首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Relative to other parvalbumin isoforms, the mammalian beta-parvalbumin (oncomodulin) displays attenuated divalent ion affinity. High-resolution structural data for the Ca(2+)-bound protein have provided little insight into the physical basis for this behavior, prompting an examination of the unliganded state. This article describes the solution structure and peptide backbone dynamics of Ca(2+)-free rat beta-parvalbumin (beta-PV). Ca(2+) removal evidently provokes significant structural alterations. Interaction between the D helix and the AB domain in the Ca(2+)-bound protein is greatly diminished in the apo-form, permitting the D helix to straighten. There is also a significant reorganization of the hydrophobic core and a concomitant remodeling of the interface between the AB and CD-EF domains. These modifications perturb the orientation of the C and D helices, and the energetic penalty associated with their reversal could contribute to the low-affinity signature of the CD site. By contrast, Ca(2+) removal causes a comparatively minor perturbation of the E and F helices, consistent with the more typical divalent ion affinity observed for the EF site. Ca(2+)-free rat beta-PV retains structural rigidity on the picosecond-nanosecond timescale. At 20 degrees C, the majority of amide vectors show no evidence for motion on timescales above 20 ps, and the average order parameter for the entire molecule is 0.92.  相似文献   

2.
The use of competitive isothermal titration calorimetry (ITC) to measure high-affinity binding constants has been largely restricted to systems with a single binding site or multiple identical sites. This study demonstrates the extension of this approach to proteins with two nonequivalent EF-hand Ca(2+)-binding sites--rat beta parvalbumin and the S55D/E59D variant of rat alpha parvalbumin. The method involves simultaneous (global) least-squares analysis of titrations with Ca(2+), with Mg(2+), with Ca(2+) in the presence of Mg(2+), and with Ca(2+) or Mg(2+) in the presence of a competitive chelator (EDTA or EGTA). The Ca(2+) and Mg(2+) binding constants obtained for rat beta agree well with estimates obtained by flow dialysis. Although the Ca(2+) affinity of alpha S55D/E59D is too high to measure by flow dialysis, it was amenable to analysis using the ITC-based approach. The combined S55D and E59D mutations increase the Ca(2+) and Mg(2+) affinities of the mutated binding site by factors of 14 and 26, respectively. This behavior is consistent with that seen previously for the rat beta S55D variant.  相似文献   

3.
Henzl MT  Tanner JJ  Tan A 《Proteins》2011,79(3):752-764
Birds express two β-parvalbumin isoforms, parvalbumin 3 and avian thymic hormone (ATH). Parvalbumin 3 from chicken (CPV3) is identical to rat β-parvalbumin (β-PV) at 75 of 108 residues. CPV3 displays intermediate Ca(2+) affinity--higher than that of rat β-parvalbumin, but lower than that of ATH. As in rat β-PV, the attenuation of affinity is associated primarily with the CD site (residues 41-70), rather than the EF site (residues 80-108). Structural data for rat α- and β-parvalbumins suggest that divalent ion affinity is correlated with the similarity of the unliganded and Ca(2+)-bound conformations. We herein present a comparison of the solution structures of Ca(2+)-free and Ca(2+)-bound CPV3. Although the structures are generally similar, the conformations of residues 47 to 50 differ markedly in the two protein forms. These residues are located in the C helix, proximal to the CD binding loop. In response to Ca(2+) removal, F47 experiences much greater solvent accessibility. The side-chain of R48 assumes a position between the C and D helices, adjacent to R69. Significantly, I49 adopts an interior position in the unliganded protein that allows association with the side-chain of L50. Concomitantly, the realignment of F66 and F70 facilitates their interaction with I49 and reduces their contact with residues in the N-terminal AB domain. This reorganization of the hydrophobic core, although less profound, is nevertheless reminiscent of that observed in rat β-PV. The results lend further support to the idea that Ca(2+) affinity correlates with the structural similarity of the apo- and bound parvalbumin conformations.  相似文献   

