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 31 毫秒
1.
Using site-directed mutagenesis we have expressed in Escherichia coli three engineered calmodulins (CaM) containing deletions in the solvent-exposed region of the central helix. These are CaM delta 84, Glu-84 removed; CaM delta 83-84, Glu-83 and Glu-84 removed; and CaM delta 81-84, Ser-81 through Glu-84 removed. The abilities of these proteins to activate skeletal muscle myosin light chain kinase, plant NAD kinase, and bovine brain calcineurin activities were determined, as were their abilities to bind a synthetic peptide based on the calmodulin-binding domain of skeletal muscle myosin light chain kinase. Similar results were obtained with all three deletion proteins. Vm values for enzymes activated by the deletion proteins are all within 10-20% of those values obtained with bacterial control calmodulin. Relative to bacterial control values, changes in Kact or Kd values associated with the deletions are all less than an order of magnitude: Kact values for NAD kinase and myosin light chain kinase are increased 5-7-fold, Kd values for binding of the synthetic peptide are increased 4-7-fold, and Kact values for calcineurin are increased only 1-3-fold. In assays of NAD kinase and myosin light chain kinase activation some differences between bovine calmodulin and bacterial control calmodulin were observed. With NAD kinase, Kact values for the bacterial control protein are increased 4-fold relative to values for bovine calmodulin, and Vm values are increased by 50%; with myosin light chain kinase, Kact values are increased 2-fold and Vm values are decreased 10-15% relative to those values obtained with bovine calmodulin. These differences between bacterial control and bovine calmodulins probably can be attributed to known differences in postranslational processing of calmodulin in bacterial and eucaryotic cells. No differences between bovine and control calmodulins were observed in assays of calcineurin activation or peptide binding. Our observations indicate that contacts with the deleted residues, Ser-81 through Glu-84, are not critical in the calmodulin-target complexes we have evaluated. Formation of these calmodulin-target complexes also does not appear to be greatly affected by the global alterations in the structure of calmodulin that are associated with the deletions. In models in which the central helix is maintained in the altered calmodulins, each deleted residue causes the two lobes of calmodulin to be twisted 100 degrees relative to one another and brought 1.5 A closer together.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

2.
A gene coding for a calmodulin was synthesized and expressed in Escherichia coli. The gene was produced by the enzymatic ligation of 61 chemically synthesized DNA fragments. The gene possesses 27 unique, regularly spaced, restriction endonuclease cleavage sites to facilitate gene mutagenesis by the replacement of specific gene segments with synthetic double-stranded DNA. An expression vector containing the calmodulin gene was used to transform E. coli. Purification and characterization of calmodulin (VU-1 calmodulin) expressed by these transformants showed that it lacks two posttranslational modifications: an amino-terminal blocking group and N epsilon, N epsilon, N epsilon-trimethyllysine at position 115. The cyclic nucleotide phosphodiesterase activator properties of VU-1, higher plant, and vertebrate calmodulins were not statistically different. However, VU-1 calmodulin was found to activate nicotinamide adenine dinucleotide (NAD) kinase to a maximal level that was at least 3-fold higher than that found with higher plant and vertebrate calmodulins. This higher level of activation is also characteristic of calmodulins from Dictyostelium discoideum and Chlamydomonas reinhardtii [Roberts, D. M., Burgess, W. H., & Watterson, D. M. (1984) Plant Physiol. 75, 796-798; Marshak, D. R., Clarke, M., Roberts, D. M., & Watterson, D. M. (1984) Biochemistry 23, 2891-2899]. The only common feature among Dictyostelium, Chlamydomonas, and VU-1 calmodulins not found in higher plant and vertebrate calmodulins is an unmethylated lysine at position 115. The results indicate that the lack of methylation of lysine-115 may contribute to the maximal level of NAD kinase activation.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
A site-directed mutagenesis study of yeast calmodulin   总被引:2,自引:0,他引:2  
A site-directed mutagenesis study was carried out in order to understand the regulatory mechanism of calmodulin. We started from the yeast (Saccharomyces cerevisiae) calmodulin gene since it has many differences in amino acid sequence and inferior functional properties compared with the vertebrate calmodulin. Recombinant yeast calmodulins were generated in Escherichia coli transformed by constructed expression plasmids. Three recombinant calmodulins were obtained. The first two were YCM61G, in which the Ca2(+)-binding site 2 (the four Ca2(+)-binding EF-hand structures in calmodulin were numbered from the N-terminus) was converted to the same as that in vertebrate calmodulin, and YCM delta 132-148, in which the C-terminal half sequence of site 4 was deleted. These two recombinant calmodulins had the same maximum Ca2+ binding (3 mol/mol) as yeast calmodulin, which indicates that site 4 of yeast calmodulin was the one losing Ca2+ binding capacity. YCM delta 132-148 could not activate target enzymes, whereas its Ca2+ binding profile was similar to those of yeast calmodulin and YCM61G. Therefore, the structure in site 4 which cannot bind Ca2+ is indispensable for the regulatory function of yeast calmodulin. The complete regulatory function of vertebrate calmodulin can be attained by the combination of 4 Ca2+ binding structures. The negative charge cluster in the central alpha-helix region is suggested to stabilize the active conformation of calmodulin, since the third yeast calmodulin mutant, YCM83E, which had the negative charge cluster, increased the maximum activation of myosin light chain kinase.  相似文献   

