Skip to main content

Protein Modelling and Surface Folding by Limiting the Degrees of Freedom

  • Chapter
  • First Online:
Book cover Computational Modeling in Tissue Engineering

Abstract

One aspect of tissue engineering represents modelling of the extracellular matrix of connective tissue as the fiber network arrangement of the matrix determines its tensile strength. In order to define the correct position of the e.g. collagen in a structure, an optimized tertiary structure must be characterized. Existing approaches of protein models consider random packing of rigid spheres. We propose an alternative strategy to model protein structure by focusing on the folding. Our model considers (a) segments of amino-acid peptides or beads, (b) hydrogen bond distances, and (c) the distance geometry as functional components rather than minimizing distances between the centers of atoms. We reduced the molecular volume by using concepts from low dimensional topology, such as braids and surfaces, via differential geometry. A braid group maintains the continuity of a sequence while the spatial minimization is performed, and guarantees the continuity during the process. We have applied this approach to different examples of known protein sequences using ab initio protocols of ProteoRubix Systems™. Sequence files of three different proteins types were tested and modeled by ProteoRubix™ and compared to models derived by other methods. ProteoRubix™ created near-identical models with minimal computational load. This model can be expanded to large, multi-molecular network structures.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 84.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 109.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 109.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

  1. Adams, C., Hildebrand, M., Weeks, J.: Hyperbolic invariants of knots and links. Trans. Amer. Math. Soc. 1, 1–56 (1991)

    Article  MathSciNet  Google Scholar 

  2. Adams, J.F.: Vector fields on spheres. Ann. Math. 75, 603–632 (1962)

    Article  MATH  Google Scholar 

  3. Andrew, C.D., Penel, S., Jones, G.R., Doig, A.J.: Stabilizing nonpolar/polar side-chain interactions in the α-helix. Proteins Struct. Funct. Genetics 45, 449–455 (2001)

    Google Scholar 

  4. Arge, E., Bruaset, A.M., Langtangen, H.P.: Modern Software tools for Scientific Computing. Birkhauser Press, Boston (1997)

    Book  MATH  Google Scholar 

  5. Baker, E.N., Hubbard, R.E.: Hydrogen bonding in globular proteins. Prog. Biophys. Mol. Biol. 179, 97–177 (1984)

    Article  Google Scholar 

  6. Balakrishnan, R., Ramasubu, N., Varughese, K., Parthasathy, R.: Crystal structure of the cooper and nickel complexes of RNAase A: metal-induced interprotein interactions and identification of a novel cooper binding motif. Proc. Natl. Acad. Sci. U. S. A. 94, 9620–9625 (1997)

    Article  Google Scholar 

  7. Biegler, T.F., Mumenthaler, C., Wener, B.: Folding proteins by energy minimization and Montecarlo simulations with hydrophobic surface area potentials. J. Mol. Model. 1, 1–10 (1995)

    Article  Google Scholar 

  8. Birman, J.S.: Recent developments in Braid and Link theory. Math. Intell. 13, 52–60 (1991)

    Google Scholar 

  9. Blaney, J.M., Dixon, J.S.: Distance Geometry in Molecular Modelling. CRC Press, Boca Raton (1993)

    Google Scholar 

  10. Bonneau, R., Tsai, J., Ruczinski, I., Chivian, D., Rohl, C., Strauss, C.E., Baker, D.: Rosetta in CASPA 4: progress in ab initio protein structure prediction. Proteins 5, 119–126 (2001)

    Article  Google Scholar 

  11. Bondi, A.: van der Waals volumes and radii. J. Phys. Chem. 68, 441–451 (1964)

    Article  Google Scholar 

  12. Bryngelson, J.D., Billing, E.M.: Interatomic interactions to protein structure. Rev. Comput. Chem. 5, 84 (1995)

    Google Scholar 

  13. Bussemaker, H.J., Thirumalia, D., Bhattacharjee, J.K.: Thermodynamic stability of folding protein against mutation. Phys. Rev. Let. 79, 3530–3533 (1997)

    Article  Google Scholar 

  14. Campbell, P.G., Cohen, A.P., Ernst, L.A., Ernsthausen, J., Farkas, D.L., Galbraith, W., Israelowitz, M.: US Patent Application. USPTO Patent Application Number 20030216867 (2003)

    Google Scholar 

  15. Carmo, M.P.: Differential Geometry of Curves and surfaces. Prentice-Hall, Englewood Cliffs (1976)

    MATH  Google Scholar 

  16. Case, D.A., Cheatham III, T.E., Darden, T., Gohlke, H., Luo, R., Merz Jr, K.M., Onufriev, A., Simmerling, C., Wang, B., Woods, R.: The Amber biomolecular simulation programs. J. Comput. Chem. 26, 1668–1688 (2005)

    Article  Google Scholar 

  17. Chen, J.M., Kung, C.E., Feairheller, S.H., Brown, E.M.: An energetic evaluation of a “Smith” collagen microfil model. J. Protein Chem. 10(5), 535–551 (1991)