4.
The gelsolin family of actin regulatory proteins is activated by Ca(2+) to sever and cap actin filaments. Gelsolin has six homologous gelsolin-like domains (G1-G6), and Ca(2+)-dependent conformational changes regulate its accessibility to actin. Caenorhabditis elegans gelsolin-like protein-1 (GSNL-1) has only four gelsolin-like domains (G1-G4) and still exhibits Ca(2+)-dependent actin filament-severing and -capping activities. We found that acidic residues (Asp-83 and Asp-84) in G1 of GSNL-1 are important for its Ca(2+) activation. These residues are conserved in GSNL-1 and gelsolin and previously implicated in actin-severing activity of the gelsolin family. We found that alanine mutations at Asp-83 and Asp-84 (D83A/D84A mutation) did not disrupt actin-severing or -capping activity. Instead, the mutants exhibited altered Ca(2+) sensitivity when compared with wild-type GSNL-1. The D83A/D84A mutation enhanced Ca(2+) sensitivity for actin severing and capping and its susceptibility to proteolytic digestion, suggesting a conformational change. Single mutations caused minimal changes in its activity, whereas Asp-83 and Asp-84 were required to stabilize Ca(2+)-free and Ca(2+)-bound conformations, respectively. On the other hand, the D83A/D84A mutation suppressed sensitivity of GSNL-1 to phosphatidylinositol 4,5-bisphosphate inhibition. The structure of an inactive form of gelsolin shows that the equivalent acidic residues are in close contact with G3, which may maintain an inactive conformation of the gelsolin family.  相似文献   

5.
Tanner JJ  Agah S  Lee YH  Henzl MT 《Biochemistry》2005,44(33):10966-10976
Simultaneous replacement of Asp-94 with serine and Gly-98 with glutamate in rat alpha-parvalbumin creates a CD-site ligand array in the context of the EF-site binding loop. Previous work has shown that, relative to the wild-type CD site, this engineered site has markedly reduced Ca(2+) affinity. Seeking an explanation for this phenomenon, we have obtained the crystal structure of the alpha D94S/G98E variant. The Ca(2+) coordination within the engineered EF site of the 94/98E variant is nearly identical to that within the CD site, suggesting that the attenuated affinity of the EF site in 94/98E is not a consequence of suboptimal coordination geometry. We have also examined the divalent ion binding properties of the alpha 94/98E variant in both Na(+)- and K(+)-containing buffers. Although the Ca(2+) and Mg(2+) affinities are higher in K(+) solution, the increases are comparable to those observed for wild-type alpha. Consistent with that finding, the apparent Na(+) stoichiometry, estimated from stability studies conducted as a function of Na(+) concentration, is 1.0 +/- 0.1, identical to that of wild-type alpha. Thus, the reduced affinity for divalent ions is evidently not the result of heightened monovalent ion competition. The thermodynamic analysis indicates that the less favorable Gibbs free energy of binding reflects a substantial enthalpic penalty. Significantly, the crystal structure reveals a steric clash between Phe-57 and the C(gamma) atom of Glu-98. The consequent displacement of Phe-57 also produces a close contact with Ser-55. Thus, steric interference may be the source of the enthalpic penalty.  相似文献   

6.
Yang W  Wilkins AL  Li S  Ye Y  Yang JJ 《Biochemistry》2005,44(23):8267-8273
The effects of Ca(2+) binding on the dynamic properties of Ca(2+)-binding proteins are important in Ca(2+) signaling. To understand the role of Ca(2+) binding, we have successfully designed a Ca(2+)-binding site in the domain 1 of rat CD2 (denoted as Ca.CD2) with the desired structure and retained function. In this study, the backbone dynamic properties of Ca.CD2 have been investigated using (15)N spin relaxation NMR spectroscopy to reveal the effect of Ca(2+) binding on the global and local dynamic properties without the complications of multiple interactive Ca(2+) binding and global conformational change. Like rat CD2 (rCD2) and human CD2 (hCD2), residues involved in the recognition of the target molecule CD48 exhibit high flexibility. Mutations N15D and N17D that introduce the Ca(2+) ligands increase the flexibility of the neighboring residues. Ca(2+)-induced local dynamic changes occur mainly at the residues proximate to the Ca(2+)-binding pocket or the residues in loop regions. The beta-strand B of Ca.CD2 that provides two Asp for the Ca(2+) undergoes an S(2) decrease upon the Ca(2+) binding, while the DE-loop that provides one Asn and one Asp undergoes an S(2) increase. Our study suggests that Ca(2+) binding has a differential effect on the rigidity of the residues depending on their flexibility and location within the secondary structure.  相似文献   

