1932

Abstract

Numerous proteins, including cytokines and chemokines, enzymes and enzyme inhibitors, extracellular matrix proteins, and membrane receptors, bind heparin. Although they are traditionally classified as heparin-binding proteins, under normal physiological conditions these proteins actually interact with the heparan sulfate chains of one or more membrane or extracellular proteoglycans. Thus, they are more appropriately classified as heparan sulfate–binding proteins (HSBPs). This review provides an overview of the various modes of interaction between heparan sulfate and HSBPs, emphasizing biochemical and structural insights that improve our understanding of the many biological functions of heparan sulfate.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-biochem-060713-035314
2014-06-02
2024-03-19
Loading full text...

Full text loading...

/deliver/fulltext/biochem/83/1/annurev-biochem-060713-035314.html?itemId=/content/journals/10.1146/annurev-biochem-060713-035314&mimeType=html&fmt=ahah

Literature Cited

  1. Guerardel Y, Czeszak X, Sumanovski LT, Karamanos Y, Popescu O. 1.  et al. 2004. Molecular fingerprinting of carbohydrate structure phenotypes of three Porifera proteoglycan-like glyconectins. J. Biol. Chem. 279:15591–603 [Google Scholar]
  2. Yamada S, Morimoto H, Fujisawa T, Sugahara K. 2.  2007. Glycosaminoglycans in Hydra magnipapillata (Hydrozoa, Cnidaria): demonstration of chondroitin in the developing nematocyst, sting organelle, and structural characterization of glycosaminoglycans. Glycobiology 17:886–94 [Google Scholar]
  3. Lawrence R, Olson SK, Steele RE, Wang L, Warrior R. 3.  et al. 2008. Evolutionary differences in glycosaminoglycan fine structure detected by quantitative glycan reductive isotope labeling. J. Biol. Chem. 283:33674–84 [Google Scholar]
  4. Ori A, Wilkinson MC, Fernig DG. 4.  2011. A systems biology approach for the investigation of the heparin/heparan sulfate interactome. J. Biol. Chem. 286:19892–904 [Google Scholar]
  5. Bishop JR, Schuksz M, Esko JD. 5.  2007. Heparan sulphate proteoglycans fine-tune mammalian physiology. Nature 446:1030–37 [Google Scholar]
  6. Lin X, Wei G, Shi ZZ, Dryer L, Esko JD. 6.  et al. 2000. Disruption of gastrulation and heparan sulfate biosynthesis in EXT1-deficient mice. Dev. Biol. 224:299–311 [Google Scholar]
  7. Stickens D, Zak BM, Rougier N, Esko JD, Werb Z. 7.  2005. Mice deficient in Ext2 lack heparan sulfate and develop exostoses. Development 132:5055–68 [Google Scholar]
  8. Maccarana M, Sakura Y, Tawada A, Yoshida K, Lindahl U. 8.  1996. J. Biol. Chem. 27117804–10
  9. Murphy KJ, Merry CL, Lyon M, Thompson JE, Roberts IS, Gallagher JT. 9.  2004. J. Biol. Chem. 27927239–45
  10. Esko JD, Selleck SB. 10.  2002. Order out of chaos: assembly of ligand binding sites in heparan sulfate. Annu. Rev. Biochem. 71:435–71 [Google Scholar]
  11. Lawrence R, Brown JR, Al-Mafraji K, Lamanna WC, Beitel JR. 11.  et al. 2012. Disease-specific non-reducing end carbohydrate biomarkers for mucopolysaccharidoses. Nat. Chem. Biol. 8:197–204 [Google Scholar]
  12. Carlsson P, Kjellén L. 12.  2012. Heparin biosynthesis. Handb. Exp. Pharmacol. 207:23–41 [Google Scholar]
  13. Kreuger J, Kjellén L. 13.  2012. Heparan sulfate biosynthesis: regulation and variability. J. Histochem. Cytochem. 60:898–907 [Google Scholar]
  14. Esko JD, Sharon N. 14.  2009. Microbial lectins: hemagglutinins, adhesins, and toxins. Essentials of Glycobiology A Varki, RD Cummings, JD Esko, HH Freeze, P Stanley , et al., pp. 489–500 Cold Spring Harbor, N.Y.: Cold Spring Harb. Lab. [Google Scholar]
  15. Esko JD, Linhardt RJ. 15.  2009. Proteins that bind sulfated glyocosaminoglycans. Essentials of Glycobiology A Varki, RD Cummings, JD Esko, HH Freeze, P Stanley , et al., pp. 501–12 Cold Spring Harbor, N.Y.: Cold Spring Harb. Lab. [Google Scholar]
  16. Ishihara M, Fedarko NS, Conrad HE. 16.  1986. Transport of heparan sulfate into the nuclei of hepatocytes. J. Biol. Chem. 261:13575–80 [Google Scholar]
