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 0 毫秒
1.
The interaction of calmodulin with amphiphilic peptides   总被引:16,自引:0,他引:16  
Calmodulin has recently been shown to form exceptionally tight, calcium-dependent complexes with several natural peptides (Kdiss greater than 10(-7) M). These peptides were demonstrated to be capable of forming basic, amphiphilic alpha-helices. To further illustrate the importance of this structural feature for calmodulin binding, several other amphiphilic alpha-helical peptides were tested for their ability to bind calmodulin. To monitor complexes of high affinity (greater than 10(8) M-1), a new competition assay was devised with Sepharose 4B-conjugated melittin. Stoichiometries were assessed by electrophoresis and equilibrium size exclusion chromatography. Three peptides, which were designed to form idealized amphiphilic alpha-helices were tested. The basic peptides, N alpha-9-fluorenylmethoxycarboxyl-(FMOC)-(Leu-Lys-Lys-Leu-Leu-Lys-L eu)1 and FMOC-(Leu-Lys-Lys-Leu-Leu-Lys-Leu)2 bind calmodulin in a 1:1 complex with dissociation constants of 150 and 3 nM, respectively. The acidic peptide, FMOC-(Leu-Glu-Glu-Leu-Leu-Glu-Leu)2 failed to bind calmodulin, even at micromolar concentrations. Complex formation between calmodulin and the 14-residue basic peptide leads to an increase in the helicity of the complex which is attributed to an increase of about 50% in the helicity of the peptide. Calmodulin also interacts with the neutral alpha-helical peptide toxin delta-hemolysin. Concomitant with binding, the fluorescence maximum of the unique Trp residue increases 2-fold and is blue-shifted. A dissociation constant could not be unambiguously estimated though, since delta-hemolysin has a strong tendency to self-aggregate. The above data support our hypothesis that a basic, amphiphilic alpha-helix is a structural feature which underlies the calmodulin-binding properties common to a variety of peptides.  相似文献   

2.
Mouse J chain was isolated from an IgM-producing hybridoma by gel filtration and ion-exchange chromatography. The sequence of the amino-terminal 25 residues was determined. At these positions, the results agree with the amino acid sequence deduced from the cDNA sequence determined previously by Koshland and co-workers and indicate that a leader sequence terminating in glycine is removed to form the mature J chain. Tryptic peptides of J chain were isolated by high pressure liquid chromatography and their amino acid compositions were compared with those expected from the cDNA sequence. The amino acid sequence of the carboxy-terminal peptide and a mixture of two other peptides was determined. The results were consistent with the cDNA sequence except that we found valine, not leucine, at position 67, and arginine, not glycine, at position 117. The presence of aspartic acid at the carboxy-terminus, as predicted from the cDNA, indicates that processing does not occur at this end of the polypeptide chain. Upon amino acid analysis, glucosamine was found in tryptic peptides 47-57 and 47-58. J chain was also cleaved at aspartylproline bonds with formic acid and the unfractionated digest was subjected to automated Edman degradation. The mixed sequence was consistent with the sequence deduced from the cDNA at positions 1 to 13, 28 to 40, 52 to 64, and 73 to 85. In conjunction with the results obtained previously by analysis of cDNA, these data show that mouse J chain is a polypeptide containing 137 amino acid residues, 93 of which are identical to residues in human J chain.  相似文献   

3.
Activation of glutamate decarboxylase (GAD) by calcium-bound calmodulin (CaM) is required for normal plant growth through regulation of gamma-aminobutyrate and glutamate metabolism. The interaction of CaM with the C-terminal domain of GAD is believed to induce dimerization of the enzyme, an event implicated for Ca(2+)-dependent enzyme activation. Here, we present the solution structure of CaM in complex with a dimer of peptides derived from the C-terminus of Petunia hybrida GAD. The 23 kDa ternary complex is pseudo-symmetrical with each domain of CaM bound to one of the two antiparallel GAD peptides, which form an X-shape with an interhelical angle of 60 degrees. To accommodate the dimeric helical GAD target, the two domains of CaM adopt an orientation markedly different from that seen in other CaM-target complexes. Although the dimeric GAD domain is much larger than previously studied CaM-binding peptides, the two CaM domains appear closer together and make a number of interdomain contacts not observed in earlier complexes. The present structure of a single CaM molecule interacting with two target peptides provides new evidence for the conformational flexibility of CaM as well as a structural basis for the ability of CaM to activate two enzyme molecules simultaneously.  相似文献   