7.
LipL32 is the most abundant outer membrane protein from pathogenic Leptospira and has been shown to bind extracellular matrix (ECM) proteins as well as Ca(2+). Recent crystal structures have been obtained for the protein in the apo- and Ca(2+)-bound forms. In this work, we produced three LipL32 mutants (D163-168A, Q67A, and S247A) and evaluated their ability to interact with Ca(2+) and with ECM glycoproteins and human plasminogen. The D163-168A mutant modifies aspartate residues involved in Ca(2+) binding, whereas the other two modify residues in a cavity on the other side of the protein structure. Loss of calcium binding in the D163-D168A mutant was confirmed using intrinsic tryptophan fluorescence, circular dichroism, and thermal denaturation whereas the Q67A and S247A mutants presented the same Ca(2+) affinity as the wild-type protein. We then evaluated if Ca(2+) binding to LipL32 would be crucial for its interaction with collagen type IV and plasma proteins fibronectin and plasminogen. Surprisingly, the wild-type protein and all three mutants, including the D163-168A variant, bound to these ECM proteins with very similar affinities, both in the presence and absence of Ca(2+) ions. In conclusion, calcium binding to LipL32 may be important to stabilize the protein, but is not necessary to mediate interaction with host extracellular matrix proteins.  相似文献   

8.
In addition to binding Ca(2+), the S100 protein S100B binds Zn(2+) with relatively high affinity as confirmed using isothermal titration calorimetry (ITC; K(d) = 94 +/- 17 nM). The Zn(2+)-binding site on Ca(2+)-bound S100B was examined further using NMR spectroscopy and site-directed mutagenesis. Specifically, ITC measurements of S100B mutants (helix 1, H15A and H25A; helix 4, C84A, H85A, and H90A) were found to bind Zn(2+) with lower affinity than wild-type S100B (from 2- to >25-fold). Thus, His-15, His-25, Cys-84, His-85, and perhaps His-90 of S100B are involved in coordinating Zn(2+), which was confirmed by NMR spectroscopy. Previous studies indicate that the binding of Zn(2+) enhances calcium and target protein-binding affinities, which may contribute to its biological function. Thus, chemical shift perturbations observed here for residues in both EF-hand domains of S100B during Zn(2+) titrations could be detecting structural changes in the Ca(2+)-binding domains of S100B that are pertinent to its increase in Ca(2+)-binding affinity in the presence of Zn(2+). Furthermore, Zn(2+) binding causes helix 4 to extend by one full turn when compared to Ca(2+)-bound S100B. This change in secondary structure likely contributes to the increased binding affinity that S100B has for target peptides (i.e., TRTK peptide) in the presence of Zn(2+).  相似文献   

9.
The solution NMR structure is reported for Ca(2+)-loaded S100B bound to a 12-residue peptide, TRTK-12, from the actin capping protein CapZ (alpha1 or alpha2 subunit, residues 265-276: TRTKIDWNKILS). This peptide was discovered by Dimlich and co-workers by screening a bacteriophage random peptide display library, and it matches exactly the consensus S100B binding sequence ((K/R)(L/I)XWXXIL). As with other S100B target proteins, a calcium-dependent conformational change in S100B is required for TRTK-12 binding. The TRTK-12 peptide is an amphipathic helix (residues W7 to S12) in the S100B-TRTK complex, and helix 4 of S100B is extended by three or four residues upon peptide binding. However, helical TRTK-12 in the S100B-peptide complex is uniquely oriented when compared to the three-dimensional structures of other S100-peptide complexes. The three-dimensional structure of the S100B-TRTK peptide complex illustrates that residues in the S100B binding consensus sequence (K4, I5, W7, I10, L11) are all involved in the S100B-peptide interface, which can explain its orientation in the S100B binding pocket and its relatively high binding affinity. A comparison of the S100B-TRTK peptide structure to the structures of apo- and Ca(2+)-bound S100B illustrates that the binding site of TRTK-12 is buried in apo-S100B, but is exposed in Ca(2+)-bound S100B as necessary to bind the TRTK-12 peptide.  相似文献   

