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 963 毫秒
1.
The cDNA fragments corresponding to the domains with four consecutive E-F hand structures in the large and small subunits of chicken and rabbit calcium-activated neutral protease (CANP) were inserted into an expression vector (pUC8 or pUC18). The resulting plasmids were used to transform E. coli, and isopropyl-1-thio-beta-D-galactoside (IPTG)-inducible expression was performed. The resulting four kinds of E-F hand structure-domains (the chicken large subunit, rabbit high- and low-calcium-requiring large subunits, and rabbit small subunit) were purified and analyzed for their calcium-binding abilities and capacities by the microscale filter assay. Most of the E-F hand structures could bind calcium and 2 or 4 mol of Ca2+ ions bound to the four consecutive E-F hand structures. The calcium-binding affinity of the E-F hand structures in the large subunit roughly corresponds to the calcium concentration required for its CANP activity.  相似文献   

2.
Calcium-activated neutral protease (rabbit mCANP), composed of large and small subunits, was converted to a lower-Ca2+-requiring form (derived microCANP) by limited autolysis in the presence of Ca2+. The NH2-terminal regions of the two subunits of mCANP were cleaved by autolysis, but the COOH-termini remained intact after autolysis. When native mCANP or derived microCANP was dissociated into subunits, the proteolytic activity of the large subunit was reduced to 2-5% of that of the native dimeric enzyme. The Ca2+-sensitivity of one hybrid CANP reconstituted from the large subunit of derived microCANP and the small subunit of native mCANP was similar to that of derived microCANP. However, the other hybrid molecule composed of the large subunit of native mCANP and the small subunit of derived microCANP required a high concentration of Ca2+ for activity, like native mCANP. These results indicate that the Ca2+-sensitivity of derived microCANP is determined by the structural change of the large subunit resulting from loss of its NH2-terminal region. The autolysis of the small subunit apparently has no effect on the reduction of the Ca2+-requirement.  相似文献   

3.
The structure of bovine intestinal calcium-binding protein (ICaBP) has been determined crystallographically at a resolution of 2.3 A and refined by a least squares technique to an R factor of 17.8%. The refined structure includes all 600 non-hydrogen protein atoms, two bound calcium ions, and solvent consisting of one sulfate ion and 36 water molecules. The molecule consists of two helix-loop-helix calcium-binding domains known as EF hands, connected by a linker containing a single turn of helix. Helix-helix interactions are primarily hydrophobic, but also include a few strategic hydrogen bonds. Most of the hydrogen bonds, however, are found in the calcium-binding loops, where they occur both within a single loop and between the two. Examination of the hydrogen bonding patterns in the calcium-binding loops of ICaBP and the related protein, parvalbumin, reveals several conserved hydrogen bonds which are evidently important for loop stabilization. The primary and tertiary structural features which promote the formation of an EF hand were originally identified from the structure of parvalbumin. They are modified in light of the ICaBP structure and considered as they apply to other calcium-binding proteins. The C-terminal domain of ICaBP is a normal EF hand, with ion binding properties similar to those of the calmodulin hands, but the N-terminal domain is a variant hand whose calcium ligands are mostly peptide carbonyls. Relative to a normal EF hand, this domain exhibits a similar KD for calcium binding but a greatly reduced affinity for calcium analogs such as cadmium and the lanthanide series. Lanthanides in particular may be inappropriate models for calcium in this system.  相似文献   

