Skip to main content
Chemistry LibreTexts

3: Calcium in Biological Systems

  • Page ID
    59622
  • \( \newcommand{\vecs}[1]{\overset { \scriptstyle \rightharpoonup} {\mathbf{#1}} } \) \( \newcommand{\vecd}[1]{\overset{-\!-\!\rightharpoonup}{\vphantom{a}\smash {#1}}} \)\(\newcommand{\id}{\mathrm{id}}\) \( \newcommand{\Span}{\mathrm{span}}\) \( \newcommand{\kernel}{\mathrm{null}\,}\) \( \newcommand{\range}{\mathrm{range}\,}\) \( \newcommand{\RealPart}{\mathrm{Re}}\) \( \newcommand{\ImaginaryPart}{\mathrm{Im}}\) \( \newcommand{\Argument}{\mathrm{Arg}}\) \( \newcommand{\norm}[1]{\| #1 \|}\) \( \newcommand{\inner}[2]{\langle #1, #2 \rangle}\) \( \newcommand{\Span}{\mathrm{span}}\) \(\newcommand{\id}{\mathrm{id}}\) \( \newcommand{\Span}{\mathrm{span}}\) \( \newcommand{\kernel}{\mathrm{null}\,}\) \( \newcommand{\range}{\mathrm{range}\,}\) \( \newcommand{\RealPart}{\mathrm{Re}}\) \( \newcommand{\ImaginaryPart}{\mathrm{Im}}\) \( \newcommand{\Argument}{\mathrm{Arg}}\) \( \newcommand{\norm}[1]{\| #1 \|}\) \( \newcommand{\inner}[2]{\langle #1, #2 \rangle}\) \( \newcommand{\Span}{\mathrm{span}}\)\(\newcommand{\AA}{\unicode[.8,0]{x212B}}\)

    Calcium, like many other "inorganic elements" in biological systems, has during the last decade become the subject of much attention both by scientists and by the general public.1 The presence and central role of calcium in mammalian bones and other mineralized tissues were recognized soon after its discovery as an element by Davy in 1808. Much later, the insight arrived that Ca2+ ions could play an important role in other tissues as well. Experiments of great historical influence were performed by the British physiologist Sidney Ringer a little over a century ago.2 He was interested in the effects of various cations on frog-heart muscle and somewhat serendipitously discovered that Ca2+ ions, everpresent in the tap water distributed in central London, in millimolar concentrations were necessary for muscle contraction and tissue survival.

    Today it is widely recognized that Ca2+ ions are central to a complex intracellular messenger system that is mediating a wide range of biological processes: muscle contraction, secretion, glycolysis and gluconeogenesis, ion transport, cell division and growth (for definitions of terms in boldface, see Appendix A in Section IX). The detailed organization of this messenger system is presently the subject of considerable scientific activity, and some details are already known. One of the links in the system is a class of highly homologous Ca2+-binding proteins, to be discussed later on in this chapter, that undergo Ca2+-dependent conformational changes and respond to transitory increases in intracellular Ca2+-ion concentrations. A prerequisite for the proper function of the calcium messenger system in higher organisms is that the cytosolic Ca2+ concentration in a "resting" cell be kept very low, on the order of 100 to 200 nM. Transitory increases in the Ca2+ concentration that may result from hormonal action on a membrane receptor must rapidly be reduced. Several transport proteins, driven either by ATP hydrolysis or by gradients of some other ion like Na+, are involved in this activity.

    Ca2+ ions are also known to play various roles outside cells. In the plant kingdom Ca2+ ions often form links between individual cells and are required for maintaining the rigidity of whole plants; some seaweeds are typical examples. In the blood plasma of mammals, in which the Ca2+ concentration exceeds the intracellular by a factor of about 104, Ca2+ ions are instrumental in joining certain proteins in the blood-clotting system with membrane surfaces of circulating cells. Many extracellular enzymes also contain Ca2+ ions, sometimes at the active site but most often at other locations. It is generally believed that Ca2+ ions confer on proteins an increased thermal stability, and indeed proteins in heat-tolerant microorganisms often hold many such ions.

    Vertebrates require much calcium in their food; in the USA the recommended daily allowance (RDA) for adult humans is 800 mg, and most other countries have comparable recommendations. During gestation in mammals, calcium must be transported across the placenta into the fetus, in particular during those phases of pregnancy when bone formation is most rapid. Interestingly, there appear to be some parallels between intestinal and placental transport that will be discussed further below. The role of calcium in biominerals is a vast subject that we can treat only superficially in this chapter.

