1932

Abstract

Disulfide bonds represent versatile posttranslational modifications whose roles encompass the structure, catalysis, and regulation of protein function. Due to the oxidizing nature of the extracellular environment, disulfide bonds found in secreted proteins were once believed to be inert. This notion has been challenged by the discovery of redox-sensitive disulfides that, once cleaved, can lead to changes in protein activity. These functional disulfides are twisted into unique configurations, leading to high strain and potential energy. In some cases, cleavage of these disulfides can lead to a gain of function in protein activity. Thus, these motifs can be referred to as switches. We describe the couples that control redox in the extracellular environment, examine several examples of proteins with switchable disulfides, and discuss the potential applications of disulfides in molecular biology.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-chembioeng-080615-033553
2016-06-07
2024-03-29
Loading full text...

Full text loading...

/deliver/fulltext/chembioeng/7/1/annurev-chembioeng-080615-033553.html?itemId=/content/journals/10.1146/annurev-chembioeng-080615-033553&mimeType=html&fmt=ahah

Literature Cited

  1. Jones DP. 1.  2006. Redefining oxidative stress. Antioxid. Redox Signal. 8:1865–79 [Google Scholar]
  2. Kemp M, Go Y-M, Jones DP. 2.  2008. Nonequilibrium thermodynamics of thiol/disulfide redox systems: A perspective on redox systems biology. Free Radic. Biol. Med. 44:921–37 [Google Scholar]
  3. Chaiswing L, Oberley TD. 3.  2010. Extracellular/microenvironmental redox state. Antioxid. Redox Signal. 13:449–65 [Google Scholar]
  4. Go YM, Jones DP. 4.  2008. Redox compartmentalization in eukaryotic cells. Biochim. Biophys. Acta 1780:1273–90 [Google Scholar]
  5. Jones DP, Park Y, Gletsu-Miller N, Liang Y, Yu T. 5.  et al. 2011. Dietary sulfur amino acid effects on fasting plasma cysteine/cystine redox potential in humans. Nutrition 27:199–205 [Google Scholar]
  6. Go YM, Jones DP. 6.  2005. Intracellular proatherogenic events and cell adhesion modulated by extracellular thiol/disulfide redox state. Circulation 111:2973–80 [Google Scholar]
  7. Jiang S, Moriarty-Craige SE, Orr M, Cai J, Sternberg P, Jones DP. 7.  2005. Oxidant-induced apoptosis in human retinal pigment epithelial cells: dependence on extracellular redox state. Investig. Ophthalmol. Vis. Sci. 46:1054–61 [Google Scholar]
  8. Ramirez A, Ramadan B, Ritzenthaler JD, Rivera HN, Jones DP, Roman J. 8.  2007. Extracellular cysteine/cystine redox potential controls lung fibroblast proliferation and matrix expression through upregulation of transforming growth factor-β. Am. J. Physiol. Lung Cell. Mol. Physiol. 293:L972–81 [Google Scholar]
  9. Hildebrandt W, Kinscherf R, Hauer K, Holm E, Dröge W. 9.  2002. Plasma cystine concentration and redox state in aging and physical exercise. Mech. Ageing Dev. 123:1269–81 [Google Scholar]
  10. Jones DP. 10.  2006. Extracellular redox state: refining the definition of oxidative stress in aging. Rejuvenation Res. 9:169–81 [Google Scholar]
  11. Iyer SS, Ramirez AM, Ritzenthaler JD, Torres-Gonzalez E, Roser-Page S. 11.  et al. 2009. Oxidation of extracellular cysteine/cystine redox state in bleomycin-induced lung fibrosis. Am. J. Physiol. Lung Cell. Mol. Physiol. 296:L37–45 [Google Scholar]
  12. Go YM, Jones DP. 12.  2011. Cysteine/cystine redox signaling in cardiovascular disease. Free Radic. Biol. Med. 50:495–509 [Google Scholar]
  13. Yang H, Zhu JW, Yuan JF, Yang HM, Wang ST. 13.  et al. 2012. Extracellular cysteine (Cys)/cystine (CySS) redox regulates metabotropic glutamate receptor 5 activity. Biochimie 94:617–27 [Google Scholar]
  14. Samiec PS, Drews-Botsch C, Flagg EW, Kurtz JC, Sternberg P. 14.  et al. 1998. Glutathione in human plasma: decline in association with aging, age-related macular degeneration, and diabetes. Free Radic. Biol. Med. 24:699–704 [Google Scholar]
