• Keine Ergebnisse gefunden

Protein crystallization

H. Kroneck a , Claus W. Heizmann b and Günter Fritz a

4.2. Structures of Zn 2+ -loaded Ca 2+ -S100B

We have determined the structures of Zn2+-Ca2+-loaded human-S100B at pH 6.5, 9.0 and 10.0 (3.4.5. Table. 2). The protein crystallized in the presence of equimolar amounts of Zn2+ (one Zn2+ per subunit) and excess amounts of Ca2+. The structures revealed a high-occupancy sites for both Zn2+ and Ca2+. There are two Zn2+ ions and four Ca2+ ions bound per dimer. The overall fold of the human Zn2+-Ca2+-S100B is similar to that of other S100 proteins like the human Zn2+-Ca2+-S100A7 and the Cu2+-Ca2+-S100A12 structure (3.4.5.Fig. 5 C). The overall folds of the three presented structures are virtually identical. Figure 1 A (3.4.5.). The Ca2+ ions are bound in the two EF-hands, whereas the Zn2+ ions are bound at interface of the subunits (3.4.5. Fig. 2 A-C). The structure of Zn2+-Ca2+-loaded human S100B is very similar to the structures of S100B in complex with Ca2+ or Ca2+ and peptides derived from target proteins like TRTK-S100B (1MWN), NDR-S100B (1PSB) and p53367-388 (1DT7) (Bhattacharya et al., 2003; Inman et al., 2002a; Rustandi et al., 2000a). However, there are some noticeable differences between the X-ray structures of Zn2+-Ca2+-S100B when compared to other structures of S100B. In a recent study, it was shown that binding of Zn2+ to Ca2+-loaded S100B leads to the formation of a kink in helixIV, close to their C-terminus and that the dimers rearrange to a denser packing, in the zinc loaded state (3.4.5. Fig. 1 C) (Wilder et al., 2005). The structures of human Zn2+-Ca2+-S100B presented in this study, revealed neither a kink formation nor a denser packing. Remarkably, the binding of Zn2+ to Ca2+-S100B results only in a minor rearrangement of the C-terminus, which induces an elongation of helix IV by 5 residues compared to Zn2+ free Ca2+-S100B (3.4.5.Fig. 1 A and 3). CD spectra of S100B in the Ca2+ loaded and Zn2+-Ca2+-loaded state were recorded and compared. In general, the CD spectra of Ca2+-S100B are very similar to those of Zn2+-Ca2+-S100B.

Nevertheless, upon addition of Zn2+ there was a minor change in the CD spectrum, indicating an increased α-helical content most likely due to the elongation in helix IV (3.4.5. Fig. 6.). This helix extension causes a reorientation of several residues, which are involved in target recognition. Particular Phe87 and Phe88, which now point towards the target-binding cleft. The rearrangement of the C-terminal residues Phe87 and Phe88 upon Zn2+ binding enlarges the hydrophobic surface of S100B. Such an rearrangement of the hydrophobic residues Phe87 and Phe88 are reasonably for the 5-fold higher affinity of bound TRTK-12 peptide from CapZ to the Zn2+-Ca2+-S100B, compared to Ca2+ -S100B (Barber et al., 1999). Besides this, the binding of Zn2+ could also affect the

binding affinity of ligands to S100B in a positive manner by stabilizing the binding site at the C-terminus of S100B and therfor stabilizing the ligand bound S100B complex. A more pronounced effect of sole Zn2+-loaded S100B is reflected in the ability of Zn2+ -S100B to induce total disassembly of microtubules in the presence of µM Zn2+

concentrations were Zn2+-S100B leads to an inhibition of τ phosphorylation by protein kinase II (Baudier and Cole, 1988), possibly by Zn2+ sequestration. It might be possible that the C-terminus of S100B form a functional Zn2+ binding site including the residues Phe87 and Phe88 which can cooperate independently from the Ca2+ binding EF-hands, therefore S100B may exert besides a Ca2+ dependent function also a Zn2+ dependent function, without Ca2+. In all these cases, Zn2+ binding may enable Ca2+-S100B to an additional way to modulate the dissociation rate for bound ligands.

