H.; van der Donk, W. A.; Lambris, J. D. Synthesis and activity of thioether-containing analogues of the complement inhibitor compstatin ACS Chem. Biol. 2011, 6, 753– 760There is no corresponding record for this reference.
- 38Smadbeck, J.; Peterson, M. B.; Khoury, G. A.; Taylor, M. S.; Floudas, C. A. Protein WISDOM: a workbench for in silico de novo design of biomolecules J. Visualized Exp. 2013, e50476– e50476There is no corresponding record for this reference.
- 39Rajgaria, R.; McAllister, S. R.; Floudas, C. A. Distance dependent centroid to centroid force fields using high resolution decoys Proteins 2008, 70, 950– 970There is no corresponding record for this reference.
- 40Güntert, P. Automated NMR structure calculation with CYANA Methods Mol. Biol. 2004, 278, 353– 37840Automated NMR structure calculation with CYANAGuntert, PeterMethods in Molecular Biology (Totowa, NJ, United States) (2004), 278 (Protein NMR Techniques), 353-378CODEN: MMBIED; ISSN:1064-3745. (Humana Press Inc.)This chapter gives an introduction to automated NMR structure calcn. with the program CYANA. Given a sufficiently complete list of assigned chem. shifts and one or several lists of cross-peak positions and columns from two-, three-, or four-dimensional nuclear Overhauser effect spectroscopy (NOESY) spectra, the assignment of the NOESY cross-peaks and the three-dimensional structure of the protein in soln. can be calcd. automatically with CYANA.
- 41Güntert, P.; Mumenthaler, C.; Wüthrich, K. Torsion angle dynamics for NMR structure calculation with the new program DYANA J. Mol. Biol. 1997, 273, 283– 29841Torsion angle dynamics for NMR structure calculation with the new program DYANAGuntert, P.; Mumenthaler, C.; Wuthrich, K.Journal of Molecular Biology (1997), 273 (1), 283-298CODEN: JMOBAK; ISSN:0022-2836. (Academic)The new program DYANA (DYnamics Algorithm for Nmr Applications) for efficient calcn. of three-dimensional protein and nucleic acid structures from distance constraints and torsion angle constraints collected by NMR expts. performs simulated annealing by mol. dynamics in torsion angle space and uses a fast recursive algorithm to integrate the equations of motions. Torsion angle dynamics can be more efficient than mol. dynamics in Cartesian coordinate space because of the reduced no. of degrees of freedom and the concomitant absence of high-frequency bond and angle vibrations, which allows for the use of longer time-steps and/or higher temps. in the structure calcn. It also represents a significant advance over the variable target function method in torsion angle space with the REDAC strategy used by the predecessor program DIANA. DYANA computation times per accepted conformer in the "bundle" used to represent the NMR structure compare favorably with those of other presently available structure calcn. algorithms, and are of the order of 160 s for a protein of 165 amino acid residues when using a DEC Alpha 8400 5/300 computer. Test calcns. starting from conformers with random torsion angle values further showed that DYANA is capable of efficient calcn. of high-quality protein structures with up to 400 amino acid residues, and of nucleic acid structures.
- 42Ponder, J. W. TINKER, Software Tools for Molecular Design. Jay Ponder Lab, Washington Unversity: St. Louis, MO, 2004.There is no corresponding record for this reference.
- 43Cornell, W. D.; Cieplak, P.; Bayly, C. I.; Gould, I. R.; Merz, K. M.; Ferguson, D. M.; Spellmeyer, D. C.; Fox, T.; Caldwell, J. W.; Kollman, P. A. A second generation force field for the simulation of proteins, nucleic acids, and organic molecules J. Am. Chem. Soc. 1995, 117, 5179– 519743A Second Generation Force Field for the Simulation of Proteins, Nucleic Acids, and Organic MoleculesCornell, Wendy D.; Cieplak, Piotr; Bayly, Christopher I.; Gould, Ian R.; Merz, Kenneth M., Jr.; Ferguson, David M.; Spellmeyer, David C.; Fox, Thomas; Caldwell, James W.; Kollman, Peter A.Journal of the American Chemical Society (1995), 117 (19), 5179-97CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The authors present the derivation of a new mol. mech. force field for simulating the structures, conformational energies, and interaction energies of proteins, nucleic acids, and many related org. mols. in condensed phases. This effective two-body force field is the successor to the Weiner et al. force field and was developed with some of the same philosophies, such as the use of a simple diagonal potential function and electrostatic potential fit atom centered charges. The need for a 10-12 function for representing hydrogen bonds is no longer necessary due to the improved performance of the new charge model and new van der Waals parameters. These new charges are detd. using a 6-31G* basis set and restrained electrostatic potential (RESP) fitting and have been shown to reproduce interaction energies, free energies of solvation, and conformational energies of simple small mols. to a good degree of accuracy. Furthermore, the new RESP charges exhibit less variability as a function of the mol. conformation used in the charge detn. The new van der Waals parameters have been derived from liq. simulations and include hydrogen parameters which take into account the effects of any geminal electroneg. atoms. The bonded parameters developed by Weiner et al. were modified as necessary to reproduce exptl. vibrational frequencies and structures. Most of the simple dihedral parameters have been retained from Weiner et al., but a complex set of .vphi. and ψ parameters which do a good job of reproducing the energies of the low-energy conformations of glycyl and alanyl dipeptides was developed for the peptide backbone.