  15. Chawla RK, Lewis FW, Kutner MH, Bate DM, Roy RG, Rudman D. 15.  1984. Plasma cysteine, cystine, and glutathione in cirrhosis. Gastroenterology 87:770–76 [Google Scholar]
  16. Suh JH, Kanathezhath B, Shenvi S, Guo H, Zhou A. 16.  et al. 2014. Thiol/redox metabolomic profiling implicates GSH dysregulation in early experimental graft versus host disease (GVHD). PLOS ONE 9:e88868 [Google Scholar]
  17. Schmidt B, Ho L, Hogg PJ. 17.  2006. Allosteric disulfide bonds. Biochemistry 45:7429–33 [Google Scholar]
  18. Thornton JM. 18.  1981. Disulphide bridges in globular proteins. J. Mol. Biol. 151:261–87 [Google Scholar]
  19. Jones DP, Go Y-M, Anderson CL, Ziegler TR, Kinkade JM, Kirlin WG. 19.  2004. Cysteine/cystine couple is a newly recognized node in the circuitry for biologic redox signaling and control. FASEB J. 18:1246–48 [Google Scholar]
  20. Jonas CR, Gu LH, Nkabyo YS, Mannery YO, Avissar NE. 20.  et al. 2003. Glutamine and KGF each regulate extracellular thiol/disulfide redox and enhance proliferation in Caco-2 cells. Am. J. Physiol. Regul. Integr. Comp. Physiol. 285:R1421–29 [Google Scholar]
  21. Nkabyo YS, Go Y-M, Ziegler TR, Jones DP. 21.  2005. Extracellular cysteine/cystine redox regulates the p44/p42 MAPK pathway by metalloproteinase-dependent epidermal growth factor receptor signaling. Am. J. Physiol. Gastrointest. Liver Physiol. 289:G70–78 [Google Scholar]
  22. Blanco RA, Ziegler TR, Carlson BA, Cheng PY, Park Y. 22.  et al. 2007. Diurnal variation in glutathione and cysteine redox states in human plasma. Am. J. Clin. Nutr. 86:1016–23 [Google Scholar]
  23. Mandal PK, Seiler A, Perisic T, Kölle P, Canak AB. 23.  et al. 2010. System xc and thioredoxin reductase 1 cooperatively rescue glutathione deficiency. J. Biol. Chem. 285:22244–53 [Google Scholar]
  24. Dröge W, Eck HP, Gmünder H, Mihm S. 24.  1991. Modulation of lymphocyte functions and immune responses by cysteine and cysteine derivatives. Am. J. Med. 91:140S–44S [Google Scholar]
  25. Banjac A, Perisic T, Sato H, Seiler A, Bannai S. 25.  et al. 2008. The cystine/cysteine cycle: a redox cycle regulating susceptibility versus resistance to cell death. Oncogene 27:1618–28 [Google Scholar]
  26. Bannai S. 26.  1984. Transport of cystine and cysteine in mammalian cells. Biochim. Biophys. Acta 779:289–306 [Google Scholar]
  27. Iwata S, Hori T, Sato N, Hirota K, Sasada T. 27.  et al. 1997. Adult T cell leukemia (ATL)-derived factor/human thioredoxin prevents apoptosis of lymphoid cells induced by l-cystine and glutathione depletion: possible involvement of thiol-mediated redox regulation in apoptosis caused by pro-oxidant state. J. Immunol. 158:3108–17 [Google Scholar]
  28. Jones DP, Carlson JL, Mody VC, Cai J, Lynn MJ, Sternberg P. 28.  2000. Redox state of glutathione in human plasma. Free Radic. Biol. Med. 28:625–35 [Google Scholar]
  29. Ballatori N, Hammond CL, Cunningham JB, Krance SM, Marchan R. 29.  2005. Molecular mechanisms of reduced glutathione transport: role of the MRP/CFTR/ABCC and OATP/SLC21A families of membrane proteins. Toxicol. Appl. Pharmacol. 204:238–55 [Google Scholar]
  30. Cole SPC, Deeley RG. 30.  2006. Transport of glutathione and glutathione conjugates by MRP1. Trends Pharmacol. Sci. 27:438–46 [Google Scholar]
  31. Meredith MJ, Reed DJ. 31.  1982. Status of the mitochondrial pool of glutathione in the isolated hepatocyte. J. Biol. Chem. 257:3747–53 [Google Scholar]
  32. Wright EC, Stern J, Ersser R, Patrick AD. 32.  1979. Glutathionuria: γ-glutamyl transpeptidase deficiency. J. Inherit. Metab. Dis. 2:3–7 [Google Scholar]
  33. Wendel A, Cikryt P. 33.  1980. The level and half-life of glutathione in human plasma. FEBS Lett. 120:209–11 [Google Scholar]
  34. Hanigan MH, Ricketts WA. 34.  1993. Extracellular glutathione is a source of cysteine for cells that express γ-glutamyl transpeptidase. Biochemistry 32:6302–6 [Google Scholar]
  35. Nakamura H, Vaage J, Valen G, Padilla CA, Björnstedt M, Holmgren A. 35.  1998. Measurements of plasma glutaredoxin and thioredoxin in healthy volunteers and during open-heart surgery. Free Radic. Biol. Med. 24:1176–86 [Google Scholar]