  17. Richardson TP, Trinkaus-Randall V, Nugent MA. 17.  2001. Regulation of heparan sulfate proteoglycan nuclear localization by fibronectin. J. Cell Sci. 114:1613–23 [Google Scholar]
  18. Chen L, Sanderson RD. 18.  2009. Heparanase regulates levels of syndecan-1 in the nucleus. PLoS ONE 4:e4947 [Google Scholar]
  19. Zong F, Fthenou E, Wolmer N, Hollósi P, Kovalszky I. 19.  et al. 2009. Syndecan-1 and FGF-2, but not FGF receptor-1, share a common transport route and co-localize with heparanase in the nuclei of mesenchymal tumor cells. PLoS ONE 4:e7346 [Google Scholar]
  20. Varki A, Etzler ME, Cummings RD, Esko JD. 20.  2009. Discovery and classification of glycan-binding proteins. Essentials of Glycobiology A Varki, RD Cummings, JD Esko, HH Freeze, P Stanley , et al., pp. 375–86 Cold Spring Harbor, N.Y: Cold Spring Harb. Lab. [Google Scholar]
  21. Radek KA, Taylor KR, Gallo RL. 21.  2009. FGF-10 and specific structural elements of dermatan sulfate size and sulfation promote maximal keratinocyte migration and cellular proliferation. Wound Repair Regen. 17:118–26 [Google Scholar]
  22. Taylor KR, Rudisill JA, Gallo RL. 22.  2005. Structural and sequence motifs in dermatan sulfate for promoting fibroblast growth factor 2 (FGF-2) and FGF-7 activity. J. Biol. Chem. 280:5300–6 [Google Scholar]
  23. Trowbridge JM, Rudisill JA, Ron D, Gallo RL. 23.  2002. Dermatan sulfate binds and potentiates activity of keratinocyte growth factor (FGF-7). J. Biol. Chem. 277:42815–20 [Google Scholar]
  24. Trowbridge JM, Gallo RL. 24.  2002. Dermatan sulfate: new functions from an old glycosaminoglycan. Glycobiology 12:R117–25 [Google Scholar]
  25. Ori A, Wilkinson MC, Fernig DG. 25.  2008. The heparanome and regulation of cell function: structures, functions and challenges. Front. Biosci. 13:4309–38 [Google Scholar]
  26. Faller B, Mely Y, Gerard D, Bieth JG. 26.  1992. Heparin-induced conformational change and activation of mucus proteinase inhibitor. Biochemistry 31:8285–90 [Google Scholar]
  27. Olson ST, Halvorson HR, Björk I. 27.  1991. Quantitative characterization of the thrombin–heparin interaction. Discrimination between specific and nonspecific binding models. J. Biol. Chem. 266:6342–52 [Google Scholar]
  28. Friedrich U, Blom AM, Dahlbäck B, Villoutreix BO. 28.  2001. Structural and energetic characteristics of the heparin-binding site in antithrombotic protein C. J. Biol. Chem. 276:24122–28 [Google Scholar]
  29. Thompson LD, Pantoliano MW, Springer BA. 29.  1994. Energetic characterization of the basic fibroblast growth factor–heparin interaction: identification of the heparin binding domain. Biochemistry 33:3831–40 [Google Scholar]
  30. Stanford KI, Bishop JR, Foley EM, Gonzales JC, Niesman IR. 30.  et al. 2009. Syndecan-1 is the primary heparan sulfate proteoglycan mediating hepatic clearance of triglyceride-rich lipoproteins in mice. J. Clin. Investig. 119:3236–45 [Google Scholar]
  31. Mascotti DP, Lohman TM. 31.  1995. Thermodynamics of charged oligopeptide–heparin interactions. Biochemistry 34:2908–15 [Google Scholar]
  32. Sheinerman FB, Norel R, Honig B. 32.  2000. Electrostatic aspects of protein–protein interactions. Curr. Opin. Struct. Biol. 10:153–59 [Google Scholar]
  33. Duchesne L, Octeau V, Bearon RN, Beckett A, Prior IA. 33.  et al. 2012. Transport of fibroblast growth factor 2 in the pericellular matrix is controlled by the spatial distribution of its binding sites in heparan sulfate. PLoS Biol. 10:e1001361 [Google Scholar]
  34. Sadir R, Imberty A, Baleux F, Lortat-Jacob H. 34.  2004. Heparan sulfate/heparin oligosaccharides protect stromal cell–derived factor 1 (SDF-1)/CXCL12 against proteolysis induced by CD26/dipeptidyl peptidase IV. J. Biol. Chem. 279:43854–60 [Google Scholar]
  35. Lortat-Jacob H, Baltzer F, Grimaud JA. 35.  1996. Heparin decreases the blood clearance of interferon-γ and increases its activity by limiting the processing of its carboxyl-terminal sequence. J. Biol. Chem. 271:16139–43 [Google Scholar]
