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.
The native dimeric form of methionyl-tRNA synthetase of Escherichia coli contains two zinc atoms per dimer, one per subunit. The bound zinc is retained upon trypsin modification which yields a monomer with one zinc atom. The enzymatic activity of both the dimeric forms is reversibly inhibited by 1,10-phenanthroline but not by its non-chelating analogues. In addition, the native enzyme binds two Mn2+ per dimer with a binding constant of approx. 70 micron but no binding is observed with the trypsin-modified monomer.  相似文献   

2.
《Biophysical journal》2020,118(5):1221-1233
Voltage-gated proton channels (HV1) are essential for various physiological tasks but are strongly inhibited by Zn2+ cations. Some determinants of Zn2+ binding have been elucidated experimentally and in computational studies. However, the results have always been interpreted under the assumption that Zn2+ binds to monomeric HV1 despite evidence that HV1 expresses as a dimer and that the dimer has a higher affinity for zinc than the monomer and experimental data that suggest coordination in the dimer interface. The results of former studies are also controversial, e.g., supporting either one single or two binding sites. Some structural determinants of the binding are still elusive. We performed a series of molecular dynamics simulations to address different structures of the human proton channel, the monomer and two plausible dimer conformations, to compare their respective potential to interact with and bind Zn2+ via the essential histidines. The series consisted of several copies of the system to generate independent trajectories and increase the significance compared to a single simulation. The amount of time simulated totals 29.9 μs for 126 simulations of systems comprising ∼59,000 to ∼187,000 atoms. Our approach confirms the existence of two binding sites in monomeric and dimeric human HV1. The dimer interface is more efficient for attracting and binding Zn2+ via the essential histidines than the monomer or a dimer with the histidines in the periphery. The higher affinity is due to the residues in the dimer interface that create an attractive electrostatic potential funneling the zinc cations toward the binding sites.  相似文献   

3.
The binding of phenosafranine (PS) to DNA was studied by a combination of spectroscopic methods (absorption and fluorescence) together with hydrodynamic measurements (sedimentation and viscosity), Analysis of spectroscopic binding curves revealed that the strength of binding of PS to DNA is generally lower than that of proflavine. These measurements enabled recognition of several modes of interaction between PS and native DNA: strong monomer binding prevailing at high DNA phosphate/dye ratios (p) comprising binding outside the DNA helix as well as intercalation; two modes of dimer binding at lower values of p; and probably also weak surface-binding of monomers as p approaches unity. Longer surface-bound aggregates of PS were not detected because of the low tendency of the dye to form aggregates, though the presence of dimeric species distinct frorn pure surface-stacked PS dimer was indicated by various observations. It occurs over a broad range of p values Starting at p ≈110 for ionic strengths 10?3–10?1. Thermal denaturation data indicate that this species is bound more strongly than pure surface-bound stacked dimer. Its dimeric character may be explained in terms of interaction of an intercalated dye molecule with an adjacent outside-bound one as suggested for acridines by Armstrong et al. Various properties of this species are discussed. Both strong and weak modes or binding of PS to DNA are sensitive to the presence of organic solvents. The effectiveness of solvents to destabilise the complexes substantially coincides with their capacity to alter the water activity. Viscometric investigations reveal that in the region of strongest bindins (p ? 15) the elongation of the DNA helix by approximately 0.18 nm per bound PS molecule is accompanied by a strong negative change in persistence length, i.e. bending. Similar bending is also found at higher levels of binding (p ? 15) induced by less lightly bound PS molecules, in which region, however, the unusually high elongation of approximately 0.34 nm per bound PS molecule is observed.  相似文献   

