• Keine Ergebnisse gefunden

X- ray diffraction by crystals

4.2 Structure and ligand-binding properties of RAGE

4.2.2 Mechanisms for activation and blocking of RAGE

Up to now it is rarely understood how RAGE is activated and functions in cellular signaling.

RAGE is composed of a large extracellular ligand-binding region, a short transmembrane helix, and a highly acidic 43 amino acid long cytoplasmic domain essential for intracellular RAGE signaling (Huttunen et al., 1999). DN-RAGE is the truncated form of the receptor that is devoid of the cytoplasmic tail but has intact transmembrane-spanning and extracellular domains. In vitro studies with transformed mouse microglial cells showed that introduction of DN-RAGE prevented RAGE mediated cellular activation by S100B, even in the presence of endogenous wild-type receptor (Hofmann et al., 1999). This dominant negative effect on RAGE-dependent signaling could also be demonstrated in numerous cell and animal studies

-CONCLUSIONS- al., 2003). Remarkably, this intracellular region of RAGE is very small and has little

persistent secondary or tertiary structure (Dattilo et al., 2007). Despite being so small and unstructured, a recent study revealed that ERK kinases bind directly to this cytoplasmic tail (Ishihara et al., 2003). But still the mechanism how the signal is transduced to the intracellular domain remains unclear. By the molecular architecture with structurally inter-dependent V and C1 domains and a structurally autonomous C2 domain with rotational freedom, it is unlikely that the signal is transduced via C2 domain and the transmembrane helix to the intracellular domain. Therefore, it is proposed that RAGE signaling is mediated by ligand-induced receptor oligomerization. This was already suggested based on the hexameric structure of S100A12 (Moroz et al., 2002). The authors propose that three dimers (six subunits) of S100A12 bind three molecules of RAGE. A similar model is described for S100B where a tetrameric form (two dimers) of S100B binds to V domain of RAGE (Ostendorp et al., 2007). Such a receptor multimerization has been established as a general mechanism for the initiation of signal transduction. Many cell-surface receptors are activated by a ligand-induced multimerization process, such as Toll-like receptors (de Bouteiller et al., 2005; Hu et al., 2004) or the receptor for tumor necrosis factor α (Bazzoni et al., 1995).

An activation mechanism via RAGE dimerization is consistent with the arrangement of V-C1 molecules in the crystal. Two molecules arrange side-by-side via C1 domain. This molecular composition involves numerous hydrogen bonds as well as polar interactions (3.5 Figure 4). Interestingly, there exists a second interaction site via a bridging Zn2+ adjacent to the site mentioned above. The Zn2+ is coordinated by His180 and Glu182 from one molecule and His185’ from the other molecule. The coordination is completed by an acetate molecule derived from the crystallization buffer. Based on this pronounced interaction of two V-C1 molecules in the crystal, SPR experiments were performed to analyze the kinetic properties and the dependence on Zn2+ of this interaction. Actually, binding of V-C1 domains to V-C1 domains was observed with, as well as without Zn2+ (3.5 Figure 1 D and E). Nevertheless, Zn2+ considerably influences the kinetic properties of the interaction. Whereas the association is similar, dissociation is much faster in the absence of Zn2+. Hence, in the presence of Zn2+

the affinity is about 4-fold higher (Kd value: ~2 µM) than in the absence of Zn2+ (Kd value:

~0.5 µM). Notably, the extracellular Zn2+ concentration can reach 0.5-1 mM (Fraker and Telford, 1997). Thus, it can be assumed that a Zn2+-dependent interaction also occurs in vivo.

Compared to previous studies this study for the first time gives evidence of a dimerization of RAGE. Analytical ultracentrifugation has been used to show that bacterially expressed

-CONCLUSIONS- consistent with the initial purification of RAGE (Schmidt et al., 1992) and has been

corroborated by studies with mouse RAGE and recombinant sRAGE from yeast (Hanford et al., 2004; Ostendorp et al., 2006). However, the kinetic analysis of the complex formation suggests that these dimers are not stable enough to be detected in solution. Based on these findings the current hypothesis for RAGE activation is that bound ligands stabilize a dimeric or even higher multimeric state of RAGE. Thereby, the cytosolic domains of RAGE approximate which induces the formation of a signaling platform (3.5 Figure 7). This complex then activates the signal transducing cascade. Furthermore, with this model blocking of RAGE can be explained as well. RAGE isoforms with a truncated C-terminus, representing the intracellular signaling domain, have a dominant negative effect on RAGE activation.