    To provide a background to the more biologically oriented sections that follow, we begin with a brief recapitulation of some basic facts about calcium. Then we continue with an outline of calcium distribution in biological tissues and organelles, and of the methods that can be used to obtain this information. After this follows a brief section on Ca2+ transport, and an account of the mechanism of intracellular Ca2+ release as it is presently understood. A discussion of some selected Ca2+-binding proteins of general interest, both intracellular and extracellular, then follows. Before we conclude the chapter, we will summarize some recent observations on Ca2+-binding proteins in prokaryotes.

    II. Basic Facts About Calcium: Its Compounds and Reactions

    III. Calcium In Living Cells: Methods For Determining Concentrations And Spatial Distributions

    1. Measurements of "Free" Calcium Concentrations

      1. Ca2+-selective microelectrodes
      2. Bioluminescence
      3. Complexing Agents with Ca2+-dependent Light Absorption or Fluorescence
      4. Complexing Agents with Ca2+-dependent NMR Spectra
    2. Summary

    Much of our present knowledge about the biological role of Ca2+ rests on detailed measurements of the concentration, distribution, and chemical nature of Ca2+ and its complexes. Concentrations of uncomplexed, or "free," Ca2+ can be measured by Ca2+-selective microelectrodes, bioluminescence and complexing agents with Ca2+-dependent light absorption, fluorescence, or NMR spectra. An outcome of such studies is that the "free" Ca2+ concentration in resting eukaryotic cells generally is very low, on the order of 100 to 200 nM. Total Ca2+ concentrations, uncomplexed and complexed, can be measured by a variety of physical techniques. Some techniques, like atomic absorption, are sensitive but give poor spatial resolution. Others involve the bombardment of the sample with electrons or charged atoms, and can yield spatial resolutions of the order of a few nm; however, there is a trade-off between detectability and resolution

    IV. The Transport and Regulation of Ca2+ Ions in Higher Organisms

    1. Ca2+ Uptake and Secretion

    2. Intracellular Ca2+ Transport

      1. The Ca2+-ATPases
      2. The Na+/Ca2+ Exchanger of the Plasma Membrane
      3. Mitochondrial Ca2+ Transport: Influx
      4. Mitochondrial Ca2+ Transport: Efflux
      5. Ca2+ Efflux from Non-mitochondrial Stores
      6. Other Voltage-gated or Receptor-activated Ca2+ Channels
    3. Inositol Trisphosphate and the Ca2+ Messenger System

    4. Summary

    The fluxes of Ca2+ ions and their regulation in higher organisms, as well as in microorganisms, depend on several transport proteins in addition to vesicular and gated processes. An important class of transport proteins are the Ca2+-ATPases, which are particularly abundant in muscle cells. These proteins translocate Ca2+ ions against large activity (or concentration) gradients through the expenditure of ATP. Transport of Ca2+ ions against activity gradients across membranes may also be accomplished by coupled transport of other ions, like Na+, with a gradient in the opposite direction.

    As a result of some external stimulus—the action of a hormone, for example—the "free" Ca2+-ion concentrations in the cytoplasm of many cell types may transiently increase several orders of magnitude. This increase largely results from the release of Ca2+ from intracellular stores (ER, SR) in response to the initial formation of a new type of messenger, 1,4,5-IP3. The activity of Ca2+-transport proteins eventually restores the Ca2+ concentration levels to resting levels. This sequence of events forms the basis for Ca2+'s role in the regulation of a wide variety of cellular activities (see Section V).