  36. Jones DP, Mody VC, Carlson JL, Lynn MJ, Sternberg P. 36.  2002. Redox analysis of human plasma allows separation of pro-oxidant events of aging from decline in antioxidant defenses. Free Radic. Biol. Med. 33:1290–300 [Google Scholar]
  37. Jonas CR, Ziegler TR, Gu LH, Jones DP. 37.  2002. Extracellular thiol/disulfide redox state affects proliferation rate in a human colon carcinoma (Caco2) cell line. Free Radic. Biol. Med. 33:1499–506 [Google Scholar]
  38. Adams JD, Lauterburg BH, Mitchell JR. 38.  1983. Plasma glutathione and glutathione disulfide in the rat: regulation and response to oxidative stress. J. Pharmacol. Exp. Ther. 227:749–54 [Google Scholar]
  39. Franco R, Cidlowski JA. 39.  2012. Glutathione efflux and cell death. Antioxid. Redox Signal. 17:1694–713 [Google Scholar]
  40. Dringen R. 40.  2000. Metabolism and functions of glutathione in brain. Prog. Neurobiol. 62:649–71 [Google Scholar]
  41. Berggren M, Dawson J, Moldéus P. 41.  1984. Glutathione biosynthesis in the isolated perfused rat lung: utilization of extracellular glutathione. FEBS Lett. 176:189–92 [Google Scholar]
  42. Brown LAS, Ping X-D, Harris FL, Gauthier TW. 42.  2007. Glutathione availability modulates alveolar macrophage function in the chronic ethanol-fed rat. Am. J. Physiol. Lung Cell. Mol. Physiol. 292:L824–32 [Google Scholar]
  43. Dahm LJ, Jones DP. 43.  2000. Rat jejunum controls luminal thiol-disulfide redox. J. Nutr. 130:2739–45 [Google Scholar]
  44. Laurent TC, Moore EC, Reichard P. 44.  1964. Enzymatic synthesis of deoxyribonucleotides. J. Biol. Chem. 239:3436–44 [Google Scholar]
  45. Powis G, Montfort WR. 45.  2001. Properties and biological activities of thioredoxins. Annu. Rev. Pharmacol. Toxicol. 41:261–95 [Google Scholar]
  46. Matsui M, Oshima M, Oshima H, Takaku K, Taketo MM. 46.  et al. 1996. Early embryonic lethality caused by targeted disruption of the mouse thioredoxin gene. Dev. Biol. 178:179–85 [Google Scholar]
  47. Rubartelli A, Bajetto A, Allavena G, Wollman E, Sitia R. 47.  1992. Secretion of thioredoxin by normal and neoplastic cells through a leaderless secretory pathway. J. Biol. Chem. 267:24161–64 [Google Scholar]
  48. Watson WH, Pohl J, Montfort WR, Stuchlik O, Reed MS. 48.  et al. 2003. Redox potential of human thioredoxin 1 and identification of a second dithiol/disulfide motif. J. Biol. Chem. 278:33408–15 [Google Scholar]
  49. Tagaya Y, Okada M, Sugie K, Kasahara T, Yodoi J. 49.  et al. 1988. IL-2 receptor(p55)/Tac-inducing factor: purification and characterization of adult T cell leukemia-derived factor. J. Immunol. 140:2614–20 [Google Scholar]
  50. Kondo N, Ishii Y, Kwon Y-W, Tanito M, Horita H. 50.  et al. 2004. Redox-sensing release of human thioredoxin from T lymphocytes with negative feedback loops. J. Immunol. 172:442–48 [Google Scholar]
  51. Jikimoto T, Nishikubo Y, Koshiba M, Kanagawa S, Morinobu S. 51.  et al. 2002. Thioredoxin as a biomarker for oxidative stress in patients with rheumatoid arthritis. Mol. Immunol. 38:765–72 [Google Scholar]
  52. Nakamura H, Matsuda M, Furuke K, Kitaoka Y, Iwata S. 52.  et al. 1994. Adult T cell leukemia-derived factor/human thioredoxin protects endothelial F-2 cell injury caused by activated neutrophils or hydrogen peroxide. Immunol. Lett. 42:213 [Google Scholar]
  53. Bertini R, Howard OM, Dong HF, Oppenheim JJ, Bizzarri C. 53.  et al. 1999. Thioredoxin, a redox enzyme released in infection and inflammation, is a unique chemoattractant for neutrophils, monocytes, and T cells. J. Exp. Med. 189:1783–89 [Google Scholar]
  54. Matsuda M, Masutani H, Nakamura H, Miyajima S, Yamauchi A. 54.  et al. 1991. Protective activity of adult T cell leukemia-derived factor (ADF) against tumor necrosis factor-dependent cytotoxicity on U937 cells. J. Immunol. 147:3837–41 [Google Scholar]
  55. Wakasugi N, Tagaya Y, Wakasugi H, Mitsui A, Maeda M. 55.  et al. 1990. Adult T-cell leukemia-derived factor/thioredoxin, produced by both human T-lymphotropic virus type I- and Epstein-Barr virus-transformed lymphocytes, acts as an autocrine growth factor and synergizes with interleukin 1 and interleukin 2. PNAS 87:8282–86 [Google Scholar]