10.
The type I dockerin domain is responsible for incorporating its associated glycosyl hydrolase into the bacterial cellulosome, a multienzyme cellulolytic complex, via its interaction with a receptor domain (cohesin domain) of the cellulosomal scaffolding subunit. The highly conserved dockerin domain is characterized by two Ca(2+)-binding sites with sequence similarity to the EF-hand motif. Here, we present the three-dimensional solution structure of the 69 residue dockerin domain of Clostridium thermocellum cellobiohydrolase CelS. Torsion angle dynamics calculations utilizing a total of 728 NOE-derived distance constraints and 79 torsion angle restraints yielded an ensemble of 20 structures with an average backbone r.m.s.d. for residues 5 to 29 and 32 to 66 of 0.54 A from the mean structure. The structure consists of two Ca(2+)-binding loop-helix motifs connected by a linker; the E helices entering each loop of the classical EF-hand motif are absent from the dockerin domain. Each dockerin Ca(2+)-binding subdomain is stabilized by a cluster of buried hydrophobic side-chains. Structural comparisons reveal that, in its non-complexed state, the dockerin fold displays a dramatic departure from that of Ca(2+)-bound EF-hand domains. A putative cohesin-binding surface, comprised of conserved hydrophobic and basic residues, is proposed, providing new insight into cellulosome assembly.  相似文献   

11.
S100B(beta beta) is a dimeric Ca2+-binding protein that interacts with p53, inhibits its phosphorylation by protein kinase C (PKC) and promotes disassembly of the p53 tetramer. Likewise, a 22 residue peptide derived from the C-terminal regulatory domain of p53 has been shown to interact with S100B(beta beta) in a Ca2+-dependent manner and inhibits its phosphorylation by PKC. Hence, structural studies of Ca2+-loaded S100B(beta beta) bound to the p53 peptide were initiated to characterize this interaction. Analysis of nuclear Overhauser effect (NOE) correlations, amide proton exchange rates, 3J(NH-H alpha) coupling constants, and chemical shift index data show that, like apo- and Ca2+-bound S100B(beta beta), S100B remains a dimer in the p53 peptide complex, and each subunit has four helices (helix 1, Glu2-Arg20; helix 2, Lys29-Asn38; helix 3, Gln50-Asp61; helix 4, Phe70-Phe87), four loops (loop 1, Glu21-His25; loop 2, Glu39-Glu49; loop 3, Glu62-Gly66; loop 4, Phe88-Glu91), and two beta-strands (beta-strand 1, Lys26-Lys28; beta-strand 2, Glu67-Asp69), which forms a short antiparallel beta-sheet. However, in the presence of the p53 peptide helix 4 is longer by five residues than in apo- or Ca2+-bound S100B(beta beta). Furthermore, the amide proton exchange rates in helix 3 (K55, V56, E58, T59, L60, D61) are significantly slower than those of Ca2+-bound S100B(beta beta). Together, these observations plus intermolecular NOE correlations between the p53 peptide and S100B(beta beta) support the notion that the p53 peptide binds in a region of S100B(beta beta), which includes residues in helix 2, helix 3, loop 2, and the C-terminal loop, and that binding of the p53 peptide interacts with and induces the extension of helix 4.  相似文献   

12.
Ran X  Miao HH  Sheu FS  Yang D 《Biochemistry》2003,42(17):5143-5150
Neurogranin/RC3 is a neuron-specific, Ca(2+)-sensitive calmodulin binding protein and a specific protein kinase C substrate. Neurogranin may function to regulate calmodulin levels at specific sites in neurons through phosphorylation at serine residue within its IQ motif, oxidation outside the IQ motif, or changes in local cellular Ca(2+) concentration. To gain insight into the functional role of neurogranin in the regulation of calmodulin-dependent activities, we investigated the structure and dynamics of a full-length rat neurogranin protein with 78 amino acids using triple resonance NMR techniques. In the absence of calmodulin or PKC, neurogranin exists in an unfolded form as evidenced by high backbone mobility and the absence of long-range nuclear Overhauser effect (NOE). Analyses of the chemical shifts (13)C(alpha), (13)C(beta), and (1)H(alpha) reveal the presence of a local alpha-helical structure for the region between residues G25-A42. Three-bond (1)H(N)-(1)H(alpha) coupling constants support the finding that the sequence between residues G25 and A42 populates a non-native helical structure in the unfolded neurogranin. Homonuclear NOE results are consistent with the conclusions drawn from chemical shifts and coupling constants. (15)N relaxation data indicate motional restrictions on a nanosecond time scale in the region from D15 to S48. Spectral densities and order parameters data further confirm that the unfolded neurogranin exists in conformation with residual secondary structures. The medium mobility of the nascent helical region may help to reduce the entropy loss when neurogranin binds to its targets, but the complex between neurogranin and calmodulin is not stable enough for structural determination by NMR. Calmodulin titration of neurogranin indicates that residues D15-G52 of neurogranin undergo significant structural changes upon binding to calmodulin.  相似文献   