    V. Molecular Aspects of Ca2+-regulated Intracellular Processes

    1. Parvalbumin and Calbindins D9K and D28K

    2. Sarcoplasmic Calcium-Binding Protein from Nereis diversicolor

    3. Membrane Cytoskeleton and Phospholipid Binding Proteins

    4. Ca2+-dependent Proteases

    5. Protein Kinase C

    6. Summary

    Many different biological processes in eukaryotic cells are regulated by intracellular Ca2+ concentration levels. Examples of such processes are muscle contraction, transport processes, cell division and growth, enzyme activities, and metabolic processes. A link in this regulatory chain is a number of intracellular Ca2+ receptors with Ca2+ affinities such that their binding sites are largely unoccupied at resting Ca2+ concentration levels, but are occupied at Ca2+ levels reached as a result of some external stimulus. This class of Ca2+ receptors is often called the "calmodulin superfamily" and includes the well-known members troponin C (regulating muscle contraction in striated muscle) and calmodulin (playing an important role in the regulation of many cellular processes). Amino-acid sequence determinations as well as x-ray and 2D 1H NMR studies have revealed a strong homology between the regulatory Ca2+-binding proteins. The Ca2+-binding sites are located in a loop flanked by two helices, and the Ca2+ ions are ligated with approximately octahedral or pentagonal bipyramidal symmetry. The ligands are six or seven oxygen atoms that are furnished by side-chain carboxylate or hydroxyl groups, backbone carbonyls, and water molecules. Pairs of these Ca2+ sites, rather than individual sites, appear to be the functional unit, and a common consequence of their arrangement is cooperative Ca2+ binding. Ca2+ binding to the intracellular receptor proteins is accompanied by structural changes that expose hydrophobic patches on their surfaces, thereby enabling them to bind to their target proteins.

    VI. Extracellular Ca2+-binding Proteins

    1. Ca2+-binding in Some Extracellular Enzymes

    2. Summary

    In higher organisms, the Ca2+ concentration in extracellular fluids generally is considerably higher than the intracellular concentrations. In mammalian body fluids, the Ca2+ concentration is typically on the order of a few mM. The extracellular concentration levels are highly regulated and undergo only minor variations. A consequence of these high levels of Ca2+ in extracellular fluids is that the binding constant need be only 103 to 104 M-1 in order for a protein site to be highly occupied by Ca2+. Several extracellular enzymes and enzyme activators have one or more Ca2+ ions as integral parts of their structures. Some Ca2+ ions are bound at, or near, the active cleft and may take part in the enzymatic reactions (e.g., phospholipase A2, \(\alpha\)-amylase). In other molecules, for example, serine proteases like trypsin and chymotrypsin, the Ca2+ ion is not essential for enzymatic activity, and may play more of a structural role. Ca2+ ions are involved in the cascade of enzymatic events that results in blood clotting in mammals. Several of the proteins in this system contain two new amino acids, \(\gamma\)-carboxyglutamic acid (Gla) and \(\beta\)-hydroxyaspartic acid (Hya), which are strongly suspected to be involved as ligands in Ca2+ binding. In the presence of Ca2+ ions, prothrombin and other Gla-containing proteins will bind to cell membranes containing acidic phospholipids, in particular, the platelet membrane. It appears likely that Ca2+ ions form a link between the protein and the membrane surface.

    VII. Calcium in Mineralized Tissues

    Summary

    Calcium is, along with iron, silicon, and the alkaline earth metals, an important constituent of mineralized biological tissues. Some Ca2+-based biominerals, like bone or mother-of-pearl, can be regarded as complex composites with microscopic crystallites embedded in a protein matrix. The formation of calcified biominerals is a highly regulated process, and human bone, for instance, is constantly being dissolved and rebuilt. When the rates of these two counteracting processes are not in balance, the result may be decalcification, or osteoporosis, which seriously reduces the strength of the bone.

    VIII. Ca2+-binding Proteins in Microorganisms: The Search for a Prokaryotic Calmodulin

    Summary

    The role of Ca2+ ions in the regulation of biological activities of procaryotic organisms is still largely unsettled. Over the last decade, however, evidence has gradually accumulated that calcium ions are involved in diverse bacterial activities, such as chemotaxis and substrate transport, sporulation, initiation of DNA replication, phospholipid synthesis, and protein phosphorylation.168 An important landmark is the recent demonstration that the intracellular Ca2+ concentration in E. coli is tightly regulated to about 100 nM, a level similar to that typical of resting eukaryotic cells.169 Furthermore, increasing numbers of calcium-binding proteins, some of which also have putative EF-hand Ca2+ sites characteristic of the calmodulin superfamily of intracellular regulatory proteins, have been isolated in bacteria.168