4.
Kv channel-interacting proteins (KChIPs) and neuronal calcium sensor-1 (NCS-1) have been shown to interact with Kv4 channel alpha-subunits to regulate the expression and/or gating of these channels. Here we examine the specificity and sites of these proteins for interaction with Kv channel proteins. Immunoprecipitation and green fluorescent protein imaging show that KChIPs (but not NCS-1) effectively bind to Kv4.3 protein and localize at the plasma membrane when channel proteins are coexpressed. Analysis with chimeric proteins between KChIP2 and NCS-1 reveals that the three regions of KChIP2 (the linker between the first and second EF hands, the one between the third and fourth EF hands, and the C-terminal peptide after the fourth EF hand) are necessary and sufficient for its effective binding to Kv4.3 protein. The chimera with these three KChIP2 portions slowed inactivation and facilitated recovery from inactivation of Kv4.3 current. These results indicate that the sequence difference in these three regions between KChIPs and NCS-1 determines the specificity and affinity for interaction with Kv4 protein. Because the three identified regions surround the large hydrophobic crevice based on the NCS-1 crystal structure, this crevice may be the association site of KChIPs for the channel protein.  相似文献   

5.
Fifteen hybridomas secreting antibodies against calcium-activated neutral protease (CANP), especially those for rabbit muscle mCANP with low calcium sensitivity, have been produced by the cell fusion technique. Eight of the monoclonal antibodies belong to the class IgG1, one to the class IgG2a, and six to the class IgG2b. The antibodies from these clones were characterized with regard to their relative binding affinities to the large subunits (80K) and the small subunits (30K) of mCANP as well as mu CANP, which is another type of CANP with high calcium sensitivity. Fourteen antibodies bound only to the 80K subunit of mCANP and one antibody bound to the 80K subunit of both mCANP and mu CANP. These antibodies recognized rat mCANP but not chicken CANP, with the exception of one antibody. Examination of the effects of these antibodies on the enzyme activity of mCANP showed that six antibodies partially inhibited the enzyme activity and the others were noninhibitory. These monoclonal antibodies should be useful for analyzing the fine structure of CANPs and the mechanism of the activation of mCANP, and also for determining the intracellular localization of mCANP.  相似文献   

6.
It is widely believed that the selectivity of voltage-dependent calcium channels is mainly controlled by amino acid residues contained within four p-loop motifs forming the pore of the channel. An examination of the amino acid sequences of high voltage-activated calcium channels reveals that their domain III S5-H5 regions contain a highly conserved motif with homology to known EF hand calcium binding proteins, hinting that this region may contribute to channel permeation. To test this hypothesis, we used site-directed mutagenesis to replace three conserved negatively charged residues in the N-type calcium channel alpha1B subunit (Glu-1321, Asp-1323, and Glu-1332) with positively charged amino acids (lysine and arginine) and studied their effect on ion selectivity using whole cell and single channel patch clamp recordings. Whereas the wild type channels conducted barium much more effectively than calcium, the mutant displayed nearly equal permeabilities for these two ions. Individual replacement of residue 1332 or a double substitution of residues 1321 and 1323 with lysine and arginine, respectively, were equally effective. Disruption of the putative EF hand motif through replacement of the central glycine residue (1326) with proline resulted in a similar effect, indicating that the responses observed with the triple mutant were not due to changes in the net charge of the channel. Overall, our data indicate that residues outside of the narrow region of the pore have the propensity to contribute to calcium channel permeation. They also raise the possibility that interactions of calcium ions with a putative calcium binding domain at the extracellular side of the channel may underlie the differential permeabilities of the channel for barium and calcium ions.  相似文献   