  36. Yan D, Lin X. 36.  2009. Shaping morphogen gradients by proteoglycans. Cold Spring Harb. Perspect. Biol. 1:a002493 [Google Scholar]
  37. Bellaiche Y, The I, Perrimon N. 37.  1998. Tout-velu is a Drosophila homologue of the putative tumour suppressor EXT-1 and is needed for Hh diffusion. Nature 394:85–88 [Google Scholar]
  38. Baeg GH, Lin X, Khare N, Baumgartner S, Perrimon N. 38.  2001. Heparan sulfate proteoglycans are critical for the organization of the extracellular distribution of Wingless. Development 128:87–94 [Google Scholar]
  39. Takei Y, Ozawa Y, Sato M, Watanabe A, Tabata T. 39.  2004. Three Drosophila EXT genes shape morphogen gradients through synthesis of heparan sulfate proteoglycans. Development 131:73–82 [Google Scholar]
  40. Han C, Yan D, Belenkaya TY, Lin X. 40.  2005. Drosophila glypicans Dally and Dally-like shape the extracellular Wingless morphogen gradient in the wing disc. Development 132:667–79 [Google Scholar]
  41. Lei J, Wan FY, Lander AD, Nie Q. 41.  2011. Robustness of signaling gradient in Drosophila wing imaginal disc. Discret. Contin. Dyn. Syst. B 16:835–66 [Google Scholar]
  42. Carmeliet P, Ng YS, Nuyens D, Theilmeier G, Brusselmans K. 42.  et al. 1999. Impaired myocardial angiogenesis and ischemic cardiomyopathy in mice lacking the vascular endothelial growth factor isoforms VEGF164 and VEGF188. Nat. Med. 5:495–502 [Google Scholar]
  43. Bulow HE, Hobert O. 43.  2006. The molecular diversity of glycosaminoglycans shapes animal development. Annu. Rev. Cell Dev. Biol. 22:375–407 [Google Scholar]
  44. Lortat-Jacob H.44.  2009. The molecular basis and functional implications of chemokine interactions with heparan sulphate. Curr. Opin. Struct. Biol. 19:543–48 [Google Scholar]
  45. Weber M, Sixt M. 45.  2013. Live cell imaging of chemotactic dendritic cell migration in explanted mouse ear preparations. Methods Mol. Biol. 1013:215–26 [Google Scholar]
  46. Goodsell DS, Olson AJ. 46.  2000. Structural symmetry and protein function. Annu. Rev. Biophys. Biomol. Struct. 29:105–53 [Google Scholar]
  47. Belov AA, Mohammadi M. 47.  2013. Molecular mechanisms of fibroblast growth factor signaling in physiology and pathology. Cold Spring Harb. Perspect. Biol. 5:a015958 [Google Scholar]
  48. Mohammadi M, Olsen SK, Ibrahimi OA. 48.  2005. Structural basis for fibroblast growth factor receptor activation. Cytokine Growth Factor Rev. 16:107–37 [Google Scholar]
  49. DiGabriele AD, Lax I, Chen DI, Svahn CM, Jaye M. 49.  et al. 1998. Structure of a heparin-linked biologically active dimer of fibroblast growth factor. Nature 393:812–17 [Google Scholar]
  50. Brown A, Robinson CJ, Gallagher JT, Blundell TL. 50.  2013. Cooperative heparin-mediated oligomerization of fibroblast growth factor 1 (FGF1) precedes recruitment of FGFR2 to ternary complexes. Biophys. J. 104:1720–30 [Google Scholar]
  51. Pellegrini L, Burke DF, Von Delft F, Mulloy B, Blundell TL. 51.  2000. Crystal structure of fibroblast growth factor receptor ectodomain bound to ligand and heparin. Nature 407:1029–34 [Google Scholar]
  52. Beenken A, Eliseenkova AV, Ibrahimi OA, Olsen SK, Mohammadi M. 52.  2012. Plasticity in interactions of fibroblast growth factor 1 (FGF1) N terminus with FGF receptors underlies promiscuity of FGF1. J. Biol. Chem. 287:3067–78 [Google Scholar]
  53. Xu D, Young JH, Krahn JM, Song D, Corbett KD. 53.  et al. 2013. Stable RAGE–heparan sulfate complexes are essential for signal transduction. Am. Chem. Soc. Chem. Biol. 8:1611–20 [Google Scholar]
  54. Xu D, Young J, Song D, Esko JD. 54.  2011. Heparan sulfate is essential for high mobility group protein 1 (HMGB1) signaling by the receptor for advanced glycation end products (RAGE). J. Biol. Chem. 286:41736–44 [Google Scholar]
  55. Thinakaran G, Koo EH. 55.  2008. Amyloid precursor protein trafficking, processing, and function. J. Biol. Chem. 283:29615–19 [Google Scholar]