    IX. Appendixes

    1. Definition of Biochemical Terms

    Antiport A transport protein that carries two ions or molecules in opposite directions across a membrane.
    Basal lateral membrane The membrane in intestinal epithelial cells that is located on the base of the cells, opposite the microvilli that face the intestinal lumen.
    Cytosol The unstructured portion of the interior of a cell—the cell nucleus excluded—in which the organelles are bathed.
    Electrogenic A biological process driven by electric field gradients.
    Endocytosis The process by which eukaryotic cells take up solutes and/or particles by enclosure in a portion of the plasma membrane to (temporarily) form cytoplasmic vesicles.
    Endoplasmic reticulum (ER) Sheets of folded membranes, within the cytoplasm of eukaryotic cells, that are the sites for protein synthesis and transport.
    Epithelial cells Cells that form the surface layer of most, if not all, body cavities (blood vessels, intestine, urinary bladder, mouth, etc.).
    Erythrocytes Red-blood corpuscles.
    Eukaryotic cells Cells with a well-definied nucleus.
    Exocytosis The process by which eukaryotic cells release packets of molecules (e.g., neurotransmitters) to the environment by fusing vesicles formed in the cytoplasm with the plasma membrane.
    Gluconeogenesis Metabolic synthesis of glucose.
    Glycolysis Metabolic degradation of glucose.
    Hydropathy A measure of the relative hydrophobic or hydrophilic character of an amino acid or amino-acid side chain.
    Lamina propria mucosae The layer of connective tissue underlying the epithelium of a mucous membrane.
    Mitrochondrion A double-membrane organelle in eukaryotic cells that is the center for aerobic oxidation processes leading to the formation of energy-rich ATP.
    Organelle A structurally distinct region of the cell that contains specific enzymes or other proteins that perform particular biological functions.
    Osteoporosis Brittle-bond disease.
    Phorbol esters Polycyclic organic molecules that act as analogues to diacylglycerol and therefore are strong activators of protein kinase C.
    Prokaryotic cells Cells lacking a well-defined nucleus.
    Sarcoplasmic reticulum The ER of muscle cells.
    Trophoblasts The cells between the maternal and fetal circulation systems.
    Tryptic digest Fragmentation of proteins as a result of treatment with the proteolytic enzyme trypsin.
    Uniporter A transport protein that carries a particular ion or molecule in one direction across a membrane.
    1. One-Letter Code for Amino-Acid Residues

    A—alanine, C—cysteine, D—aspartate, E—glutamate, F—phenylalanine, G—glycine, H—histidine, I—isoleucine, K—Iysine, L—Ieucine, M—methionine, N—asparagine, P—proline, Q—glutamine, R—arginine, S—serine, T—threonine, V—valine, W—tryptophan, Y—tyrosine.

    1. The Activity of a Transport Protein

    This is usually described in terms of the classical Michaelis-Menten scheme:

    \[V (= transport\; rate) = V_{max} \cdotp \frac{[S]}{[S] + K_{m}},\]

    where [S] is the concentration of the solute to be transported and Km = (k-1+k2)/k1 is the Michaelis constant (dimension "concentration") for the reaction

    \[E+S \xrightleftharpoons[k_{-1}]{k_{1}} ES \xrightarrow{k_{2}} R \ldotp\]

    Approximated as the reciprocal ratio between on- and off-rate constants relevant to the solute-protein complex, 1/Km = k1/k-1 may be taken as a lower limit of the affinity of the protein for the solute.