7.
C1-Tetrahydrofolate synthase is a multifunctional enzyme which catalyzes three reactions in 1-carbon metabolism: 10-formyltetrahydrofolate synthetase; 5,10-methenyltetrahydrofolate cyclohydrolase; 5,10-methylenetetrahydrofolate dehydrogenase. A rapid 1-day purification procedure has been developed which gives 40 mg of pure enzyme from 10 rabbit livers. The 10-formyltetrahydrofolate synthetase activity of this trifunctional enzyme has a specific activity that is 4-fold higher than the enzyme previously purified from rabbit liver. Conditions have been developed for the rapid isolation of a tryptic fragment of the enzyme which contains the methylenetetrahydrofolate dehydrogenase and methenyltetrahydrofolate cyclohydrolase activities. This fragment is a monomer exhibiting a subunit and native molecular weight of 36,000 in most buffers. However, in phosphate buffers the native molecular weight suggests that the fragment is a dimer. Conditions are also given whereby chymotryptic digestion allows the simultaneous isolation from the native enzyme of a large fragment containing the 10-formyltetrahydrofolate synthetase activity and a smaller fragment containing the dehydrogenase and cyclohydrolase activities. The large fragment is a dimer with a subunit molecular weight of 66,000. The small fragment retains all of the dehydrogenase and cyclohydrolase activities of the native enzyme. The large fragment is unstable but retains most of the 10-formyltetrahydrofolate synthetase activity. Km values of substrates for the two fragments are the same as the values for the native enzyme. The 10-formyltetrahydrofolate synthetase activity of the native enzyme requires ammonium or potassium ions for expression of full catalytic activity. The effect of these two ions on the catalytic activity of the large chymotryptic fragment is the same as with the native enzyme. We have shown by differential scanning calorimetry that the native enzyme contains two protein domains which show thermal transitions at 47 and 60 degrees C. Evidence is presented that the two domains are related to the two protein fragments generated by proteolysis of the native enzyme. The larger of the two domains contains the active site for the 10-formyltetrahydrofolate synthetase activity while the smaller domain contains the active site which catalyzes the dehydrogenase and cyclohydrolase reactions. Replacement of sodium ion buffers with either ammonium or potassium ions results in an increase in stability of the large domain of the native enzyme. This change in stability is not accompanied by a change in the quaternary structure of the enzyme.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

8.
We examined the structure and the distribution of binding activities within bacterially produced fragments of Drosophila alpha spectrin. By electron microscopy, purified spectrin fragments resembled the corresponding regions of native spectrin. The contour lengths of recombinant spectrin molecules were proportional to the length of their coding sequences, which is consistent with current models of spectrin structure in which individual segments of the polypeptide contribute independently to the structure of the native molecule. We localized two sites at which calcium may regulate spectrin function. First, a site responsible for calmodulin binding to Drosophila alpha spectrin was identified near the junction of repetitive segments 14 and 15. Second, a domain of Drosophila alpha spectrin that includes two EF hand calcium-binding sequences bound 45Ca in blot overlay assays. EF hand sequences from a homologous domain of Drosophila alpha actinin did not bind calcium under the same conditions.  相似文献   

9.
The interaction of uranyl ions with inorganic pyrophosphatase from baker's yeast was investigated by measurement of their effect on the protein fluorescence. Fluorescence titrations of the native enzyme with uranyl nitrate show that there is a specific binding of uranyl ions to the enzyme. It was deduced that each subunit of the enzyme binds one uranyl ion. The binding constant was estimated to be in the order of 10(7) M-1. The enzyme which contains a small number of chemically modified carboxyl groups was not able to bind uranyl ions specifically. The modification of carboxyl groups was carried out by use of a water soluble carbodiimide and the nucleophilic reagent N-(2,4-dinitro-phenyl)-hexamethylenediamine. The substrate analogue calcium pyrophosphate displaced the uranyl ions from their binding sites at the enzyme. From the results it is concluded that carboxyl groups of the active site are the ligands for the binding of uranyl ions.  相似文献   