Comparison of the interhelical angles within the three Zn2+-Ca2+-S100B structures indicates that the angles and distances of the flanking helices of the S100-specific EF-hand are virtually identical (I-II = 137.7 ± 0.4°, 15 Å; III-IV = 106,2 ± 0.8°, 15 Å Table 4 3.4.5.). A difference was detected for the flanking helices III and IV of the classical EF-hand, were the comparison of the human Ca2+-S100B crystal structure (2H61) with the human Zn2+-Ca2+-S100B crystal structures shows only a minor effect. Here, the addition of Zn2+ to the Ca2+-S100B induces only a small change to a more open conformation, shown by an increase of the interhelical angles between helices III and IV of 2°-3°.

Combined, Zn2+ binding does not induce any noteworthy change in the interhelical angles of both EF-hands but it favors an enlargement of the hydrophobic surface at the C-terminus and therefore may contribute to the binding affinity of Zn2+-Ca2+-S100B to its ligands.

In previous studies, the divalent cation binding properties of the human S100B was studied by Baudier et. al. and it was found that S100B binds four Ca2+ and 6-8 Zn2+ (two high affinity plus 4-6 low affinity) atoms per dimer. The affinity of S100B for Zn2+ is considerably higher than that for Ca2+. The binding constant for the Zn2+ ion was measured and determined to ~100 nM. (Baudier et al., 1986; Rustandi et al., 1998). The strong binding of the Zn2+ to S100B is achieved through four residues of a preformed binding site (3.4.5. Fig.2 A-C). The distance between the two Zn2+ ions in the structures is approximately 30.4 Å. Interestingly, alignments of the main-chain atoms (residues 1-89) of human Ca2+-S100B subunits with the obtained Zn2+-Ca2+-S100B structures show no significant changes upon binding of Zn2+, neither in the overall structure (not shown) nor in the EF-hand Ca2+ binding loops (3.4.5. Fig. 5 B). Crystal structures obtained at pH

-CONCLUSIONS-

124

6.5 and pH 10.0 show two histidines His15, His25 from one subunit and one histidine, His85´ and a Glu89´ from the second subunit involved in coordinating Zn2+. The Zn2+

coordination is obtained by the imidazoles of the histidines and the terminal carboxylate oxygen atom of the glutamate. The O-Zn and N-Zn distances at the Zn2+ binding sites are comparable to other Zn2+ binding sites (3.4.5. Table: 4) (Harding, 2000).

Furthermore, the Zn2+ binding site ties the two subunits more together, since two of the coordinating ligands of at Zn2+ site originate from one subunit and the other ligands are derived from the second subunit (3.4.5. Fig. 2 A-C). Remarkably, in the crystal structure obtained at pH 9.0, the Zn2+ ion is coordinated by four histidines (His15, His25, His85´

and His90´). In all three Zn2+-Ca2+-S100B structures, the Zn2+ ligand is coordinated in a slightly distorted tetraeder (3.4.5. Fig. 2 A-C). A similar coordination was shown for Cu2+ in the Cu2+-Ca2+-S100A12 structure by three histidines His15, His85´ and His89´

and Asp25 (Moroz et al., 2003) and for the Zn2+-Ca2+-loaded S100A7 were Zn2+ is coordinated by His15, His85´ and His89´ and Asp25 (Fig 4 A-C). It is proposed that S100B binds Cu2+ at the same site. Sequence alignments showed that only the calgranulins S100A8, S100A9 and S100A12, as well as S100B and S100A7 have such a common Zn2+-/Cu2+-binding motif which is highly conserved (Moroz et al., 2003).

Interestingly, these Zn2+ binding S100-proteins have an extracellular function. There they can exert Zn2+ dependent function distinct from intracellular functions, in both circumstances, Zn2+ binding promotes ligand-binding affinity of S100-proteins. In general, once formed, Zn2+ binding site seems to be very stable in S100-proteins, which were shown by previous, and our recent study (pH range 6.5 to 10.0).