- 44Lilien, R. H.; Stevens, B. W.; Anderson, A. C.; Donald, B. R. A novel ensemble-based scoring and search algorithm for protein redesign and its application to modify the substrate specificity of the gramicidin synthetase a phenylalanine adenylation enzyme J. Comput. Biol. 2005, 12, 740– 76144A Novel Ensemble-Based Scoring and Search Algorithm for Protein Redesign and Its Application to Modify the Substrate Specificity of the Gramicidin Synthetase A Phenylalanine Adenylation EnzymeLilien, Ryan H.; Stevens, Brian W.; Anderson, Amy C.; Donald, Bruce R.Journal of Computational Biology (2005), 12 (6), 740-761CODEN: JCOBEM; ISSN:1066-5277. (Mary Ann Liebert, Inc.)Realization of novel mol. function requires the ability to alter mol. complex formation. Enzymic function can be altered by changing enzyme-substrate interactions via modification of an enzyme's active site. A redesigned enzyme may either perform a novel reaction on its native substrates or its native reaction on novel substrates. A no. of computational approaches have been developed to address the combinatorial nature of the protein redesign problem. These approaches typically search for the global min. energy conformation among an exponential no. of protein conformations. We present a novel algorithm for protein redesign, which combines a statistical mechanics-derived ensemble-based approach to computing the binding const. with the speed and completeness of a branch-and-bound pruning algorithm. In addn., we developed an efficient deterministic approxn. algorithm, capable of approximating our scoring function to arbitrary precision. In practice, the approxn. algorithm decreases the execution time of the mutation search by a factor of ten. To test our method, we examd. the Phe-specific adenylation domain of the nonribosomal peptide synthetase gramicidin synthetase A (GrsA-PheA). Ensemble scoring, using a rotameric approxn. to the partition functions of the bound and unbound states for GrsA-PheA, is first used to predict binding of the wild-type protein and a previously described mutant (selective for leucine), and second, to switch the enzyme specificity toward leucine, using two novel active site sequences computationally predicted by searching through the space of possible active site mutations. The top scoring in silico mutants were created in the wet-lab. and dissocn./binding consts. were detd. by fluorescence quenching. These tested mutations exhibit the desired change in specificity from Phe to Leu. Our ensemble-based algorithm, which flexibly models both protein and ligand using rotamer-based partition functions, has application in enzyme redesign, the prediction of protein-ligand binding, and computer-aided drug design.
- 45Lee, M. R.; Baker, D.; Kollman, P. A. 2.1 and 1.8 Å average Cα rmsd structure predictions on two small proteins, HP-36 and S15 J. Am. Chem. Soc. 2001, 123, 1040– 1046There is no corresponding record for this reference.
- 46Rohl, C. A.; Baker, D. De novo determination of protein backbone structure from residual dipolar couplings using Rosetta J. Am. Chem. Soc. 2002, 124, 2723– 2729There is no corresponding record for this reference.
- 47Rohl, C. A.; Strauss, C. E. M.; Misura, K. M. S.; Baker, D. Protein structure prediction using Rosetta. In Methods in Enzymology; Elsevier: San Diego, CA, 2004; Vol. 383, pp 66– 93.There is no corresponding record for this reference.