10.
We performed kinetic studies on the reactions of a fluorescent ATP analog, 2'-(5-dimethyl-aminonaphthalene-1-sulfonyl) amino-2'-deoxyATP (DNS-ATP), with E. coli F1-ATPase (EF1) and its subunits, to clarify the role of each subunit in the ATPase reaction. The following results were obtained. 1. One mol of EF1, which contains nonexchangeable 2 mol ATP and 0.5 mol ADP, binds 3 mol of DNS-ATP. The apparent dissociation constant, in the presence of Mg2+, was 0.23 microM. Upon binding, the fluorescence intensity of DNS-ATP at 520 nm increased exponentially with t1/2 of 35 s, and reached 3.5 times the original fluorescence level. Following the fluorescence increase, DNS-ATP was hydrolyzed, and the fluorescence intensity maintained its enhanced level. 2. The addition of an excess of ATP over the EF1-DNS-nucleotide complex, in the presence of Mg2+, decreased the fluorescence intensity rapidly, indicating the acceleration of DNS-nucleotide release from EF1. ADP and GTP also decreased the fluorescence intensity. 3. DCCD markedly inhibited the accelerating effect of ATP on DNS-nucleotide release from EF1 and the EF1-DNS-ATPase or -ATPase activity in a steady state. On the other hand, DCCD only slightly inhibited the fluorescence increase of DNS-ATP, due to its binding to EF1, and the rate of single cleavage of 1 mol of DNS-ATP per mol of alpha subunit of EF1. 4. In the presence of Mg2+, 0.65-0.82 mol of DNS-ATP binds to 1 mol of the isolated alpha subunit of EF1 with an apparent dissociation constant of 0.06-0.07 microM. Upon binding, the fluorescence intensity of DNS-ATP at 520 nm increased 1.55 fold very rapidly (t1/2 less than 1 s). No hydrolysis of DNS-ATP was observed upon the addition of the isolated alpha subunit. The fluorescence intensity of DNS-ATP was unaffected by the addition of the isolated beta subunit. DNS-ATP was also unhydrolyzed by the isolated beta subunit. 5. EF1-ATPase was reconstituted from alpha, beta, and gamma subunits in the presence of Mg2+ and ATP. The kinetic properties of the fluorescence change of DNS-ATP in the reaction with the reconstituted EF1-ATPase were quite similar to those of native EF1. Most of our findings are consistent with a simple mechanism that the high affinity catalytic site and low affinity regulatory site exist in the alpha subunit and beta subunit, respectively. However, the findings mentioned in (4) suggest that the binding of the alpha and beta subunit, which is mediated by the gamma subunit, induces conformational change(s) in the ATP binding site located probably in the alpha subunit, and that the conformational change(s) is essential to exert the full hydrolyzing activity.  相似文献   

11.
Bacterial translation initiation factor IF2 complexed with GTP binds to the 30S ribosomal subunit, promotes ribosomal binding of fMet‐tRNA, and favors the joining of the small and large ribosomal subunits yielding a 70S initiation complex ready to enter the translation elongation phase. Within the IF2 molecule subdomain G3, which is believed to play an important role in the IF2‐30S interaction, is positioned between the GTP‐binding G2 and the fMet‐tRNA binding C‐terminal subdomains. In this study the solution structure of subdomain G3 of Geobacillus stearothermophilus IF2 has been elucidated. G3 forms a core structure consisting of two β‐sheets with each four anti‐parallel strands, followed by a C‐terminal α‐helix. In line with its role as linker between G3 and subdomain C1, this helix has no well‐defined orientation but is endowed with a dynamic nature. The structure of the G3 core is that of a typical OB‐fold module, similar to that of the corresponding subdomain of Thermus thermophilus IF2, and to that of other known RNA‐binding modules such as IF2‐C2, IF1 and subdomains II of elongation factors EF‐Tu and EF‐G. Structural comparisons have resulted in a model that describes the interaction between IF2‐G3 and the 30S ribosomal subunit.  相似文献   

12.
Recombinant calcineurin heterodimer with the full length delta-isoform of the catalytic subunit (CaN(500)) was expressed in insect cells using the baculovirus system and compared to native bovine brain enzyme in its response to divalent metal ions, redox reagents, and enzymatic modification of arginine residues. The response to various metal ions showed essentially the same profile as bovine brain calcineurin, although Co2+ and Zn2+ did not support recombinant activity as well. Kinetic analysis showed that metal ion and substrate binding were not independent, as found for the bovine brain calcineurin. Incubation with DTT or ascorbate alone caused similar effects on the activity of both enzymes, but different responses were observed when incubated with both DTT and ascorbate; only the recombinant enzyme showed activation. Arginine deimination of recombinant calcineurin by peptidylarginine deiminase resulted in the loss of 60-80% of its phosphatase activity with protection observed if calmodulin was present. Recombinant calcineurin was reactivated by treatment with the protease clostripain, suggesting that deimination of an arginine in the carboxyl terminal domain may be responsible for the loss of phosphatase activity and decreased calmodulin binding [Arch. Biochem. Biophys. 318 (1995) 370]. Supporting this conclusion, a truncated variant of the catalytic subunit lacking the carboxyl terminus showed no loss of phosphatase activity compared to full length calcineurin subunit and contained lower amounts of citrulline than the full length subunit after deimination. These different responses of recombinant calcineurin are consistent with conformational differences compared to bovine brain calcineurin and raise questions about its utility for studying the mechanism of calcineurin.  相似文献   