13.
Peptidylarginine deiminase 4 (PAD4) is a Ca(2+)-dependent enzyme that catalyzes the conversion of protein arginine residues to citrulline. Its gene is a susceptibility locus for rheumatoid arthritis. Here we present the crystal structure of Ca(2+)-free wild-type PAD4, which shows that the polypeptide chain adopts an elongated fold in which the N-terminal domain forms two immunoglobulin-like subdomains, and the C-terminal domain forms an alpha/beta propeller structure. Five Ca(2+)-binding sites, none of which adopt an EF-hand motif, were identified in the structure of a Ca(2+)-bound inactive mutant with and without bound substrate. These structural data indicate that Ca(2+) binding induces conformational changes that generate the active site cleft. Our findings identify a novel mechanism for enzyme activation by Ca(2+) ions, and are important for understanding the mechanism of protein citrullination and for developing PAD-inhibiting drugs for the treatment of rheumatoid arthritis.  相似文献   

14.
S-modulin in frog or its bovine homologue, recoverin, is a 23-kDa EF-hand Ca(2+)-binding protein found in rod photoreceptors. The Ca(2+)-bound form of S-modulin binds to rhodopsin kinase (Rk) and inhibits its activity. Through this regulation, S-modulin is thought to modulate the light sensitivity of a rod. In the present study, we tried to identify the interaction site of the Ca(2+)-bound form of S-modulin to Rk. First, we mapped roughly the interaction regions by using partial peptides of S-modulin. The result suggested that a specific region near the amino terminus is the interaction site of S-modulin. We then identified the essential amino acid residues in this region by using S-modulin mutant proteins: four amino acid residues (Phe(22), Glu(26), Phe(55), and Thr(92)) were suggested to interact with Rk. These residues are located in a small closed pocket in the Ca(2+)-free, inactive form of S-modulin, but exposed to the surface of the molecule in the Ca(2+)-bound, active form of S-modulin. Two additional amino acid residues (Tyr(108) and Arg(150)) were found to be crucial for the Ca(2+)-dependent conformational changes of S-modulin.  相似文献   

15.
16.
Calcyclin is a homodimeric protein belonging to the S100 subfamily of EF-hand Ca(2+)-binding proteins, which function in Ca(2+) signal transduction processes. A refined high-resolution solution structure of Ca(2+)-bound rabbit calcyclin has been determined by heteronuclear solution NMR. In order to understand the Ca(2+)-induced structural changes in S100 proteins, in-depth comparative structural analyses were used to compare the apo and Ca(2+)-bound states of calcyclin, the closely related S100B, and the prototypical Ca(2+)-sensor protein calmodulin. Upon Ca(2+) binding, the position and orientation of helix III in the second EF-hand is altered, whereas the rest of the protein, including the dimer interface, remains virtually unchanged. This Ca(2+)-induced structural change is much less drastic than the "opening" of the globular EF-hand domains that occurs in classical Ca(2+) sensors, such as calmodulin. Using homology models of calcyclin based on S100B, a binding site in calcyclin has been proposed for the N-terminal domain of annexin XI and the C-terminal domain of the neuronal calcyclin-binding protein. The structural basis for the specificity of S100 proteins is discussed in terms of the variation in sequence of critical contact residues in the common S100 target-binding site.  相似文献   

17.
Wilder PT  Varney KM  Weiss MB  Gitti RK  Weber DJ 《Biochemistry》2005,44(15):5690-5702
The EF-hand calcium-binding protein S100B also binds one zinc ion per subunit with a relatively high affinity (K(d) approximately 90 nM) [Wilder et al., (2003) Biochemistry 42, 13410-13421]. In this study, the structural characterization of zinc binding to calcium-loaded S100B was examined using high-resolution NMR techniques, including structural characterization of this complex in solution at atomic resolution. As with other S100 protein structures, the quaternary structure of Zn(2+)-Ca(2+)-bound S100B was found to be dimeric with helices H1, H1', H4, and H4' forming an X-type four-helix bundle at the dimer interface. NMR data together with mutational analyses are consistent with Zn(2+) coordination arising from His-15 and His-25 of one S100B subunit and from His-85 and Glu-89 of the other subunit. The addition of Zn(2+) was also found to extend helices H4 and H4' three to four residues similar to what was previously observed with the binding of target proteins to S100B. Furthermore, a kink in helix 4 was observed in Zn(2+)-Ca(2+)-bound S100B that is not in Ca(2+)-bound S100B. These structural changes upon Zn(2+)-binding could explain the 5-fold increase in affinity that Zn(2+)-Ca(2+)-bound S100B has for peptide targets such as the TRTK peptide versus Ca(2+)-bound S100B. There are also changes in the relative positioning of the two EF-hand calcium-binding domains and the respective helices comprising these EF-hands. Changes in conformation such as these could contribute to the order of magnitude higher affinity that S100B has for calcium in the presence of Zn(2+).  相似文献   