4.
Calcium binding to tryptic fragments of calmodulin   总被引:2,自引:0,他引:2  
Fragments of scallop testis calmodulin were prepared by tryptic digestion. One peptide consisted of 75 amino acid residues from N-acetylalanine to lysine at position 75 (F12) and the other of 71 residues from aspartic acid at position 78 to C-terminal lysine (F34). Flow dialysis and equilibrium dialysis experiments revealed the existence of two Ca2+ binding sites in each fragment. Half-saturating concentrations of the Ca2+ titration curves were 11 microM for F12 and 3.2 microM for F34, and Hill coefficients were obtained as 1.14 and 1.84, respectively. The results indicate that the high-affinity sites for Ca2+ are located on the C-terminal region of the calmodulin. The sum of the two Ca2+ titration curves of F12 and F34 fits well to the curves of Ca2+ binding to intact calmodulin. This shows that the characteristic of Ca2+ bindings in intact calmodulin did not change after separation of the whole molecule into two domains, F12 and F34. The domains corresponding to F12 and F34 may exist independently from each other in the intact calmodulin molecule.  相似文献   

5.
A novel photoreactive amino acid has been incorporated synthetically into two model peptides and the calmodulin-binding domain from myosin light chain kinase. Cross-linked photoadducts of each peptide with calmodulin have been prepared and digested by chemical and/or enzymatic methods to determine the site of label attachment. Depending on the position of the photoprobe in the peptide sequence, either Met-144 or Met-71 is photolabeled. These results are discussed in relation to the three-dimensional structure of calmodulin obtained crystallographically and the known solution properties of calmodulin.  相似文献   

6.
Binding of simple peptides, hormones, and neurotransmitters by calmodulin   总被引:7,自引:0,他引:7  
D A Malencik  S R Anderson 《Biochemistry》1982,21(14):3480-3486
We have prepared a fluorescent conjugate of porcine calmodulin with 5-(dimethylamino)-1-naphthalene-sulfonyl chloride that is highly sensitive to both calcium binding and protein binding. We have used the fluorescence of this conjugate in addition to the intrinsic peptide fluorescence to show that adrenocorticotropic hormone (ACTH), beta-endorphin, glucagon, and substance P undergo calcium-dependent binding by calmodulin, with competition for common binding sites. The dissociation constants determined in the presence of 0.85 mM CaCl2 and 0.2 N KC1, pH 7.3 at 25 degrees C, range from 1.5 muM to 3.4 muM. The alpha-melanocyte-stimulating hormone, bombesin, and somatostatin also bind, with dissociation constants between 60 muM and 90 muM. Angiotensins I and III, bradykinin, neurotensin, physalaemin, substance P octapeptide, insulin, and Leu- and Met-enkephalin show little or no binding. Sequence comparisons show that the peptides that bind calmodulin well contain regions structurally similar to the recognition sequence for the cAMP-dependent protein kinase and to the sequences surrounding phosphorylated serine residues in several calmodulin binding proteins. This result suggests that modification of calmodulin binding sites in calmodulin-dependent proteins is one of the functions of protein kinase. Calcium has a dual role in peptide binding by calmodulin. The occupation of calcium binding sites having a pK approximately 4 results in a 2-fold increase in peptide binding affinity.  相似文献   

7.
Binding of a spin-labelled chlorpromazine analogue to calmodulin.   总被引:1,自引:1,他引:0       下载免费PDF全文
The binding of a spin-labelled derivative of chlorpromazine to calmodulin was investigated by e.s.r. spectrometry. The completion of the spectroscopic changes requires the presence of 4 Ca2+ ions per calmodulin molecule. The influences of various physicochemical factors (pH, ionic strength) are discussed in relation to the nature (hydrophobic and polar) of the interactions that hold the drug-calmodulin complex together.  相似文献   