- 48DiMaggio, P. A.; McAllister, S. R.; Floudas, C. A.; Feng, X.-J.; Rabinowitz, J. D.; Rabitz, H. A. Biclustering via optimal re-ordering of data matrices in systems biology: rigorous methods and comparative studies BMC Bioinf. 2008, 9, 45848Biclustering via optimal re-ordering of data matrices in systems biology: rigorous methods and comparative studiesDiMaggio Peter A Jr; McAllister Scott R; Floudas Christodoulos A; Feng Xiao-Jiang; Rabinowitz Joshua D; Rabitz Herschel ABMC bioinformatics (2008), 9 (), 458 ISSN:.BACKGROUND: The analysis of large-scale data sets via clustering techniques is utilized in a number of applications. Biclustering in particular has emerged as an important problem in the analysis of gene expression data since genes may only jointly respond over a subset of conditions. Biclustering algorithms also have important applications in sample classification where, for instance, tissue samples can be classified as cancerous or normal. Many of the methods for biclustering, and clustering algorithms in general, utilize simplified models or heuristic strategies for identifying the "best" grouping of elements according to some metric and cluster definition and thus result in suboptimal clusters. RESULTS: In this article, we present a rigorous approach to biclustering, OREO, which is based on the Optimal RE-Ordering of the rows and columns of a data matrix so as to globally minimize the dissimilarity metric. The physical permutations of the rows and columns of the data matrix can be modeled as either a network flow problem or a traveling salesman problem. Cluster boundaries in one dimension are used to partition and re-order the other dimensions of the corresponding submatrices to generate biclusters. The performance of OREO is tested on (a) metabolite concentration data, (b) an image reconstruction matrix, (c) synthetic data with implanted biclusters, and gene expression data for (d) colon cancer data, (e) breast cancer data, as well as (f) yeast segregant data to validate the ability of the proposed method and compare it to existing biclustering and clustering methods. CONCLUSION: We demonstrate that this rigorous global optimization method for biclustering produces clusters with more insightful groupings of similar entities, such as genes or metabolites sharing common functions, than other clustering and biclustering algorithms and can reconstruct underlying fundamental patterns in the data for several distinct sets of data matrices arising in important biological applications.
- 49DiMaggio, P. A., Jr; McAllister, S. R.; Floudas, C. A.; Feng, X.-J.; Rabinowitz, J. D.; Rabitz, H. A. A network flow model for biclustering via optimal re-ordering of data matrices J. Global Optim. 2010, 47, 343– 354There is no corresponding record for this reference.
- 50Gray, J. J.; Moughon, S.; Wang, C.; Schueler-Furman, O.; Kuhlman, B.; Rohl, C. A.; Baker, D. Protein-protein docking with simultaneous optimization of rigid-body displacement and side-chain conformations J. Mol. Biol. 2003, 331, 281– 29950Protein-protein docking with simultaneous optimization of rigid-body displacement and side-chain conformationsGray, Jeffrey J.; Moughon, Stewart; Wang, Chu; Schueler-Furman, Ora; Kuhlman, Brian; Rohl, Carol A.; Baker, DavidJournal of Molecular Biology (2003), 331 (1), 281-299CODEN: JMOBAK; ISSN:0022-2836. (Elsevier Science Ltd.)Protein-protein docking algorithms provide a means to elucidate structural details for presently unknown complexes. Here, we present and evaluate a new method to predict protein-protein complexes from the coordinates of the unbound monomer components. The method employs a low-resoln., rigid-body, Monte Carlo search followed by simultaneous optimization of backbone displacement and side-chain conformations using Monte Carlo minimization. Up to 105 independent simulations are carried out, and the resulting "decoys" are ranked using an energy function dominated by van der Waals interactions, an implicit solvation model, and an orientation-dependent hydrogen bonding potential. Top-ranking decoys are clustered to select the final predictions. Small-perturbation studies reveal the formation of binding funnels in 42 of 54 cases using coordinates derived from the bound complexes and in 32 of 54 cases using independently detd. coordinates of one or both monomers. Exptl. binding affinities correlate with the calcd. score function and explain the predictive success or failure of many targets. Global searches using one or both unbound components predict at least 25% of the native residue-residue contacts in 28 of the 32 cases where binding funnels exist. The results suggest that the method may soon be useful for generating models of biol. important complexes from the structures of the isolated components, but they also highlight the challenges that must be met to achieve consistent and accurate prediction of protein-protein interactions.