    X. References

    1. See, for example, "Going Crazy over Calcium" in the Health and Fitness section of Time, February 23, 1987, p. 49, or C. Garland et al., The Calcium Diet, Penguin Books, 1990.
    2. S. Ringer, J. Physiol. 3 (1883), 195.
    3. Handbook of Chemistry and Physics, 64th ed., CRC Press, 1984.
    4. R. D. Shannon, Acta Cryst. A, 32 (1976), 751.
    5. R. J. P. Williams, in Calcium in Biological Systems, Cambridge Univ. Press, 1976, p. 1.
    6. B. A. Levine and R. J. P. Williams, in L. J. Anghileri and A. M. T. Anghileri, eds., Role of Calcium in Biological Systems, CRC Press, 1982, pp. 3-26.
    7. A. K. Campbell, Intracellular Calcium: Its Universal Role as Regulator, Wiley, 1983.
    8. A. P. Somlyo, M. Bond, and A. V. Somlyo, Nature 314 (1985), 622.
    9. T. B. Johansson, R. Akselsson, and S. A. E. Johansson, Nucl. Inst. Methods 84 (1970), 141.
    10. R. B. Martin, in H. Sigel, ed., Metal Ions in Biological Systems, Dekker, 17 (1984), I.
    11. H. Einspahr and C. E. Bugg, in ibid., pp. 52-97.
    12. L. G. Sillen and A. E. Martell, eds., Stability Constants of Metal Ion Complexes, Chemical Soc., London, 1964.
    13. A. E. Martell and R. M. Smith, eds., Critical Stability Constants, Plenum Press, 1, 1975.
    14. J. D. Potter and J. Gergely, J. Biol. Chem. 250 (1975), 4628.
    15. W. Marki, M. Oppliger, and R. Schwyzer, Helvetica Chemica Acta 60 (1977), 807.
    16. C. M. Frey and J. Stuehr, in H. Sigel, ed., Metal Ions in Biological Systems, Dekker, 1 (1974), 51.
    17. W. J. Moore, Physical Chemistry, Longman, 5th ed., 1972, Chapter 10.
    18. M. V. Thomas, Techniques in Calcium Research, Academic Press, London, 1982.
    19. T. J. Rink, Pure Appl. Chem. 55 (1988), 1977.
    20. W. Simon et al., Ann. N.Y. Acad. Sci. 307 (1987), 5269.
    21. R. C. Thomas, Ion-Sensitive Intracellular Microelectrodes, Academic Press, London, 1987.
    22. J. R. Blinks et al., Prog. Biophys. Mol. Biol. 4 (1983), I.
    23. R. Y. Tsien, Biochemistry 19 (1980), 2396.
    24. G. Grynkiewicz, M. Poenie, and R. Y. Tsien, J. Biol. Chem. 260 (1985), 3440.
    25. A. Minta, J. P. Y. Kao, and R. Y. Tsien, J. Biol. Chem 264 (1989), 8171.
    26. R. Y. Tsien and M. Poenie, Trends Biochem. Sci. 11 (1986), 450.
    27. E. Chiancone et al., J. Biol. Chem. 26 (1986), 16306.
    28. G. A. Smith et al., Proc. Natl. Acad. Sci. USA 80 (1983), 7178.
    29. J. C. Metcalfe, T. R. Hesketh, and G. A. Smith, Cell Calcium 6 (1985), 183.
    30. A. P. Somlyo, Cell Calcium 6 (1985), 197.
    31. G. W. Grime et al., Trends Biochem. Sci. 10 (1985), 6.
    32. G. H. Morrison and G. Slodzian, Anal. Chem. 47 (1975), 932A.
    33. S. Chandra and G. H. Morrison, Science 228 (1985), 1543.
    34. E. Murphy et al., Circulation Research 68 (1991), 1250.
    35. R. H. Wasserman and C.S. Fullmer, in W. Y. Cheung, ed., Calcium and Cell Function, Academic Press, 2 (1982), 176.
    36. H. Rasmussen, O. Fontaine, and T. Matsumoto, Ann. N. Y. Acad. Sci. 77 (1981), 518.
    37. S. Linse et al., Biochemistry 26 (1987), 6723.
    38. D. T. W. Bryant and P. Andrews, Biochem. J. 219 (1984), 287.
    39. R. H. Kretsinger, J. E. Mann, and J. G. Simmonds, in A. W. Norman et al., eds., Vitamin D, Chemical, Biochemical, and Clinical Endocrinology of Calcium Metabolism, W. de Gruyter, pp. 232-248.
    40. J. J. Feher, Am. J. Physiol. 244 (1983), 303.
    41. I. Nemere, V. Leathers, and A. W. Norman, J. Biol. Chem. 261 (1986), 16106.