13.
The regulatory light chains (RLCs) located on the myosin head, regulate the interaction of myosin with actin in response to either Ca2+ or phosphorylation signals. The RLCs belong to a family of calcium binding proteins and are composed of four "EF hand" ancestral calcium binding motifs (numbered I to IV). To determine the role of the first EF hand (EF hand I) in the regulatory process, chimaeric light chains were constructed by protein engineering, by switching this region between smooth muscle and skeletal muscle myosin RLCs. For example, chimaera G(I)S consisted of EF hand I of the smooth muscle (gizzard) RLC and EF hands II to IV of the skeletal muscle RLC, whereas chimaera S(I)G consisted of EF hand I of the skeletal muscle RLC and EF hands II to IV of the smooth muscle RLC. The chimaeric RLCs were expressed in Escherichia coli using the pLcII expression system, and after isolation and purification their regulatory properties were compared with those of wild-type smooth and skeletal muscle myosin RLCs. The chimaeric RLCs bound to the myosin heads in scallop striated muscle myofibrils from which the endogenous RLCs had been removed ("desensitized" myofibrils) with similar affinities to those of the wild-type smooth and skeletal muscle RLCs. Both chimaeric RLCs were able to regulate the actin-activated Mg(2+)-ATPase activity of scallop myosin: G(I)S inhibited the ATPase in the presence and absence of Ca2+, like the wild-type skeletal muscle RLC, while S(I)G inhibited the myosin ATPase in the absence of Ca2+, and this inhibition was relieved on Ca2+ addition, in the same way as the wild-type smooth muscle RLC. Thus the type of regulation that the RLCs confer on the myosin is determined by the source of EF hands II to IV rather than that of EF hand I.  相似文献   

14.
The helix-loop-helix (i.e., EF-hand) Ca(2+) ion binding motif is characteristic of a large family of high-affinity calcium ion binding proteins, including the parvalbumins, oncomodulins and calmodulins. In this work we describe a set of molecular dynamics computations on the major parvalbumin from the silver hake (SHPV-B) and on functional fragments of this protein, consisting of the first four helical regions (the ABCD fragment), and the internal helix-loop- helix region (the CD fragment). In both whole protein and protein fragments (i.e., ABCD and CD fragments), the 9th loop residue in the calcium ion binding site in the CD helix-loop-helix region (the so-called "gateway" position) has been mutated from glutamic acid to aspartic acid. Aspartic acid is one of the most common residues found at the gateway position in other (non-parvalbumin) EF- hand proteins, but has never been found at the gateway position of any parvalbumin. (Interestingly, aspartic acid does occur at the gateway position in the closely related rat and human oncomodulins.) Consistent with experimental observations, the results of our molecular dynamics simulations show that incorporation of aspartic acid at the gateway position is very disruptive to the structural integrity of the calcium ion coordination site in the whole protein. The aspartic acid mutation is somewhat less disruptive to the calcium ion coordination sites in the two parvalbumin fragments (i.e., the ABCD and CD fragments), presumably due to the higher degree of motional freedom allowable in these protein fragments. One problem associated with the E59D whole protein variant is a prohibitively close approach of the aspartate carboxyl group to the CD calcium ion observed in the energy-minimized (pre-molecular dynamics) structure. This steric situation does not emerge during energy-minimization of the wild-type protein. The damage to the structural integrity of the calcium ion coordination site in the whole protein E59D variant is not relieved during the molecular dynamics simulation. In fact, during the course of the 300 picosecond simulation, all of the calcium ion ligands leave the primary coordination sphere. In addition, the conserved hydrogen- bonds (in the short beta-sheet structure) that links the CD site to the symmetry-related EF site (in the non-mutated whole protein) is also somewhat disrupted in the E59D whole protein variant. These results suggest that the Ca(2+) ion binding deficiencies in the CD loop are related, at least in part, to the unique interaction that exists between the paired CD and EF hands in the whole protein. Our theoretical results correlate well with previous studies on engineered EF-hand proteins and with all of our experimental evidence on whole silver hake parvalbumin and enzymatically-generated parvalbumin fragments.  相似文献   