4.
The number of artificial protein supramolecules has been increasing; however, control of protein oligomer formation remains challenging. Cytochrome c′ from Allochromatium vinosum (AVCP) is a homodimeric protein in its native form, where its protomer exhibits a four‐helix bundle structure containing a covalently bound five‐coordinate heme as a gas binding site. AVCP exhibits a unique reversible dimer–monomer transition according to the absence and presence of CO. Herein, domain‐swapped dimeric AVCP was constructed and utilized to form a tetramer and high‐order oligomers. The X‐ray crystal structure of oxidized tetrameric AVCP consisted of two monomer subunits and one domain‐swapped dimer subunit, which exchanged the region containing helices αA and αB between protomers. The active site structures of the domain‐swapped dimer subunit and monomer subunits in the tetramer were similar to those of the monomer subunits in the native dimer. The subunit–subunit interactions at the interfaces of the domain‐swapped dimer and monomer subunits in the tetramer were also similar to the subunit–subunit interaction in the native dimer. Reduced tetrameric AVCP dissociated to a domain‐swapped dimer and two monomers upon CO binding. Without monomers, the domain‐swapped dimers formed tetramers, hexamers, and higher‐order oligomers in the absence of CO, whereas the oligomers dissociated to domain‐swapped dimers in the presence of CO, demonstrating that the domain‐swapped dimer maintains the CO‐induced subunit dissociation behavior of native ACVP. These results suggest that protein oligomer formation may be controlled by utilizing domain swapping for a dimer–monomer transition protein.  相似文献   

5.
Studying the dimeric RNA structural organization is a step toward the understanding of retroviral genomic RNA dimerization. A kissing loop dimer is rearranged into an extended dimer during maturation of the virus particle. The extended dimer formation may be inhibited by ligands interacting with the RNA kissing loop dimer. A study was made of the interaction of dimeric RNA with paromomycin and magnesium ions. RNA dimers were formed from two hairpin RNAs having complementary sequences in the loop. The structural features of RNA dimers and the influence of the ligands were inferred from the fluorescence of 2-aminopurine (2-AP) incorporated in one of the two RNA hairpin sequences. As dimeric RNA interacted with paromomycin, 2-AP fluorescence increased. The increase was explained by a flipping of the fluorescent base out of the RNA structure. The binding constants and stoichiometry were estimated for dimeric RNA binding with paromomycin. An RNA dimer was found to interact with two paromomycin molecules; the binding constant was approximately the same (about 3 × 105 M−1) for both types of dimers. It was observed that the antibiotic and Mg2+ ions compete for binding to the hairpin RNA dimer and that one paromomycin molecule is displaced by one Mg2+ ion.  相似文献   

6.
The transient kinetics of aldehyde reduction by NADH catalyzed by liver alcohol dehydrogenase consist of two kinetic processes. This biphasic rate behavior is consistent with a model in which one of the two identical subunits in the enzyme is inactive during the reaction at the adjacent protomer. Alternatively, enzyme heterogeneity could result in such biphasic behavior. We have prepared liver alcohol dehydrogenase containing a single major isozyme; and the transient kinetics of this purified enzyme are biphasic.Addition of two [14C]carboxymethyl groups per dimer to the two “reactive” sulfhydryl groups (Cys46) yields enzyme which is catalytically inactive toward alcohol oxidation. Alkylated enzyme, as initially isolated by gel filtration chromatography at pH 7·5, forms an NAD+-pyrazole complex. However, the ability to bind NAD+-pyrazole is rapidly lost in pH 8·75 buffer; therefore, our alkylated preparations, as isolated by chromatography at pH 8·75, are inactive toward NAD+-pyrazole complex formation. We have prepared partially inactivated enzyme by allowing iodoacetic acid to react with liver alcohol dehydrogenase until 50% of the NAD+-pyrazole binding capacity remains; under these reaction conditions one [14C]carboxymethyl group is added per dimer. This partially alkylated enzyme preparation is isolated by gel filtration and has been aged sufficiently to lose NAD+-pyrazole binding ability at alkylated subunits. When solutions of native liver alcohol dehydrogenase and partially alkylated liver alcohol dehydrogenase containing the same number of unmodified active sites are allowed to react with substrate under single turnover conditions, partially alkylated enzyme is only half as reactive as native enzyme. This indicates that some molecular species in partially alkylated liver alcohol dehydrogenase that react with pyrazole and NAD+ during the active site titration do not react with substrate. These data are consistent with a model in which a subunit adjacent to an alkylated protomer in the dimeric enzyme is inactive toward substrate. In addition, NAD+-pyrazole binding at the protomers adjacent to alkylated subunits is slowly lost so that 75% of the enzyme-NAD+-pyrazole binding capacity is lost in 50% alkylated enzyme. These data supply strong evidence for subunit interactions in liver alcohol dehydrogenase.Binding experiments performed on partially alkylated liver alcohol dehydrogenase indicate that coenzyme binding is normal at a subunit adjacent to an alkylated protomer even though active ternary complexes cannot be formed. One hypothesis consistent with these results is the unavailability of zinc for substrate binding at the active site in subunits adjacent to alkylated protomers in monoalkylated dimer.  相似文献   