    42. G. E. Lester. Feder. Proc. 45 (1986), 2524.
    43. P. Marche, C. LeGuern, and P. Cassier, Cell Tissue Res. 197 (1979), 69.
    44. M. Warembourg, C. Perret, and M. Thomasset, Endocrinol. 119 (1986), 176.
    45. H. J. Schatzman, Experientia 22 (1966), 364.
    46. V. Niggli, J. T. Penniston, and E. Carafoli, J. Biol. Chem. 254 (1979), 9955.
    47. E. Carafoli, M. Zurini, and G. Benaim, in Calcium and the Cell, CIBA Foundation Symposium no. 122, Wiley, 1986, pp. 58~65.
    48. D. MacLennan, J. Biol. Chem. 245 (1970), 4508.
    49. N. M. Green et al., in Reference 47, pp. 93.
    50. D. M. Clarke et aI., Nature 339 (1989), 476.
    51. G. Inesi, Annu. Rev. Physiol. 47 (1985), 573.
    52. R. J. P. Williams, Eur. J. Biochem. 150 (1985), 231.
    53. C. Tanford, Proc. Natl. Acad. Sci. USA 79 (1982), 6527.
    54. D. H. MacLennan, K. P. Campbell, and R. A. Reithmeier, in Reference 35, pp. 151-173.
    55. A. Maurer et al., Proc. Natl. Acad. Sci. USA 82 (1985), 4036.
    56. M. P. Blaustein and M. T. Nelson, in E. Carafoli, ed., Membrane Transport of Calcium, Academic Press, 1982, pp. 217-236.
    57. P. F. Baker, in Reference 47, pp. 73-84.
    58. E. Carafoli, G. Inesi, and B. P. Rosen, in Reference 10, pp. 140-143.
    59. G. Fiskum, in Reference 10, pp. 187-214.
    60. A. L. Lehninger et al., in F. Bronner and M. Peterlik, eds., Calcium and Phosphate Transport Across Membranes, Academic Press, 1981, pp. 73-78.
    61. E. Carafoli, in Reference 56, pp. 109-139.
    62. M. J. Berridge, in Reference 47, pp. 39-49.
    63. R. F. Irvine, Brit. Med. Bull. 42 (1986), 369.
    64. E. W. McCleskey et al., J. Exp. BioI. 124 (1986), 177.
    65. D. J. Adams, T. Dwyer, and B. Hille, J. Gen. Physiol. 75 (1980), 493.
    66. A. B. MacDermott et al., Nature 321 (1986), 519.
    67. R. J. Miller, Science 235 (1987), 46.
    68. R. W. Tsien, Annu. Rev. Physiol. 45 (1983), 341.
    69. E. W. Sutherland, Science 177 (1972), 401.
    70. M. J. Berridge and R. F. Irvine, Nature 312 (1984), 315.
    71. M. J. Berridge, Biochem. J. 212 (1983), 849.
    72. M. J. Berridge, J. Exp. Biol. 124 (1986), 323.
    73. C. C. Chadwick, A. Saito, and S. Fleischer, Proc. Natl. Acad. Sci. USA 87 (1990), 2132.
    74. A. R. Hughes and J. W. Putney, Envir. Health Perspec. 84 (1990), 141.
    75. A. S. Manalan and C. B. Klee, in Adv. Cycl. Nucl. Prot. Phospho Res. (1984), 227.
    76. J. D. Potter and J. D. Johnson, in Reference 35, pp. 145.
    77. (a) R. H. Kretsinger, Cold Spring Harbor Symp. Quant. Biol. 52 (1987), 499; (b) C. W. Heizmann and K. Braun, Trends in Neurosciences 15 (1992), 259.
    78. A. Marks et al., J. Neurochem. 41 (1983), 107.
    79. P. B. Moore and J. R. Dedman, in H. Hidaka and P. J. Hartshorne, eds., Calmodulin Antagonists and Cellular Physiology, Academic Press, 1985, pp. 483-494.
    80. Y. Nishizuka, Phil. Trans. Roy. Soc. Lond. (B) 302 (1983), 101.
    81. M. J. Geisow, FEBS Lett. 203 (1986) 99.
    82. M. J. Crumpton and J. R. Dedman, Nature 345 (1990), 212.
    83. Y. S. Babu, C. E. Bugg, and W. J. Cook, J. Mol. Biol. 204 (1988), 191.
    84. R. H. Kretsinger and C. E. Nockolds, J. Biol. Chem. 248 (1973), 3313.
    85. S. Forsén, H. J. Vogel, and T. Drakenberg, in Reference 35, 6 (1986) 113.
    86. M. D. Tsai et aI., Biochemistry 26 (1987), 3635.
    87. S. R. Martin et al., Eur. J. Biochem. 151 (1985), 543.
    88. A. Teleman, T. Drakenberg, and S. Forsén, Biochim. Biophys. Acta 873, (1986), 204.