    Article  Google Scholar 

  18. Cheatham III, T.E., Young, M.A.: Molecular dynamics simulation of nucleic acids: successes, limitations and promise. Biopolymers 56, 232–256 (2001)

    Article  Google Scholar 

  19. Choi, H.K., Laursen, R.A., Allen, K.N.: The 2.1 Å structure of a cysteine protease with proline specificity from ginger rhizome, Zingiber officinale. Biochemistry 38, 11624–11633 (1999)

    Article  Google Scholar 

  20. Connolly, M.L.: Computation of molecular volume. J. Am. Chem. Soc. 107, 1118–1124 (1985)

    Google Scholar 

  21. Connolly, M.L.: Adjoin volumes. J. Math. Chem. 15, 339–352 (1994)

    Article  MathSciNet  Google Scholar 

  22. Costantini, S., Colonna, G., Facchiano, A.M.: Amino acid propensities for secondary structures are influenced by the protein structural class. Biochem. Biophys. Res. Commun. 342, 441–451 (2006)

    Article  Google Scholar 

  23. Cramer, C.J.: Essentials of computational chemistry: theories and models. Wiley, West Sussex (2004)

    Google Scholar 

  24. Das, A.K., Cohen, P.W., Barford, D.: The structure of the tetratricopeptide repeats of protein phosphatase 5: implications for TPR-mediated protein–protein interactions. EMBO J. 17, 1192–1199 (1998)

    Article  Google Scholar 

  25. DeLano, W.L.: The PyMOL molecular graphics system on World Wide Web (2002). http://www.pymol.org

  26. Eaton, W.A., Munoz, V., Thompson, P.A., Henry, E.R., Hofrichter, J.: Kinetics and dynamics loops, α-helices, β-haipins, and fast-folding proteins. Acc. Chem. Res. 31, 745–753 (1998)

    Article  Google Scholar 

  27. Eğe, S.: Organic Chemistry, pp. 18–71. D.C. Heath and Company, Lexington (1984)

    Google Scholar 

  28. Fernandez, A., Sinanoglu, O.: Denaturation of proteins in methanol/water mixtures. Biophys. Chem. 21, 163–164 (1985)

    Article  Google Scholar 

  29. Frömmel, C., Gille, C., Goede, A., Gröpl, C., Hougardy, S., Nierhoff, T., Preissner, R., Thimm, M.: Accelerating screening of 3D protein data with a graph theoretical approach. Bioinformatics 19, 2442–2447 (2003)

    Article  Google Scholar 

  30. Garnier, J., Osguthorpe, D.J., Robson, B.: Analysis of the accuracy and implications of simple methods for predicting the secondary structure of globular proteins. J. Mol. Biol. 120, 97–12 (1978)

    Google Scholar 

  31. Garrett, R., Grisham, C.M.: Biochemistry, p. 150. Brooks/Cole, Belmont (2005)

    Google Scholar 

  32. Gibson, K.D., Scheraga, H.A.: An algorithm for packing multistrand polypeptide structures by energy minimization. J. Comput. Chem. 15, 1414–1428 (1994)

    Article  MathSciNet  Google Scholar 

  33. Gille, C., Lorenzen, S., Michalsky, E., Frömmel, C.: KISS for STRAP: user extensions for a protein alignment editor. Bioinformatics 19, 2489–2491 (2003)

    Article  Google Scholar 

  34. Gille, C.: Structural interpretation of mutations and SPNs using STRAP-NT. Protein Sci. 15, 208–210 (2006)

    Article  Google Scholar 

  35. Gong, H., Porter, L.L., Rose, G.: Counting peptide-water hydrogen bonds in unfolded proteins. Protein Sci. 574, 417–427 (2011)

    Article  Google Scholar 

  36. Havel, T.F.: An evaluation of computational strategies for use in the determination of protein structure from distance constraints obtained by nuclear magnetic resonance. Prog. Biophys. Mol. Biol. 56, 43–78 (1991)

    Article  Google Scholar 

  37. Hill, B.R., Raleigh, D.P., Lombardi, A., Degrado, W.F.: De novo design of helical bundles as models for understanding protein folding and function. Acc. Chem. Res. 33, 745–754 (2000)

    Google Scholar 

  38. Hummer, G., Garde, S., García, A.E., Paulaitis, M.E., Pratt, L.R.: The pressure dependence of hydrophobic interactions is consistent with the observed pressure denaturation of proteins. Proc. Natl. Acad. Sci. U. S. A. 95, 1522–1555 (1998)

    Article  Google Scholar 

  39. Hunt, A.J., Gittes, F., Howard, J.: The force exerted by a kinesin molecule against a viscous load. Biophys. J. 67, 766–781 (1994)

    Article  Google Scholar 

  40. Israelowitz, M., Rizvi, S.W.H., Kramer, J., von Schroeder, H.P.: Computational modelling of type I Collagen fibers to determine the extracellular matrix structure of connective tissues. Protein Eng. Des. Sel. 18, 329–335 (2005)

    Article  Google Scholar 

  41. Jacoby, S.L.S., Kowalik, J.S., Pizzo, J.T.: Interactive methods for nonlinear optimization problems. Prentice-Hall, Englewood Cliffs (1972)