4.
Calmodulin's calculated electrostatic potential surface is asymmetrically distributed about the molecule. Concentrations of uncompensated negative charge are localized near certain alpha-helices and calcium-binding loops. Further calculations suggest that these charge features of calmodulin can be selectively perturbed by changing clusters of phylogenetically conserved acidic amino acids in helices to lysines. When these cluster charge reversals are actually produced by using cassette-based site-specific mutagenesis of residues 82-84 or 118-120, the resulting proteins differ in their interaction with two distinct calmodulin-dependent protein kinases, myosin light chain kinase and calmodulin-dependent protein kinase II. Each calmodulin mutant can be purified to apparent chemical homogeneity by an identical purification protocol that is based on conservation of its overall properties, including calcium binding. Although cluster charge reversals result in localized perturbations of the computed negative surface, single amino acid changes would not be expected to alter significantly the distribution of the negative surface because of the relatively high density of uncompensated negative charge in the region around residues 82-84 and 118-120. However, this does not preclude the possibility of single amino acid charge perturbations having a functional effect on the more intimate, catalytically active complex. The electrostatic surface of calmodulin described in this report may be a feature that would be altered only by cluster charge reversal mutations. Overall, the results suggest that the charge properties of calmodulin are one of several properties that are important for the efficient assembly of calmodulin-protein kinase signal transduction complexes in eukaryotic cells.  相似文献   

5.
The crystal structure of calmodulin (Mr 16,700, 148 residues) from Drosophila melanogaster as expressed in a bacterial system has been determined and refined at 2.2-A resolution. Starting with the structure of mammalian calmodulin, we produced an extensively refitted and refined model with a conventional crystallographic R value of 0.197 for the 5,239 reflections (F greater than or equal to 2 sigma (F)) within the 10.0-2.2-A resolution range. The model includes 1,164 protein atoms, 4 calcium ions, and 78 water molecules and has root mean square deviations from standard values of 0.018 A for bond lengths and 0.043 A for angle distances. The overall structure is similar to mammalian calmodulin, with a seven-turn central helix connecting the two calcium-binding domains. The "dumb-bell" shaped molecule contains seven alpha-helices and four "EF hand" calcium-binding sites. Although the amino acid sequences of mammalian and Drosophila calmodulins differ by only three conservative amino acid changes, the refined model reveals a number of significant differences between the two structures. Superimposition of the structures yields a root mean square deviation of 1.22 A for the 1,120 equivalent atoms. The calcium-binding domains have a root mean square deviation of 0.85 A for the 353 equivalent atoms. There are also differences in the amino terminus, the bend of the central alpha-helix, and the orientations of some of the side chains.  相似文献   