7.
d-lactate dehydrogenase has been purified from horseshoe crab (Limulus polyphemus) skeletal muscle and the seaworm (Nereis virens). The purified Limulus dehydrogenase was shown to be a dimer, with a molecular weight of approximately 70 000. Sephadex gel filtration and equilibrium sedimentation yield molecular weights of about 80 000 and 70 000 respectively. Acid dissociation yields a molecular weight species of about 35 000. The native enzyme has an so20w of 3.95. Extrapolation of para-hydroxymercuribenzoate inhibition curves to 100% inhibition corresponds to two molecules of para-hydroxymercuribenzoate bound per molecule of enzyme. Studies on the stoichiometric binding of reduced coenzyme show two molecules bound per molecule of enzyme. The number of tryptic peptides has been found to be one-half that expected from the amino acid composition. The electrophoretic pattern of isoenzymic forms can be best interpreted as suggesting that the enzyme is dimeric. In vitro high salt, freeze-thaw hybridizations of the isolated Limulus muscle isoenzymes yield the electrophoretic pattern predicted by a dimeric structure.The physical properties ot Nereis lactate dehydrogenase have been found to be similar to those for the Limulus muscle lactate dehydrogenase.  相似文献   

8.
Curcuma longa rhizome lectin, of non-seed origin having antifungal, antibacterial and α-glucosidase inhibitory activities, forms a homodimer with high thermal stability as well as acid tolerance. Size exclusion chromatography and dynamic light scattering show it to be a dimer at pH 7, but it converts to a monomer near pH 2. Circular dichroism spectra and fluorescence emission maxima are virtually indistinguishable from pH 7 to 2, indicating secondary and tertiary structures remain the same in dimer and monomer within experimental error. The tryptophan environment as probed by acrylamide quenching data yielded very similar data at pH 2 and pH 7, implying very similar folding for monomer and dimer. Differential scanning calorimetry shows a transition at 350.3 K for dimer and at 327.0 K for monomer. Thermal unfolding and chemical unfolding induced by guanidinium chloride for dimer are both reversible and can be described by two-state models. The temperatures and the denaturant concentrations at which one-half of the protein molecules are unfolded, are protein concentration-dependent for dimer but protein concentration-independent for monomer. The free energy of unfolding at 298 K was found to be 5.23 Kcal mol−1 and 14.90 Kcal mol−1 for the monomer and dimer respectively. The value of change in excess heat capacity upon protein denaturation (ΔCp) is 3.42 Kcal mol−1 K−1 for dimer. The small ΔCp for unfolding of CLA reflects a buried hydrophobic core in the folded dimeric protein. These unfolding experiments, temperature dependent circular dichroism and dynamic light scattering for the dimer at pH 7 indicate its higher stability than for the monomer at pH 2. This difference in stability of dimeric and monomeric forms highlights the contribution of inter-subunit interactions in the former.  相似文献   