    89. L. Sjölin, Acta Cryst. B46 (1990), 209.
    90. S. R. Martin and P. M. Bayley, Biochem. J. 238 (1986), 485.
    91. D. B. Heidorn and J. Trewhella, Biochemistry 27 (1988), 909.
    92. D. K. Blumenthal et al., Proc. Natl. Acad. Sci. USA 82 (1985), 3187.
    93. R. E. KIevit et al., Biochemistry 24 (1985), 8152.
    94. S. Linse, T. Drakenberg, and S. Forsén, FEBS Lett. 199 (1986), 28.
    95. S. H. Seeholzer et al., Proc. Natl. Acad. Sci. USA 83 (1986), 3634.
    96. A. Persechini and R. H. Kretsinger, J. Cardiovasc. Pharm. 12 (suppl. 5, 1988), S1.
    97. G. Barbato et al., Biochemistry, in press.
    98. M. Ikura et al., Biochemistry 30 (1991), 5498.
    99. C. H. Keller et al., Biochemistry 21 (1982), 156.
    100. O. Herzberg and M. N. G. James, Nature 313 (1985), 653.
    101. K. A. Satyshur et al., J. Biol. Chem. 263 (1988), 1628.
    102. E. D. McCubbin and C. M. Kay, Ace. Chem. Res. 13 (1980), 185.
    103. M. T. Hincke, W. D. McCubbin, and C. M. Kay, Can. J. Biochem. 56 (1978), 384.
    104. P. C. Leavis et al., J. BioI. Chem. 253 (1978), 5452.
    105. O. Teleman et al., Eur. J. Biochem. 134 (1983), 453.
    106. H. J. Vogel and S. Forsén, in L. J. Berliner and J. Reuben, eds., Biological Magnetic Resonance, Plenum Press, 7 (1987), 247.
    107. S. Forsén et al., in B. de Bernard et al., eds., Calcium Binding Proteins 1983, Elsevier, 1984, pp. 121-131.
    108. T. Drakenberg et al., J. Biol. Chem. 262 (1987), 672.
    109. K. B. Seamon and R. H. Kretsinger, in T. G. Spiro, ed., Calcium in Biology, Wiley, 6 (1983), 1; review of the literature.
    110. B. A. Levine and D. C. Dalgamo, Biochim. Biophys. Acta 726 (1983), 187.
    111. O. Herzberg, J. Moult, and M. N. G. James, J. Biol. Chem. 261 (1986), 2638.
    112. W. Chazin et al., personal communication.
    113. M. L. Greaser and J. Gergely, J. Biol. Chem. 246 (1971), 4226.
    114. W. Wnuk, J. A. Cox, and E. A. Stein, in Reference 35, pp. 243-278.
    115. C. W. Heizmann, in Reference 109, pp. 61-63.
    116. R. H. Kretsinger, CRC Crit. Rev. Biochem. 8 (1980), 119; R. H. Kretsinger and C. D. Barry, Biochim. Biophys. Acta 405 (1975), 4051.
    117. W. Hunziker, Proc. Natl. Acad. Sci. USA 83 (1986), 7578.
    118. A. L. Swain, R. H. Kretsinger, and E. L. Arnma, J. Biol. Chem. 264 (1988), 16620.
    119. M. F. Summers, Coord. Chem. Rev. 86 (1988), 43; review of the literature.
    120. J. M. Gillis et al., J. Muscl. Res. Cell. Mot. 3 (1982), 377.
    121. D. M. E. Szebenyi and K. Moffat, J. Biol. Chem. 261 (1986), 8761.
    122. M. Akke, T. Drakenberg, and W. J. Chazin, Biochemistry 31 (1992), 1011.
    123. S. Linse et al., Biochemistry 30 (1991), 154.
    124. S. Linse et al., Nature 335 (1988), 651.
    125. M. Akke and S. Forsén, Proteins 8 (1990), 23.
    126. W. Chazin et al., Proc. Natl. Acad. Sci. USA 86 (1989), 2195.
    127. W. Wnuk, J. A. Cox, and E. A. Stein, in Reference 35, pp. 243-278.
    128. W. J. Cook et al., J. Biol. Chem. 266 (1991), 652.
    129. J. J. Geisow and H. H. Walker, Trends Biochem. Sci. 11 (1986), 420.
    130. R. H. Kretsinger and C. E. Creutz, Nature 320 (1986), 573.
    131. M. R. Crompton, S. E. Moss, and M. J. Crumpton, Cell 55(1988), 1.
    132. J. Kay in A. Heidland and W. H. Horl, eds., Protease Role in Health and Disease, 1984, pp. 519- 570.
    133. S. Ohno et al., Nature 312 (1984), 566.
    134. D. E. Croall and G. N. DeMartino, Physiol. Rev. 71 (1991), 813.