6.
Recent evidence proposes that the calcium-binding protein, calmodulin, plays a crucial role in the regulation or modulation of the calcium-dependent potassium conductance in Paramecium tetraurelia (Hinrichsen, R.D., Burgess-Cassler, A., Soltvedt, B.C., Hennessey, T. and Kung, C. (1986) Science 323, 503-506). We purified the calmodulins from both the wild type and pantophobiac A (a mutant lacking the above-mentioned conductance and whose phenotypic defect is traceable to its calmodulin) by hydrophobic interaction and immunoaffinity chromatographies, and examined them biochemically. In this paper we address the preliminary characterization of the two calmodulins and discuss the consequences of the genetic alteration. The differences described here are in their electrophoretic mobilities in polyacrylamide gel electrophoresis and in their binding characteristics to monoclonal antibodies raised against calmodulin from wild-type paramecia. Also, we present data which indicate a difference in the stimulation of the calmodulin-dependent enzyme bovine brain phosphodiesterase under certain conditions.  相似文献   

7.
We have studied the calcium-binding properties of two high affinity calcium-binding proteins from squid optic lobes: one, squid calmodulin (SCaM), similar to bovine brain calmodulin (BCaM), the other, squid calcium-binding protein (SCaBP), distinct (Head, J.F., Spielberg, S., and Kaminer, B. (1983) Biochem J. 209, 797-802). Equilibrium dialysis measurements on the squid proteins (and BCaM) were made at 100 mM KCl in the presence and absence of 3 mM Mg2+, and at 400 mM KCl in the presence of 3 mM Mg2+, which more closely resembles the conditions in the squid. SCaM, SCaBP, and BCaM each bind a maximum of 4 Ca2+ ions/molecule of protein under the ionic conditions tested. SCaBP has a higher affinity than SCaM or BCaM for Ca2+ at 100 mM KCl in the absence of Mg2+. However, in the presence of Mg2+, half-maximal binding to SCaBP occurs at a similar pCa value to that observed with calmodulin. Increasing the KCl concentration reduces the affinity of all three proteins for Ca2+. UV absorption measurements showed that the binding of 4 Ca2+ ions/molecule is necessary to complete spectral changes in SCaBP, compared to two for the calmodulins. While Ca2+ causes perturbations in aromatic chromophores in SCaM and SCaBP, Mg2+ causes a significant perturbation only in SCaBP. These Mg2+-induced changes differ qualitatively from those induced by Ca2+.  相似文献   

8.
In the preceding paper (Lukas, Iverson, Schleicher, Watterson 1984 Plant Physiol 75: 788-795), we reported that the amino acid sequence of spinach calmodulin has at least 13 amino acid sequence differences from vertebrate calmodulin. In the present study, we investigated the effect of these amino acid sequence substitutions on the enzyme activator properties of vertebrate and plant calmodulins. Calmodulins from spinach and the green alga Chlamydomonas reinhardtii activate chicken gizzard myosin light chain kinase in a manner similar but not identical to chicken calmodulin. In contrast, these calmodulins have very different NAD kinase activator properties. The concentration required for half-maximal activation of pea seedling NAD kinase by spinach calmodulin (3-4 nanomolar) is lower than the corresponding concentrations of chicken (20 nanomolar) and Chlamydomonas (40 nanomolar) calmodulins. However, the maximum level of activation obtained with Chlamydomonas calmodulin is 4- to 6-fold higher than spinach or chicken calmodulin. These data indicate that the limited structural heterogeneity among calmodulins have differential effects on their biochemical activities.  相似文献   