Treatment with exogenous sRAGE has been shown to effectively block RAGE-dependent signaling (Park et al., 1998). This can be explained by ligand binding to wild-type RAGE and a C-terminally truncated isoform of RAGE. Such a complex can not form the required signaling platform and signal transduction is aborted (3.5 Figure 7). This clearly demonstrates the potential of RAGE as a therapeutic target in a various number of diseases (Bierhaus et al., 2006; Hudson and Schmidt, 2004), by blocking the association of RAGE.

-REFERENCES-

5. R EFERENCES

Ababou, A., Shenvi, R. A., and Desjarlais, J. R. (2001). Long-range effects on calcium binding and conformational change in the N-domain of calmodulin. Biochemistry 40, 12719-26.

Alberts, B., Johnson, A., Lewis, J., Raff, M., Roberts, K., and Walter, P. (2002). "The Cell,"

Garland Science, New York.

Allbritton, N. L., Oancea, E., Kuhn, M. A., and Meyer, T. (1994). Source of nuclear calcium signals. Proc Natl Acad Sci U S A 91, 12458-62.

Allen, F. H. (2002). The Cambridge Structural Database: a quarter of a million crystal structures and rising. Acta Crystallogr B 58, 380-8.

Anderson, C. W., and Appella, E. (2003). Signaling to the p53 Tumor Suppressor through Pathways Activated by Genotoxic and Nongenotoxic Stresses. In "Handbook of Cell Signaling" (R. Bradshaw and E. Dennis, eds.), Vol. 3. Academic Press, San Diego.

Andersson, M., Malmendal, A., Linse, S., Ivarsson, I., Forsen, S., and Svensson, L. A. (1997).

Structural basis for the negative allostery between Ca2+- and Mg2+-binding in the intracellular Ca2+-receptor calbindin D9K. Protein Sci 6, 1139-47.

Andre, I., and Linse, S. (2002). Measurement of Ca2+-binding constants of proteins and presentation of the CaLigator software. Anal Biochem 305, 195-205.

Arnesano, F., Banci, L., Bertini, I., Fantoni, A., Tenori, L., and Viezzoli, M. S. (2005).

Structural interplay between calcium(II) and copper(II) binding to S100A13 protein.

Angew Chem Int Ed Engl 44, 6341-4.

Atar, D., Backx, P. H., Appel, M. M., Gao, W. D., and Marban, E. (1995). Excitation-transcription coupling mediated by zinc influx through voltage-dependent calcium channels. J Biol Chem 270, 2473-7.

Auld, D. S. (2001). Zinc sites in Metalloenzymes and Related Proteins. In "Handbook on Metalloproteins" (I. Bertini, A. Sigel and H. Sigel, eds.), pp. 881-959. Marcel Dekker Inc., New York.

Bairoch, A., and Apweiler, R. (2000). The SWISS-PROT protein sequence database and its supplement TrEMBL in 2000. Nucleic Acids Res 28, 45-8.

Baker, N. A., Sept, D., Joseph, S., Holst, M. J., and McCammon, J. A. (2001). Electrostatics of nanosystems: application to microtubules and the ribosome. Proc Natl Acad Sci U S A 98, 10037-41.

Banci, L., Bertini, I., Cantini, F., Felli, I. C., Gonnelli, L., Hadjiliadis, N., Pierattelli, R., Rosato, A., and Voulgaris, P. (2006). The Atx1-Ccc2 complex is a metal-mediated protein-protein interaction. Nat Chem Biol 2, 367-8.

Barnes, P. J., and Karin, M. (1997). Nuclear factor-κB: a pivotal transcription factor in chronic inflammatory diseases. N Engl J Med 336, 1066-71.