9.
Cyanide binding to Chromatium vinosum ferricytochrome c′ has been studied to further investigate possible allosteric interactions between the subunits of this dimeric protein. Cyanide binding to C. vinosum cytochrome c′ appears to be cooperative. However, the cyanide binding reaction is unusual in that the overall affinity of cyanide increases as the concentration of cytochrome c′ decreases and that cyanide binding causes the ligated dimer to dissociate to monomers as shown by gel-filtration chromatography. Therefore, the cyanide binding properties of C. vinosum ferricytochrome c′ are complicated by a cyanide-linked dimer to monomer dissociation equilibrium of the complexed protein. The dimer to monomer dissociation constant is 20-fold smaller than that for CO linked dissociation constant of ferrocytochrome c′. Furthermore, the pH dependence of both the intrinsic equilibrium binding constant and the dimer to monomer equilibrium dissociation constant was investigated over the pH range of 7.0 to 9.2 to examine the effect of any ionizable groups. The equilibrium constants did not exhibit a significant pH dependence over this pH range.  相似文献   

10.
The dimeric nature of triosephosphate isomerases (TIMs) is maintained by an extensive surface area interface of more than 1600 Å2. TIMs from Trichomonas vaginalis (TvTIM) are held in their dimeric state by two mechanisms: a ball and socket interaction of residue 45 of one subunit that fits into the hydrophobic pocket of the complementary subunit and by swapping of loop 3 between subunits. TvTIMs differ from other TIMs in their unfolding energetics. In TvTIMs the energy necessary to unfold a monomer is greater than the energy necessary to dissociate the dimer. Herein we found that the character of residue I45 controls the dimer-monomer equilibrium in TvTIMs. Unfolding experiments employing monomeric and dimeric mutants led us to conclude that dimeric TvTIMs unfold following a four state model denaturation process whereas monomeric TvTIMs follow a three state model. In contrast to other monomeric TIMs, monomeric variants of TvTIM1 are stable and unexpectedly one of them (I45A) is only 29-fold less active than wild-type TvTIM1. The high enzymatic activity of monomeric TvTIMs contrast with the marginal catalytic activity of diverse monomeric TIMs variants. The stability of the monomeric variants of TvTIM1 and the use of cross-linking and analytical ultracentrifugation experiments permit us to understand the differences between the catalytic activities of TvTIMs and other marginally active monomeric TIMs. As TvTIMs do not unfold upon dimer dissociation, herein we found that the high enzymatic activity of monomeric TvTIM variants is explained by the formation of catalytic dimeric competent species assisted by substrate binding.  相似文献   

11.
Nucleotide pyrophosphatase of mung bean seedlings has earlier been isolated in our laboratory in a dimeric form (Mr 65,000) and has been shown to be converted to a tetramer by AMP and to a monomer by p-hydroxymercuribenzoate. All the molecular forms were enzymatically active with different kinetic properties. By a modified procedure using blue-Sepharose affinity chromatography, we have now obtained a dimeric form of the enzyme which is desensitized to AMP interaction. The molecular weight of the desensitized form of the enzyme was found to be the same as that of the native dimeric enzyme. However, the desensitized enzyme functioned with a linear time course, contrary to the biphasic time course exhibited by the native enzyme. In addition, it was not converted to a tetramer on the addition of AMP, had only one binding site for adenine nucleotides, and p-hydroxy-mercuribenzoate had no effect on the time course of the reaction or on the molecular weight of the enzyme. The temperature optimum of the desensitized enzyme was found to be 67 °C in contrast to the optimum of 49 °C for the native dimer. Fifty percent of the tryptophan residues of the desensitized enzyme were not accessible for quenching by iodide. Fluorescence studies gave Kd values of 0.34, 2.2, and 0.8 mm for AMP, ADP, and ATP, which were close to the Ki values of 0.12, 2.2, and 0.9 mm, respectively, for these nucleotides. The binding and inhibition studies with AMP and its analogs showed that the 6-amino group and the 5′-phosphate group were essential for the inhibition of the enzyme activity.  相似文献   