9.
The flip-flop model is a mechanistic model proposed to describe how calmodulin activates enzymes. One prediction based upon this model is that calmodulin-activated enzymes would contain a calmodulin-like binding site which, among other attributes, would bind the peptide melittin. Five purified calmodulin-activated enzymes, namely calcineurin, myosin light chain kinase, phosphorylase b kinase, phosphodiesterase, and NAD kinase, were all found to bind biotinylated melittin and to also bind an antimelittin antibody and biotinylated calmodulins. Using gel blots of crude tissue extracts (rat brain and Arabidopsis), most proteins did not bind any of the probes and thus do not have these characteristics. However, among those which bind any of these probes, a strong correlation was found between those proteins which bind biotinylated calmodulins and those which bind melittin and antimelittin. Gel blots of phosphorylase b kinase demonstrate that the alpha, beta, and gamma subunits all bind calmodulin and melittin. A putative calmodulin-like binding site sequence was identified in eight enzymes or subunits which may play an important role in both melittin binding and calmodulin-dependent regulation of these enzymes.  相似文献   

10.
An engineered calmodulin differs from vertebrate calmodulin in its ability to activate Bordetella pertussis adenylate cyclase, and this difference has been utilized as the basis for a new purification protocol for the adenylate cyclase. VU-8 calmodulin, in which 3 glutamic acid residues (residues 82-84) have been substituted with 3 lysine residues, has a 1000-fold lower apparent affinity for the adenylate cyclase, compared to vertebrate calmodulin, and decreased maximal activity. Because of the relatively calcium-independent nature of the interaction between calmodulin and the cyclase, the use of calmodulin-Sepharose conjugates in the purification of the cyclase requires the use of chaotropic agents for elution. However, when immobilized VU-8 calmodulin was tested as a calcium-dependent, affinity-based, adsorption chromatography step in the purification of the cyclase from culture media or bacterial extracts, the enzyme bound to the column in a calcium-dependent manner, and a nearly homogeneous enzyme was obtained in high yield. These results demonstrate the feasibility of using engineered calmodulins that have selective differences in activity for the rational design of rapid purification protocols for calmodulin-binding proteins as well as indicate the importance of the conserved negative charge cluster at residues 82-84 of calmodulin for activation of this cyclase.  相似文献   

11.
S100b is a calcium-binding protein that will bind to many calmodulin target molecules in a Ca2+-dependent manner. In order to study the Ca2+-dependent binding properties of S100b, its interaction with a calmodulin antagonist, trifluoperazine (TFP), was investigated using [19F]- and [1H]-NMR and UV-difference spectroscopy. It was estimated from [19F]-NMR that in the absence of Ca2+, thek 1/2 value of TFP was 130 µM, while itsk 1/2 value decreased to 28 µM in the presence of Ca2+. The addition of KCl was not antagonistic to the Ca2+-dependent interaction of TFP to S100b. The chemical exchange rate of TFP with Ca2+-bound S100b was estimated to be 9×102 sec–1. By comparison with TFP-calmodulin exchange rates, it is suggested that the TFP-binding site on S100b is structurally different from its binding sites on calmodulin. Proton NMR resonance broadening in the range 6.8–7.2 ppm, corresponding to phenylalanine nuclei of S100b, indicates that these residues may be involved in TFP binding. Addition of Ca2+ to a 1:1 mixture of S100b and TFP resulted in a red-shifted UV-difference spectrum, while no significant difference spectrum was detected when Mg2+ was added to a S100b-TFP solution. Thus, we suggest that Ca2+ induces the exposure of a hydrophobic domain on S100b containing one or more phenylalanine residues that will bind TFP but that this domain is different from the hydrophobic domain on calmodulin.  相似文献   