Basta, G., Lazzerini, G., Massaro, M., Simoncini, T., Tanganelli, P., Fu, C., Kislinger, T., Stern, D. M., Schmidt, A. M., and De Caterina, R. (2002). Advanced glycation end products activate endothelium through signal-transduction receptor RAGE: a mechanism for amplification of inflammatory responses. Circulation 105, 816-22.

Baudier, J., Glasser, N., and Gerard, D. (1986). Ions binding to S100 proteins. I. Calcium- and

-REFERENCES- S100b (beta beta) protein: Zn2+ regulates Ca2+ binding on S100b protein. J Biol Chem

261, 8192-203.

Bazzoni, F., Alejos, E., and Beutler, B. (1995). Chimeric tumor necrosis factor receptors with constitutive signaling activity. Proc Natl Acad Sci U S A 92, 5376-80.

Beard, N. A., Laver, D. R., and Dulhunty, A. F. (2004). Calsequestrin and the calcium release channel of skeletal and cardiac muscle. Prog Biophys Mol Biol 85, 33-69.

Bellomo, G., Vairetti, M., Stivala, L., Mirabelli, F., Richelmi, P., and Orrenius, S. (1992).

Demonstration of nuclear compartmentalization of glutathione in hepatocytes. Proc Natl Acad Sci U S A 89, 4412-6.

Berg, J. M., and Shi, Y. (1996). The galvanization of biology: a growing appreciation for the roles of zinc. Science 271, 1081-5.

Berridge, M. J. (2005). Unlocking the secrets of cell signaling. Annu Rev Physiol 67, 1-21.

Berridge, M. J., Bootman, M. D., and Roderick, H. L. (2003). Calcium signalling: dynamics, homeostasis and remodelling. Nat Rev Mol Cell Biol 4, 517-29.

Berridge, M. J., Lipp, P., and Bootman, M. D. (2000). The versatility and universality of calcium signalling. Nat Rev Mol Cell Biol 1, 11-21.

Bertini, I., Luchinat, C., and Monnanni, R. (1985). Zinc Enzymes. J Chem Educ 62, 924-7.

Beyersmann, D., and Haase, H. (2001). Functions of zinc in signaling, proliferation and differentiation of mammalian cells. Biometals 14, 331-41.

Bhattacharya, S., Bunick, C. G., and Chazin, W. J. (2004). Target selectivity in EF-hand calcium binding proteins. Biochim Biophys Acta 1742, 69-79.

Bhattacharya, S., Large, E., Heizmann, C. W., Hemmings, B., and Chazin, W. J. (2003).

Structure of the Ca2+/S100B/NDR kinase peptide complex: insights into S100 target specificity and activation of the kinase. Biochemistry 42, 14416-26.

Bierhaus, A., Hofmann, M. A., Ziegler, R., and Nawroth, P. P. (1998). AGEs and their interaction with AGE-receptors in vascular disease and diabetes mellitus. I. The AGE concept. Cardiovasc Res 37, 586-600.

Bierhaus, A., Humpert, P. M., Morcos, M., Wendt, T., Chavakis, T., Arnold, B., Stern, D. M., and Nawroth, P. P. (2005). Understanding RAGE, the receptor for advanced glycation end products. J Mol Med 83, 876-86.

Bierhaus, A., Schiekofer, S., Schwaninger, M., Andrassy, M., Humpert, P. M., Chen, J., Hong, M., Luther, T., Henle, T., Kloting, I., Morcos, M., Hofmann, M., Tritschler, H., Weigle, B., Kasper, M., Smith, M., Perry, G., Schmidt, A. M., Stern, D. M., Haring, H. U., Schleicher, E., and Nawroth, P. P. (2001). Diabetes-associated sustained activation of the transcription factor nuclear factor-κB. Diabetes 50, 2792-808.

Bierhaus, A., Stern, D. M., and Nawroth, P. P. (2006). RAGE in inflammation: a new therapeutic target? Curr Opin Investig Drugs 7, 985-91.

Blanchard, H., Grochulski, P., Li, Y., Arthur, J. S., Davies, P. L., Elce, J. S., and Cygler, M.