- 51Kuhlman, B.; Baker, D. Native protein sequences are close to optimal for their structures Proc. Natl. Acad. Sci. U.S.A. 2000, 97, 10383– 1038851Native protein sequences are close to optimal for their structuresKuhlman, Brian; Baker, DavidProceedings of the National Academy of Sciences of the United States of America (2000), 97 (19), 10383-10388CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)How large is the vol. of sequence space that is compatible with a given protein structure. Starting from random sequences, low free energy sequences were generated for 108 protein backbone structures by using a Monte Carlo optimization procedure and a free energy function based primarily on Lennard-Jones packing interactions and the Lazaridis-Karplus implicit solvation model. Remarkably, in the designed sequences 51% of the core residues and 27% of all residues were identical to the amino acids in the corresponding positions in the native sequences. The lowest free energy sequences obtained for ensembles of native-like backbone structures were also similar to the native sequence. Furthermore, both the individual residue frequencies and the covariances between pairs of positions obsd. in the very large SH3 domain family were recapitulated in core sequences designed for SH3 domain structures. Taken together, these results suggest that the vol. of sequence space optimal for a protein structure is surprisingly restricted to a region around the native sequence.
- 52Drew, K.; Renfrew, P. D.; Craven, T. W.; Butterfoss, G. L.; Chou, F.-C.; Lyskov, S.; Bullock, B. N.; Watkins, A.; Labonte, J. W.; Pacella, M.; Kilambi, K. P.; Leaver-Fay, A.; Kuhlman, B.; Gray, J. J.; Bradley, P.; Kirshenbaum, K.; Arora, P. S.; Das, R.; Bonneau, R. Adding diverse noncanonical backbones to Rosetta: enabling peptidomimetic design PLoS One 2013, 8, e67051There is no corresponding record for this reference.
- 53Renfrew, P. D.; Choi, E. J.; Bonneau, R.; Kuhlman, B. Incorporation of noncanonical amino acids into Rosetta and use in computational protein-peptide interface design PLoS One 2012, 7, e3263753Incorporation of noncanonical amino acids into Rosetta and use in computational protein-peptide interface designRenfrew, P. Douglas; Choi, Eun Jung; Bonneau, Richard; Kuhlman, BrianPLoS One (2012), 7 (3), e32637CODEN: POLNCL; ISSN:1932-6203. (Public Library of Science)Noncanonical amino acids (NCAAs) can be used in a variety of protein design contexts. For example, they can be used in place of the canonical amino acids (CAAs) to improve the biophys. properties of peptides that target protein interfaces. We describe the incorporation of 114 NCAAs into the protein-modeling suite Rosetta. We describe our methods for building backbone dependent rotamer libraries and the parameterization and construction of a scoring function that can be used to score NCAA contg. peptides and proteins. We validate these addns. to Rosetta and our NCAA-rotamer libraries by showing that we can improve the binding of a calpastatin derived peptides to calpain-1 by substituting NCAAs for native amino acids using Rosetta. Rosetta (executables and source), auxiliary scripts and code, and documentation can be found at online.
- 54Mills, J. H.; Khare, S. D.; Bolduc, J. M.; Forouhar, F.; Mulligan, V. K.; Lew, S.; Seetharaman, J.; Tong, L.; Stoddard, B. L.; Baker, D. Computational design of an unnatural amino acid dependent metalloprotein with atomic level accuracy J. Am. Chem. Soc. 2013, 135, 13393– 1339954Computational Design of an Unnatural Amino Acid Dependent Metalloprotein with Atomic Level AccuracyMills, Jeremy H.; Khare, Sagar D.; Bolduc, Jill M.; Forouhar, Farhad; Mulligan, Vikram Khipple; Lew, Scott; Seetharaman, Jayaraman; Tong, Liang; Stoddard, Barry L.; Baker, DavidJournal of the American Chemical Society (2013), 135 (36), 13393-13399CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Genetically encoded unnatural amino acids could facilitate the design of proteins and enzymes of novel function, but correctly specifying sites of incorporation and the identities and orientations of surrounding residues represents a formidable challenge. Computational design methods have been used to identify optimal locations for functional sites in proteins and design the surrounding residues but have not incorporated unnatural amino acids in this process. The authors extended the Rosetta design methodol. to design metalloproteins in which the amino acid (2,2'-bipyridin-5-yl)-alanine (Bpy-Ala) is a primary ligand of a bound metal ion. Following initial results that indicated the importance of buttressing the Bpy-Ala amino acid, the authors designed a buried metal binding site with octahedral coordination geometry consisting of Bpy-Ala, two protein-based metal ligands, and two metal-bound water mols. Exptl. characterization revealed a Bpy-Ala-mediated metalloprotein with the ability to bind divalent cations including Co2+, Zn2+, Fe2+, and Ni2+, with a Kd for Zn2+ of ∼40 pM. X-ray crystal structures of the designed protein bound to Co2+ and Ni2+ have RMSDs to the design model of 0.9 and 1.0 Å resp. over all atoms in the binding site.