Binding of Zn2+ induces no conformational change in both EF-hands, (3.4.5. Fig.5 B-C).

Although His25 is located at the S100-specific N-terminal EF-hand the conformation of this EF-hand, remains essentially the same upon Zn2+ binding. The 10-fold increase in Ca2+ affinity to S100B (Baudier et al., 1986) in the presence of Zn2+ is explained in that way that Zn2+ binding leads to a pre positioning of the N-terminal EF-hand loop which ease the followed Ca2+-binding (3.4.5. Fig. 6 A). One can presume that the N-terminal EF-hand loop is brought into a virtual Ca2+ loaded position and therefore promotes the higher binding affinity for Ca2+. Alignment of the N-terminal EF-hand of the Cu2+-Ca2+ -S100A12 with the Zn2+-Ca2+-S100B showed that in both structures the EF-hand loops are kept in the very same position (3.4.5. Fig. 6 C). The same was shown for the EF-hand region in the S100A7 Zn2+-loaded structure (Brodersen et al., 1999). Besides this there was also no structural change observed in the C-terminal classical EF-hand loop. This

indicates that Zn2+ binding does not induce any conformational change at the EF-hand regions. So one can presume that the Ca2+-loaded N-terminal EF-hand exhibits already an ideal topology and geometry for Ca2+ binding and that Zn2+ binding occurs mainly because of stability and less because of structural reasons. It is coherent that the binding of Zn2+ stabilizes the structure at the Ca2+-loaded N-terminal EF-hand, by inducing a more rigid loop conformation that might exhibit a Ca2+-loaded like topology, which promotes calcium binding.

5 R EFERENCES

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

Alison, J.E.G., Keir, G. and Thomson, E.J. (1997) A specific and sensitive ELSIA for measuring S-100b in cerebrospinal fluid. J. Immunol. Methodes, 35-41.

Ambartsumian, N., Klingelhofer, J., Grigorian, M., Christensen, C., Kriajevska, M., Tulchinsky, E., Georgiev, G., Berezin, V., Bock, E., Rygaard, J., Cao, R., Cao, Y. and Lukanidin, E. (2001) The metastasis-associated Mts1(S100A4) protein could act as an angiogenic factor. Oncogene, 20, 4685-4695.

Arancio, O., Zhang, H.P., Chen, X., Lin, C., Trinchese, F., Puzzo, D., Liu, S., Hegde, A., Yan, S.F., Stern, A., Luddy, J.S., Lue, L.F., Walker, D.G., Roher, A., Buttini, M., Mucke, L., Li, W., Schmidt, A.M., Kindy, M., Hyslop, P.A., Stern, D.M. and Du Yan, S.S. (2004) RAGE potentiates Abeta-induced perturbation of neuronal function in transgenic mice. Embo J, 23, 4096-4105.

Arumugam, T., Simeone, D.M., Schmidt, A.M. and Logsdon, C.D. (2004) S100P stimulates cell proliferation and survival via receptor for activated glycation end products (RAGE). J Biol Chem, 279, 5059-5065.

Baudier, J., Briving, C., Deinum, J., Haglid, K., Sorskog, L. and Wallin, M. (1982) Effect of S-100 proteins and calmodulin on Ca2+-induced disassembly of brain microtubule proteins in vitro. FEBS Lett, 147, 165-168.

Baudier, J. and Cole, R.D. (1988) Interactions between the microtubule-associated tau proteins and S100b regulate tau phosphorylation by the Ca2+/calmodulin-dependent protein kinase II. J Biol Chem, 263, 5876-5883.

Baudier, J., Glasser, N. and Gerard, D. (1986) Ions binding to S100 proteins. I. Calcium- and zinc-binding properties of bovine brain S100 alpha alpha, S100a (alpha beta), and S100b (beta beta) protein: Zn2+ regulates Ca2+

binding on S100b protein. J Biol Chem, 261, 8192-81203.

