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.
A group of cyclic amidases, including hydantoinase, allantoinase, dihydropyrimidinase, and dihydroorotase, catalyze the reversible hydrolysis of cyclic ureides, such as 5-monosubstituted hydantoins and dihydropyrimidines. These four enzymes carry hydrophobic patches to form dimers. With the exception of dihydroorotase, these enzymes are further dimerized to form tetramers by hydrophobic interactions. This leads us to speculate that the hydrophobic interaction domain may be a significant factor in the catalytic property of these oligomeric cyclic amidases, for which activities are not allosterically regulated. We generated a dimeric D-hydantoinase by mutating five residues in the hydrophobic alpha-helical interface of a tetramer and analyzed the kinetic properties of the dimeric form of D-hydantoinase. The specific activity of the dimeric D-hydantoinase corresponds to 5.3% of the activity of tetrameric D-hydantoinase. This low specific activity of the dimeric D-hydantoinase indicates that the dimeric interaction to form a tetramer has a significant effect on the catalytic activity of this non-allosteric tetramer.  相似文献   

2.
To assess the importance of the C-terminal tail in the structure of aldehyde dehydrogenases (ALDH), mutants of tetrameric ALDH1 were generated by adding a tail of 5 amino acids (ALDH1-5aa) or the tail from the class 3 enzyme. A mutant of dimeric ALDH3 was made, where 17 amino acids from the C-terminus were deleted to generate ALDH3DeltaTail. The expression and solubility of the ALDH1 mutants was slightly lower than the wild type. Expression of ALDH3DeltaTail mutant was similar to wild type, but the solubility was only about 30%. The activity of ALDH1-5aa mutant was 30%, while ALDH1-H3Tail mutant was 60% active, compared to the wild type. The activity of the class 3 mutant was similar to the activity of the parent ALDH3 enzyme. Analysis of stability against temperature demonstrated that ALDH1-5aa was more stable than ALDH1 wild type, while the ALDH1-H3Tail mutant was considerably less stable than ALDH1, showing a stability similar to ALDH3. However, native gel and size exclusion analysis, showed no changes in the oligomerization state of these mutants. ALDH3DeltaTail mutant was more stable than wild type; the stability against temperature was similar to ALDH1. The ALDH3DeltaTail mutant showed an elution similar to that of ALDH1 from the size exclusion column, indicating that it was possibly a tetramer. These results show that the tail in ALDH3, is involved in the determination of the quaternary structure of ALDH3, but has no effect on the ALDH1 enzyme; the absence of the C-terminal tail is not the only factor participating in holding the dimers together. Thus, the interaction between single residues, or interactions with the N-terminal region might be more important for maintaining stable tetramers.  相似文献   

3.
To assess the importance of the C-terminal tail in the structure of aldehyde dehydrogenases (ALDH), mutants of tetrameric ALDH1 were generated by adding a tail of 5 amino acids (ALDH1-5aa) or the tail from the class 3 enzyme. A mutant of dimeric ALDH3 was made, where 17 amino acids from the C-terminus were deleted to generate ALDH3ΔTail. The expression and solubility of the ALDH1 mutants was slightly lower than the wild type. Expression of ALDH3ΔTail mutant was similar to wild type, but the solubility was only about 30%. The activity of ALDH1-5aa mutant was 30%, while ALDH1-H3Tail mutant was 60% active, compared to the wild type. The activity of the class 3 mutant was similar to the activity of the parent ALDH3 enzyme. Analysis of stability against temperature demonstrated that ALDH1-5aa was more stable than ALDH1 wild type, while the ALDH1-H3Tail mutant was considerably less stable than ALDH1, showing a stability similar to ALDH3. However, native gel and size exclusion analysis, showed no changes in the oligomerization state of these mutants. ALDH3ΔTail mutant was more stable than wild type; the stability against temperature was similar to ALDH1. The ALDH3ΔTail mutant showed an elution similar to that of ALDH1 from the size exclusion column, indicating that it was possibly a tetramer. These results show that the tail in ALDH3, is involved in the determination of the quaternary structure of ALDH3, but has no effect on the ALDH1 enzyme; the absence of the C-terminal tail is not the only factor participating in holding the dimers together. Thus, the interaction between single residues, or interactions with the N-terminal region might be more important for maintaining stable tetramers.  相似文献   