    135. D. Carpenter, T. Jackson, and M. J. Hanley, Nature 325 (1987), 107.
    136. S. Ohno et al., Nature 325 (1987), 161.
    137. J. F. Harper et al., Science 252 (1991), 951.
    138. H. Rasmussen, Am. J. Med. 50 (1971), 567.
    139. W. Bode and P. Schwager, J. Mol. Biol. 98 (1975), 693.
    140. E. Chiancone et al., J. Mol. Biol. 185 (1985), 201.
    141. B. W. Dijkstra et al., J. Mol. Biol. 147 (1983), 97.
    142. D. L. Scott et al., Science 250 (1990), 1541.
    143. T. Drakenberg et al., Biochemistry 23 (1984), 2387.
    144. R. J. P. Williams, in Reference 47, pp. 144-159.
    145. D. I. Stuart et al., Nature 324 (1986), 84.
    146. G. L. Nelsestuen, in Reference 10, pp. 354-380.
    147. M. Soriano-Garcia et al., Biochemistry 28 (1989), 6805.
    148. D. W. Deerfield et al., J. Biol. Chem. 262 (1987), 4017.
    149. K. Harlos et al., Nature 330 (1987), 82.
    150. J. Lawler and R. O. Hynes, J. Cell. Biol. 103 (1986), 1635.
    151. V. M. Dixit et aI., J. Biol. Chem. 261 (1986), 1962.
    152. J. M. K. Slane, D. F. Mosher, and C.-S. Lai, FEBS Lett. 229 (1988), 363.
    153. R. E. Wuthier, in Reference 10, pp. 411-472.
    154. F. G. E. Pautard and R. J. P. Williams, in Chemistry in Britain, pp. 188-193 (1982).
    155. W. E. Brown and L. C. Chow, Annu. Rev. Material Sci. 6 (1976), 213.
    156. O. R. Bowen, Publ. Health Reports Suppl. Sept.-Oct. 1989, pp. 11-13.
    157. B. Alberts et al., Molecular Biology of the Cell, Garland, 1983, pp. 933-938.
    158. T. N. Davis et al., Cell 47 (1986), 423.
    159. S. Inouye, T. T. Franceschini, and M. Inouye, Proc. Natl. Acad. Sci. USA 80 (1983), 6829.
    160. G. Wistow, L. Summers, and T. L. Blundell, Nature 315 (1985), 771.
    161. P. F. Leadlay, G. Roberts, and J. E. Walker, FEBS Lett. 178 (1984), 157; D. G. Swan et al., Nature 329 (1987), 84.
    162. N. Bylsma et al., FEBS Lett. 299 (1992), 44.
    163. J. A. Cox and A. Bairoch, Nature 331 (1988), 491.
    164. M. A. Stevenson and S. K. Calderwood, Mol. Cell Biol. 10 (1990), 1234.
    165. K. T. O'Neil and W. F. DeGrado, Trends Biochem. Sci. 170 (1990), 59.
    166. V. Norris et al., Mol. Microbiol. 5 (1991), 775.
    167. M. R. Knight et al., FEBS Lett. 282 (1991), 405.
    168. S. Supattapone et al., J. Biol. Chem. 263 (1988), 1530.
    169. C. D. Ferris et al., Nature 342 (1989), 87.
    170. R. Huber et al., J. Mol. Biol. 223 (1992), 683.
    171. (a) M. Ikura et al., Science 256 (1992), 632; (b) W. E. Meador, A. R. Means, and F. A. Quiocho, Science 257 (1992), 1251.
    172. M. R. Knight et al., FEBS Lett. 282 (1991), 405.
    173. R. W. Tsieu and R. Y. Tsien, Annu. Rev. Cell Biol. 6 (1990), 715.
    174. T. E. Gunter and D. R. Pfeiffer, Am. J. Physiol. 258 (1990), c755.
    175. The authors would like to express their warm gratitude to the many students, colleagues, and coworkers who, during the preparation of this chapter, have supplied helpful comments, preprints of unpublished work, background material for figures, etc. Their encouragement is much appreciated. Special thanks are due to Drs. R. J. P. Williams and G. B. Jameson, who critically read and commented on an early version of the chapter.

    Contributors and Attributions

    • Sture Forsén (University of Lund, Chemical Centre, Physical Chemistry 2)
    • Johan Kördel (University of Lund, Chemical Centre, Physical Chemistry 2)

    3: Calcium in Biological Systems is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

    • Was this article helpful?