Baudier, J., Delphin, C., Grunwald, D., Khochbin, S. and Lawrence, J.J. (1992) Characterization of the tumor suppressor protein p53 as a protein kinase C substrate and a S100b-binding protein. Proc Natl Acad Sci U S A, 89, 11627-11631.

Barber KR, McClintock KA, Jamieson GA Jr, Dimlich RV and GS., S. (1999) Specificity and Zn2+ enhancement of the S100B binding epitope TRTK-12. J Biol Chem., 274(3), 1502-1508.

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, 5376-5380.

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.

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, 1695-1704.

Berridge, M.J. (1997) Elementary and global aspects of calcium signalling. J Exp Biol, 200, 315-319.

Berridge, M.J., Bootman, M.D. and Roderick, H.L. (2003) Calcium signalling: dynamics, homeostasis and remodelling. Nat Rev Mol Cell Biol, 4, 517-529.

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.

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, 14416-14426.

-REFERENCES-

128

Blanchard, H., Grochulski, P., Li, Y., Arthur, J.S., Davies, P.L., Elce, J.S. and Cygler, M. (1997) Structure of a calpain Ca(2+)-binding domain reveals a novel EF-hand and Ca(2+)-induced conformational changes. Nat Struct Biol, 4, 532-538.

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

Böhm, G. (1997) CDNN 2.1 CD Spectra Deconvolution. Halle.

Böhm, G., Muhr, R. and Jaenicke, R. (1992) Quantitative analysis of protein far UV circular dichroism spectra by neural networks. Protein Eng, 5, 191-195.

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

Brünger, A.T. (1992) Free R value: a novel statistical quantity for assessing the accuracy of crystal structures. Nature, 355, 472-475.

Brünger, A.T. (1993) Assessment of phase accuracy by cross validation: the free R value. Methods and applications.

Acta Cryst., D49, 24-36.

Brünger, A.T., Adams, P.D., Clore, G.M., DeLano, W.L., Gros, P., Grosse-Kunstleve, R.W., Jiang, J.S., Kuszewski, J., Nilges, M., Pannu, N.S., Read, R.J., Rice, L.M., Simonson, T. and Warren, G.L. (1998) Crystallography &

NMR system: A new software suite for macromolecular structure determination. Acta Crystallogr., D54, 905-921.

Brückner, F.N.A., Skerra, A., Korndoerfer, I.P. (2005) The crystal structure of the complex between human S100A8 and S100A9 reveals the reasons underlying heterotetramer formation To be Published, pdb code 1XK4.

Businaro, R., Leone, S., Fabrizi, C., Sorci, G., Donato, R., Lauro, G.M. and Fumagalli, L. (2006) S100B protects LAN-5 neuroblastoma cells against Abeta amyloid-induced neurotoxicity via RAGE engagement at low doses but increases Abeta amyloid neurotoxicity at high doses. J. Neurosci. Res., 83, 897-906.

Carafoli, E. (2002) Calcium signaling: a tale for all seasons. Proc Natl Acad Sci U S A, 99, 1115-1122.

Carafoli, E. (2003) The calcium-signalling saga: tap water and protein crystals. Nat Rev Mol Cell Biol, 4, 326-332.

Carafoli, E. (2004a) Calcium signaling: a historical account. Biol Res, 37, 497-505.

Carafoli, E. (2004b) Calcium-mediated cellular signals: a story of failures. Trends Biochem Sci, 29, 371-379.

Carafoli, E. (2005) Calcium - a universal carrier of biological signals. FEBS Journal, 272, 1073 - 1089.

CCP4. (1994) The CCP4 suite: programs for protein crystallography. Acta Crystallogr. D Biol. Crystallogr., 50, 760-763.

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, 1507-1515.

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, 1219-1225.

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

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, 30819-30826.

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, 38133-38145.

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, 38133-38145.

DeLano, W. (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, 28929-28936.