18.
The Ca(2+) titration of the (15)N-labeled mutant V136G calmodulin has been monitored using (1)H-(15)N HSQC NMR spectra. Up to a [Ca(2+)]/[CaM] ratio of 2, the Ca(2+) ions bind predominantly to sites I and II on the N-domain in contrast with the behavior of the wild-type calmodulin where the C-terminal domain has the higher affinity for Ca(2+). Surprisingly, the Ca(2+)-binding affinity for the N-domain in the mutant calmodulin is greater than that for the N-domain in the wild-type protein. The mutated C-domain is observed as a mixture of unfolded, partially folded (site III occupied), and native-like folded (sites III and IV occupied) conformations, with relative populations dependent on the [Ca(2+)]/[CaM] ratio. The occupancy of site III independently of site IV in this mutant shows that the cooperativity of Ca(2+) binding in the C-domain is mediated by the integrity of the domain structure. Several NH signals from residues in the Ca(2+)-bound N-domain appear as two signals during the Ca(2+) titration indicating separate species in slow exchange, and it can be deduced that these result from the presence and absence of interdomain interactions in the mutant. It is proposed that an unfolded part of the mutated C-domain interacts with sites on the N-domain that normally bind to target proteins. This would also account for the increase in the Ca(2+) affinity for the N-domain in the mutant compared with the wild-type calmodulin. The results therefore show the wide-ranging effects of a point mutation in a single Ca(2+)-binding site, providing details of the involvement of individual residues in the calcium-induced folding reactions.  相似文献   

19.
Relative to the same site in oncomodulin, the CD ion-binding domain of rat parvalbumin exhibits much greater affinity for Ca2+ and Mg2+. As part of an effort to understand the structural basis for these differences, site-specific variants of oncomodulin have been prepared in which the amino acid residues at positions 52, 54, 57, 59, and 60 have been replaced with the residues present at the corresponding positions in rat parvalbumin. The proteins resulting from the single-site substitutions at residues 52, 54, and 57 are indistinguishable from the wild-type protein on the basis Eu3+ luminescence spectroscopy, and none of the three variants displays increased affinity for Ca2+. By contrast, the substitutions at residues 59 and 60 perturb both the Eu3+ luminescence parameters and the Ca2+ and Mg2+ affinities, and these differences are amplified when both replacements are simultaneously incorporated into the protein. The Eu3+ 7F0----5D0 spectrum of the double variant (D59E/G60E) at pH 5.0, with a maximum at 5796 A and pronounced shoulder at 5791 A, strongly resembles that obtained with pike parvalbumin. Consistent with this increased parvalbumin-like character, KCa is decreased from 0.78 microM (for the wild-type protein) to 0.41 microM, and KMg is decreased from 3.5 to 0.74 mM. Nevertheless, the affinity of the CD ion-binding domain in D59E/G60E for Ca2+ remains almost 2 orders of magnitude lower than the corresponding site in rat parvalbumin, strongly suggesting that residues besides those present in the binding loop are involved in dictating the metal ion-binding properties of the oncomodulin CD site.  相似文献   

20.
S100A4 takes part in control of tumour cell migration and contributes to metastatic spread in in vivo models. In the active dimeric Ca(2+)-bound state it interacts with multiple intracellular targets. Conversely, oligomeric forms of S100A4 are linked with the extracellular function of this protein. We report the 1.5A X-ray crystal structure of Ca(2+)-bound S100A4 and use it to identify the residues involved in target recognition and to derive a model of the oligomeric state. We applied stopped-flow analysis of tyrosine fluorescence to derive kinetics of S100A4 activation by Ca(2+) (k(on)=3.5 microM(-1)s(-1), k(off)=20s(-1)).  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号