  56. Herms J, Anliker B, Heber S, Ring S, Fuhrmann M. 56.  et al. 2004. Cortical dysplasia resembling human type 2 lissencephaly in mice lacking all three APP family members. EMBO J. 23:4106–15 [Google Scholar]
  57. Soba P, Eggert S, Wagner K, Zentgraf H, Siehl K. 57.  et al. 2005. Homo- and heterodimerization of APP family members promotes intercellular adhesion. EMBO J. 24:3624–34 [Google Scholar]
  58. Dyrks T, Weidemann A, Multhaup G, Salbaum JM, Lemaire HG. 58.  et al. 1988. Identification, transmembrane orientation and biogenesis of the amyloid A4 precursor of Alzheimer's disease. EMBO J. 7:949–57 [Google Scholar]
  59. Xue Y, Lee S, Wang Y, Ha Y. 59.  2011. Crystal structure of the E2 domain of amyloid precursor protein–like protein 1 in complex with sucrose octasulfate. J. Biol. Chem. 286:29748–57 [Google Scholar]
  60. Gralle M, Botelho MG, Wouters FS. 60.  2009. Neuroprotective secreted amyloid precursor protein acts by disrupting amyloid precursor protein dimers. J. Biol. Chem. 284:15016–25 [Google Scholar]
  61. Mok SS, Sberna G, Heffernan D, Cappai R, Galatis D. 61.  et al. 1997. Expression and analysis of heparin-binding regions of the amyloid precursor protein of Alzheimer's disease. FEBS Lett. 415:303–7 [Google Scholar]
  62. Lee S, Xue Y, Hu J, Wang Y, Liu X. 62.  et al. 2011. The E2 domains of APP and APLP1 share a conserved mode of dimerization. Biochemistry 50:5453–64 [Google Scholar]
  63. Rossjohn J, Cappai R, Feil SC, Henry A, McKinstry WJ. 63.  et al. 1999. Crystal structure of the N-terminal, growth factor–like domain of Alzheimer amyloid precursor protein. Nat. Struct. Biol. 6:327–31 [Google Scholar]
  64. Dahms SO, Hoefgen S, Roeser D, Schlott B, Gührs KH, Than ME. 64.  2010. Structure and biochemical analysis of the heparin-induced E1 dimer of the amyloid precursor protein. Proc. Natl. Acad. Sci. USA 107:5381–86 [Google Scholar]
  65. Xue Y, Lee S, Ha Y. 65.  2011. Crystal structure of amyloid precursor-like protein 1 and heparin complex suggests a dual role of heparin in E2 dimerization. Proc. Natl. Acad. Sci. USA 108:16229–34 [Google Scholar]
  66. Proudfoot AE, Handel TM, Johnson Z, Lau EK, LiWang P. 66.  et al. 2003. Glycosaminoglycan binding and oligomerization are essential for the in vivo activity of certain chemokines. Proc. Natl. Acad. Sci. USA 100:1885–90 [Google Scholar]
  67. Salanga CL, Handel TM. 67.  2011. Chemokine oligomerization and interactions with receptors and glycosaminoglycans: the role of structural dynamics in function. Exp. Cell Res. 317:590–601 [Google Scholar]
  68. Hoogewerf AJ, Kuschert GS, Proudfoot AE, Borlat F, Clark-Lewis I. 68.  et al. 1997. Glycosaminoglycans mediate cell surface oligomerization of chemokines. Biochemistry 36:13570–78 [Google Scholar]
  69. Lortat-Jacob H, Grosdidier A, Imberty A. 69.  2002. Structural diversity of heparan sulfate binding domains in chemokines. Proc. Natl. Acad. Sci. USA 99:1229–34 [Google Scholar]
  70. Kuschert GS, Hoogewerf AJ, Proudfoot AE, Chung CW, Cooke RM. 70.  et al. 1998. Identification of a glycosaminoglycan binding surface on human interleukin-8. Biochemistry 37:11193–201 [Google Scholar]
  71. Zhang X, Chen L, Bancroft DP, Lai CK, Maione TE. 71.  1994. Crystal structure of recombinant human platelet factor 4. Biochemistry 33:8361–66 [Google Scholar]
  72. Ziarek JJ, Veldkamp CT, Zhang F, Murray NJ, Kartz GA. 72.  et al. 2013. Heparin oligosaccharides inhibit chemokine (CXC motif) ligand 12 (CXCL12) cardioprotection by binding orthogonal to the dimerization interface, promoting oligomerization, and competing with the chemokine (CXC motif) receptor 4 (CXCR4) N terminus. J. Biol. Chem. 288:737–46 [Google Scholar]
  73. Lau EK, Paavola CD, Johnson Z, Gaudry JP, Geretti E. 73.  et al. 2004. Identification of the glycosaminoglycan binding site of the CC chemokine, MCP-1: implications for structure and function in vivo. J. Biol. Chem. 279:22294–305 [Google Scholar]