4.
Nucleoside diphosphate kinase (NDPK) catalyzes the transfer of gamma-phosphate from nucleoside triphosphates to nucleoside diphosphates. The subunit folding and the dimeric basic structural unit are remarkably the same for available structures but, depending on species, dimers self-associate to form hexamers or tetramers. The crystal structure of the Escherichia coli NDPK reveals a new tetrameric quaternary structure for this protein family. The two tetramers differ by the relative orientation of interacting dimers, which face either the convex or the concave side of their central sheet as in either Myxococcus xanthus (type I) or E. coli (type II), respectively. In the type II tetramer, the subunits interact by a new interface harboring a zone called the Kpn loop as in hexamers, but by the opposite face of this loop. The evolutionary conservation of the interface residues indicates that this new quaternary structure seems to be the most frequent assembly mode in bacterial tetrameric NDP kinases.  相似文献   

5.
The three-dimensional structure of four malate dehydrogenases (MDH) from thermophilic and mesophilic phototropic bacteria have been determined by X-ray crystallography and the corresponding structures compared. In contrast to the dimeric quaternary structure of most MDHs, these MDHs are tetramers and are structurally related to tetrameric malate dehydrogenases from Archaea and to lactate dehydrogenases. The tetramers are dimers of dimers, where the structures of each subunit and the dimers are similar to the dimeric malate dehydrogenases. The difference in optimal growth temperature of the corresponding organisms is relatively small, ranging from 32 to 55 degrees C. Nevertheless, on the basis of the four crystal structures, a number of factors that are likely to contribute to the relative thermostability in the present series have been identified. It appears from the results obtained, that the difference in thermostability between MDH from the mesophilic Chlorobium vibrioforme on one hand and from the moderate thermophile Chlorobium tepidum on the other hand is mainly due to the presence of polar residues that form additional hydrogen bonds within each subunit. Furthermore, for the even more thermostable Chloroflexus aurantiacus MDH, the use of charged residues to form additional ionic interactions across the dimer-dimer interface is favored. This enzyme has a favorable intercalation of His-Trp as well as additional aromatic contacts at the monomer-monomer interface in each dimer. A structural alignment of tetrameric and dimeric prokaryotic MDHs reveal that structural elements that differ among dimeric and tetrameric MDHs are located in a few loop regions.  相似文献   

6.
Cytoplasmic serine hydroxymethyltransferase (cSHMT) is a tetrameric, pyridoxal phosphate (PLP)-dependent enzyme that catalyzes the reversible interconversion of serine and tetrahydrofolate to glycine and methylenetetrahydrofolate. The enzyme has four active sites and is best described as a dimer of obligate dimers. Each monomeric subunit within the obligate dimer contributes catalytically important amino acid residues to both active sites. To investigate the interchange of subunits among cSHMT tetramers, a dominant-negative human cSHMT enzyme (DNcSHMT) was engineered by making three amino acid substitutions: K257Q, Y82A, and Y83F. Purified recombinant DNcSHMT protein was catalytically inactive and did not bind 5-formyltetrahydrofolate. Coexpression of the cSHMT and DNcSHMT proteins in bacteria resulted in the formation of heterotetramers with a cSHMT/DNcSHMT subunit ratio of 1. Characterization of the cSHMT/DNcSHMT heterotetramers indicates that DNcSHMT and cSHMT monomers randomly associate to form tetramers and that cSHMT/DNcSHMT obligate dimers are catalytically inactive. Incubation of recombinant cSHMT protein with recombinant DNcSHMT protein did not result in the formation of hetero-oligomers, indicating that cSHMT subunits do not exchange once the tetramer is assembled. However, removal of the active site PLP cofactor does permit exchange of obligate dimers among preformed cSHMT and DNcSHMT tetramers, and the formation of heterotetramers from cSHMT and DNcSHMT homodimers does not affect the activity of the cSHMT homodimers. The results of these studies demonstrate that PLP inhibits dimer exchange among cSHMT tetramers and suggests that cellular PLP concentrations may influence the stability of cSHMT protein in vivo.  相似文献   

7.
Yeast NAD(+)-specific isocitrate dehydrogenase (IDH) is an octameric enzyme composed of four heterodimers of regulatory IDH1 and catalytic IDH2 subunits. The crystal structure suggested that the interactions between tetramers in the octamer are restricted to defined regions in IDH1 subunits from each tetramer. Using truncation and mutagenesis, we constructed three tetrameric forms of IDH. Truncation of five residues from the amino terminus of IDH1 did not alter the octameric form of the enzyme, but this truncation with an IDH1 G15D or IDH1 D168K residue substitution produced tetrameric enzymes as assessed by sedimentation velocity ultracentrifugation. The IDH1 G15D substitution in the absence of any truncation of IDH1 was subsequently found to be sufficient for production of a tetrameric enzyme. The tetrameric forms of IDH exhibited ~50% reductions in V(max) and in cooperativity with respect to isocitrate relative to those of the wild-type enzyme, but they retained the property of allosteric activation by AMP. The truncated (-5)IDH1/IDH2 and tetrameric enzymes were much more sensitive than the wild-type enzyme to inhibition by the oxidant diamide and concomitant formation of a disulfide bond between IDH2 Cys-150 residues. Binding of ligands reduced the sensitivity of the wild-type enzyme to diamide but had no effect on inhibition of the truncated or tetrameric enzymes. These results suggest that the octameric structure of IDH has in part evolved for regulation of disulfide bond formation and activity by ensuring the proximity of the amino terminus of an IDH1 subunit of one tetramer to the IDH2 Cys-150 residues in the other tetramer.  相似文献   