    Google Scholar 

  42. Tang,C.J.K., Alexandrov, V.: Relaxed Monte Carlo linear solver. In: Alexandrov, V.N., Dongarra, J.J., Juliano, B.A., Tan, C.J.K. (eds.) Lecture Notes in Computer Science, vol. 2073, p. 1289. Springer, Heidelberg (2001)

    Google Scholar 

  43. King, G., Brown, E.M., Chen, J.M.: Computer model of a bovine type I collagen microfribil. Protein Eng. 1, 43–49 (1996)

    Article  Google Scholar 

  44. Kuntz, I.D., Thomason, J.F., Oshiro, C.M.: Distance geometry. In: Openheimer N.J., James T.L. (eds.) Methods in Enzymology, vol. 177, pp. 159–204. Academic press, New York (1989)

    Google Scholar 

  45. Lyngsø, R.B., Pedersen, C.N.: RNA pseudoknot prediction in energy-based models. J. Comput. Biol. 7, 409–427 (2000)

    Article  Google Scholar 

  46. Liu, W., Chou, K.: Prediction of protein secondary structure content. Protein Eng. 12, 1041–1050 (1999)

    Article  Google Scholar 

  47. Macdonald, J.R., Johnson Jr, W.C.: Enviromental features are important in determining protein secondary structure. Protein Sci. 10, 1172–1177 (2001)

    Article  Google Scholar 

  48. Maritan, A., Micheletti, C., Triovato, A., Banava, J.B.: Optimal shapes of compact strings. Nature 406, 287–290 (2000)

    Article  Google Scholar 

  49. Marashi, S.A., Behrouzi, R., Pezeshk, H.: Adaptation of proteins to different environments: a comparison of proteome structural properties in Bacillus subtilis and Escherichia coli. J. Theor. Biol. 244(1), 127–132 (2007)

    Article  Google Scholar 

  50. MacCallum, P.H., Poet, R., Milner-White, J.E.: Coulombic interaction between partially charged main-chain atoms hydrogen-bonded to each other influence the confirmations of α–helices and antiparallel β–sheet. A new method for analysing the forces between hydrogen bonding groups in proteins includes all the coulombic interactions. J. Mol. Biol. 248, 361–373 (1995)

    Google Scholar 

  51. Maiti, R., von Domseleear, G.H., Zang H., Wisshart DS.: Super pose: a simple server sophisticated structural superposition. Nucleic Acid Res. 32, W590–W594 (2004)

    Google Scholar 

  52. Mount, D.M.: Bioinformatics: Sequence and Genome Analysis, vol. 2. Cold Spring Harbor Laboratory Press, Cold Spring Harbor (2004)

    Google Scholar 

  53. More, J.J., Wu, Z.: Issues in large scale global minimization. In: Biegler, L.T., Coleman, T. F., Conn, A.R., Santosa, F.N. (eds.) Large-scale optimization with applications, Part III, p. 99. Springer-Verlag, New York (1997)

    Google Scholar 

  54. Morgan, D., Ceder, G., Curtarolo, S.: High-throughput and data mining with ab initio methods. Meas. Sci. Technol. 16, 296–301 (2005)

    Article  Google Scholar 

  55. Nelder, J.A., Mead, R.: A simplex method for function minimization. Comput. J. 7, 308–313 (1965)

    MATH  Google Scholar 

  56. Nemethy, G., Gibson, K.D., Palmer, K.A., Yoon, C.N., Paterlini, G., Zagari, A., Rumsey, S., Scheraga, H.A.: Energy parameters in polypeptides. 10. Improved geometric parameters and nonbonded interactions for use in the ECEPP/3 algorithm, with application to proline-containing peptides. J. Phys. Chem. 96, 6472–6484 (1992)

    Article  Google Scholar 

  57. Nicholls, A., Sharp, K.A., Honig, B.: Protein folding and association: insights from the interfacial and thermodynamic properties of hydrocarbons. Proteins 11, 281–296 (1991)

    Article  Google Scholar 

  58. Pavanï, R., Ranghino, G.: A method to compute the volume of a molecule. Comput. Chem. 6, 133–135 (1982)

    Google Scholar 

  59. Pearlman, D.A., Case, D.A., Caldwell, J.W., Ross, W.R., Cheatham III, T.E., DeBolt, S., Ferguson, D., Seibe,l G., Kollman, P.: AMBER, a computer program for applying molecular mechanics, normal mode analysis, molecular dynamics and free energy calculations to elucidate the structures and energies of molecules. Comput. Phys. Commun. 91, 1–41 (1995)

    Google Scholar 

  60. Pham, T.H., Satou, K., Ho, T.B.: Support vector machines for prediction and analysis of beta and gamma-turns in proteins. J. Bioinform. Comput. Biol. 3, 343–358 (2005)

    Article  Google Scholar 

  61. Polanowska-Grabowska, R., Simon Jr, C.G., Shabanowitz, J., Hunt, D.F., Gear, A.R.L.: Platelet adhesion to collagen under flow causes dissociation of a phosphoprotein complex of heat-shock proteins and protein phosphatase 1. Blood 90, 1516–1526 (1997)