12.
The equilibrium unfolding transitions of Cro repressor variants, dimeric variant Cro F58W and monomer Cro K56[DGEVK]F58W, have been studied by urea and guanidine hydrochloride to probe the folding mechanism. The unfolding transitions of a dimeric variant are well described by a two state process involving native dimer and unfolded monomer with a free energy of unfolding, DeltaG(0,un)(0), of approximately 10-11 kcal/mol. The midpoint of transition curves is dependent on total protein concentration and DeltaG(0,un)(0) is independent of protein concentration, as expected for this model. Unfolding of Cro monomer is well described by the standard two state model. The stability of both forms of protein increases in the presence of salt but decreases with the decrease in pH. Because of the suggested importance of a N2<-->2F dimerization process in DNA binding, we have also studied the effect of sodium perchlorate, containing the chaotropic perchlorate anion, on the conformational transition of Cro dimer by CD, fluorescence and NMR (in addition to urea and guanidine hydrochloride) in an attempt both to characterize the thermodynamics of the process and to identify conditions that lead to an increase in the population of the folded monomers. Data suggest that sodium perchlorate stabilizes the protein at low concentration (<1.5 M) and destabilizes the protein at higher perchlorate concentration with the formation of a "significantly folded" monomer. The tryptophan residue in the "significantly folded" monomer induced by perchlorate is more exposed to the solvent than in native dimer.  相似文献   

13.
The anthranilate phosphoribosyltransferase from Sulfolobus solfataricus (ssAnPRT) forms a homodimer with a hydrophobic subunit interface. To elucidate the role of oligomerisation for catalytic activity and thermal stability of the enzyme, we loosened the dimer by replacing two apolar interface residues with negatively charged residues (mutations I36E and M47D). The purified double mutant I36E+M47D formed a monomer with wild-type catalytic activity but reduced thermal stability. The single mutants I36E and M47D were present in a monomer-dimer equilibrium with dissociation constants of about 1 μM and 20 μM, respectively, which were calculated from the concentration-dependence of their heat inactivation kinetics. The monomeric form of M47D, which is populated at low subunit concentrations, was as thermolabile as monomeric I36E+M47D. Likewise, the dimeric form of I36E, which was populated at high subunit concentrations, was as thermostable as dimeric wild-type ssAnPRT. These findings show that the increased stability of wild-type ssAnPRT compared to the I36E+M47D double mutant is not caused by the amino acid exchanges per se but by the higher intrinsic stability of the dimer compared to the monomer. In accordance with the negligible effect of the mutations on catalytic activity and stability, the X-ray structure of M47D contains only minor local perturbations at the dimer interface. We conclude that the monomeric double mutant resembles the individual wild-type subunits, and that ssAnPRT is a dimer for stability but not for activity reasons.  相似文献   

14.
The 21st amino acid, selenocysteine (Sec), is incorporated translationally into proteins and is synthesized on its specific tRNA (tRNASec). In Bacteria, the selenocysteine synthase SelA converts Ser-tRNASec, formed by seryl-tRNA synthetase, to Sec-tRNASec. SelA, a member of the fold-type-I pyridoxal 5′-phosphate-dependent enzyme superfamily, has an exceptional homodecameric quaternary structure with a molecular mass of about 500 kDa. Our previously determined crystal structures of Aquifex aeolicus SelA complexed with tRNASec revealed that the ring-shaped decamer is composed of pentamerized SelA dimers, with two SelA dimers arranged to collaboratively interact with one Ser-tRNASec. The SelA catalytic site is close to the dimer–dimer interface, but the significance of the dimer pentamerization in the catalytic site formation remained elusive. In the present study, we examined the quaternary interactions and demonstrated their importance for SelA activity by systematic mutagenesis. Furthermore, we determined the crystal structures of “depentamerized” SelA variants with mutations at the dimer–dimer interface that prevent pentamerization. These dimeric SelA variants formed a distorted and inactivated catalytic site and confirmed that the pentamer interactions are essential for productive catalytic site formation. Intriguingly, the conformation of the non-functional active site of dimeric SelA shares structural features with other fold-type-I pyridoxal 5′-phosphate-dependent enzymes with native dimer or tetramer (dimer-of-dimers) quaternary structures.  相似文献   