(1997). Structure of a calpain Ca2+-binding domain reveals a novel EF-hand and Ca2+ -induced conformational changes. Nat Struct Biol 4, 532-8.

Blasie, C. A., and Berg, J. M. (2002). Structure-based thermodynamic analysis of a coupled metal binding-protein folding reaction involving a zinc finger peptide. Biochemistry 41, 15068-73.

Blow, D. (2002). "Outline of crystallography for biologists," Oxford University Press, Oxford.

Böhm, G., Muhr, R., and Jaenicke, R. (1992). Quantitative analysis of protein far UV circular dichroism spectra by neural networks. Protein Eng 5, 191-5.

Branden, C., and Tooze, J. (1991). "Introduction to Protein Structure," Garland Publishing, Inc., New York.

-REFERENCES- Brodersen, D. E., Nyborg, J., and Kjeldgaard, M. (1999). Zinc-binding site of an S100 protein

revealed. Two crystal structures of Ca2+-bound human psoriasin (S100A7) in the Zn2+ -loaded and Zn2+-free states. Biochemistry 38, 1695-704.

Bruker (2005). XPREP. Bruker AXS Inc., Madison, Wisonsin, USA.

Canzoniero, L. M., Turetsky, D. M., and Choi, D. W. (1999). Measurement of intracellular free zinc concentrations accompanying zinc-induced neuronal death. J Neurosci 19, RC31.

Carafoli, E. (2002). Calcium signaling: a tale for all seasons. Proc Natl Acad Sci U S A 99, 1115-22.

Carafoli, E. (2003). The calcium-signalling saga: tap water and protein crystals. Nat Rev Mol Cell Biol 4, 326-32.

Carafoli, E. (2004). Calcium-mediated cellular signals: a story of failures. Trends Biochem Sci 29, 371-9.

Carafoli, E. (2005). Calcium-a universal carrier of biological signals. Febs J 272, 1073-89.

Chattopadhyaya, R., Meador, W. E., Means, A. R., and Quiocho, F. A. (1992). Calmodulin structure refined at 1.7 A resolution. J Mol Biol 228, 1177-92.

Chavakis, T., Bierhaus, A., Al-Fakhri, N., Schneider, D., Witte, S., Linn, T., Nagashima, M., Morser, J., Arnold, B., Preissner, K. T., and Nawroth, P. P. (2003). The pattern recognition receptor (RAGE) is a counterreceptor for leukocyte integrins: a novel pathway for inflammatory cell recruitment. J Exp Med 198, 1507-15.

Chavakis, T., Bierhaus, A., and Nawroth, P. P. (2004). RAGE (receptor for advanced glycation end products): a central player in the inflammatory response. Microbes Infect 6, 1219-25.

Chen, X., Chu, M., and Giedroc, D. P. (2000). Spectroscopic characterization of Co(II)-, Ni(II)-, and Cd(II)-substituted wild-type and non-native retroviral-type zinc finger peptides. J Biol Inorg Chem 5, 93-101.

Chin, D., and Means, A. R. (2000). Calmodulin: a prototypical calcium sensor. Trends Cell Biol 10, 322-8.

Cho, W., Bittova, L., and Stahelin, R. V. (2001). Membrane binding assays for peripheral proteins. Anal Biochem 296, 153-61.

Choi, D. W., and Koh, J. Y. (1998). Zinc and brain injury. Annu Rev Neurosci 21, 347-75.

Cormier, M. J. (1983). Calmodulin: The Regulation of Cellular Function. In "Calcium in Biology", Vol. 6, pp. 53-106. John Wiley & Sons, New York.

Cousins, R. J., Liuzzi, J. P., and Lichten, L. A. (2006). Mammalian zinc transport, trafficking, and signals. J Biol Chem 281, 24085-9.

Coyle, J. T., and Puttfarcken, P. (1993). Oxidative stress, glutamate, and neurodegenerative disorders. Science 262, 689-95.

Dattilo, B. M., Fritz, G., Leclerc, E., Kooi, C. W., Heizmann, C. W., and Chazin, W. J.

(2007). The Extracellular Region of the Receptor for Advanced Glycation End Products Is Composed of Two Independent Structural Units. Biochemistry.