    Google Scholar 

  62. Ponder, J.W., Case, D.A.: Force fields for protein simulations. Adv. Protein Chem. 66, 27–85 (2003)

    Article  Google Scholar 

  63. Ochsebein, F., Gilquin, B.: NMR for protein analysis. CLEFS CEA 56, 52–55 (2008)

    Google Scholar 

  64. Ott, R., Bijma, J., Hemleben, C.: A computer method for estimating volume and surface areas of complex structure consisting of overlapping spheres 16, 83–98 (1992)

    MathSciNet  MATH  Google Scholar 

  65. Radzicka, A., Wolfeden, R.: Comparing the polarities of the amino acids: side-chain distribution coefficients between the vapor phase, cyclohexane, 1-octanol, and neutral aqueous solution. Biochemistry 27, 1664–1670 (1988)

    Article  Google Scholar 

  66. Ramachandra, G.N., Ramakrishnan, C., Sasisekharan, V.: Stereochemistry of polypeptide chain configurations. J. Mol. Biol. 7, 95–99 (1963)

    Article  Google Scholar 

  67. Raspanti, M.: Different architectures of collagen fibrils enforce different fibrillogenesis mechanisms. J. Biomed. Sci. Eng. 3, 1169–1174 (2010)

    Google Scholar 

  68. Rojnuckarin, A., Santae, K., Shankar, S.: Brownian dynamics simulations of protein folding: Access to milliseconds time scale and beyond. PNAS 68, 4288–4292 (1998)

    Article  Google Scholar 

  69. Roux, B.: Perspective in molecular dynamics and computational method. J. Cell Biol. 135, 547–548 (2010)

    Google Scholar 

  70. Sanjeev, A., Barak, B.: Computational Complexity: A Modern Approach, pp. 50–59. Cambridge University Press, Cambridge (2009)

    Google Scholar 

  71. Sanchez, R., Pieper, U., Mirkovic, N., de Bakker, P.I.W., Wittenstein, E., Sali, A.: MODBase, a database of annotated comparative protein structure models. Nucleic Acids Res. 28, 250–253 (2000)

    Article  Google Scholar 

  72. Sharman, G.J., Searle, M.S.: Cooperative interaction between the three strands of a designed antiparallel β-sheet. J. Am. Chem. Soc. 120, 5291–5300 (1998)

    Article  Google Scholar 

  73. Schenck, H.L., Gelmman, S.H.: Use of a designed tripled-stranded antiparallel β-sheet to probe β–sheet cooperativity in aqueous solution. J. Am. Chem. Soc. 120, 4869–4870 (1998)

    Article  Google Scholar 

  74. Shakhnovich, E.I., Farztdinov, G., Gutin, A.M., Karplus, M.: Protein folding bottlenecks: a lattice Monte Carlo simulation. Phys. Rev. Lett. 67, 1665–1668 (1991)

    Article  Google Scholar 

  75. Shreraga, H., Gibson, K.D.: An algorithm for packing regular multistrand polypeptide structures by energy minimization. J. Comput. Chem. 15, 1414–1428 (1994)

    Article  Google Scholar 

  76. Sinaglou, O.: Microscopic surface tension down to molecular dimensions and microthermodynamic surface areas of molecules or clusters. J. Chem. Phys. 1, 463–468 (1981)

    Article  Google Scholar 

  77. Srinavasan, R.: Helix length distribution from protein crystallographic data. Indian J. Biochem. Biophys. 13, 192–193 (1976)

    Google Scholar 

  78. Thompson, J.D., Higgins, D.G., Gibson, T.J.: CLUSTAL W: improving the sensitivity of progressive multiple sequence alignment through sequence weighting, position-specific gap penalties and weight matrix choice. Nucleic Acids Res. 22, 4673–4680 (1994)

    Article  Google Scholar 

  79. Ting, C.-K.: On the mean convergence time of multi-parent genetic algorithms without selection. In: Freitas, A.A., Bentley, P.J., Johnson, C.G., Timmis, J. (eds.) Advances in artificial life, p. 403 Springer-Verlag, Berlin (2005)

    Google Scholar 

  80. Tiraboschi, G., Gresh, N., Giessner-Prettre, C., Pedersen, L.G., Deerfield II, D.W.: A joint ab initio and molecular mechanics investigation of polycoordinated Zn(II) complexes with model hard and soft ligands. Variations of the binding energy and of its components with the number and the charges of the ligands. J. Comput. Chem. 21, 1011–1039 (2000)

    Article  Google Scholar 

  81. Torda, A.E., Van Gunsteren, W.F.: Molecular modelling using nuclear magnetic resonance data. In: Lipkowitz, K.B., Boy, D.B. (eds.) Reviews in Computational Chemistry, vol 3, pp. 143–172. VCH Publishers, New York (1992)