- 55Yu, H.; Daura, X.; van Gunsteren, W. F. Molecular dynamics simulations of peptides containing an unnatural amino acid: dimerization, folding, and protein binding Proteins 2004, 54, 116– 127There is no corresponding record for this reference.
- 56Daura, X.; van Gunsteren, W. F.; Mark, A. E. Folding-unfolding thermodynamics of a beta-heptapeptide from equilibrium simulations Proteins 1999, 34, 269– 280There is no corresponding record for this reference.
- 57Daura, X.; Gademann, K.; Schäfer, H.; Jaun, B.; Seebach, D.; van Gunsteren, W. F. The β-peptide hairpin in solution: conformational study of a β-hexapeptide in methanol by NMR spectroscopy and MD simulation J. Am. Chem. Soc. 2001, 123, 2393– 240457The β-Peptide Hairpin in Solution: Conformational Study of a β-Hexapeptide in Methanol by NMR Spectroscopy and MD SimulationDaura, Xavier; Gademann, Karl; Schaefer, Heiko; Jaun, Bernhard; Seebach, Dieter; van Gunsteren, Wilfred F.Journal of the American Chemical Society (2001), 123 (10), 2393-2404CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The structural and thermodn. properties of a 6-residue β-peptide I that was designed to form a hairpin conformation have been studied by NMR spectroscopy and mol. dynamics simulation in methanol soln. The predicted hairpin would be characterized by a 10-membered hydrogen-bonded turn involving residues 3 and 4, and two extended antiparallel strands. The interproton distances and backbone torsional dihedral angles derived from the NMR expts. at room temp. are in general terms compatible with the hairpin conformation. Two trajectories of system configurations from 100-ns mol.-dynamics simulations of the peptide in soln. at 298 and 340 K have been analyzed. In both simulations, reversible folding to the hairpin conformation is obsd. Interestingly, there is a significant conformational overlap between the unfolded state of the peptide at each of the temps. As already obsd. in previous studies of peptide folding, the unfolded state is composed of a (relatively) small no. of predominant conformers and in this case lacks any type of secondary-structure element. The trajectories provide an excellent ground for the interpretation of the NMR-derived data in terms of ensemble avs. and distributions as opposed to single-conformation interpretations. From this perspective, a relative population of the hairpin conformation of 20% to 30% would suffice to explain the NMR-derived data. Surprisingly, however, the ensemble of structures from the simulation at 340 K reproduces more accurately the NMR-derived data than the ensemble from the simulation at 298 K, and this point needs further investigation.
- 58Schäfer, H.; Daura, X.; Mark, A. E.; van Gunsteren, W. F. Entropy calculations on a reversibly folding peptide: changes in solute free energy cannot explain folding behavior Proteins 2001, 43, 45– 56There is no corresponding record for this reference.
- 59Rathore, N.; Gellman, S. H.; de Pablo, J. J. Thermodynamic stability of β-peptide helices and the role of cyclic residues Biophys. J. 2006, 91, 3425– 3435There is no corresponding record for this reference.
- 60McGovern, M.; Abbott, N.; de Pablo, J. J. Dimerization of helical β-peptides in solution Biophys. J. 2012, 102, 1435– 1442There is no corresponding record for this reference.