  74. Goodger SJ, Robinson CJ, Murphy KJ, Gasiunas N, Harmer NJ. 74.  et al. 2008. Evidence that heparin saccharides promote FGF2 mitogenesis through two distinct mechanisms. J. Biol. Chem. 283:13001–8 [Google Scholar]
  75. Tan K, Duquette M, Liu JH, Shanmugasundaram K, Joachimiak A. 75.  et al. 2008. Heparin-induced cis- and trans-dimerization modes of the thrombospondin-1 N-terminal domain. J. Biol. Chem. 283:3932–41 [Google Scholar]
  76. Lietha D, Chirgadze DY, Mulloy B, Blundell TL, Gherardi E. 76.  2001. Crystal structures of NK1–heparin complexes reveal the basis for NK1 activity and enable engineering of potent agonists of the MET receptor. EMBO J. 20:5543–55 [Google Scholar]
  77. Vander Kooi CW, Jusino MA, Perman B, Neau DB, Bellamy HD, Leahy DJ. 77.  2007. Structural basis for ligand and heparin binding to neuropilin B domains. Proc. Natl. Acad. Sci. USA 104:6152–57 [Google Scholar]
  78. Coles CH, Shen Y, Tenney AP, Siebold C, Sutton GC. 78.  et al. 2011. Proteoglycan-specific molecular switch for RPTPσ clustering and neuronal extension. Science 332:484–88 [Google Scholar]
  79. Shaw AS, Filbert EL. 79.  2009. Scaffold proteins and immune-cell signalling. Nat. Rev. Immunol. 9:47–56 [Google Scholar]
  80. Jackson CM, Nemerson Y. 80.  1980. Blood coagulation. Annu. Rev. Biochem. 49:765–811 [Google Scholar]
  81. Rau JC, Beaulieu LM, Huntington JA, Church FC. 81.  2007. Serpins in thrombosis, hemostasis and fibrinolysis. J. Thromb. Haemost. 5:Suppl. 1102–15 [Google Scholar]
  82. Björk I, Lindahl U. 82.  1982. Mechanism of the anticoagulant action of heparin. Mol. Cell Biochem. 48:161–82 [Google Scholar]
  83. Olson ST, Björk I. 83.  1991. Predominant contribution of surface approximation to the mechanism of heparin acceleration of the antithrombin–thrombin reaction. Elucidation from salt concentration effects. J. Biol. Chem. 266:6353–64 [Google Scholar]
  84. Li W, Johnson DJ, Esmon CT, Huntington JA. 84.  2004. Structure of the antithrombin–thrombin–heparin ternary complex reveals the antithrombotic mechanism of heparin. Nat. Struct. Mol. Biol. 11:857–62 [Google Scholar]
  85. Bray B, Lane DA, Freyssinet JM, Pejler G, Lindahl U. 85.  1989. Anti-thrombin activities of heparin. Effect of saccharide chain length on thrombin inhibition by heparin cofactor II and by antithrombin. Biochem. J. 262:225–32 [Google Scholar]
  86. Johnson DJ, Li W, Adams TE, Huntington JA. 86.  2006. Antithrombin–S195A factor Xa-heparin structure reveals the allosteric mechanism of antithrombin activation. EMBO J. 25:2029–37 [Google Scholar]
  87. Carter WJ, Cama E, Huntington JA. 87.  2005. Crystal structure of thrombin bound to heparin. J. Biol. Chem. 280:2745–49 [Google Scholar]
  88. Dementiev A, Petitou M, Herbert JM, Gettins PG. 88.  2004. The ternary complex of antithrombin–anhydrothrombin–heparin reveals the basis of inhibitor specificity. Nat. Struct. Mol. Biol. 11:863–67 [Google Scholar]
  89. Neese LL, Wolfe CA, Church FC. 89.  1998. Contribution of basic residues of the D and H helices in heparin binding to protein C inhibitor. Arch. Biochem. Biophys. 355:101–8 [Google Scholar]
  90. Li W, Adams TE, Nangalia J, Esmon CT, Huntington JA. 90.  2008. Molecular basis of thrombin recognition by protein C inhibitor revealed by the 1.6-Å structure of the heparin-bridged complex. Proc. Natl. Acad. Sci. USA 105:4661–66 [Google Scholar]
  91. Yang L, Manithody C, Rezaie AR. 91.  2002. Contribution of basic residues of the 70–80-loop to heparin binding and anticoagulant function of activated protein C. Biochemistry 41:6149–57 [Google Scholar]
  92. Li W, Huntington JA. 92.  2008. The heparin binding site of protein C inhibitor is protease-dependent. J. Biol. Chem. 283:36039–45 [Google Scholar]
  93. Ibrahimi OA, Zhang F, Hrstka SC, Mohammadi M, Linhardt RJ. 93.  2004. Kinetic model for FGF, FGFR, and proteoglycan signal transduction complex assembly. Biochemistry 43:4724–30 [Google Scholar]