    Google Scholar 

  82. Voelz, V.A., Bowman, G.R., Beauchamp, K., Pande, V.S.: Molecular simulation of ab initio protein folding for a millisecond folder NTL9 (1–39). J. Am. Chem. 132, 1526–1528 (2010)

    Article  Google Scholar 

  83. Wade, L.G.: Structure and Stereochemistry of Alkanes: Organic Chemistry, 6th edn, pp. 103–122. Pearson Prentice Hall, Upper Saddle River (2006)

    Google Scholar 

  84. Wiltscheck, R., Kammerer, R.A., Dames, S.A., Schulthess, T., Blommers, M.J., Engel, J., Alexandrescu, A.T.: NMR assignments and secondary structure of the coiled coil trimerization domain from cartilage matrix protein in oxidized and reduced forms. Protein Sci. 6, 1734–1745 (1997)

    Article  Google Scholar 

  85. William, P., Teukolsky, S.A., Vetterling, W.T., Flannery, B.P.: Numerical Recipes in Fortran, 2nd edn., pp. 312–326. Cambridge University Press, Cambridge (1992)

    Google Scholar 

  86. Word, J.M., Lovell, S.C., LaBean, T.H., Taylor, H.C., Zalis, M.E., Presley, B.K., Richardson, J.S., Richardson, D.C.: Visualizing and quantifying molecular goodness-of-fit: small-probe contact dots with explicit hydrogens. J. Mol. Biol. 285, 1711–1733 (1999)

    Article  Google Scholar 

  87. Zimmermann, O., Hansmann, U.H.: Support vector machines for prediction of dihedral angle regions. Bioinformatics 22, 3009–3015 (2006)

    Article  Google Scholar 

  88. Zhang, Y., Skolnick, J.: TM-align: a protein structure alignment algorithm based on the TM-score. Nucleic Acids Res. 33(7), 2302–2309 (2005)

    Google Scholar 

  89. Zhang, X.Q., Jansem, A.P.: Kinetic Monte Carlo method for simulating reactions in Solution. Phy. Rev. E Stat. Nonlin. Softw. Matter Phys. 82, 046704 (2010)

    Google Scholar 

  90. Zhong, L., Johnson Jr, W.C.: Environment affects amino acid preference for secondary structure. PNAS 89, 4462–4465 (1992)

    Article  Google Scholar 

Download references

Acknowledgments

The authors thank: Isabella Verdinelli and Lauren Ernst for the discussions related to the model and Troy Wymore from the Pittsburgh Supercomputing Center for his suggestions in verifying the models and a special thanks to Dr. Alex Cohen from ProteoRubix Systems who developed the minimizer and for discussion and development of ProteoRubix™ software for the Geometry Modelling. Chris Holm for editing the manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Herbert P. von Schroeder .

Editor information

Editors and Affiliations

Appendices

Appendix 1

1.1 NP

Suppose the bead involves \( d \) dihedral angles. Let \( \phi^{ * } = \left( {\phi_{1}^{ * } , \ldots ,\phi_{n}^{ * } } \right) \in \left[ {0,360} \right]^{d} \) be an optimal solution to the constrained optimization problem

(1.1) \( \mathop {\min }\limits_{{\phi \in \left[ {0,360} \right]^{d} }} \left\{ {v(\phi :\phi \; {\text{is arotation about the bond i}})} \right\} \)

Then there is a maximal number \( n > 0 \).

(1.2) The \( p^{n} \) hard problem of an exhaustive search over the angles \( 0 \le \phi_{i}^{1} < \cdots < \phi_{i}^{P} \le 360 \) to find an approximate optimizer \( \overline{\phi } \) to \( \phi^{ * } \) may be possible for a modern computer.

(1.3) There is exactly one solution to (4) in \( \left| {\overline{{\phi_{i} }} - p,\overline{{\phi_{i} }} + p} \right| \) which would be \( \phi^{ * } \) and can be approximated using a given constrained optimization algorithm (B 1).

The convergence ball for the constrained optimization algorithm provides a candidate for p in the proposition. Using this proposition, we can obtain an acceptable initial condition for a constrained optimization algorithm.

Appendix 2

1.1 Boundary Determination to Prevent Overlap

A path is a one-dimensional sub-manifold \( {\text{M}} \) of \( {\text{R}}^{ 3} \), so that, for any point \( {\text{x}} \in {\text{M}} \) there is a local parameterization near x. \( {\text{C}}^{\text{k}} \left( {{\text{k}} \ge 2} \right) \) denotes the curvature of the path and \( {\text{D}} \) denotes the coordinates identifying the path. The output of each iteration is a set of coordinates in three dimensions, \( \left( {{\text{D}} = {\text{x}}_{ 1} , {\text{x}}_{ 2} ,\ldots , {\text{x}}_{\text{n}} } \right) \) identifying a path. We denote by length bond is the polygonal arc around the path (Fig. 15). The curvature \( {\text{C}}^{\text{k}} \) and the arc-length are non-regular. Let \( {\text{x}} = {\text{x(t),}} \) with \( {\text{a}} \le {\text{t}} \le {\text{b}} \) and consider a partition [15]:

\( {\text{a}} = {\text{t}}_{ 0} < {\text{t}}_{ 1} < \cdots < {\text{t}}_{\text{n}} = {\text{b}} \), of an interval (a, b).