8.
Efficient electron transfer from reductase domain to oxygenase domain in nitric oxide synthase (NOS) is dependent on the binding of calmodulin (CaM). Rate constants for the binding of CaM to NOS target peptides was only determined previously by surface plasmon resonance (SPR) (Biochemistry 35, 8742-8747, 1996) suggesting that the binding of CaM to NOSs is slow and does not support the fast electron transfer in NOSs measured in previous and this studies. To resolve this contradiction, the binding rates of holo Alexa 350 labeled T34C/T110W CaM (Alexa-CaM) to target peptides from three NOS isozymes were determined using fluorescence stopped-flow. All three target peptides exhibited fast kon constants at 4.5 °C: 6.6 × 108 M− 1 s− 1 for nNOS726-749, 2.9 × 108 M− 1 s− 1 for eNOS492-511 and 6.1 × 108 M− 1 s− 1 for iNOS507-531, 3-4 orders of magnitude faster than those determined previously by SPR. Dissociation rates of NOS target peptides from Alexa-CaM/peptide complexes were measured by Ca2+ chelation with ETDA: 3.7 s− 1 for nNOS726-749, 4.5 s− 1 for eNOS492-511, and 0.063 s− 1 for iNOS507-531. Our data suggest that the binding of CaM to NOS is fast and kinetically competent for efficient electron transfer and is unlikely rate-limiting in NOS catalysis. Only iNOS507-531 was able to bind apo Alexa-CaM, but in a very different conformation from its binding to holo Alexa-CaM.  相似文献   

9.
In kidney the nickel ion exists primarily as soluble cytoplasmic complexes. We have recently identified a major component of these complexes in the human kidney as a Ni(II) complex of a low molecular weight anionic peptide (Templeton, D.M. and Sarkar, B. (1985) Biochem. J. 230, 35-42). We have now purified a small amount of this peptide to homogeneity and developed an HPLC technique to study its metal-binding properties on sub-nanomole quantities. We are able to demonstrate a binding stoichiometry of one Ni atom per molecule of peptide, with an apparent dissociation constant of 1.1 X 10(-5) M. A similar site exists for Cd. The site for Ni persists after trypsinization, and is localized in the 20-residue C-terminal tryptic fragment of the peptide.  相似文献   

10.
We have analysed the structural and physical properties of the carboxy-terminal stalk region of a kinesin-II, Xenopus kinesin-like protein 3A/B (Xklp3A/B), which we showed to be essential for heterodimerization in a previous work (De Marco et al., 2001). We expressed the corresponding A-stalk and B-stalk fragments and investigated their modes of interaction by analytical ultracentrifugation (AUC), circular dichroism spectroscopy, denaturation assays and electron microscopy. Co-expression of the A-stalk and B-stalk produced the properly folded, hetero-dimeric coiled coil at high yields. The dimeric nature of the complex was confirmed by AUC. We also found that the isolated A-stalk fragment forms a stable helix by itself and shows a significant tendency towards homodimer and higher-order complex formation. In the absence of the corresponding A-stalk fragment, the isolated B-stalk fragment remains partially unfolded, which suggests that the A-stalk provides a template structure for the B-stalk in order to recompose the complete heterodimeric coiled coil.  相似文献   

11.
Intracellular localization plays an important role in the functional regulation of the cell cycle inhibitor p21. We have previously shown that calmodulin binds to p21 and that calmodulin is essential for the nuclear accumulation of p21. Here, we analyze the mechanism of this regulation. We show that calmodulin inhibits in vitro phosphorylation of p21 by protein kinase C (PKC) and that this inhibition is dependent upon calmodulin binding to p21. Two-dimensional electrophoresis analysis of cells expressing the p21 wild type or p21S153A, a nonphosphorylatable mutant of p21 at position 153, indicates that Ser153 of p21 is a phosphorylable residue in vivo. Furthermore, Western blot analysis using phospho-Ser153-specific antibodies indicates that Ser153 phosphorylation in vivo is induced when PKC is activated and calmodulin is inhibited. The mutation of Ser153 to aspartate, a pseudophosphorylated residue, inhibits the nuclear accumulation of p21. Finally, whereas wild-type p21 translocates to the cytoplasm after PKC activation in the presence of calmodulin inhibitors, p21 carrying a nonphosphorylatable residue at position 153 remains in the nucleus. We propose that calmodulin binding to p21 prevents its phosphorylation by PKC at Ser153 and consequently allows its nuclear localization. When phosphorylated at Ser153, p21 is located at the cytoplasm and disrupts stress fibers.  相似文献   