- 61Khoury, G. A.; Thompson, J. P.; Smadbeck, J.; Kieslich, C. A.; Floudas, C. A. Forcefield_PTM: ab initio charge and AMBER forcefield parameters for frequently occurring post-translational modifications J. Chem. Theory Comput. 2013, 9, 5653– 567461Forcefield_PTM: Ab Initio Charge and AMBER Forcefield Parameters for Frequently Occurring Post-Translational ModificationsKhoury, George A.; Thompson, Jeff P.; Smadbeck, James; Kieslich, Chris A.; Floudas, Christodoulos A.Journal of Chemical Theory and Computation (2013), 9 (12), 5653-5674CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)The authors introduce Forcefield_PTM, a set of AMBER forcefield parameters consistent with ff03 for 32 common post-translational modifications. Partial charges were calcd. through ab initio calcns. and a two-stage RESP-fitting procedure in an ether-like implicit solvent environment. The charges are generally consistent with others previously reported for phosphorylated amino acids, and trimethyllysine, using different parametrization methods. Pairs of modified structures and their corresponding unmodified structures were curated from the PDB for both single and multiple modifications. Background structural similarity was assessed in the context of secondary and tertiary structures from the global data set. Next, the charges derived for Forcefield_PTM were tested on a macroscopic scale using unrestrained all-atom Langevin mol. dynamics simulations in AMBER for 34 (17 pairs of modified/unmodified) systems in implicit solvent. Assessment was performed in the context of secondary structure preservation, stability in energies, and correlations between the modified and unmodified structure trajectories on the aggregate. As an illustration of their utility, the parameters were used to compare the structural stability of the phosphorylated and dephosphorylated forms of OdhI. Microscopic comparisons between quantum and AMBER single point energies along key χ torsions on several PTMs were performed, and corrections to improve their agreement in terms of mean-squared errors and squared correlation coeffs. were parametrized. This forcefield for post-translational modifications in condensed-phase simulations can be applied to a no. of biol. relevant and timely applications including protein structure prediction, protein and peptide design, and docking and to study the effect of PTMs on folding and dynamics. The authors make the derived parameters and an assocd. interactive webtool capable of performing post-translational modifications on proteins using Forcefield_PTM available at http://selene.princeton.edu/FFPTM.
- 62Khoury, G. A.; Smadbeck, J.; Tamamis, P.; Vandris, A. C.; Kieslich, C. A.; Floudas, C. A. Forcefield_NCAA: ab initio charge parameters to aid in the discovery and design of therapeutic proteins and peptides with unnatural amino acids and their application to complement inhibitors of the compstatin family. ACS Synth. Biol. [Online early access]. DOI: DOI: 10.1021/sb400168u. Published Online: January 6, 2014.There is no corresponding record for this reference.
- 63Mills, J. E.; Dean, P. M. Three-dimensional hydrogen-bond geometry and probability information from a crystal survey J. Comput.-Aided Mol. Des. 1996, 10, 607– 62263Three-dimensional hydrogen-bond geometry and probability information from a crystal surveyMills, J.E.J.; Dean, P.M.Journal of Computer-Aided Molecular Design (1996), 10 (6), 607-622CODEN: JCADEQ; ISSN:0920-654X. (ESCOM)An extensive crystal survey of the Cambridge Structural Database has been carried out to provide hydrogen-bond data for use in drug-design strategies. Previous crystal surveys have generated 1D frequency distributions of hydrogen-bond distances and angles, which are not sufficient to model the hydrogen bond as a ligand-receptor interaction. For each hydrogen-bonding group of interest to the drug designer, geometric hydrogen-bond criteria have been derived. The 3D distribution of complementary atoms about each hydrogen-bonding group has been ascertained by dividing the space about each group into bins of equal vol. and continuing the no. of obsd. hydrogen-bonding contacts in each bin. Finally, the propensity of each group to form a hydrogen bond has been calcd. Together, these data can be used to predict the potential site points with which a ligand could interact and therefore could be used in mol.-similarity studies, pharmacophore query searching of databases, or de novo design algorithms.
- 64Pettersen, E. F.; Goddard, T. D.; Huang, C. C.; Couch, G. S.; Greenblatt, D. M.; Meng, E. C.; Ferrin, T. E. UCSF Chimera: a visualization system for exploratory research and analysis J. Comput. Chem. 2004, 25, 1605– 161264UCSF Chimera-A visualization system for exploratory research and analysisPettersen, Eric F.; Goddard, Thomas D.; Huang, Conrad C.; Couch, Gregory S.; Greenblatt, Daniel M.; Meng, Elaine C.; Ferrin, Thomas E.Journal of Computational Chemistry (2004), 25 (13), 1605-1612CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)The design, implementation, and capabilities of an extensible visualization system, UCSF Chimera, are discussed. Chimera is segmented into a core that provides basic services and visualization, and extensions that provide most higher level functionality. This architecture ensures that the extension mechanism satisfies the demands of outside developers who wish to incorporate new features. Two unusual extensions are presented: Multiscale, which adds the ability to visualize large-scale mol. assemblies such as viral coats, and Collab., which allows researchers to share a Chimera session interactively despite being at sep. locales. Other extensions include Multalign Viewer, for showing multiple sequence alignments and assocd. structures; ViewDock, for screening docked ligand orientations; Movie, for replaying mol. dynamics trajectories; and Vol. Viewer, for display and anal. of volumetric data. A discussion of the usage of Chimera in real-world situations is given, along with anticipated future directions. Chimera includes full user documentation, is free to academic and nonprofit users, and is available for Microsoft Windows, Linux, Apple Mac OS X, SGI IRIX, and HP Tru64 Unix from http://www.cgl.ucsf.edu/chimera/.