  94. Kalinina J, Dutta K, Ilghari D, Beenken A, Goetz R. 94.  et al. 2012. The alternatively spliced acid box region plays a key role in FGF receptor autoinhibition. Structure 20:77–88 [Google Scholar]
  95. Powell AK, Fernig DG, Turnbull JE. 95.  2002. Fibroblast growth factor receptors 1 and 2 interact differently with heparin/heparan sulfate: implications for dynamic assembly of a ternary signaling complex. J. Biol. Chem. 277:28554–63 [Google Scholar]
  96. Schlessinger J, Plotnikov AN, Ibrahimi OA, Eliseenkova AV, Yeh BK. 96.  et al. 2000. Crystal structure of a ternary FGF-FGFR-heparin complex reveals a dual role for heparin in FGFR binding and dimerization. Mol. Cell 6:743–50 [Google Scholar]
  97. Robinson CJ, Harmer NJ, Goodger SJ, Blundell TL, Gallagher JT. 97.  2005. Cooperative dimerization of fibroblast growth factor 1 (FGF1) upon a single heparin saccharide may drive the formation of 2:2:1 FGF1·FGFR2c·heparin ternary complexes. J. Biol. Chem. 280:42274–82 [Google Scholar]
  98. Olsen SK, Li JY, Bromleigh C, Eliseenkova AV, Ibrahimi OA. 98.  et al. 2006. Structural basis by which alternative splicing modulates the organizer activity of FGF8 in the brain. Genes Dev. 20:185–98 [Google Scholar]
  99. Fukuhara N, Howitt JA, Hussain SA, Hohenester E. 99.  2008. Structural and functional analysis of slit and heparin binding to immunoglobulin-like domains 1 and 2 of Drosophila Robo. J. Biol. Chem. 283:16226–34 [Google Scholar]
  100. del Sol A, Tsai CJ, Ma B, Nussinov R. 100.  2009. The origin of allosteric functional modulation: multiple pre-existing pathways. Structure 17:1042–50 [Google Scholar]
  101. Tsai CJ, del Sol A, Nussinov R. 101.  2008. Allostery: absence of a change in shape does not imply that allostery is not at play. J. Mol. Biol. 378:1–11 [Google Scholar]
  102. Langdown J, Belzar KJ, Savory WJ, Baglin TP, Huntington JA. 102.  2009. The critical role of hinge-region expulsion in the induced-fit heparin binding mechanism of antithrombin. J. Mol. Biol. 386:1278–89 [Google Scholar]
  103. Jin L, Abrahams JP, Skinner R, Petitou M, Pike RN, Carrell RW. 103.  1997. The anticoagulant activation of antithrombin by heparin. Proc. Natl. Acad. Sci. USA 94:14683–88 [Google Scholar]
  104. Johnson DJ, Langdown J, Huntington JA. 104.  2010. Molecular basis of factor IXa recognition by heparin-activated antithrombin revealed by a 1.7-Å structure of the ternary complex. Proc. Natl. Acad. Sci. USA 107:645–50 [Google Scholar]
  105. Al-Mohanna F, Parhar R, Kotwal GJ. 105.  2001. Vaccinia virus complement control protein is capable of protecting xenoendothelial cells from antibody binding and killing by human complement and cytotoxic cells. Transplantation 71:796–801 [Google Scholar]
  106. Khan S, Nan R, Gor J, Mulloy B, Perkins SJ. 106.  2012. Bivalent and co-operative binding of complement factor H to heparan sulfate and heparin. Biochem. J. 444:417–28 [Google Scholar]
  107. Ganesh VK, Smith SA, Kotwal GJ, Murthy KH. 107.  2004. Structure of Vaccinia complement protein in complex with heparin and potential implications for complement regulation. Proc. Natl. Acad. Sci. USA 101:8924–29 [Google Scholar]
  108. Murthy KHM, Smith SA, Ganesh VK, Judge KW, Mullin N. 108.  et al. 2001. Crystal structure of a complement control protein that regulates both pathways of complement activation and binds heparan sulfate proteoglycans. Cell 104:301–11 [Google Scholar]
  109. Meri S, Pangburn MK. 109.  1990. Discrimination between activators and nonactivators of the alternative pathway of complement: regulation via a sialic acid/polyanion binding site on factor H. Proc. Natl. Acad. Sci. USA 87:3982–86 [Google Scholar]
  110. Kreuger J, Spillmann D, Li JP, Lindahl U. 110.  2006. Interactions between heparan sulfate and proteins: the concept of specificity. J. Cell Biol. 174:323–27 [Google Scholar]
  111. Lindahl U, Li JP. 111.  2009. Interactions between heparan sulfate and proteins—design and functional implications. Int. Rev. Cell Mol. Biol. 276:105–59 [Google Scholar]