  56. Oblong JE, Berggren M, Gasdaska PY, Powis G. 56.  1994. Site-directed mutagenesis of active site cysteines in human thioredoxin produces competitive inhibitors of human thioredoxin reductase and elimination of mitogenic properties of thioredoxin. J. Biol. Chem. 269:11714–20 [Google Scholar]
  57. Gasdaska JR, Berggren M, Powis G. 57.  1995. Cell growth stimulation by the redox protein thioredoxin occurs by a novel helper mechanism. Cell Growth Differ. 6:1643–50 [Google Scholar]
  58. Söderberg A, Sahaf B, Rosén A, Rose A. 58.  2000. Thioredoxin reductase, a redox-active selenoprotein, is secreted by normal and neoplastic cells: presence in human plasma. Cancer Res. 60:2281–89 [Google Scholar]
  59. Lundström J, Holmgren A. 59.  1993. Determination of the reduction-oxidation potential of the thioredoxin-like domains of protein disulfide-isomerase from the equilibrium with glutathione and thioredoxin. Biochemistry 32:6649–55 [Google Scholar]
  60. Jiang XM, Fitzgerald M, Grant CM, Hogg PJ. 60.  1999. Redox control of exofacial protein thiols/disulfides by protein disulfide isomerase. J. Biol. Chem. 274:2416–23 [Google Scholar]
  61. Chen K, Detwiler TC, Essex DW. 61.  1995. Characterization of protein disulphide isomerase released from activated platelets. Br. J. Haematol. 90:425–31 [Google Scholar]
  62. Barbouche R, Miquelis R, Jones IM, Fenouillet E. 62.  2003. Protein-disulfide isomerase-mediated reduction of two disulfide bonds of HIV envelope glycoprotein 120 occurs post-CXCR4 binding and is required for fusion. J. Biol. Chem. 278:3131–36 [Google Scholar]
  63. Cho J, Furie BC, Coughlin SR, Furie B. 63.  2008. A critical role for extracellular protein disulfide isomerase during thrombus formation in mice. J. Clin. Investig. 118:1123–31 [Google Scholar]
  64. Butera D, Cook K, Chiu J. 64.  2014. Control of blood proteins by functional disulfide bonds. Blood 123:2000–7 [Google Scholar]
  65. Kim K, Hahm E, Li J, Holbrook LM, Sasikumar P. 65.  et al. 2013. Platelet protein disulfide isomerase is required for thrombus formation but not for hemostasis in mice. Blood 122:1052–61 [Google Scholar]
  66. Chang TSK, Morton B. 66.  1975. Epididymal sulfhydryl oxidase: a sperm-protective enzyme from the male reproductive tract. Biochem. Biophys. Res. Commun. 66:309–15 [Google Scholar]
  67. Hoober KL, Joneja B, White HB III, Thorpe C. 67.  1996. A sulfhydryl oxidase from chicken egg. J. Biol. Chem. 271:30510–16 [Google Scholar]
  68. Benayoun B, Esnard-Féve, Castella S, Courty Y, Esnard F. 68.  2001. Rat seminal vesicle FAD-dependent sulfhydryl oxidase: biochemical characterization and molecular cloning of a member of the new sulfhydryl oxidase/quiescin Q6 gene family. J. Biol. Chem. 276:13830–37 [Google Scholar]
  69. Codding JA, Israel BA, Thorpe C. 69.  2012. Protein substrate discrimination in the quiescin-sulfhydryl oxidase (QSOX) family. Biochemistry 51:4226–35 [Google Scholar]
  70. Ilani T, Alon A, Grossman I, Horowitz B, Kartvelishvily E. 70.  et al. 2013. A secreted disulfide catalyst controls extracellular matrix composition and function. Science 341:74–76 [Google Scholar]
  71. Israel BA, Jiang L, Gannon SA, Thorpe C. 71.  2014. Disulfide bond generation in mammalian blood serum: detection and purification of quiescin-sulfhydryl oxidase. Free Radic. Biol. Med. 69:129–35 [Google Scholar]
  72. Matthias LJ, Azimi I, Tabrett CA, Hogg PJ. 72.  2010. Reduced monomeric CD4 is the preferred receptor for HIV. J. Biol. Chem. 285:40793–99 [Google Scholar]
  73. Matthias LJ, Yam PTW, Jiang X-M, Vandegraaff N, Li P. 73.  et al. 2002. Disulfide exchange in domain 2 of CD4 is required for entry of HIV-1. Nat. Immunol. 3:727–32 [Google Scholar]
  74. Cerutti N, Killick M, Jugnarain V, Papathanasopoulos M, Capovilla A. 74.  2014. Disulfide reduction in CD4 domain 1 or 2 is essential for interaction with HIV glycoprotein 120 (gp120), which impairs thioredoxin-driven CD4 dimerization. J. Biol. Chem. 289:10455–65 [Google Scholar]