The sequence (a, b) are the boundaries of a single coil) gives an approximation to the polygon arc C. As illustrated the length between two points (a, b), where \( {\text{D}} \) are segments of arc-length given by:

$$ \lambda ( {\text{D)}} = \sum\limits_{{{\text{j}} = 1}}^{\text{n}} {{\text{D}}_{\text{j}} = } \sum\limits_{{{\text{i}} = 1}}^{\text{n}} {\left\| {{\text{x}}_{\text{i}} - {\text{x}}_{{{\text{i}} - 1}} } \right\|} = \sum\limits_{{{\text{i}} = 1}}^{\text{n}} {\left\| {{\text{x}}\left( {{\text{t}}_{\text{i}} } \right) - {\text{x}}\left( {{\text{t}}_{{{\text{i}} - 1}} } \right)} \right\|} $$
(2.1)

The arc-length can be bounded from above and from below. The upper bound is given by:

$$ \rho_{ + } \left( {{\text{K}},{\text{D}}} \right){ = }\frac{ 1}{{\lambda \left( {\text{D}} \right)}}\sum\limits_{{\left( {{\text{K}} \circ {\text{D}}{}_{\text{j}}} \right) \cap {\text{D}} \ne \emptyset }} {\lambda \left( {{\text{K}} \circ {\text{D}}_{\text{j}} } \right)} $$
(2.2)

And the lower bound is:

$$ \rho_{ - } \left( {{\text{K}},{\text{D}}} \right) = \frac{ 1}{{\lambda \left( {\text{D}} \right)}}\sum\limits_{{\left( {{\text{K}} \circ {\text{D}}_{\text{j}} } \right) \subset {\text{D}}}} {\lambda \left( {{\text{K}} \circ {\text{D}}_{\text{j}} } \right)} $$
(2.3)

where \( {{\uprho}}_{ + } ( {\text{K,D)}} \) is the ratio of the total measure of the set in the system \( {\rm K} \) (is the volume minimization) so that the transformation ° (projection) of the segments and the curve \( {\text{C}} \) give the lower and the upper bound\( \left( {\text{a,b}} \right) \).

$$ b = \rho_{ + } = \mathop {\lim }\limits_{\lambda \left( D \right) \to \infty } \sup \rho_{ + } \left( {K,D} \right) = \mathop {\lim }\limits_{\lambda \to \infty } {\text{Sup}}\rho_{\begin{subarray}{l} + \\ \lambda \left( D \right) \ge \lambda \end{subarray} } \left( {K,D} \right) $$
(2.4)
$$ a = \rho_{ - } = \mathop {\lim }\limits_{\lambda \left( D \right) \to \infty } \inf \rho_{ - } \left( {K,D} \right) = \mathop {\lim }\limits_{\lambda \to \infty } \mathop {\inf }\limits_{\lambda \left( D \right) \ge \lambda } \rho_{ - } \left( {K,D} \right) $$
(2.5)

Hence the boundaries of \( {\text{C}} \) are given in (2.4) and in (2.5).

Fig. 15
figure 15

The sequence [\( \left( {a,b} \right) \) is length of a single coil] gives an approximation to the polygon arc \( P \); the length between two points (a, b), where \( D \) is segments of arc-length

Appendix 3

The geometry structure of the protein is defined by a braid (see B 2 for a description of a chain as a collection of beads forming a braid). The jth molecule of the chain is fitted to a conveniently shaped open bead Sj (see B 3) with is 0 center located at the center of the bead and the radius \( {\text{r}}_{\text{i}} \) has size such that the ith bead does not overlap with the jth bead when \( {\text{i}} \ne {\text{j}}. \)

The radii in Fig. 16 \( {\text{r}}_{\text{i}} \) are chosen so that the intersection of the closure of any two beads \( \mathop {{\text{S}}_{\text{i}} }\limits^{ - } \) and \( \mathop {{\text{S}}_{\text{j}} }\limits^{ - } \) is a single point \( {\text{p}}_{\text{ij}} , \) (see B 3). The point \( {\text{p}}_{\text{ij}} , \) is the origin of a right and a left vector \( {\text{v}}_{\text{iR}} , {\text{v}}_{\text{jL}} \). In this process it is important to translate (projection) and rotate these vectors. The mathematics of this construction justifies geometry of the bead construction.