12.
Calmodulin has been isolated and characterized from the gill of the bay scallop aequipecten irradians. Quantitative electrophoretic analysis of epithelial cell fractions show most of the calmodulin to be localized in the cilia, specifically in the detergent- solubilized membrane-matrix fraction. Calmodulin represents 2.2 +/- 0.3 percent of the membrane-matrix protein or 0.41 +/- 0.5 percent of the total ciliary protein. Its concentration is at least 10(-4) M if distributed uniformly within the matrix. Extraction in the presence of calcium suggests that the calmodulin is not bound to the axoneme proper. The ciliary protein is identified as a calmodulin on the basis of its calcium- dependent binding to a fluphenazine-sepharose affinity column and its comigration with bovine brain calmodulin on alkaline-urea and SDS polyacrylamide gels in both the presence and absence of calcium. Scallop ciliary calmodulin activates bovine brain phosphodiesterase to the same extent as bovine brain and chicken gizzard calmodulins. Containing trimethyllysine and lacking cysteine and tryptophan, the amino acid composition of gill calmodulin is typical of known calmodulins, except that it is relatively high in serine and low in methionine. Its composition is less acidic than other calmodulins, in agreement with an observed isoelectric point approximately 0.2 units higher than that of bovine brain. Comparative tryptic peptide mapping of scallop gill ciliary and bovine brain calmodulins indicates coincidence of over 75 percent of the major peptides, but at least two major peptides in each show no near-equivalency. Preliminary results using ATP-reactivated gill cell models show no effect of calcium at micromolar levels on ciliary beat or directionality of the lateral cilia, the cilia which constitute the vast majority of those isolated. However, ciliary arrest will occur at calcium levels more than 150 muM. Because calmodulin usually functions in the micromolar range, its role in this system is unclear. Scallop gill ciliary calmodulin may be involved in the direct regulation of dyneintubule sliding, or it may serve some coupled calcium transport function. At the concentration in which it is found, it must also at least act as a calcium buffer.  相似文献   

13.
S100b is a calcium-binding protein that will bind to many calmodulin target molecules in a Ca2+-dependent manner. In order to study the Ca2+-dependent binding properties of S100b, its interaction with a calmodulin antagonist, trifluoperazine (TFP), was investigated using [19F]- and [1H]-NMR and UV-difference spectroscopy. It was estimated from [19F]-NMR that in the absence of Ca2+, thek 1/2 value of TFP was 130 µM, while itsk 1/2 value decreased to 28 µM in the presence of Ca2+. The addition of KCl was not antagonistic to the Ca2+-dependent interaction of TFP to S100b. The chemical exchange rate of TFP with Ca2+-bound S100b was estimated to be 9×102 sec?1. By comparison with TFP-calmodulin exchange rates, it is suggested that the TFP-binding site on S100b is structurally different from its binding sites on calmodulin. Proton NMR resonance broadening in the range 6.8–7.2 ppm, corresponding to phenylalanine nuclei of S100b, indicates that these residues may be involved in TFP binding. Addition of Ca2+ to a 1:1 mixture of S100b and TFP resulted in a red-shifted UV-difference spectrum, while no significant difference spectrum was detected when Mg2+ was added to a S100b-TFP solution. Thus, we suggest that Ca2+ induces the exposure of a hydrophobic domain on S100b containing one or more phenylalanine residues that will bind TFP but that this domain is different from the hydrophobic domain on calmodulin.  相似文献   