  75. Bourgeois R, Mercier J, Paquette-Brooks I, Cohen EA. 75.  2006. Association between disruption of CD4 receptor dimerization and increased human immunodeficiency virus type 1 entry. Retrovirology 3:31 [Google Scholar]
  76. Wong JWH, Ho SYW, Hogg PJ. 76.  2011. Disulfide bond acquisition through eukaryotic protein evolution. Mol. Biol. Evol. 28:327–34 [Google Scholar]
  77. Katz BA, Kossiakoff A. 77.  1986. The crystallographically determined structures of atypical strained disulfides engineered into subtilisin. J. Biol. Chem. 261:15480–85 [Google Scholar]
  78. Kokkola R, Andersson Å, Mullins G, Östberg T, Treutiger CJ. 78.  et al. 2005. RAGE is the major receptor for the proinflammatory activity of HMGB1 in rodent macrophages. Scand. J. Immunol. 61:1–9 [Google Scholar]
  79. Park JS, Svetkauskaite D, He Q, Kim J-Y, Strassheim D. 79.  et al. 2004. Involvement of Toll-like receptors 2 and 4 in cellular activation by high mobility group box 1 protein. J. Biol. Chem. 279:7370–77 [Google Scholar]
  80. Lotze MT, Tracey KJ. 80.  2005. High-mobility group box 1 protein (HMGB1): nuclear weapon in the immune arsenal. Nat. Rev. Immunol. 5:331–42 [Google Scholar]
  81. Janko C, Filipović M, Munoz LE, Schorn C, Schett G. 81.  et al. 2014. Redox modulation of HMGB1-related signaling. Antioxid. Redox Signal. 20:1075–85 [Google Scholar]
  82. Kazama H, Ricci JE, Herndon JM, Hoppe G, Green DR, Ferguson TA. 82.  2008. Induction of immunological tolerance by apoptotic cells requires caspase-dependent oxidation of high-mobility group box-1 protein. Immunity 29:21–32 [Google Scholar]
  83. Ohndorf UM, Rould MA, He Q, Pabo CO, Lippard SJ. 83.  1999. Basis for recognition of cisplatin-modified DNA by high-mobility-group proteins. Nature 399:708–12 [Google Scholar]
  84. Sahu D, Debnath P, Takayama Y, Iwahara J. 84.  2008. Redox properties of the A-domain of the HMGB1 protein. FEBS Lett. 582:3973–78 [Google Scholar]
  85. Yang H, Lundback P. 85.  2012. Redox modification of cysteine residues regulates the cytokine activity of high mobility group box-1. Mol. Med. 18:250–59 [Google Scholar]
  86. Yang H, Hreggvidsdottir HS, Palmblad K, Wang H, Ochani M. 86.  et al. 2010. A critical cysteine is required for HMGB1 binding to Toll-like receptor 4 and activation of macrophage cytokine release. PNAS 107:11942–47 [Google Scholar]
  87. Venereau E, Casalgrandi M, Schiraldi M, Antoine DJ, Cattaneo A. 87.  et al. 2012. Mutually exclusive redox forms of HMGB1 promote cell recruitment or proinflammatory cytokine release. J. Exp. Med. 209:1519–28 [Google Scholar]
  88. Schiraldi M, Raucci A, Munoz LM, Livoti E, Celona B. 88.  et al. 2012. HMGB1 promotes recruitment of inflammatory cells to damaged tissues by forming a complex with CXCL12 and signaling via CXCR4. J. Exp. Med. 209:551–63 [Google Scholar]
  89. Hoppe G, Talcott KE, Bhattacharya SK, Crabb JW, Sears JE. 89.  2006. Molecular basis for the redox control of nuclear transport of the structural chromatin protein HMGB1. Exp. Cell Res. 312:3526–38 [Google Scholar]
  90. Cobo E, Chadee K. 90.  2013. Antimicrobial human β-defensins in the colon and their role in infectious and non-infectious diseases. Pathogens 2:177–92 [Google Scholar]
  91. Taylor K, Barran PE, Dorin JR. 91.  2007. Structure–activity relationships in β-defensin peptides. Biopolymers 90:1–7 [Google Scholar]
  92. Bensch KW, Raida M, Mägert H-J, Schulz-Knappe P, Forssmann W-G. 92.  1995. hBD-1: a novel β-defensin from human plasma. FEBS Lett. 368:331–35 [Google Scholar]
  93. Schroeder BO, Wu Z, Nuding S, Groscurth S, Marcinowski M. 93.  et al. 2011. Reduction of disulphide bonds unmasks potent antimicrobial activity of human β-defensin 1. Nature 469:419–23 [Google Scholar]
  94. Jaeger SU, Schroeder BO, Meyer-Hoffert U, Courth L, Fehr SN. 94.  et al. 2013. Cell-mediated reduction of human β-defensin 1: a major role for mucosal thioredoxin. Mucosal Immunol. 6:1179–90 [Google Scholar]
