Residues crucial for maintaining short paths in network communication mediate signaling in proteins

Here, we represent protein structures as residue interacting networks, which are assumed to involve a permanent flow of information between amino acids. By removal of nodes from the protein network, we identify fold centrally conserved residues, which are crucial for sustaining the shortest pathways and thus play key roles in long‐range interactions. Analysis of seven protein families (myoglobins, G‐protein‐coupled receptors, the trypsin class of serine proteases, hemoglobins, oligosaccharide phosphorylases, nuclear receptor ligand‐binding domains and retroviral proteases) confirms that experimentally many of these residues are important for allosteric communication. The agreement between the centrally conserved residues, which are key in preserving short path lengths, and residues experimentally suggested to mediate signaling further illustrates that topology plays an important role in network communication. Protein folds have evolved under constraints imposed by function. To maintain function, protein structures need to be robust to mutational events. On the other hand, robustness is accompanied by an extreme sensitivity at some crucial sites. Thus, here we propose that centrally conserved residues, whose removal increases the characteristic path length in protein networks, may relate to the system fragility.

[1]  J. Bonaventura,et al.  ANIONIC CONTROL OF HEMOGLOBIN FUNCTION , 1978 .

[2]  I. Kuntz,et al.  Cavities in proteins: structure of a metmyoglobin-xenon complex solved to 1.9 A. , 1984, Biochemistry.

[3]  S. Bouvier,et al.  Systematic mutation of bacteriophage T4 lysozyme. , 1991, Journal of molecular biology.

[4]  L. Johnson,et al.  Glycogen phosphorylase: control by phosphorylation and allosteric effectors , 1992, FASEB journal : official publication of the Federation of American Societies for Experimental Biology.

[5]  W. Yamamoto,et al.  AY's Neuroanatomy of C. elegans for Computation , 1992 .

[6]  W. Rutter,et al.  Converting trypsin to chymotrypsin: residue 172 is a substrate specificity determinant. , 1994, Biochemistry.

[7]  G Schreiber,et al.  Energetics of protein-protein interactions: analysis of the barnase-barstar interface by single mutations and double mutant cycles. , 1995, Journal of molecular biology.

[8]  R. Fletterick,et al.  Role of the Active Site Gate of Glycogen Phosphorylase in Allosteric Inhibition and Substrate Binding* , 1996, The Journal of Biological Chemistry.

[9]  M. Paoli,et al.  Crystal structure of T state haemoglobin with oxygen bound at all four haems. , 1996, Journal of molecular biology.

[10]  D. Perez,et al.  Activation of the α1b-Adrenergic Receptor Is Initiated by Disruption of an Interhelical Salt Bridge Constraint* , 1996, The Journal of Biological Chemistry.

[11]  H. Khorana,et al.  Structure and function in rhodopsin: correct folding and misfolding in point mutants at and in proximity to the site of the retinitis pigmentosa mutation Leu-125-->Arg in the transmembrane helix C. , 1996, Proceedings of the National Academy of Sciences of the United States of America.

[12]  T. Sakmar,et al.  Functional Interaction of Transmembrane Helices 3 and 6 in Rhodopsin , 1996, The Journal of Biological Chemistry.

[13]  S. Withers,et al.  Ternary complex crystal structures of glycogen phosphorylase with the transition state analogue nojirimycin tetrazole and phosphate in the T and R states. , 1996, Biochemistry.

[14]  M. Saji,et al.  Phe576 plays an important role in the secondary structure and intracellular signaling of the human luteinizing hormone/chorionic gonadotropin receptor. , 1997, The Journal of clinical endocrinology and metabolism.

[15]  Converting trypsin to elastase: substitution of the S1 site and adjacent loops reconstitutes esterase specificity but not amidase activity. , 1998, Protein engineering.

[16]  S. O. Smith,et al.  Constitutive activation of opsin by mutation of methionine 257 on transmembrane helix 6. , 1998, Biochemistry.

[17]  M Paoli,et al.  The stereochemical mechanism of the cooperative effects in hemoglobin revisited. , 1998, Annual review of biophysics and biomolecular structure.

[18]  Duncan J. Watts,et al.  Collective dynamics of ‘small-world’ networks , 1998, Nature.

[19]  L. Gráf,et al.  The three-dimensional structure of Asp189Ser trypsin provides evidence for an inherent structural plasticity of the protease. , 1999, European journal of biochemistry.

[20]  H. B. Schock,et al.  Non-active Site Changes Elicit Broad-based Cross-resistance of the HIV-1 Protease to Inhibitors* , 1999, The Journal of Biological Chemistry.

[21]  R. Ranganathan,et al.  Evolutionarily conserved pathways of energetic connectivity in protein families. , 1999, Science.

[22]  Millard H. Lambert,et al.  Asymmetry in the PPARγ/RXRα Crystal Structure Reveals the Molecular Basis of Heterodimerization among Nuclear Receptors , 2000 .

[23]  M. Lambert,et al.  Structural basis for autorepression of retinoid X receptor by tetramer formation and the AF-2 helix. , 2000, Genes & development.