Donato, R. (2001) S100: a multigenic family of calcium-modulated proteins of the EF-hand type with intracellular and extracellular functional roles. Int J Biochem Cell Biol, 33, 637-668.

Donato, R. (2003) Intracellular and extracellular roles of S100 proteins. Microsc Res Tech, 60, 540-551.

Drohat, A.C, Amburgey, J.C, Abildgaard, F., Starich, M.R, Baldisseri, D., Weber, D.J Solution structure of rat apo-S100B(bb) as determined by NMR spectroscopy (1996) Biochemistry 35, 11577-1588.

Drohat, A.C., Baldisseri, D.M., Rustandi, R.R. and Weber, D.J. (1998) Solution structure of calcium-bound rat S100B (ββ) as determined by nuclear magnetic resonance spectroscopy. Biochemistry, 37, 2729-2740.

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, 800-809.

Drum, C.L., Yan, S.Z., Bard, J., Shen, Y.Q., Lu, D., Soelaiman, S., Grabarek, Z., Bohm, A. and Tang, W.J. (2002) Structural basis for the activation of anthrax adenylyl cyclase exotoxin by calmodulin. Nature, 415, 396-402.

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, 5296-5301.

Ebashi, S., Ebashi, F. and Kodama, A. (1967) Troponin as the Ca2+-receptive protein in the contractile system. J Biochem (Tokyo), 62, 137-138.

Falke, J.J., Drake, S.K., Hazard, A.L. and Peersen, O.B. (1994) Molecular tuning of ion binding to calcium signaling proteins. Q Rev Biophys, 27, 219-290.

Fernández Chacón, R., Shin, O., Koenigstorfer, A., Matos, M.F., Meyer, A.C., Garcia, J., Gerber, S.H., Rizo, J., Suedhof, T.C. and and Rosenmund, C. (2002) Structure/Function Analysis of Ca2+ Binding to the C2A Domain of Synaptotagmin 1. J. Neurosci., 22, 8438–8446.

Fiser, A. and Sali, A. (2003) ModLoop: automated modeling of loops in protein structures. Bioinformatics, 19, 2500-2501.

Föhr, UG, Heizmann, CW, Engelkamp, D, Schäfer, BW, Cox, JA, Purification and cation binding properties of the recombinant human S100 calcium-binding protein A3, an EF-hand motif protein with high affinity for zinc.

J Biol Chem, 1995 270, 21056–21061.

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, 425-433.

Fritz, G. and Heizmann, C.W. (2004) 3D structures of the calcium and zinc binding S100 proteins. In Messerschmidt, A., Bode, W. and Cygler, M. (eds.), Handbook of Metalloproteins. John Wiley & Sons, Inc., Chichester, Vol.

3, pp. 529-540.

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, 264-276.

Fritz, G., Mittl, P.R., Vasak, M., Grütter, 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, 33092-33098.

Garbuglia, M., Verzini, M., Hofmann, A., Huber, R. and Donato, R. (2000) S100A1 and S100B interactions with annexins. Biochim Biophys Acta, 1498, 192-206.

Glusker, J.P. (1991) Structural Aspects of Metal Liganding to Functional Groups in Proteins. Advances in Protein Chemistry, 42, 1-78(19).

-REFERENCES-

130

Green, D.W., Ingram, V.M. and Perutz, M.F. (1954) The structure of hemoglobin. IV. Sign determination by the isomorphous replacement method. Proc. Roy. Soc. A, 225, 287-307.

Griffin, W.S., Stanley, L.C., Ling, C., White, L., MacLeod, V., Perrot, L.J., White, C.L., 3rd and Araoz, C. (1989) Brain interleukin 1 and S-100 immunoreactivity are elevated in Down syndrome and Alzheimer disease.

Proc Natl Acad Sci U S A, 86, 7611-7615.