  95. Gromer S, Arscott LD, Williams CH, Schirmer RH, Becker K. 95.  1998. Human placenta thioredoxin reductase. J. Biol. Chem. 273:20096–101 [Google Scholar]
  96. Lee KP, Jun JY, Chang I-Y, Suh S-H, So I, Kim KW. 96.  2005. TRPC4 is an essential component of the nonselective cation channel activated by muscarinic stimulation in mouse visceral smooth muscle cells. Mol. Cells 20:435–41 [Google Scholar]
  97. Tsvilovskyy VV, Zholos AV, Aberle T, Philipp SE, Dietrich A. 97.  et al. 2009. Deletion of TRPC4 and TRPC6 in mice impairs smooth muscle contraction and intestinal motility in vivo. Gastroenterology 137:1415–24 [Google Scholar]
  98. Ahmmed GU, Malik AB. 98.  2005. Functional role of TRPC channels in the regulation of endothelial permeability. Pflügers Arch. 451:131–42 [Google Scholar]
  99. Xu S-Z, Sukumar P, Zeng F, Li J, Jairaman A. 99.  et al. 2008. TRPC channel activation by extracellular thioredoxin. Nature 451:69–72 [Google Scholar]
  100. Strübing C, Krapivinsky G, Krapivinsky L, Clapham DE. 100.  2003. Formation of novel TRPC channels by complex subunit interactions in embryonic brain. J. Biol. Chem. 278:39014–19 [Google Scholar]
  101. Venkatachalam K, Zheng F, Gill DL. 101.  2003. Regulation of canonical transient receptor potential (TRPC) channel function by diacylglycerol and protein kinase C. J. Biol. Chem. 278:29031–40 [Google Scholar]
  102. Liu X, Wang W, Singh BB, Lockwich T, Jadlowiec J. 102.  et al. 2000. TRP1, a candidate protein for the store-operated Ca2+ influx mechanism in salivary gland cells. J. Biol. Chem. 275:3403–11 [Google Scholar]
  103. Ordaz B, Tang J, Xiao R, Salgado A, Sampieri A. 103.  et al. 2005. Calmodulin and calcium interplay in the modulation of TRPC5 channel activity: identification of a novel C-terminal domain for calcium/calmodulin-mediated facilitation. J. Biol. Chem. 280:30788–96 [Google Scholar]
  104. Long SB, Campbell EB, Mackinnon R. 104.  2005. Voltage sensor of Kv1.2: structural basis of electro-mechanical coupling. Science 309:903–9 [Google Scholar]
  105. Beech DJ, Sukumar P. 105.  2007. Channel regulation by extracellular redox protein. Channels 1:400–3 [Google Scholar]
  106. Jung S, Mühle A, Schaefer M, Strotmann R, Schultz G, Plant TD. 106.  2003. Lanthanides potentiate TRPC5 currents by an action at extracellular sites close to the pore mouth. J. Biol. Chem. 278:3562–71 [Google Scholar]
  107. Yoshida T, Inoue R, Morii T, Takahashi N, Yamamoto S. 107.  et al. 2006. Nitric oxide activates TRP channels by cysteine S-nitrosylation. Nat. Chem. Biol. 2:596–607 [Google Scholar]
  108. Takahashi N, Mori Y. 108.  2011. TRP channels as sensors and signal integrators of redox status changes. Front. Pharmacol. 2:1–11 [Google Scholar]
  109. Shi T, Iverson GM, Qi JC, Cockerill KA, Linnik MD. 109.  et al. 2004. β2-glycoprotein I binds factor XI and inhibits its activation by thrombin and factor XIIa: loss of inhibition by clipped β2-glycoprotein I. PNAS 101:3939–44 [Google Scholar]
  110. Maiti SN, Balasubramanian K, Ramoth JA, Schroit AJ. 110.  2008. β-2-glycoprotein 1-dependent macrophage uptake of apoptotic cells: binding to lipoprotein receptor-related protein receptor family members. J. Biol. Chem. 283:3761–66 [Google Scholar]
  111. De Groot PG, Meijers JCM. 111.  2011. β2-glycoprotein I: evolution, structure and function. J. Thromb. Haemost. 9:1275–84 [Google Scholar]
  112. McNeil HP, Simpson RJ, Chesterman CN, Krilis SA. 112.  1990. Anti-phospholipid antibodies are directed against a complex antigen that includes a lipid-binding inhibitor of coagulation: β2-glycoprotein I (apolipoprotein H). PNAS 87:4120–24 [Google Scholar]
  113. Ioannou Y, Zhang J, Passam F, Rahgozar S, Qi J. 113.  et al. 2010. Naturally occurring free thiols within β2-glycoprotein I in vivo: functional implications in the regulation of oxidative stress–induced endothelial cell injury. Blood 116:1961–71 [Google Scholar]
  114. Passam FH, Rahgozar S, Qi M, Raftery MJ, Wong JWH. 114.  et al. 2010. Redox control of β2-glycoprotein I–von Willebrand factor interaction by thioredoxin-1. J. Thromb. Haemost. 8:1754–62 [Google Scholar]