[24]  L. Johnson,et al.  A new allosteric site in glycogen phosphorylase b as a target for drug interactions. , 2000, Structure.

[25]  P. Garriga,et al.  Mutations at position 125 in transmembrane helix III of rhodopsin affect the structure and signalling of the receptor. , 2001, European journal of biochemistry.

[26]  J. Ballesteros,et al.  Structural mimicry in G protein-coupled receptors: implications of the high-resolution structure of rhodopsin for structure-function analysis of rhodopsin-like receptors. , 2001, Molecular pharmacology.

[27]  A. Barabasi,et al.  Lethality and centrality in protein networks , 2001, Nature.

[28]  S. Wuchty Scale-free behavior in protein domain networks. , 2001, Molecular biology and evolution.

[29]  T. Sakmar,et al.  Rhodopsin: structural basis of molecular physiology. , 2001, Physiological reviews.

[30]  H Frauenfelder,et al.  The role of structure, energy landscape, dynamics, and allostery in the enzymatic function of myoglobin , 2001, Proceedings of the National Academy of Sciences of the United States of America.

[31]  D E Wemmer,et al.  Two-state allosteric behavior in a single-domain signaling protein. , 2001, Science.

[32]  Y. Lu,et al.  The role of redox-active amino acids on compound I stability, substrate oxidation, and protein cross-linking in yeast cytochrome C peroxidase. , 2001, Biochemistry.

[33]  M Karplus,et al.  Small-world view of the amino acids that play a key role in protein folding. , 2002, Physical review. E, Statistical, nonlinear, and soft matter physics.

[34]  E. Shakhnovich,et al.  Topological determinants of protein folding , 2002, Proceedings of the National Academy of Sciences of the United States of America.

[35]  A. Barabasi,et al.  Hierarchical Organization of Modularity in Metabolic Networks , 2002, Science.

[36]  A. Horovitz,et al.  Mapping pathways of allosteric communication in GroEL by analysis of correlated mutations , 2002, Proteins.

[37]  Richard A Goldstein,et al.  Why are proteins so robust to site mutations? , 2002, Journal of molecular biology.

[38]  Rama Ranganathan,et al.  Allosteric determinants in guanine nucleotide-binding proteins , 2003, Proceedings of the National Academy of Sciences of the United States of America.

[39]  D. Kern,et al.  The role of dynamics in allosteric regulation. , 2003, Current opinion in structural biology.

[40]  Gürol M. Süel,et al.  Evolutionarily conserved networks of residues mediate allosteric communication in proteins , 2003, Nature Structural Biology.

[41]  H Frauenfelder,et al.  Myoglobin: The hydrogen atom of biology and a paradigm of complexity , 2003, Proceedings of the National Academy of Sciences of the United States of America.

[42]  Tcoffee@igs: A web server for computing, evaluating and combining multiple sequence alignments. , 2003, Nucleic acids research.

[43]  Victoria A. Higman,et al.  Uncovering network systems within protein structures. , 2003, Journal of molecular biology.

[44]  Krzysztof Palczewski,et al.  Role of the conserved NPxxY(x)5,6F motif in the rhodopsin ground state and during activation , 2003, Proceedings of the National Academy of Sciences of the United States of America.

[45]  Tal Pupko,et al.  Structural Genomics , 2005 .

[46]  L. Gráf,et al.  Three dimensional structures of S189D chymotrypsin and D189S trypsin mutants: the effect of polarity at site 189 on a protease-specific stabilization of the substrate-binding site. , 2003, Journal of molecular biology.

[47]  Allostery and Coupled Sequence Variation in Nuclear Hormone Receptors , 2004, Cell.

[48]  Rama Ranganathan,et al.  Structural Determinants of Allosteric Ligand Activation in RXR Heterodimers , 2004, Cell.

[49]  Olivier Poirot,et al.  3DCoffee@igs: a web server for combining sequences and structures into a multiple sequence alignment , 2004, Nucleic Acids Res..

[50]  J. Mccammon,et al.  HIV‐1 protease molecular dynamics of a wild‐type and of the V82F/I84V mutant: Possible contributions to drug resistance and a potential new target site for drugs , 2004, Protein science : a publication of the Protein Society.

[51]  O. Lichtarge,et al.  Evolutionary Trace of G Protein-coupled Receptors Reveals Clusters of Residues That Determine Global and Class-specific Functions* , 2004, Journal of Biological Chemistry.

[52]  R. Nussinov,et al.  Is allostery an intrinsic property of all dynamic proteins? , 2004, Proteins.

[53]  Gil Amitai,et al.  Network analysis of protein structures identifies functional residues. , 2004, Journal of molecular biology.

[54]  Dan S. Tawfik,et al.  The 'evolvability' of promiscuous protein functions , 2005, Nature Genetics.

[55]  Veronica Rotemberg,et al.  CoC: a database of universally conserved residues in protein folds , 2005, Bioinform..

[56]  A. del Sol,et al.  Small‐world network approach to identify key residues in protein–protein interaction , 2004, Proteins.

[57]  Simon Byrne,et al.  Switching between allosteric and dimerization inhibition of HIV-1 protease. , 2005, Chemistry & biology.