8.
Abstract: Acetylcholinesterase (AChE) is secreted from muscle and nerve cells and associates as multimers through intermolecular covalent and noncovalent bonds. The amino acid sequence of the C-terminus is thought to play an important role in these interactions. We generated mutants in the C-terminus of the catalytic T-subunit of chicken AChE to determine the importance of this region to oligomerization and to the amphipathic character of the protein. Wild-type recombinant chicken AChE secreted from human embryonic kidney 293 cells was assembled into dimers and tetramers exclusively. Mutants lacking the C-terminal Cys764, the only cysteine involved in interchain disulfide bonds, showed lower but significant levels of the secreted dimeric and tetrameric forms. A truncated mutant, lacking the C-terminal 39 amino acids, exhibited a severe decrease in content of the multimeric forms, yet small amounts of the dimer were detectable. The amphipathic character was dependent on the state of oligomerization. When analyzed by sucrose gradients, the sedimentation of tetramers was not affected by detergent, but monomers and dimers sedimented more slowly in the presence of detergent. Most of the recombinant wild-type enzyme, shown to be dimeric and tetrameric by sedimentation analysis, was monomeric when analyzed by sodium dodecyl sulfate-polyacrylamide gel electrophoresis under nonreducing conditions, indicating that much of the secreted oligomeric AChE was not disulfide bonded. These data suggest that disulfide bonding of Cys764 is not required for the catalytic subunit of chicken AChE to form oligomers and that regions outside of the C-terminus contribute to the hydrophobic interactions that are important for stabilizing the oligomeric forms.  相似文献   

9.
L-Malate (MalDH) and L-lactate (LDH) dehydrogenases belong to the same family of NAD-dependent enzymes. LDHs are tetramers, whereas MalDHs can be either dimeric or tetrameric. To gain insight into molecular relationships between LDHs and MalDHs, we studied folding intermediates of a mutant of the LDH-like MalDH (a protein with LDH-like structure and MalDH enzymatic activity) from the halophilic archaeon Haloarcula marismortui (Hm MalDH). Crystallographic analysis of Hm MalDH had shown a tetramer made up of two dimers interacting mainly via complex salt bridge clusters. In the R207S/R292S Hm MalDH mutant, these salt bridges are disrupted. Its structural parameters, determined by neutron scattering and analytical centrifugation under different conditions, showed the protein to be a tetramer in 4 M NaCl. At lower salt concentrations, stable oligomeric intermediates could be trapped at a given pH, temperature, or NaCl solvent concentration. The spectroscopic properties and enzymatic behavior of monomeric, dimeric, and tetrameric species were thus characterized. The properties of the dimeric intermediate were compared to those of dimeric intermediates of LDH and dimeric MalDHs. A detailed analysis of the putative dimer-dimer contact regions in these enzymes provided an explanation of why some can form tetramers and others cannot. The study presented here makes Hm MalDH the best characterized example so far of an LDH-like MalDH.  相似文献   

10.
Tetrameric D-glyceraldehyde-3-phosphate dehydrogenase (EC 1.2.1.12) isolated from rabbit skeletal muscle was covalently bound to CNBr-activated Sepharose 4B via a single subunit. Catalytically active immobilized dimer and monomeric forms of the enzyme were prepared after urea-induced dissociation of the tetramer. A study of the coenzyme-binding properties of matrix-bound tetrameric, dimeric and monomeric species has shown that: (1) an immobilized tetramer binds NAD+ with negative cooperativity, the dissociation constants being 0.085 microM for the first two coenzyme molecules and 1.3 microM for the third and the fourth one; (2) coenzyme binding to the dimeric enzyme form also displays negative cooperativity with Kd values of 0.032 microM and 1.1 microM for the first and second sites, respectively; (3) the binding of NAD+ to a monomer can occur with a dissociation constant of 1.6 microM which is close to the Kd value for low-affinity coenzyme binding sites of the tetrameric or dimeric enzyme forms. In the presence of NAD+ an immobilized monomer acquires a stability which is not inferior to that of a holotetramer. The catalytic properties of monomeric and tetrameric enzyme forms were compared and found to be different under certain conditions. Thus, the monomers of rabbit muscle D-glyceraldehyde-3-phosphate dehydrogenase displayed a hyperbolic kinetic saturation curve for NAD+, whereas the tetramers exhibited an intermediary plateau region corresponding to half-saturating concentrations of NAD+. At coenzyme concentrations below half-saturating a monomer is more active than a tetramer. This difference disappears at saturating concentrations of NAD+. Immobilized monomeric and tetrameric forms of D-glyceraldehyde-3-phosphate dehydrogenase from baker's yeast were also used to investigate subunit interactions in catalysis. The rate constant of inactivation due to modification of essential arginine residues in the holoenzyme decreased in the presence of glyceraldehyde 3-phosphate, probably as a result of conformational changes accompanying catalysis. This effect was similar for monomeric and tetrameric enzyme forms at saturating substrate concentrations, but different for the two enzyme species under conditions in which about one-half of the active centers remained unsaturated. Taken together, the results indicate that association of D-glyceraldehyde-3-phosphate dehydrogenase monomers into a tetramer imposes some constraints on the functioning of the active centers.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