Davey, G. E., Murmann, P., and Heizmann, C. W. (2001). Intracellular Ca2+ and Zn2+ levels regulate the alternative cell density-dependent secretion of S100B in human

glioblastoma cells. J Biol Chem 276, 30819-26.

de Bouteiller, O., Merck, E., Hasan, U. A., Hubac, S., Benguigui, B., Trinchieri, G., Bates, E.

E., and Caux, C. (2005). Recognition of double-stranded RNA by human toll-like receptor 3 and downstream receptor signaling requires multimerization and an acidic pH. J Biol Chem 280, 38133-45.

Deane, R., Du Yan, S., Submamaryan, R. K., LaRue, B., Jovanovic, S., Hogg, E., Welch, D., Manness, L., Lin, C., Yu, J., Zhu, H., Ghiso, J., Frangione, B., Stern, A., Schmidt, A.

-REFERENCES- Stern, D., and Zlokovic, B. (2003). RAGE mediates amyloid-β peptide transport

across the blood-brain barrier and accumulation in brain. Nat Med 9, 907-13.

Delano, W. L. (2002). The PyMol Molecular Graphics System.

Dell'Angelica, E. C., Schleicher, C. H., and Santome, J. A. (1994). Primary structure and binding properties of calgranulin C, a novel S100-like calcium-binding protein from pig granulocytes. J Biol Chem 269, 28929-36.

Dempsey, A. C., Walsh, M. P., and Shaw, G. S. (2003). Unmasking the annexin I interaction from the structure of Apo-S100A11. Structure 11, 887-97.

Devinney, M. J., 2nd, Reynolds, I. J., and Dineley, K. E. (2005). Simultaneous detection of intracellular free calcium and zinc using fura-2FF and FluoZin-3. Cell Calcium 37, 225-32.

Donato, R. (2003). Intracellular and extracellular roles of S100 proteins. Microsc Res Tech 60, 540-51.

Drohat, A. C., Tjandra, N., Baldisseri, D. M., and Weber, D. J. (1999). The use of dipolar couplings for determining the solution structure of rat apo-S100B(betabeta). Protein Sci 8, 800-9.

Du Yan, S., Zhu, H., Fu, J., Yan, S. F., Roher, A., Tourtellotte, W. W., Rajavashisth, T., Chen, X., Godman, G. C., Stern, D., and Schmidt, A. M. (1997). Amyloid-beta peptide-receptor for advanced glycation endproduct interaction elicits neuronal expression of macrophage-colony stimulating factor: a proinflammatory pathway in Alzheimer disease. Proc Natl Acad Sci U S A 94, 5296-301.

Eklund, H., Nordstrom, B., Zeppezauer, E., Soderlund, G., Ohlsson, I., Boiwe, T., and Branden, C. I. (1974). The structure of horse liver alcohol dehydrogenase. FEBS Lett 44, 200-4.

Emsley, P., and Cowtan, K. (2004). Coot: model-building tools for molecular graphics. Acta Crystallogr D Biol Crystallogr 60, 2126-32.

Faber, H. R., Groom, C. R., Baker, H. M., Morgan, W. T., Smith, A., and Baker, E. N. (1995).

1.8 A crystal structure of the C-terminal domain of rabbit serum haemopexin.

Structure 3, 551-9.

Fages, C., Nolo, R., Huttunen, H. J., Eskelinen, E., and Rauvala, H. (2000). Regulation of cell migration by amphoterin. J Cell Sci 113 ( Pt 4), 611-20.

Faller, P., and Vasak, M. (1997). Distinct metal-thiolate clusters in the N-terminal domain of neuronal growth inhibitory factor. Biochemistry 36, 13341-8.

Feng, G., Xu, X., Youssef, E. M., and Lotan, R. (2001). Diminished expression of S100A2, a putative tumor suppressor, at early stage of human lung carcinogenesis. Cancer Res 61, 7999-8004.

Férnandez-Chacón, R., Shin, O. H., Konigstorfer, A., Matos, M. F., Meyer, A. C., Garcia, J., Gerber, S. H., Rizo, J., Südhof, T. C., and Rosenmund, C. (2002). Structure/function analysis of Ca2+ binding to the C2A domain of synaptotagmin 1. J Neurosci 22, 8438-46.