12.
Automated carboxy-terminal sequence analysis of peptides.   总被引:1,自引:1,他引:0       下载免费PDF全文
Proteins and peptides can be sequenced from the carboxy-terminus with isothiocyanate reagents to produce amino acid thiohydantoin derivatives. Previous studies in our laboratory have focused on solution phase conditions for formation of the peptidylthiohydantoins with trimethylsilylisothiocyanate (TMS-ITC) and for hydrolysis of these peptidylthiohydantoins into an amino acid thiohydantoin derivative and a new shortened peptide capable of continued degradation (Bailey, J. M. & Shively, J. E., 1990, Biochemistry 29, 3145-3156). The current study is a continuation of this work and describes the construction of an instrument for automated C-terminal sequencing, the application of the thiocyanate chemistry to peptides covalently coupled to a novel polyethylene solid support (Shenoy, N. R., Bailey, J. M., & Shively, J. E., 1992, Protein Sci. I, 58-67), the use of sodium trimethylsilanolate as a novel reagent for the specific cleavage of the derivatized C-terminal amino acid, and the development of methodology to sequence through the difficult amino acid, aspartate. Automated programs are described for the C-terminal sequencing of peptides covalently attached to carboxylic acid-modified polyethylene. The chemistry involves activation with acetic anhydride, derivatization with TMS-ITC, and cleavage of the derivatized C-terminal amino acid with sodium trimethylsilanolate. The thiohydantoin amino acid is identified by on-line high performance liquid chromatography using a Phenomenex Ultracarb 5 ODS(30) column and a triethylamine/phosphoric acid buffer system containing pentanesulfonic acid. The generality of our automated C-terminal sequencing methodology was examined by sequencing model peptides containing all 20 of the common amino acids. All of the amino acids were found to sequence in high yield (90% or greater) except for asparagine and aspartate, which could be only partially removed, and proline, which was found not be capable of derivatization. In spite of these current limitations, the methodology should be a valuable new tool for the C-terminal sequence analysis of peptides.  相似文献   

13.
Recent reports which show that several chemokines can act as direct microbicidal agents have drawn renewed attention to these chemotactic signalling proteins. Here we present a structure-function analysis of peptides derived from the human chemokines macrophage inflammatory protein-3α (MIP-3α/CCL20), interleukin-8 (IL-8), neutrophil activating protein-2 (NAP-2) and thrombocidin-1 (TC-1). These peptides encompass the C-terminal α-helices of these chemokines, which have been suggested to be important for the direct antimicrobial activities. Far-UV CD spectroscopy showed that the peptides are unstructured in aqueous solution and that a membrane mimetic solvent is required to induce a helical secondary structure. A co-solvent mixture was used to determine solution structures of the peptides by two-dimensional 1H-NMR spectroscopy. The highly cationic peptide, MIP-3α51-70, had the most pronounced antimicrobial activity and displayed an amphipathic structure. A shorter version of this peptide, MIP-3α59-70, remained antimicrobial but its structure and mechanism of action were unlike that of the former peptide. The NAP-2 and TC-1 proteins differ in their sequences only by the deletion of two C-terminal residues in TC-1, but intact TC-1 is a very potent antimicrobial while NAP-2 is inactive. The corresponding C-terminal peptides, NAP-250-70 and TC-150-68, had very limited and no bactericidal activity, respectively. This suggests that other regions of TC-1 contribute to its bactericidal activity. Altogether, this work provides a rational structural basis for the biological activities of these peptides and proteins and highlights the importance of experimental characterization of peptide fragments as distinct entities because their activities and structural properties may differ substantially from their parent proteins.  相似文献   

14.
The binding of Ca2+ to calmodulin and its two tryptic fragments has been studied using microcalorimetry. The binding process is accompanied by the uptake or release of protons, depending on the ionic strength. With no added salt, the total enthalpy change for the binding of four calcium ions to calmodulin is -41 kJ mol-1 but in the presence of 0.15 mM KCl delta Htot is +17 kJ mol-1. The mode of binding of Ca2+ is also completely different with and without added salt. It is also shown that for the C-terminal fragment of calmodulin, TR2C, the drastic reduction in delta Gtot for the binding process on increasing the ionic strength is largely an enthalpic effect. Domain interactions in calmodulin are indicated by the fact that the sum of the enthalpies of calcium binding to the two tryptic fragments is not the same as the total binding enthalpy to calmodulin itself. The binding of Ca2+ to calmodulin has also been studied calorimetrically at different temperatures in the range 21-37 degrees C. delta Cp is large and negative in this interval.  相似文献   