- 65Hubbart, S. J.; Thornton, J. M.NACCESS (computer program); Department of Biochemistry and Molecular Biology, University College London: London, 1993.There is no corresponding record for this reference.
- 66Bellows, M. L.; Taylor, M. S.; Cole, P. A.; Shen, L.; Siliciano, R. F. Discovery of entry inhibitors for HIV-1 via a new de novo protein design framework Biophys. J. 2010, 99, 3445– 3453There is no corresponding record for this reference.
- 67Bellows, M. L.; Floudas, C. A. Computational methods for de novo protein design and its applications to the human immunodeficiency virus 1, purine nucleoside phosphorylase, ubiquitin specific protease 7, and histone demethylases Curr. Drug Targets 2010, 11, 264– 27867Computational methods for de novo protein design and its applications to the human immunodeficiency virus 1, purine nucleoside phosphorylase, ubiquitin specific protease 7, and histone demethylasesBellows, M. L.; Floudas, C. A.Current Drug Targets (2010), 11 (3), 264-278CODEN: CDTUAU; ISSN:1389-4501. (Bentham Science Publishers Ltd.)A review. This paper provides an overview of computational de novo protein design methods, highlighting recent advances and successes. Four protein systems are described that are important targets for drug design: human immunodeficiency virus 1, purine nucleoside phosphorylase, ubiquitin specific protease 7, and histone demethylases. Target areas for drug design for each protein are described, along with known inhibitors, focusing on peptidic inhibitors, but also describing some small-mol. inhibitors. Computational design methods that have been employed in elucidating these inhibitors for each protein are outlined, along with steps that can be taken in order to apply computational protein design to a system that has mainly used exptl. methods to date.
- 68Bellows-Peterson, M. L.; Fung, H. K.; Floudas, C. A.; Kieslich, C. A.; Zhang, L.; Morikis, D.; Wareham, K. J.; Monk, P. N.; Hawksworth, O. A.; Woodruff, T. M. De novo peptide design with C3a receptor agonist and antagonist activities: theoretical predictions and experimental validation J. Med. Chem. 2012, 55, 4159– 4168There is no corresponding record for this reference.
- 69Bellows, M. L.; Fung, H. K.; Floudas, C. A. Recent advances in de novo protein design. In Molecular Systems Engineering; Adjiman, C. S.; Galindo, A., Eds.; Process Systems Engineering, Vol. 6; Wiley-VCH: Weinheim, Germany, 2010; pp 207– 232.There is no corresponding record for this reference.
- 70Case, D. A.; Darden, T. A.; Cheatham, T. E., III; Simmerling, C. L.; Wang, J.; Duke, R. E.; Luo, R.; Walker, R. C.; Zhang, W.; Merz, K. M.; Roberts, B.; Wang, B.; Hayik, S.; Roitberg, A.; Seabra, G.; Kolossváry, I.; Wong, K. F.; Paesani, F.; Vanicek, J.; Liu, J.; Wu, X.; Brozell, S. R.; Steinbrecher, T.; Gohlke, H.; Cai, Q.; Ye, X.; Wang, J.; Hsieh, M.-J.; Cui, G.; Roe, D. R.; Mathews, D. H.; Seetin, M. G.; Sagui, C.; Babin, V.; Luchko, T.; Gusarov, S.; Kovalenko, A.; Kollman, P. A.AMBER 11; University of California: San Francisco, CA, 2010.There is no corresponding record for this reference.
- 71Onufriev, A.; Bashford, D.; Case, D. A. Modification of the generalized Born model suitable for macromolecules J. Phys. Chem. B 2000, 104, 3712– 372071Modification of the Generalized Born Model Suitable for MacromoleculesOnufriev, Alexey; Bashford, Donald; Case, David A.Journal of Physical Chemistry B (2000), 104 (15), 3712-3720CODEN: JPCBFK; ISSN:1089-5647. (American Chemical Society)The analytic generalized Born approxn. is an efficient electrostatic model that describes mols. in soln. Here it is modified to permit a more accurate description of large macromols., while its established performance on small compds. is nearly unaffected. The modified model is also adapted to describe mols. with an interior dielec. const. not equal to unity. The model was tested by computations of pK shifts for a no. of titratable residues in lysozyme, myoglobin, and bacteriorhodopsin. In general, except for some deeply buried residues of bacteriorhodopsin, the results show reasonable agreement with both exptl. data and calcns. based on numerical soln. of the Poisson-Boltzmann equation. A very close agreement between the two models is obtained in prediction of the pK shifts assocd. with conformational change. The calcns. based on this version of the generalized Born approxn. are much faster than finite difference solns. of the Poisson-Boltzmann equation, which makes the present method useful for a variety of other applications where computational time is a crit. factor. The model may also be integrated into mol. dynamics programs to replace explicit solvent simulations which are particularly time-consuming for large mols.