  112. Richard B, Swanson R, Olson ST. 112.  2009. The signature 3-O-sulfo group of the anticoagulant heparin sequence is critical for heparin binding to antithrombin but is not required for allosteric activation. J. Biol. Chem. 284:27054–64 [Google Scholar]
  113. Turnbull JE, Fernig DG, Ke Y, Wilkinson MC, Gallagher JT. 113.  1992. Identification of the basic fibroblast growth factor binding sequence in fibroblast heparan sulfate. J. Biol. Chem. 267:10337–41 [Google Scholar]
  114. Lortat-Jacob H, Turnbull JE, Grimaud JA. 114.  1995. Molecular organization of the interferon γ–binding domain in heparan sulphate. Biochem. J. 310:497–505 [Google Scholar]
  115. Spillmann D, Witt D, Lindahl U. 115.  1998. Defining the interleukin-8-binding domain of heparan sulfate. J. Biol. Chem. 273:15487–93 [Google Scholar]
  116. Kamimura K, Koyama T, Habuchi H, Ueda R, Masu M. 116.  et al. 2006. Specific and flexible roles of heparan sulfate modifications in Drosophila FGF signaling. J. Cell Biol. 174:773–78 [Google Scholar]
  117. 117.  Deleted in proof
  118. Liu J, Shworak NW, Sinay P, Schwartz JJ, Zhang L. 118.  et al. 1999. Expression of heparan sulfate D-glucosaminyl 3-O-sulfotransferase isoforms reveals novel substrate specificities. J. Biol. Chem. 274:5185–92 [Google Scholar]
  119. Neugebauer JM, Cadwallader AB, Amack JD, Bisgrove BW, Yost HJ. 119.  2013. Differential roles for 3-OSTs in the regulation of cilia length and motility. Development 140:3892–902 [Google Scholar]
  120. Jemth P, Kreuger J, Kusche-Gullberg M, Sturiale L, Giménez-Gallego G, Lindahl U. 120.  2002. Biosynthetic oligosaccharide libraries for identification of protein-binding heparan sulfate motifs—exploring the structural diversity by screening for fibroblast growth factor (FGF) 1 and FGF2 binding. J. Biol. Chem. 277:30567–73 [Google Scholar]
  121. Faham S, Hileman RE, Fromm JR, Linhardt RJ, Rees DC. 121.  1996. Heparin structure and interactions with basic fibroblast growth factor. Science 271:1116–20 [Google Scholar]
  122. Pye DA, Vives RR, Turnbull JE, Hyde P, Gallagher JT. 122.  1998. J. Biol. Chem. 27322936–42
  123. Jastrebova N, Vanwildemeersch M, Lindahl U, Spillmann D. 123.  2010. J. Biol. Chem. 28526842–51
  124. Luo Y, Ye S, Kan M, McKeehan WL. 124.  2006. J. Cell. Biochem. 971241–58
  125. Dulaney SB, Huang X. 125.  2012. Strategies in synthesis of heparin/heparan sulfate oligosaccharides: 2000–present. Adv. Carbohydr. Chem. Biochem. 67:95–136 [Google Scholar]
  126. Turnbull JE, Gallagher JT. 126.  1991. Sequence analysis of heparan sulphate indicates defined location of N-sulphated glucosamine and iduronate 2-sulphate residues proximal to the protein-linkage region. Biochem. J. 277:297–303 [Google Scholar]
  127. Stringer SE, Gallagher JT. 127.  1997. Specific binding of the chemokine platelet factor 4 to heparan sulfate. J. Biol. Chem. 272:20508–14 [Google Scholar]
  128. Stringer SE, Forster MJ, Mulloy B, Bishop CR, Graham GJ, Gallagher JT. 128.  2002. Characterization of the binding site on heparan sulfate for macrophage inflammatory protein 1α. Blood 100:1543–50 [Google Scholar]
  129. Axelsson J, Xu D, Na Kang B, Nussbacher JK, Handel TM. 129.  et al. 2012. Inactivation of heparan sulfate 2-O-sulfotransferase accentuates neutrophil infiltration during acute inflammation in mice. Blood 120:1742–51 [Google Scholar]
  130. Carter NM, Ali S, Kirby JA. 130.  2003. Endothelial inflammation: the role of differential expression of N-deacetylase/N-sulphotransferase enzymes in alteration of the immunological properties of heparan sulphate. J. Cell Sci. 116:3591–600 [Google Scholar]
  131. Krenn EC, Wille I, Gesslbauer B, Poteser M, van Kuppevelt TH, Kungl AJ. 131.  2008. Glycanogenomics: a qPCR-approach to investigate biological glycan function. Biochem. Biophys. Res. Commun. 375:297–302 [Google Scholar]
  132. Catlow KR, Deakin JA, Wei Z, Delehedde M, Fernig DG. 132.  et al. 2008. Interactions of hepatocyte growth factor/scatter factor with various glycosaminoglycans reveal an important interplay between the presence of iduronate and sulfate density. J. Biol. Chem. 283:5235–48 [Google Scholar]