14.
Murine cDNA that encodes neuromodulin, a neurospecific calmodulin binding protein, was inserted into the plasmid pKK223-3 for expression in Escherichia coli. After being transformed into E. coli strain SG20252 (lon-), the expression vector directed the synthesis of a protein that was recognized by polyclonal antibodies raised against bovine neuromodulin. The recombinant protein expressed in E. coli was found to be tightly associated with insoluble cell material and was extractable only with guanidine hydrochloride or sodium dodecyl sulfate. Following solubilization with guanidine hydrochloride, the protein was purified to apparent homogeneity by a single CaM-Sepharose affinity column step with a yield of 0.2 mg of protein/L of E. coli culture. The availability of the purified recombinant neuromodulin made it possible to answer several specific questions concerning the structure and function of the protein. Despite the fact that murine neuromodulin is 12 amino acid residues shorter than the bovine protein and the recombinant protein expressed in E. coli may lack any posttranslational modifications, the two proteins displayed similar biochemical properties in almost all respects examined. They both had higher affinity for CaM-Sepharose in the absence of Ca2+ than in its presence; they were both phosphorylated in vitro by protein kinase C in a Ca2+- and phospholipid-dependent manner; neither form of the proteins was autophosphorylated, and the phosphorylated form of the proteins did not bind calmodulin. The recombinant neuromodulin and neuromodulin purified from bovine brain had similar, but not identical, affinities of calmodulin, indicating that the palmitylation of the protein that occurs in animal cells is not crucial for calmodulin interactions.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
Xenopus membrane skeleton protein 4.1 is expressed constitutively during embryonic development and accumulates to high levels within the retina during normal morphogenesis. There exists a high degree of amino acid identity between Xenopus protein 4.1 and human protein 4.1, suggesting that the mechanisms known to modulate the function(s) of human protein 4.1 may also serve to regulate Xenopus protein 4.1. Calmodulin (CaM) is one regulatory protein known to affect membrane-cytoskeletal interactions. An in vitro binding assay was used to test the ability of Xenopus protein 4.1 to bind CaM. Two independent approaches, involving protein 4.1 synthesized in vitro from synthetic RNA or a partial length protein 4.1 fusion protein expressed in Escherichia coli, demonstrate calcium-dependent, CaM binding. Both approaches demonstrate that the CaM-binding site is within the amino-terminal region of Xenopus protein 4.1. Results of this calmodulin binding activity suggest a possible regulatory mechanism by which calcium and calmodulin may affect the function of protein 4.1 during development.  相似文献   

16.
Calmodulin is phosphorylated by casein kinase II on Thr-79, Ser-81, Ser-101 and Thr-117. To determine the consensus sequences for casein kinase II in intact calmodulin, we examined casein kinase II-mediated phosphorylation of engineered calmodulins with 1-4 deletions in the central helical region (positions 81-84). Total casein kinase II-catalyzed phosphate incorporation into all deleted calmodulins was similar to control calmodulin. Neither CaM delta 84 (Glu-84 deleted) nor CaM delta 81-84 (Ser-81 to Glu-84 deleted) has phosphate incorporated into Thr-79 or Ser-81, but both exhibit increased phosphorylation of residues Ser-101 and Thr-117. These data suggest that phosphoserine in the +2 position may be a specificity determinant for casein kinase II in intact proteins and/or secondary structures are important in substrate recognition by casein kinase II.  相似文献   

17.
Ophiobolin A, a fungal toxin that affects maize and rice, has previously been shown to inhibit calmodulin by reacting with the lysine (Lys) residues in the calmodulin. In the present study we mutated Lys-75, Lys-77, and Lys-148 in the calmodulin molecule by site-directed mutagenesis, either by deleting them or by changing them to glutamine or arginine. We found that each of these three Lys residues could bind one molecule of ophiobolin A. Normally, only Lys-75 and Lys-148 bind ophiobolin A. Lys-77 seemed to be blocked by the binding of ophiobolin A to Lys-75. Lys-75 is the primary binding site and is responsible for all of the inhibition of ophiobolin A. When Lys-75 was removed, Lys-77 could then react with ophiobolin A to produce inhibition. Lys-148 was shown to be a binding site but not an inhibition site. The Lys-75 mutants were partially resistant to ophiobolin A. When both Lys 75 and Lys-77 or all three Lys residues were mutated, the resulting calmodulins were very resistant to ophiobolin A. Furthermore, Lys residues added in positions 86 and/or 143 (which are highly conserved in plant calmodulins) did not react with ophiobolin A. None of the mutations seemed to affect the properties of calmodulin. These results show that ophiobolin A reacts quite specifically with calmodulin.  相似文献   