- 72Onufriev, A.; Bashford, D.; Case, D. A. Exploring protein native states and large-scale conformational changes with a modified generalized Born model Proteins 2004, 55, 383– 39472Exploring protein native states and large-scale conformational changes with a modified Generalized Born modelOnufriev, Alexey; Bashford, Donald; Case, David A.Proteins: Structure, Function, and Bioinformatics (2004), 55 (2), 383-394CODEN: PSFBAF ISSN:. (Wiley-Liss, Inc.)Implicit solvation models provide, for many applications, a reasonably accurate and computationally effective way to describe the electrostatics of aq. solvation. Here, a popular anal. Generalized Born (GB) solvation model is modified to improve its accuracy in calcg. the solvent polarization part of free energy changes in large-scale conformational transitions, such as protein folding. In contrast to an earlier GB model (implemented in the AMBER-6 program), the improved version does not overstabilize the native structures relative to the finite-difference Poisson-Boltzmann continuum treatment. In addn. to improving the energy balance between folded and unfolded conformers, the algorithm (available in the AMBER-7 and NAB mol. modeling packages) is shown to perform well in more than 50 ns of native-state mol. dynamics (MD) simulations of thioredoxin, protein-A, and ubiquitin, as well as in a simulation of Barnase/Barstar complex formation. For thioredoxin, various combinations of input parameters have been explored, such as the underlying gas-phase force fields and the at. radii. The best performance is achieved with a previously proposed modification to the torsional potential in the Amber ff99 force field, which yields stable native trajectories for all of the tested proteins, with back-bone root-mean-square deviations from the native structures being ∼ 1.5 Å after 6 ns of simulation time. The structure of Barnase/Barstar complex is regenerated, starting from an unbound state, to within 1.9 Å relative to the crystal structure of the complex.
- 73Johnson, L. V.; Forest, D. L.; Banna, C. D.; Radeke, C. M.; Maloney, M. A.; Hu, J.; Spencer, C. N.; Walker, A. M.; Tsie, M. S.; Bok, D.; Radeke, M. J.; Anderson, D. H. Cell culture model that mimics drusen formation and triggers complement activation associated with age-related macular degeneration Proc. Natl. Acad. Sci. U.S.A. 2011, 1, 18277– 18282There is no corresponding record for this reference.
- 74Henchoz, Y.; Bard, B.; Guillarme, D.; Carrupt, P.-A.; Veuthey, J.-L.; Martel, S. Analytical tools for the physicochemical profiling of drug candidates to predict absorption/distribution Anal. Bioanal. Chem. 2009, 394, 707– 72974Analytical tools for the physicochemical profiling of drug candidates to predict absorption/distributionHenchoz, Yveline; Bard, Bruno; Guillarme, Davy; Carrupt, Pierre-Alain; Veuthey, Jean-Luc; Martel, SophieAnalytical and Bioanalytical Chemistry (2009), 394 (3), 707-729CODEN: ABCNBP; ISSN:1618-2642. (Springer)A review. The measurement of physicochem. properties at an early phase of drug discovery and development is crucial to reduce attrition rates due to poor biopharmaceutical properties. Among these properties, ionization, lipophilicity, soly. and permeability are mandatory to predict the pharmacokinetic behavior of NCEs (new chem. entities). Due to the high no. of NCEs, the anal. tools used to measure these properties are automated and progressively adapted to high-throughput technologies. The present review is dedicated to exptl. methods applied in the early drug discovery process for the detn. of soly., ionization consts., lipophilicity and permeability of small mols. The principles and exptl. conditions of the different methods are described, and important enhancements in terms of throughput are highlighted.
Supporting Information
Supporting Information
Tables and figures of peptide design rankings, compstatin peptide inhibitory activity (from ELISAs, hemolytic assays, and cell based RPE assays), and solubility/lipophilicity data. This material is available free of charge via the Internet at http://pubs.acs.org.
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.