Finkel, T., and Holbrook, N. J. (2000). Oxidants, oxidative stress and the biology of ageing.

Nature 408, 239-47.

Fiser, A., and Sali, A. (2003). Modeller: generation and refinement of homology-based protein structure models. Methods Enzymol 374, 461-91.

Folkers, G. E., Hanzawa, H., and Boelens, R. (2001). Zinc Finger Proteins. In "Handbook on Metalloproteins" (I. Bertini, A. Sigel and H. Sigel, eds.), pp. 961-1000. Marcel Dekker Inc., New York.

Fortelle, E. d. l., and Bricogne, G. (1997). Maximum-Likelihood Heavy-Atom Parameter Refinement for Multiple Isomorphous Replacement and Multiwavelength Anomalous

-REFERENCES- Fraker, P. J., and Telford, W. G. (1997). A reappraisal of the role of zinc in life and death

decisions of cells. Proc Soc Exp Biol Med 215, 229-36.

Franz, C., Durussel, I., Cox, J. A., Schäfer, B. W., and Heizmann, C. W. (1998). Binding of Ca2+ and Zn2+ to human nuclear S100A2 and mutant proteins. J Biol Chem 273, 18826-34.

Frederickson, C. (2003). Imaging zinc: old and new tools. Sci STKE 2003, pe18.

Frederickson, C. J., Burdette, S. C., Sensi, S. L., Weiss, J. H., Yin, H. Z., Balaji, R. V., Truong-Tran, A. Q., Bedell, E., Prough, D. S., and Lippard, S. J. (2004). Method for identifying neuronal cells suffering zinc toxicity by use of a novel fluorescent sensor.

J Neurosci Methods 139, 79-89.

Freigang, J., Proba, K., Leder, L., Diederichs, K., Sonderegger, P., and Welte, W. (2000). The crystal structure of the ligand binding module of axonin-1/TAG-1 suggests a zipper mechanism for neural cell adhesion. Cell 101, 425-33.

Fritz, G., and Heizmann, C. W. (2004). 3D structures of the calcium and zinc binding S100 proteins. In "Handbook of Metalloproteins" (A. Messerschmidt, W. Bode and M.

Cygler, eds.), Vol. 3, pp. 529-40. John Wiley & Sons, Inc., Chichester.

Fritz, G., Heizmann, C. W., and Kroneck, P. M. (1998). Probing the structure of the human Ca2+- and Zn2+-binding protein S100A3: spectroscopic investigations of its transition metal ion complexes, and three-dimensional structural model. Biochim Biophys Acta 1448, 264-76.

Fritz, G., Mittl, P. R., Vasak, M., Grutter, M. G., and Heizmann, C. W. (2002). The crystal structure of metal-free human EF-hand protein S100A3 at 1.7-A resolution. J Biol Chem 277, 33092-8.

Gimona, M., Lando, Z., Dolginov, Y., Vandekerckhove, J., Kobayashi, R., Sobieszek, A., and Helfman, D. M. (1997). Ca2+-dependent interaction of S100A2 with muscle and nonmuscle tropomyosins. J Cell Sci 110 ( Pt 5), 611-21.

Glenney, J. R., Jr., Kindy, M. S., and Zokas, L. (1989). Isolation of a new member of the S100 protein family: amino acid sequence, tissue, and subcellular distribution. J Cell Biol 108, 569-78.

Grabarek, Z. (2006). Structural basis for diversity of the EF-hand calcium-binding proteins. J Mol Biol 359, 509-25.

Greenamyre, J. T., and Hastings, T. G. (2004). Biomedicine. Parkinson's--divergent causes, convergent mechanisms. Science 304, 1120-2.

Gribenko, A. V., and Makhatadze, G. I. (1998). Oligomerization and divalent ion binding properties of the S100P protein: a Ca2+/Mg2+-switch model. J Mol Biol 283, 679-94.

Guex, N., and Peitsch, M. C. (1997). SWISS-MODEL and the Swiss-PdbViewer: an environment for comparative protein modeling. Electrophoresis 18, 2714-23.