Griffin, W.S., Yeralan, O., Sheng, J.G., Boop, F.A., Mrak, R.E., Rovnaghi, C.R., Burnett, B.A., Feoktistova, A. and Van Eldik, L.J. (1995) Overexpression of the neurotrophic cytokine S100β in human temporal lobe epilepsy.

J. Neurochem., 65, 228-233.

Guex, N. and Peitsch, M.C. (1997) SWISS-MODEL and the Swiss-PdbViewer: an environment for comparative protein modeling. Electrophoresis, 18, 2714-2723.

Hagiwara, M., Ochiai, M., Owada, K., Tanaka, T. and Hidaka, H. (1988) Modulation of tyrosine phosphorylation of p36 and other substrates by the S-100 protein. J Biol Chem, 263, 6438-6441.

Harding, M.M. (2000) The geometry of metal-ligand interactions relevant to proteins. II. Angles at the metal atom, additional weak metal-donor interactions. Acta Crystallogr D Biol Crystallogr, 56 ( Pt 7), 857-867.

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, 383-397.

Heizmann, C.W., Fritz, G. and Schäfer, B.W. (2002) S100 proteins: structure, functions and pathology. Front Biosci, 7, d1356-1368.

Heizmann, C.W., Schäfer, B.W. and Fritz, G. (2003) The family of S100 Cell Signaling Proteins. In Bradshaw, E.A.

and Dennis, E.A. (eds.), Handbook of Cell Signaling. Academic Press, Boston, Vol. 2, pp. 87-93.

Hendrickson, W.A., Smith, J.L., Phizackerley, R.P. and Merrit, E.A. (1988) Crystallographic structure analysis of lamprey hemoglobin from anomalous dispersion of synchrotron radiation. Proteins, 4, 77-88.

Hilt, D.C. and Kligman, D. (1991) ''The S100 Protein Family: A Biochemical and Functional Overview''. In Heizmann, C.W. (ed.), Novel Calcium-Binding Proteins. Springer-Verlag, Berlin, pp. 63-103.

Hoeflich, K.P. and Ikura, M. (2002) Calmodulin in action: diversity in target recognition and activation mechanisms.

Cell, 108, 739-742.

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.

Hoppe, W. (1957) Die Faltmolekülmethode: eine neue Methode zur Bestimmung der Kristallstruktur bei ganz oder teilweise bekannten Molekülstrukturen. Acta Cryst., 10, 750-751.

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, 25752-25761.

Hsieh, H.L., Schafer, B.W., Weigle, B. and Heizmann, C.W. (2004) S100 protein translocation in response to extracellular S100 is mediated by receptor for advanced glycation endproducts in human endothelial cells.

Biochem Biophys Res Commun, 316, 949-959.

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, 9369-9374.

Huber, R. (1965) Die automatisierte Faltmolekülmethode. Acta Cryst., 19, 353-356.

Humphrey, W., Dalke, A. and Schulten, K. (1996) VMD: visual molecular dynamics. J. Mol. Graph., 14, 33-38, 27-38.

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, 40096-40105.

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, 2357-2365.

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, 1003-1014.

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, 15536-15542.

Johnsson, B., Lofas, S. and Lindquist, G. (1991) Immobilization of proteins to a carboxymethyldextran-modified gold surface for biospecific interaction analysis in surface plasmon resonance sensors. Anal. Biochem., 198, 268-277.

Jones, D.T. (1999) GenTHREADER: an efficient and reliable protein fold recognition method for genomic sequences.

J. Mol. Biol., 287, 797-815.

Jones, T.A., Zou, J.Y., Cowan, S.W. and Kjeldgaard. (1991a) Improved methods for building protein models in electron density maps and the location of errors in these models. Acta Crystallogr. A, 47, 110-119.

Kabsch, W. (1988) Automatic indexing of rotation diffraction patterns. J Appl. Cryst., 21, 67-71.

Kabsch, W. (1993) Automatic processing of rotation diffraction data from crystals of initially unknown symmetry and

Kabsch, W. (1993) Automatic processing of rotation diffraction data from crystals of initially unknown symmetry and