  115. Bouma B, de Groot PG, van den Elsen JM, Ravelli RB, Schouten A. 115.  et al. 1999. Adhesion mechanism of human β2-glycoprotein I to phospholipids based on its crystal structure. EMBO J. 18:5166–74 [Google Scholar]
  116. Schwarzenbacher R, Zeth K, Diederichs K, Gries A, Kostner GM. 116.  et al. 1999. Crystal structure of human β2-glycoprotein I: implications for phospholipid binding and the antiphospholipid syndrome. EMBO J. 18:6228–39 [Google Scholar]
  117. Moake JL, Turner NA, Stathopoulos NA, Nolasco LH, Hellums JD. 117.  1986. Involvement of large plasma von Willebrand factor (vWF) multimers and unusually large vWF forms derived from endothelial cells in shear stress-induced platelet aggregation. J. Clin. Investig. 78:1456–61 [Google Scholar]
  118. Federici AB, Bader R, Pagani S, Colibretti ML, De Marco L, Mannucci PM. 118.  1989. Binding of von Willebrand factor to glycoproteins Ib and IIb/IIIa complex: affinity is related to multimeric size. Br. J. Haematol. 73:93–99 [Google Scholar]
  119. Furlan M. 119.  1996. Von Willebrand factor: molecular size and functional activity. Ann. Hematol. 72:341–48 [Google Scholar]
  120. Xie L, Chesterman CN, Hogg PJ. 120.  2000. Reduction of von Willebrand factor by endothelial cells. Thromb. Haemost. 84:506–13 [Google Scholar]
  121. Xie L, Chesterman CN, Hogg PJ. 121.  2001. Control of von Willebrand factor multimer size by thrombospondin-1. J. Exp. Med. 193:1341–49 [Google Scholar]
  122. Molberg Ø, Mcadam SN, Körner R, Quarsten H, Sollid LM. 122.  et al. 1998. Tissue transglutaminase selectively modifies gliadin peptides that are recognized by gut-derived T cells in celiac disease. Nat. Med. 4:713–17 [Google Scholar]
  123. van de Wal Y, Kooy Y, van Veelen P, Peña S, Mearin L. 123.  et al. 1998. Cutting edge: selective deamidation by tissue transglutaminase strongly enhances gliadin-specific T cell reactivity. J. Immunol. 161:1585–88 [Google Scholar]
  124. Kumar S, Mehta K. 124.  2013. Tissue transglutaminase, inflammation, and cancer: How intimate is the relationship?. Amino Acids 44:81–88 [Google Scholar]
  125. Citron BA, Zoloty JE, Suo Z, Festoff BW. 125.  2005. Tissue transglutaminase during mouse central nervous system development: lack of alternative RNA processing and implications for its role(s) in murine models of neurotrauma and neurodegeneration. Brain Res. Mol. Brain Res. 135:122–33 [Google Scholar]
  126. Grosso H, Mouradian MM. 126.  2012. Transglutaminase 2: biology, relevance to neurodegenerative diseases and therapeutic implications. Pharmacol. Ther. 133:392–410 [Google Scholar]
  127. Eckert RL, Kaartinen MT, Nurminskaya M, Belkin AM, Colak G. 127.  et al. 2014. Transglutaminase regulation of cell function. Physiol. Rev. 94:383–417 [Google Scholar]
  128. Pinkas DM, Strop P, Brunger AT, Khosla C. 128.  2007. Transglutaminase 2 undergoes a large conformational change upon activation. PLOS Biol. 5:e327 [Google Scholar]
  129. Király R, Csosz E, Kurtán T, Antus S, Szigeti K. 129.  et al. 2009. Functional significance of five noncanonical Ca2+-binding sites of human transglutaminase 2 characterized by site-directed mutagenesis. FEBS J. 276:7083–96 [Google Scholar]
  130. Siegel M, Strnad P, Watts RE, Choi K, Jabri B. 130.  et al. 2008. Extracellular transglutaminase 2 is catalytically inactive, but is transiently activated upon tissue injury. PLOS ONE 3:e1861 [Google Scholar]
  131. Stamnaes J, Pinkas DM, Fleckenstein B, Khosla C, Sollid LM. 131.  2010. Redox regulation of transglutaminase 2 activity. J. Biol. Chem. 285:25402–9 [Google Scholar]
  132. Iversen R, Mysling S, Hnida K, Jørgensen TJD, Sollid LM. 132.  2014. Activity-regulating structural changes and autoantibody epitopes in transglutaminase 2 assessed by hydrogen/deuterium exchange. PNAS 111:17146–51 [Google Scholar]
  133. Jin X, Stamnaes J, Klöck C, DiRaimondo TR, Sollid LM, Khosla C. 133.  2011. Activation of extracellular transglutaminase 2 by thioredoxin. J. Biol. Chem. 286:37866–73 [Google Scholar]