Fig. 16
figure 16

The radii are chosen so that the intersection of the closure of any two beads \( \mathop {S_{i} }\limits^{ - } \) and \( \mathop {S_{j} }\limits^{ - } \) is a single point \( p_{ij} , \). The point \( p_{ij} , \) is the origin of a right and a left vector \( v_{iR} ,v_{jL} \)

Appendix 4

With our model of collagen in mind, we next introduced the concept of the braid group. The braid was defined as the union of the backbones creating a string representing the amino acids. The collagen has three strands (as a group) or coils and each strand has a back bone, represented as the union of all points x (ti − 1, ti) that are generated:

$$ {\text{Bonds}} = \mathop \cup \limits_{{{\text{n}} = 1}}^{\text{N}} \left\{ {{\text{x}}\left( {{\text{t}}_{{{\text{i}}\, - \, 1}} , {\text{t}}_{\text{i}} } \right)} \right\} $$
(4.1)

A braid is a collection of beads for which two operators \( \left( { \circ , = } \right) \) can be defined. The bead in the collection can be projected using least of the squares. Let B denote this collection of beads, so \( {\text{B}} = \left( {\text{braids}} \right) \), and \( \left( {{\text{B,}} \circ } \right) \) is a group. We are checking the segments of the radius of bead of a single braid. The enclosed volume shrinks driven by minimization and through the homoeopathy is guaranty [2] (see B 5). We are modelling three coils, and their geometrical configuration has an equivalence class denoted by \( {{\upsigma}}_{\text{i}} \) and \( {{\upsigma}}_{\text{i}}^{ - 1} \). A braid is equivalent and it is called isotope if the three coils cannot pass each other or themselves without intersecting [8] Fig. 17.

Fig. 17
figure 17

A braid is equivalent and it is called isotope if the three coils cannot pass each other or themselves without intersecting

\( {{\upsigma}}_{\text{i}} {{\upsigma}}_{{{\text{i}} + 1}} {{\upsigma}}_{\text{i}} = {{\upsigma}}_{{{\text{i}} + 1}} {{\upsigma}}_{\text{i}} {{\upsigma}}_{{{\text{i}} + 1}} \) if \( 1\le {\text{i}} \le {\text{n}} - 2 \) [1]

Appendix 5

The distances of the projection to \( P \) is given by \( \left\| {b - r} \right\| \), where \( v\left( {x - p} \right) = 0 \) and by Pythagorean gives us that \( b^{2} = c^{2} - a^{2} \), where, \( a^{2} = \left\| {\frac{b}{{\left\| b \right\|_{2} }} \times \left( {Q - P} \right)} \right\|_{2}^{2} \), \( c^{2} = \left\| {Q - P} \right\|_{2}^{2} \) or \( \frac{{b\left( {Q - P} \right)^{2} }}{{\left\| v \right\|_{2} }} \) shown in Fig. 18a.

Fig. 18
figure 18

The projection of the rotation of the rotation angles. The enclosed volume shrinks by the minimization and via homoeopathy which is guaranteed

1.1 B 1

Let \( \mathop \phi \limits^{ - } \) the solution to \( \phi^{ * } = \left( {\phi_{1}^{ * } , \ldots ,\phi_{n}^{ * } } \right) \in \left[ {0,360} \right]^{d} \) and \( d \) dihedral angle, \( \phi_{n}^{*} = \sum^{*} \to N \), \( 1 \le n \le k \); Let \( q \), \( r \) be polynomial such \( \phi_{n}^{*} \left( I \right) \le q\left( {\left| I \right|} \right) \), where \( I \) is the instance of the angle in our problem. Then test instance construction system for all the angles of our problem \( \left( {TICA} \right) \), then \( P = NP \).

Conversion We know that \( \phi^{ * } = \left( {\phi_{1}^{ * } , \ldots ,\phi_{n}^{ * } } \right) \in \left[ {0,360} \right]^{d} \) is the optimal solution, where \( d \) dihedral angles then \( \delta = \frac{1}{2}\mathop {\min }\limits_{{\phi \in \left[ {0,360} \right]^{d} }} \left\{ {v(\phi :\phi \; is \, arotation \, \;about\; \, the\; \, bond\; \, i)} \right\} \)\( n \) is the maximum number of angles, \( n > 0 \) and \( \delta > 0 \). Let \( \in > 0 \) be given. Where \( \phi^{ * } \) is continuous, there is a point \( p \in \phi^{*} \), \( \phi^{*} \le \frac{1}{2}\phi \left( p \right) \) where implies \( \left| {\overline{{\phi_{i} }} - p,\overline{{\phi_{i} }} + p} \right| < \in \) and \( v\left( p \right) \le \frac{1}{2}\phi \left( p \right) \), we have \( \left| {\overline{{\phi_{i} }} - p,\overline{{\phi_{i} }} + p} \right| \le \phi^{*} + \left| {v\left( p \right)} \right| < \delta + \frac{1}{2}\phi \left( p \right) < \in \)

Uniqueness using the existence and uniqueness theorem, we know that \( \phi^{ * } \) is continuous, in the interval \( \left| {\overline{{\phi_{i} }} - p,\overline{{\phi_{i} }} + p} \right| \) then converges.