15.
 Using a corrected molar extinction coefficient for yeast apo copper-zinc superoxide dismutase (CuZnSOD), we have confirmed that the metal binding properties of this protein in vitro differ greatly from those of the bovine and human CuZnSOD enzymes. Thus yeast apo CuZnSOD was found to bind only one Co2+ per protein dimer under the conditions in which the bovine and human CuZnSOD apoenzymes readily bind two per dimer. The spectroscopic properties characteristic of the two Cu2+ plus two Co2+ per dimer or four Cu2+ per dimer metal-substituted bovine apo CuZnSOD derivatives were obtained for the yeast apoprotein but by the addition of only half of the appropriate metals, i.e., one Cu2+ plus one Co2+ per dimer or two Cu2+ per dimer. This half-metallated yeast CuZnSOD has been characterized by UV-visible and EPR spectroscopy as well as by native polyacrylamide gel electrophoresis. We conclude that yeast apo CuZnSOD, unlike the bovine and human apoproteins, cannot be reconstituted fully with metal ions under the same conditions. Instead, only one subunit of the homodimer, the "normal" subunit, can be remetalled in a fashion reminiscent of the well-characterized bovine protein. The other "phantom" subunit is not competent to bind metals in this fashion. Furthermore, we have shown that CuZnSOD protein isolated from Saccharomyces cerevisiae that lacks the gene coding for the copper chaperone, Lys7p, contains only one metal ion, Zn2+, per protein dimer. The possibility that yeast CuZnSOD can exist in multiple conformational states may represent an increased propensity of the yeast protein to undergo changes that can occur in all CuZnSODs, and may have implications for amyotrophic lateral sclerosis. Received: 8 June 1998 / Accepted: 9 September 1998  相似文献   

16.
CzcE is encoded by the most distal gene of the czc determinant that allows Cupriavidus metallidurans CH34 to modulate its internal concentrations of cobalt, zinc and cadmium by regulation of the expression of the efflux pump CzcCBA. We have overproduced and purified CzcE. CzcE is a periplasm-located dimeric protein able to bind specifically 4 Cu-equivalent per dimer. Spectrophotometry and EPR are indicative of type II copper with typical d-d transitions. Re-oxidation of fully reduced CzcE led to the formation of an air stable semi-reduced form binding both 2 Cu(I) and 2 Cu(II) ions. The spectroscopic characteristics of the semi-reduced form are different of those of the oxidized one, suggesting a change in the environment of Cu(II).  相似文献   

17.
Streptococcal pyrogenic exotoxin A (SpeA1) is a bacterial superantigen associated with scarlet fever and streptococcal toxic shock syndrome (STSS). SpeA1 is found in both monomeric and dimeric forms, and previous work suggested that the dimer results from an intermolecular disulfide bond between the cysteines at positions 90 of each monomer. Here, we present the crystal structure of the dimeric form of SpeA1. The toxin crystallizes in the orthorhombic space group P212121, with two dimers in the crystallographic asymmetric unit. The final structure has a crystallographic R-factor of 21.52% for 7248 protein atoms, 136 water molecules, and 4 zinc atoms (one zinc atom per molecule). The implications of SpeA1 dimer on MHC class II and T-cell receptor recognition are discussed.  相似文献   