  134. DiRaimondo TR, Plugis NM, Jin X, Khosla C. 134.  2013. Selective inhibition of extracellular thioredoxin by asymmetric disulfides. J. Med. Chem. 56:1301–10 [Google Scholar]
  135. Kim S-H, Oh J, Choi J-Y, Jang J-Y, Kang M-W, Lee C-E. 135.  2008. Identification of human thioredoxin as a novel IFN-γ-induced factor: mechanism of induction and its role in cytokine production. BMC Immunol. 9:64 [Google Scholar]
  136. Laragione T, Bonetto V, Casoni F, Massignan T, Bianchi G. 136.  et al. 2003. Redox regulation of surface protein thiols: identification of integrin α-4 as a molecular target by using redox proteomics. PNAS 100:14737–41 [Google Scholar]
  137. Berman HM, Henrick K, Nakamura H. 137.  2003. Announcing the worldwide Protein Data Bank. Nat. Struct. Biol. 10:980 [Google Scholar]
  138. Schmidt B, Hogg PJ. 138.  2007. Search for allosteric disulfide bonds in NMR structures. BMC Struct. Biol. 7:49 [Google Scholar]
  139. Wang G, Zhang Z-T, Jiang B, Zhang X, Li C, Liu M. 139.  2014. Recent advances in protein NMR spectroscopy and their implications in protein therapeutics research. Anal. Bioanal. Chem. 406:2279–88 [Google Scholar]
  140. Azimi I, Wong JWH, Hogg PJ. 140.  2011. Control of mature protein function by allosteric disulfide bonds. Antioxid. Redox Signal. 14:113–26 [Google Scholar]
  141. Burlina F, Sagan S, Bolbach G, Chassaing G. 141.  2006. A direct approach to quantification of the cellular uptake of cell-penetrating peptides using MALDI-TOF mass spectrometry. Nat. Protoc. 1:200–5 [Google Scholar]
  142. Yin L, Ding J, He C, Cui L, Tang C, Yin C. 142.  2009. Drug permeability and mucoadhesion properties of thiolated trimethyl chitosan nanoparticles in oral insulin delivery. Biomaterials 30:5691–5700 [Google Scholar]
  143. Jha D, Mishra R, Gottschalk S, Wiesmüller KH, Ugurbil K. 143.  et al. 2011. CyLoP-1: a novel cysteine-rich cell-penetrating peptide for cytosolic delivery of cargoes. Bioconjug. Chem. 22:319–28 [Google Scholar]
  144. Torres AG, Fabani MM, Vigorito E, Williams D, Al-Obaidi N. 144.  et al. 2012. Chemical structure requirements and cellular targeting of microRNA-122 by peptide nucleic acids anti-miRs. Nucleic Acids Res. 40:2152–67 [Google Scholar]
  145. Jiao CY, Delaroche D, Burlina F, Alves ID, Chassaing G, Sagan S. 145.  2009. Translocation and endocytosis for cell-penetrating peptide internalization. J. Biol. Chem. 284:33957–65 [Google Scholar]
  146. Hazes B, Dijkstra BW. 146.  1988. Model building of disulfide bonds in proteins with known three-dimen-sional structure. Protein Eng. 2:119–25 [Google Scholar]
  147. Craig DB, Dombkowski A. 147.  2013. Disulfide by Design 2.0: a web-based tool for disulfide engineering in proteins. BMC Bioinform. 14:346 [Google Scholar]
  148. Matsumura M, Matthews BW. 148.  1989. Control of enzyme activity by an engineered disulfide bond. Science 243:792–94 [Google Scholar]
  149. Shen Y, Joachimiak A, Rosner MR, Tang W-J. 149.  2006. Structures of human insulin-degrading enzyme reveal a new substrate recognition mechanism. Nature 443:870–74 [Google Scholar]
  150. Loo TW, Clarke DM. 150.  2014. Cysteines introduced into extracellular loops 1 and 4 of human P-glycoprotein that are close only in the open conformation spontaneously form a disulfide bond that inhibits drug efflux and ATPase activity. J. Biol. Chem. 289:24749–58 [Google Scholar]
  151. Rost J, Rapoport S. 151.  1964. Reduction-potential of glutathione. Nature 4915:185 [Google Scholar]
  152. Clark WM. 152.  1960. Oxidation-Reduction Potentials of Organic Systems Baltimore, MD: Williams & Wilkins
  153. Tohgi H, Abe T, Saheki M, Hamato F, Sasaki K, Takahashi S. 153.  1995. Reduced and oxidized forms of glutathione and α-tocopherol in the cerebrospinal fluid of parkinsonian patients: comparison between before and after l-dopa treatment. Neurosci. Lett. 184:21–24 [Google Scholar]
/content/journals/10.1146/annurev-chembioeng-080615-033553
Loading
/content/journals/10.1146/annurev-chembioeng-080615-033553
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