15.
The moderate evolution rate of apomyoglobins may be the support of a simplified strategy for determining unknown covalent structures within the order of Carnivora, taking the badger apomyoglobin as a model. The CNBr cleavage was followed by the isolation of three polypeptide fragments which were subsequently submitted to trypsin digestion. The fingerprints of the three hydrolysates as may be obtained from seven Carnivora species, show a fairly constant number of spots, often corresponding to identical or closely related peptides, espcially in the case of the N-terminal and C-terminal fragments.  相似文献   

16.
The interaction of melittin with calmodulin and its tryptic fragments   总被引:4,自引:0,他引:4  
Melittin has been found to interact with both the N- and C-terminal half-molecules of calmodulin, as well as the intact molecule, in the presence of Ca2+. The interaction results in a major change in the microenvironment of Trp-19, which is in a more nonpolar, solvent-shielded, and immobilized microenvironment in the complex. The properties of Tyr-99 and Tyr-138 of calmodulin are altered by complex formation. From measurements of the efficiencies of radiationless energy transfer from Trp-19 to the nitro derivatives of Tyr-99 and/or Tyr-138, it is concluded that Trp-19 is located in proximity to the C-terminal lobe of calmodulin in the complex.  相似文献   

17.
A carboxy-terminal, 18 kD fragment of colicin Ia, a bacterial toxin, forms ion channels in artificial phospholipid bilayers. This fragment, which comprises a quarter of the intact 70 kD molecule, is resistant to extensive protease digestion and probably constitutes a structural domain of the protein. The ion channels formed by the 18 kD fragment are functionally heterogeneous, having conductances that range from 15 to 30 pS at positive voltages and from 70 to 250 pS at negative voltages, and open lifetimes that range from at least 25 msec to 5 sec. In contrast, ion channels formed by whole colicin Ia open only at negative voltages, at which their conductances range from 6 to 30 pS, and their open lifetimes range from 1 sec to 3 min. Additionally, the open state of the 18 kD fragment channel is characterized by noisy fluctuations in current, while the open state of the whole molecule ion channel is often marked by numerous, stable subconductance states. Since the properties of the fragment channel differ substantially from those of the whole molecule channel, we suggest that portions of the molecule outside of the 18 kD fragment are involved in forming the whole molecule ion channel.  相似文献   

18.
Using chromatography and preparative polyacrylamide gel electrophoresis, tryptic peptides TP 1 (residues 47-83), TP 2 (residues 84-118) and TP 3 (residues 119-161) were isolated in a highly homogeneous state from cardiac troponin C. Peptides TP 1, TP 2 and TP 3 were found to contain isolated cation-binding sites II, III and IV of cardiac troponin C. The interaction of these peptides with troponins I and T was studied. It was found that only peptide TP 2 could interact with troponin I. Neither of the peptides isolated interacted with troponin T. The cation-binding properties and structural peculiarities of peptide TP 1 were investigated. It was shown that despite its small size (37 amino acid residues), peptide TP 1 retained its ability to bind Ca2+ which caused conformational changes in the peptide structure. This was accompanied by changes in the electrophoretic mobility and absorption of TP 1 on phenyl-Sepharose.  相似文献   

19.
20.

Background

Recent studies have shown that fluorescently labeled antibodies can be dissociated from their antigen by illumination with laser light. The mechanism responsible for the photounbinding effect, however, remains elusive. Here, we give important insights into the mechanism of photounbinding and show that the effect is not restricted to antibody/antigen binding.

Methodology/Principal Findings

We present studies of the photounbinding of labeled calmodulin (CaM) from a set of CaM-binding peptides with different affinities to CaM after one- and two-photon excitation. We found that the photounbinding effect becomes stronger with increasing binding affinity. Our observation that photounbinding can be influenced by using free radical scavengers, that it does not occur with either unlabeled protein or non-fluorescent quencher dyes, and that it becomes evident shortly after or with photobleaching suggest that photounbinding and photobleaching are closely linked.

Conclusions/Significance

The experimental results exclude surface effects, or heating by laser irradiation as potential causes of photounbinding. Our data suggest that free radicals formed through photobleaching may cause a conformational change of the CaM which lowers their binding affinity with the peptide or its respective binding partner.  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