18.
The enzyme β-galactosidase from Bacillus megaterium, strain KM has been purified by affinity chromatography. The enzyme was found to have a dimeric subunit structure, with the monomer having a molecular weight of 120,000. The Keq of the monomer-dimer equilibrium was strongly shifted towards dissociation in the isolated state. Inclusion of 5% sucrose in the buffer (and maintenance of the temperature at 5 °) minimized this dissociation. Molecularly homogeneous monomer and dimer could be prepared on sucrose gradients. The dimer was determined to have an S20,w of 8, while the monomer had an S20,w of 3. The amino acid composition was found to be similar to that of the E. coli β-galactosidase although significant differences occur. The activity of the monomer was studied by both urea-denaturation experiments and by immobilization of the monomer on Sepharose-4B. The monomer, bound to Sepharose-4B, was found to be inactive but still capable of binding the inhibitor thio-methyl galactoside. Activity was reconstituted by adding free monomer, in 8 M urea, to the Sepharose-bound monomer, followed by removal of the urea by dialysis. In addition, free monomers from E. coli β-galactosidase were found to form active hybrids with Sepharose-bound B. megaterium β-galactosidase monomers. We conclude on the basis of these studies that the free monomer is inactive, and that the dimer is the active species, in marked contrast to E. coli β-galactosidase where only the tetrameric form is active.  相似文献   

19.
Matrix enzymes are imported into peroxisomes and glyoxysomes, a subclass of peroxisomes involved in lipid mobilization. Two peroxisomal targeting signals (PTS), the C-terminal PTS1 and the N-terminal PTS2, mediate the translocation of proteins into the organelle. PTS2 processing upon import is conserved in higher eukaryotes, and in watermelon the glyoxysomal processing protease (GPP) was shown to catalyse PTS2 processing. GPP and its ortholog, the peroxisomal DEG protease from Arabidopsis thaliana (AtDEG15), belong to the Deg/HtrA family of ATP-independent serine proteases with Escherichia coli DegP as their prototype. GPP existes in monomeric and dimeric forms. Their equilibrium is shifted towards the monomer upon Ca2+-removal and towards the dimer upon Ca2+-addition, which is accompanied by a change in substrate specificity from a general protease (monomer) to the specific cleavage of the PTS2 (dimer). We describe the Ca2+/calmodulin (CaM) mediated dimerization of AtDEG15. Dimerization is mediated by the CaM-like protein AtCML3 as shown by yeast two and three hybrid analyses. The binding of AtCML3 occurs within the first 25 N-terminal amino acids of AtDEG15, a domain containing a predicted CaM-binding motif. Biochemical analysis of AtDEG15 deletion constructs in planta support the requirement of the CaM-binding domain for PTS2 processing. Phylogenetic analyses indicate that the CaM-binding site is conserved in peroxisomal processing proteases of higher plants (dicots, monocots) but not present in orthologs of animals or cellular slime molds. Despite normal PTS2 processing activity, an atcml3 mutant exhibited reduced 2,4-DB sensitivity, a phenotype previously reported for the atdeg15 mutant, indicating similarly impaired peroxisome metabolism.  相似文献   

20.
Reversible dissociation of the dimeric structure of brain pyridoxal kinase into subunits was attained by addition of guanidinium HCl (2 M). The molecular mass of the subunits (40 kDa) was determined by HPLC chromatography. Separation of the processes of refolding and association of the monomeric species was achieved by attaching the protein subunits to a rigid matrix (Affi-gel 15). The matrix-bound monomer is catalytically competent. The reaction of the crosslinking reagent 4,4'-dimaleimidestilbene 2,2'-disulfonate (DMDS), a derivatized stilbene, with the dimeric structure of pyridoxal kinase resulted in the formation of an oligomeric species of 80 kDa detectable by SDS-PAGE. The crosslinked subunits exhibit the same catalytic parameters as the native enzyme. The presence of two nucleotide-binding sites per dimer was determined by fluorimetric titrations using pyridoxyl-ATP, a strong competitive inhibitor with respect to ATP. The ATP analog binds with a Kd = 5 microM to each nucleotide site of the dimeric enzyme. The mode of binding pyridoxyl-ATP to the kinase is discussed in reference to a model which assumes the presence of two binding domains per subunit.  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