15.
To identify and characterize Schistosoma mansoni proteins that are recognized by infected hosts, we have used a pool of sera from infected humans to screen cDNA libraries constructed from poly(A)+ mRNA of adult S. mansoni. The deduced amino acid sequences of the three isolated clones showed a high degree of similarity to the large subunit of calcium-activated neutral proteinase (CANP) from humans and chicken. These overlapping clones, which include a nearly full-length clone with an open reading frame of 758 amino acid residues, together encode the entire large subunit of CANP. The deduced sequence of this S. mansoni protein can be divided into four domains (I-IV) that include the two domains characteristic of other large subunits of CANP: a thiol-protease domain (II) and a calcium-binding domain (IV) containing EF hand motifs. However, the schistosome protein is unique in having only three EF hand motifs in the calcium-binding domain and in having an additional EF hand motif that is shared between domains II and III. We have shown that these EF hand motifs are capable of binding 45Ca2+. Furthermore, the large subunit is S. mansoni contains an NH2-terminal sequence of 28 residues that is absent from the mammalian CANPs and has a high degree of similarity to the presumed receptor binding sequence of colicin Ia and Ib.  相似文献   

16.
The large and small subunits of ribulose bisphosphate carboxylase from Chromatium vinosum were dissociated and separated at pH 9.6 by sucrose density gradient centrifugation. After further purification by gel filtration, the small subunit fraction contained no carboxylase activity. The large subunit fraction was highly depleted of small subunit based on analysis by denaturing polyacrylamide gel electrophoresis. Carboxylase activity of the large subunit fraction was approximately 1% of the untreated native enzyme. Addition of purified small subunit to the large subunit fraction yielded increases of up to 67-fold in carboxylase activity, further indicating that both subunit types are required for catalysis by this enzyme. The isolated large subunit was fully capable of high-affinity activator 14CO2 binding in the presence of Mg2+ and 2-carboxyarabinitol bisphosphate, indicating that the activator and catalytic sites were not grossly denatured by the depletion of small subunit. Kinetic constants of the native C. vinosum enzyme defined a new class of ribulose bisphosphate carboxylase, which permits the detection of possible kinetic differences if the large and small subunits can be favorably reassembled with those of another kinetic class. From experiments with the enzymes from tobacco and spinach leaves it is concluded that the enzyme from higher plant sources is not suitable for such dissociation/reconstitution-type experiments.  相似文献   

17.
A calcium-activated neutral cysteine protease was purified to homogeneity from Dicentrarchus labrax white muscle using three steps: hydrophobic interaction, anion exchange, and gel filtration chromatographies. The purified enzyme showed a native molecular weight of 124 kDa with an oligomeric structure (large subunit of 80 kDa and small subunit of 24 kDa). It has been classified as a milli-calpain from its calcium sensitivity. Activity was maximal at pH 7.0, 24°C in Tris buffer without NaCl as determined by means of a two-level experimental design and response surface methodology. Sea bass calpain is neither glycosylated nor phosphorylated and shared some common cleavage specificities and activation and autolysis mechanisms with other typical mammalian or invertebrates calpains. Calcium-induced activation and autolysis of calpain has been characterized together with the effect of the strontium cation acting as a calcium analog. On the basis of its in vitro properties, the contribution of the sea bass milli-calpain to the process of postmortem deterioration of fish muscle is discussed, even though further information such as in vivo regulation or in vitro effects on myofibrils is required. Received March 1, 2001; accepted July 9, 2001.  相似文献   