Gupta, S., Hussain, T., MacLennan, G. T., Fu, P., Patel, J., and Mukhtar, H. (2003).

Differential expression of S100A2 and S100A4 during progression of human prostate adenocarcinoma. J Clin Oncol 21, 106-12.

Haase, H., Hebel, S., Engelhardt, G., and Rink, L. (2006). Flow cytometric measurement of labile zinc in peripheral blood mononuclear cells. Anal Biochem 352, 222-30.

Haase, H., and Maret, W. (2003). Intracellular zinc fluctuations modulate protein tyrosine phosphatase activity in insulin/insulin-like growth factor-1 signaling. Exp Cell Res 291, 289-98.

Haiech, J., Moulhaye, S. B., and Kilhoffer, M. C. (2004). The EF-Handome: combining comparative genomic study using FamDBtool, a new bioinformatics tool, and the network of expertise of the European Calcium Society. Biochim Biophys Acta 1742, 179-83.

-REFERENCES- Hanford, L. E., Enghild, J. J., Valnickova, Z., Petersen, S. V., Schaefer, L. M., Schaefer, T.

M., Reinhart, T. A., and Oury, T. D. (2004). Purification and characterization of mouse soluble receptor for advanced glycation end products (sRAGE). J Biol Chem 279, 50019-24.

Harding, M. M. (2002). Metal-ligand geometry relevant to proteins and in proteins: sodium and potassium. Acta Crystallogr D Biol Crystallogr 58, 872-4.

Harding, M. M. (2006). Small revisions to predicted distances around metal sites in proteins.

Acta Crystallogr D Biol Crystallogr 62, 678-82.

Heinz, U., Kiefer, M., Tholey, A., and Adolph, H. W. (2005). On the competition for available zinc. J Biol Chem 280, 3197-207.

Heizmann, C. W., and Cox, J. A. (1998). New perspectives on S100 proteins: a multi-functional Ca2+-, Zn2+- and Cu2+-binding protein family. Biometals 11, 383-97.

Heizmann, C. W., Fritz, G., and Schäfer, B. W. (2002). S100 proteins: structure, functions and pathology. Front Biosci 7, d1356-68.

Heizmann, C. W., Schäfer, B. W., and Fritz, G. (2003). The Family of S100 Cell Signaling Proteins. In "Handbook of Cell Signaling" (R. Bradshaw and E. Dennis, eds.), Vol. 2.

Academic Press, San Diego.

Hendrickson, W. A., and Teeter, M. M. (1981). Structure of the hydrophobic protein crambin determined directly from the anomalous scattering of sulphur. Nature 290, 107-13.

Henehan, C. J., Pountney, D. L., Zerbe, O., and Vasak, M. (1993). Identification of cysteine ligands in metalloproteins using optical and NMR spectroscopy: cadmium-substituted rubredoxin as a model [Cd(CysS)4]2- center. Protein Sci 2, 1756-64.

Hibi, K., Fujitake, S., Takase, T., Kodera, Y., Ito, K., Akiyama, S., Shirane, M., and Nakao, A. (2003). Identification of S100A2 as a target of the DeltaNp63 oncogenic pathway.

Clin Cancer Res 9, 4282-5.

Hofmann, M. A., Drury, S., Fu, C., Qu, W., Taguchi, A., Lu, Y., Avila, C., Kambham, N., Bierhaus, A., Nawroth, P., Neurath, M. F., Slattery, T., Beach, D., McClary, J., Nagashima, M., Morser, J., Stern, D., and Schmidt, A. M. (1999). RAGE mediates a novel proinflammatory axis: a central cell surface receptor for S100/calgranulin polypeptides. Cell 97, 889-901.

Hope, H. (1988). Cryocrystallography of biological macromolecules: a generally applicable method. Acta Crystallogr B 44 ( Pt 1), 22-6.

Hopfner, K. P., Craig, L., Moncalian, G., Zinkel, R. A., Usui, T., Owen, B. A., Karcher, A., Henderson, B., Bodmer, J. L., McMurray, C. T., Carney, J. P., Petrini, J. H., and Tainer, J. A. (2002). The Rad50 zinc-hook is a structure joining Mre11 complexes in DNA recombination and repair. Nature 418, 562-6.