  133. Krilleke D, Ng YS, Shima DT. 133.  2009. The heparin-binding domain confers diverse functions of VEGF-A in development and disease: a structure–function study. Biochem. Soc. Trans. 37:1201–6 [Google Scholar]
  134. Shao C, Zhang F, Kemp MM, Linhardt RJ, Waisman DM. 134.  et al. 2006. Crystallographic analysis of calcium-dependent heparin binding to annexin A2. J. Biol. Chem. 281:31689–95 [Google Scholar]
  135. Capila I, VanderNoot VA, Mealy TR, Seaton BA, Linhardt RJ. 135.  1999. Interaction of heparin with annexin V. FEBS Lett. 446:327–30 [Google Scholar]
  136. Norgard-Sumnicht KE, Varki NM, Varki A. 136.  1993. Calcium-dependent heparin-like ligands for L-selectin in nonlymphoid endothelial cells. Science 261:480–83 [Google Scholar]
  137. Koenig A, Norgard-Sumnicht K, Linhardt R, Varki A. 137.  1998. Differential interactions of heparin and heparan sulfate glycosaminoglycans with the selectins. Implications for the use of unfractionated and low molecular weight heparins as therapeutic agents. J. Clin. Investig. 101:877–89 [Google Scholar]
  138. Ricard-Blum S, Beraud M, Raynal N, Farndale RW, Ruggiero F. 138.  2006. Structural requirements for heparin/heparan sulfate binding to type V collagen. J. Biol. Chem. 281:25195–204 [Google Scholar]
  139. Chakravarty L, Rogers L, Quach T, Breckenridge S, Kolattukudy PE. 139.  1998. Lysine 58 and histidine 66 at the C-terminal α-helix of monocyte chemoattractant protein 1 are essential for glycosaminoglycan binding. J. Biol. Chem. 273:29641–47 [Google Scholar]
  140. Sheng GJ, Oh YI, Chang SK, Hsieh-Wilson LC. 140.  2013. Tunable heparan sulfate mimetics for modulating chemokine activity. J. Am. Chem. Soc. 135:10898–901 [Google Scholar]
  141. Hanoulle X, Melchior A, Sibille N, Parent B, Denys A. 141.  et al. 2007. Structural and functional characterization of the interaction between cyclophilin B and a heparin-derived oligosaccharide. J. Biol. Chem. 282:34148–58 [Google Scholar]
  142. Cardin AD, Weintraub HJ. 142.  1989. Molecular modeling of protein–glycosaminoglycan interactions. Arteriosclerosis 9:21–32 [Google Scholar]
  143. Sobel M, Soler DF, Kermode JC, Harris RB. 143.  1992. Localization and characterization of a heparin binding domain peptide of human von Willebrand factor. J. Biol. Chem. 267:8857–62 [Google Scholar]
  144. Margalit H, Fischer N, Ben-Sasson SA. 144.  1993. Comparative analysis of structurally defined heparin binding sequences reveals a distinct spatial distribution of basic residues. J. Biol. Chem. 268:19228–31 [Google Scholar]
  145. Munoz EM, Linhardt RJ. 145.  2004. Heparin-binding domains in vascular biology. Arterioscler. Thromb. Vasc. Biol. 24:1549–57 [Google Scholar]
  146. Moon AF, Xu Y, Woody SM, Krahn JM, Linhardt RJ. 146.  et al. 2012. Dissecting the substrate recognition of 3-O-sulfotransferase for the biosynthesis of anticoagulant heparin. Proc. Natl. Acad. Sci. USA 109:5265–70 [Google Scholar]
  147. Xu Y, Masuko S, Takieddin M, Xu H, Liu R. 147.  et al. 2011. Chemoenzymatic synthesis of homogeneous ultralow molecular weight heparins. Science 334:498–501 [Google Scholar]
  148. Peterson SB, Liu J. 148.  2013. Multi-faceted substrate specificity of heparanase. Matrix Biol. 32:223–27 [Google Scholar]
  149. Schuksz M, Fuster MM, Brown JR, Crawford BE, Ditto DP. 149.  et al. 2008. Surfen, a small molecule antagonist of heparan sulfate. Proc. Natl. Acad. Sci. USA 105:13075–80 [Google Scholar]
/content/journals/10.1146/annurev-biochem-060713-035314
Loading
/content/journals/10.1146/annurev-biochem-060713-035314
Loading

Data & Media loading...

  • Article Type: Review Article
This is a required field
Please enter a valid email address
Approval was a Success
Invalid data
An Error Occurred
Approval was partially successful, following selected items could not be processed due to error