18.
Understanding the roles of metal ions in restriction enzymes has been complicated by both the presence of two metal ions in many active sites and their homodimeric structure. Using a single-chain form of the wild-type restriction enzyme PvuII (scWT) in which subunits are fused with a short polypeptide linker (Simoncsits et al. in J. Mol. Biol. 309:89–97, 2001), we have characterized metal ion and DNA binding behavior in one subunit and examined the effects of the linker on dimer behavior. scWT exhibits heteronuclear single quantum coherence NMR spectra similar to those of native wild-type PvuII (WT). For scWT, isothermal titration calorimetry data fit to two Ca(II) sites per subunit with low-millimolar K ds. The variant scWT|E68A, in which metal ion binding in one subunit is abolished by mutation, also binds two Ca(II) ions in the WT subunit with low-millimolar K ds. When there are no added metal ions, DNA binding affinity for scWT is tenfold stronger than that of the native WT, but tenfold weaker at saturating Ca(II) concentration. In the presence of Ca(II), scWT|E68A binds target DNA similarly to scWT, indicating that high-affinity substrate binding can be carried energetically by one metal-ion-binding subunit. Global analysis of DNA binding data for scWT|E68A suggests that the metal-ion-dependent behaviors observed for WT are reflective of independent subunit behavior. This characterization provides an understanding of subunit contributions in a homodimeric context.  相似文献   

19.
Two distinct calcium-dependent neutral proteases (CANPs) with different sensitivities to calcium ions were purified concurrently by almost the same procedures from rabbit skeletal muscle and their enzymatic properties were compared (sensitivity to various divalent metal ions, the pH dependency and heat-stability of the activity, and the hydrolytic activity towards various substrates). They were further compared chemically in terms of the state of thiol groups, the amino acid compositions of subunits and the peptide fragments by digestion with S. aureus V8 protease. The low calcium requiring form of CANP (microCANP) was more sensitive to other divalent metal ions such as Sr2+ and Ba2+ than the high calcium requiring form of CANP (mCANP). The comparison of the pH dependency of these CANP activities showed that microCANP was active in a broader pH range than mCANP and the former was more heat-stable than the latter. Both CANPs had similar affinity to various substrates, but the hydrolytic velocity was several times higher with microCANP than with mCANP. Although they were inhibited by thiol protease inhibitors to the same extent, the states of thiol groups in them were quite different. The thiol group involved in the catalytic activity of the enzyme was exposed without adding Ca2+ in microCANP, whereas the group in mCANP became exposed only when sufficient Ca2+ was added. The large subunits of these two CANPs were different in their amino acid compositions and in the peptide fragment patterns produced by S. aureus V8 protease but the small subunits were indistinguishable from each other. These results led us to conclude that these two CANPs are quite different in nature and are not in a simple relationship, i.e., one of them is not derived from the other by autolysis or modification.  相似文献   

20.
Using a sordarin derivative, an antifungal drug, it was possible to determine the structure of a eukaryotic ribosome small middle dotEF2 complex at 17.5 A resolution by three-dimensional (3D) cryo-electron microscopy. EF2 is directly visible in the 3D map and the overall arrangement of the complex from Saccharomyces cerevisiae corresponds to that previously seen in Escherichia coli. However, pronounced differences were found in two prominent regions. First, in the yeast system the interaction between the elongation factor and the stalk region of the large subunit is much more extensive. Secondly, domain IV of EF2 contains additional mass that appears to interact with the head of the 40S subunit and the region of the main bridge of the 60S subunit. The shape and position of domain IV of EF2 suggest that it might interact directly with P-site-bound tRNA.  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