Hoppe, W. Z. (1957). Die Faltmolekülmethode: eine neue Methode zur Bestimmung der Kristallstruktur bei ganz oder teilweise bekannten Molekülstrukturen. Acta Crystallogr 10, 750-51.

Hori, O., Brett, J., Slattery, T., Cao, R., Zhang, J., Chen, J. X., Nagashima, M., Lundh, E. R., Vijay, S., Nitecki, D., and et al. (1995). The receptor for advanced glycation end products (RAGE) is a cellular binding site for amphoterin. Mediation of neurite outgrowth and co-expression of rage and amphoterin in the developing nervous system. J Biol Chem 270, 25752-61.

Houdusse, A., Love, M. L., Dominguez, R., Grabarek, Z., and Cohen, C. (1997). Structures of four Ca2+-bound troponin C at 2.0 A resolution: further insights into the Ca2+-switch in the calmodulin superfamily. Structure 5, 1695-711.

Hu, X., Yagi, Y., Tanji, T., Zhou, S., and Ip, Y. T. (2004). Multimerization and interaction of Toll and Spatzle in Drosophila. Proc Natl Acad Sci U S A 101, 9369-74.

-REFERENCES- Huang, M., Krepkiy, D., Hu, W., and Petering, D. H. (2004). Zn-, Cd-, and Pb-transcription

factor IIIA: properties, DNA binding, and comparison with TFIIIA-finger 3 metal complexes. J Inorg Biochem 98, 775-85.

Huber, R. (1965). Die automatisierte Faltmolekülmethode. Acta Crystallogr 19, 353-56.

Hudson, B. I., Bucciarelli, L. G., Wendt, T., Sakaguchi, T., Lalla, E., Qu, W., Lu, Y., Lee, L., Stern, D. M., Naka, Y., Ramasamy, R., Yan, S. D., Yan, S. F., D'Agati, V., and Schmidt, A. M. (2003). Blockade of receptor for advanced glycation endproducts: a new target for therapeutic intervention in diabetic complications and inflammatory disorders. Archives of Biochemistry & Biophysics 419, 80-8.

Hudson, B. I., and Schmidt, A. M. (2004). RAGE: a novel target for drug intervention in diabetic vascular disease. Pharm Res 21, 1079-86.

Huttunen, H. J., Fages, C., and Rauvala, H. (1999). Receptor for advanced glycation end products (RAGE)-mediated neurite outgrowth and activation of NF-kappaB require the cytoplasmic domain of the receptor but different downstream signaling pathways.

J Biol Chem 274, 19919-24.

Huttunen, H. J., Kuja-Panula, J., Sorci, G., Agneletti, A. L., Donato, R., and Rauvala, H.

(2000). Coregulation of neurite outgrowth and cell survival by amphoterin and S100 proteins through receptor for advanced glycation end products (RAGE) activation. J Biol Chem 275, 40096-105.

Ikemoto, N., Nagy, B., Bhatnagar, G. M., and Gergely, J. (1974). Studies on a metal-binding protein of the sarcoplasmic reticulum. J Biol Chem 249, 2357-65.

Inman, K. G., Baldisseri, D. M., Miller, K. E., and Weber, D. J. (2001). Backbone dynamics of the calcium-signaling protein apo-S100B as determined by 15N NMR relaxation.

Biochemistry 40, 3439-48.

Inman, K. G., Yang, R., Rustandi, R. R., Miller, K. E., Baldisseri, D. M., and Weber, D. J.

(2002). Solution NMR structure of S100B bound to the high-affinity target peptide TRTK-12. J Mol Biol 324, 1003-14.

Ishida, H., Nakashima, K., Kumaki, Y., Nakata, M., Hikichi, K., and Yazawa, M. (2002). The solution structure of apocalmodulin from Saccharomyces cerevisiae implies a

mechanism for its unique Ca2+ binding property. Biochemistry 41, 15536-42.

mechanism for its unique Ca2+ binding property. Biochemistry 41, 15536-42.