18.
A Tripathy  L Xu  G Mann    G Meissner 《Biophysical journal》1995,69(1):106-119
The calmodulin-binding properties of the rabbit skeletal muscle Ca2+ release channel (ryanodine receptor) and the channel's regulation by calmodulin were determined at < or = 0.1 microM and micromolar to millimolar Ca2+ concentrations. [125I]Calmodulin and [3H]ryanodine binding to sarcoplasmic reticulum (SR) vesicles and purified Ca2+ release channel preparations indicated that the large (2200 kDa) Ca2+ release channel complex binds with high affinity (KD = 5-25 nM) 16 calmodulins at < or = 0.1 microM Ca2+ and 4 calmodulins at 100 microM Ca2+. Calmodulin-binding affinity to the channel showed a broad maximum at pH 6.8 and was highest at 0.15 M KCl at both < or = 0.1 MicroM and 100 microM Ca2+. Under condition closely related to those during muscle contraction and relaxation, the half-times of calmodulin dissociation and binding were 50 +/- 20 s and 30 +/- 10 min, respectively. SR vesicle-45Ca2+ flux, single-channel, and [3H]ryanodine bind measurements showed that, at < or = 0.2 microM Ca2+, calmodulin activated the Ca2+ release channel severalfold. Ar micromolar to millimolar Ca2+ concentrations, calmodulin inhibited the Ca(2+)-activated channel severalfold. Hill coefficients of approximately 1.3 suggested no or only weak cooperative activation and inhibition of Ca2+ release channel activity by calmodulin. These results suggest a role for calmodulin in modulating SR Ca2+ release in skeletal muscle at both resting and elevated Ca2+ concentrations.  相似文献   

19.
Studies of ligand binding to arrestin   总被引:1,自引:0,他引:1  
A striking homology is observed between the regions 70-83 and 361-374 of the sequence of bovine arrestin and the calcium-binding loops of calmodulin and troponin C. However, the predicted alpha-helices flanking the calcium-binding site in calmodulin and troponin C are not present in arrestin. Direct measurements therefore were made in order to assess whether arrestin can bind calcium. We found that arrestin does not bind Ca2+ at physiological ionic strength, as determined by equilibrium dialysis, gel filtration, and fluorescence spectroscopy. Rapid and quantitative precipitation of arrestin occurs with Tb3+. The precipitation is reversed by EDTA and blocked by Mg2+ but not by Ca2+. Prompted by several reports, we also investigated whether nucleotides bind to arrestin. Neither ATP nor GTP binds under the conditions tested. Binding of arrestin to photolyzed, phosphorylated rhodopsin also does not influence the binding of calcium or nucleotides.  相似文献   

20.
1. Calmodulin-like proteins were purified from the fruiting bodies of higher (basidiomycete) fungi and barley (Hordeum sp.) shoots. 2. These calmodulins have electrophoretic mobilities on 10% (w/v) polyacrylamide gels at pH 8.3 in the presence of 6 M-urea and at pH 8.3 in the presence of 0.1% sodium dodecyl sulphate similar to that of bovine brain calmodulin. They interacted with rabbit skeletal-muscle troponin I in the presence of Ca2+. 3. Barley and fungal calmodulins activated myosin light-chain kinase and phosphodiesterase in the presence of Ca2+, although the amounts needed were at least an order of magnitude greater than is required to produce the same effect with mammalian calmodulin. 4. Amino acid analyses indicated a number of differences from the mammalian protein, most notably the absence of trimethyl-lysine. 5. By using 125I-labelled calmodulin, a small amount of calmodulin-binding protein was detected in homogenates of barley and fungi. 6. No protein corresponding to calmodulin could be found in Escherichia coli or yeast, although a relatively high concentration of a protein that bound calmodulin was detected in E. coli by this technique.  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