1.2 B 2

\( {\text{D}} \) is said to be covering itself if \( \bigcup\limits_{\text{j}} {{\text{D}}_{\text{j}} } \supset {\text{D}} \) and each elements of at least one of \( {\text{D}} \) belongs to \( {\text{d}}_{\text{j}} \). The system \( {\text{D}}_{\text{j}} \) is packing if \( {\text{D}}_{\text{i}} \cap {\text{D}}_{\text{j}} = \emptyset \)\( ( {\text{i}} \ne {\text{j),}} \)\( \bigcup\limits_{\text{j}} {{\text{D}}_{\text{j}} } \supset {\text{D}} \)

If two sets \( {\text{D}}_{ 1} , {\text{D}}_{ 2} ,\ldots \) have the same elements in common then each element \( {\text{D}}_{ 1} , {\text{D}}_{ 2} ,\ldots \) belong to \( {\text{D}} . \)

1.3 B 3

Each segment can be treated as open beads, as such the coordinates belong to a set \( {\text{X}} \) and for any point \( {\text{p}} \subset {\text{D}}_{\text{j}} \) and \( {{\updelta}} = {\text{D}}_{\text{j}} \) where the measure is positive.

So, the definition of the bead is:

$$ {\text{D}} = \left\{ {{\text{x:d(p,x)}} < {{\updelta}}} \right\} $$

1.4 B 4

Let \( A \) and \( {\text{B}} \) be a disjoint convex set in a convex space, then

\( {\text{A}} = \left\{ {{\text{x:}}\left( {{\text{x}} - {\text{D}}_{\text{i}} } \right)^{ 2} < {\text{r}}_{\text{i}} } \right\} \) and \( {\text{B}} = \left\{ {{\text{x:}}\left( {{\text{x}} - {\text{D}}_{\text{j}} } \right)^{ 2} < {\text{r}}_{\text{j}} } \right\} \), the distance is given by:

\( {\text{dis(D}}_{\text{i}} , {\text{D}}_{\text{j}} )= {\text{r}}_{\text{i}} + {\text{r}}_{j} \). The closure of \( {\text{B}} \) is given by \( \mathop {\text{B}}\limits^{{\text{\_}}} = \left\{ {{\text{x:}}\left( {{\text{x}} - {\text{D}}_{\text{j}} } \right)^{ 2} \le {\text{r}}_{\text{j}} } \right\} \) then \( {\text{A}} \cap \mathop {{\text{B}} = \emptyset }\limits^{{\text{\_}}} \).

\( A \) is an open set by construction. A & B are the convex hull, also by construction, and then:

\( \exists {\text{l(x)}} = {\text{a}} \) if

$$ \begin{aligned} {\text{x}} &\in {\text{A}}\;{\text{l(x)}} \le {\text{a}} \\ {\text{x}} &\in {\text{B}}\;{\text{l(x)}} \ge {\text{a}} \\ \end{aligned} $$

where \( {\text{a}} \) is

$$ \begin{array}{*{20}c} {{\text{v}}_{\text{j}} = \left( {{\text{a}} - {\text{D}}_{\text{j}} } \right)^{ 2} } \\ {{\text{v}}_{\text{i}} = ( {\text{a}} - {\text{D}}_{\text{i}} )^{ 2} } \\ \end{array} $$

where

$$ {\text{D}}_{\text{i}} = {\text{dist}}\left( {{\text{a}} - {\text{v}}_{\text{Ri}} } \right)\;{\text{and}}\;{\text{D}}_{\text{j}} = {\text{dist}}\left( {{\text{a}} - {\text{v}}_{\text{Lj}} } \right) $$

1.5 B 5

Let \( {\text{x}} \in {\text{S}}\left( {{\text{r,x}}_{ 0} } \right) \), \( {\text{S}} \in \Re^{\text{n}} \) and \( {\text{x}}_{ 0} \ne 0 \) i.e. \( p ( {\text{x)}} = {\text{x}}_{ 0} + {\text{r}}\frac{\text{x}}{{\left\| {\text{x}} \right\|}} \) (Fig. 18b) then; \( r = \left\| {{\text{p}} - {\text{x}}_{ 0} } \right\| = \left\| {{\text{x}}_{ 0} - {\text{r}}\frac{\text{x}}{{\left\| {\text{x}} \right\|}} - {\text{x}}_{ 0} } \right\| = \frac{\text{r}}{{\left\| {\text{x}} \right\|}}\left\| {\text{x}} \right\| = {\text{r}} \) hence \( {\text{p}} \in {\text{S}}\left( {{\text{r,x}}_{ 0} } \right) \) [2].

Rights and permissions

Reprints and permissions

Copyright information

© 2012 Springer-Verlag Berlin Heidelberg

About this chapter

Cite this chapter

Israelowitz, M., Weyand, B., Rizvi, S.W.H., Gille, C., von Schroeder, H.P. (2012). Protein Modelling and Surface Folding by Limiting the Degrees of Freedom. In: Geris, L. (eds) Computational Modeling in Tissue Engineering. Studies in Mechanobiology, Tissue Engineering and Biomaterials, vol 10. Springer, Berlin, Heidelberg. https://doi.org/10.1007/8415_2012_141

Download citation

  • DOI: https://doi.org/10.1007/8415_2012_141

  • Published:

  • Publisher Name: Springer, Berlin, Heidelberg

  • Print ISBN: 978-3-642-32562-5

  • Online ISBN: 978-3-642-32563-2

  • eBook Packages: EngineeringEngineering (R0)

Publish with us

Policies and ethics