11.
Two critical steps controlling serine recombinase activity are the remodeling of dimers into the chemically active synaptic tetramer and the regulation of subunit rotation during DNA exchange. We identify a set of hydrophobic residues within the oligomerization helix that controls these steps by the Hin DNA invertase. Phe105 and Met109 insert into hydrophobic pockets within the catalytic domain of the same subunit to stabilize the inactive dimer conformation. These rotate out of the catalytic domain in the dimer and into the subunit rotation interface of the tetramer. About half of residue 105 and 109 substitutions gain the ability to generate stable synaptic tetramers and/or promote DNA chemistry without activation by the Fis/enhancer element. Phe106 replaces Phe105 in the catalytic domain pocket to stabilize the tetramer conformation. Significantly, many of the residue 105 and 109 substitutions support subunit rotation but impair ligation, implying a defect in rotational pausing at the tetrameric conformer poised for ligation. We propose that a ratchet-like surface involving Phe105, Met109 and Leu112 within the rotation interface functions to gate the subunit rotation reaction. Hydrophobic residues are present in analogous positions in other serine recombinases and likely perform similar functions.  相似文献   

12.
In mitochondria, the hydrolytic activity of ATP synthase is regulated by an inhibitor protein, IF(1). Its binding to ATP synthase depends on pH, and below neutrality, IF(1) is dimeric and forms a stable complex with the enzyme. At higher pH values, IF(1) forms tetramers and is inactive. In the 2.2 A structure of the bovine IF(1) described here, the four monomers in the asymmetric unit are arranged as a dimer of dimers. Monomers form dimers via an antiparallel alpha-helical coiled coil in the C-terminal region. Dimers are associated into oligomers and form long fibres in the crystal lattice, via coiled-coil interactions in the N-terminal and inhibitory regions (residues 14-47). Therefore, tetramer formation masks the inhibitory region, preventing IF(1) binding to ATP synthase.  相似文献   

13.
Phenylalanine hydroxylase is regulated in a complex manner, including activation by phosphorylation. It is normally found as an equilibrium of dimeric and tetrameric species, with the tetramer thought to be the active form. We converted the protein to the dimeric form by deleting the C-terminal 24 residues and show that the truncated protein remains active and regulated by phosphorylation. This indicates that changes in the tetrameric quaternary structure of phenylalanine hydroxylase are not required for enzyme activation. Truncation also facilitates crystallization of both phosphorylated and dephosphorylated forms of the enzyme.  相似文献   

14.
We show that human and bovine dopamine beta-hydroxylases (DBH) exist under three main molecular forms: a soluble nonamphiphilic form and two amphiphilic forms. Sedimentation in sucrose gradients and electrophoresis under nondenaturing conditions, by comparison with acetylcholinesterase (AChE), suggest that the three forms are tetramers of the DBH catalytic subunit and bind either no detergent, one detergent micelle, or two detergent micelles. By analogy with the Gna4 and Ga4 AChE forms, we propose to call the nonamphiphilic tetramer Dna4 and the amphiphilic tetramers Da4I and Da4II. In addition to the major tetrameric forms, DBH dimers occur as very minor species, both amphiphilic and nonamphiphilic. Reduction under nondenaturing conditions leads to a partial dissociation of tetramers into dimers, retaining their amphiphilic character. This suggests that the hydrophobic domain is not linked to the subunits through disulfide bonds. The two amphiphilic tetramers are insensitive to phosphatidylinositol phospholipase C, but may be converted into soluble DBH by proteolysis in a stepwise manner; Da4II----Da4I----Dna4. Incubation of soluble DBH with various phospholipids did not produce any amphiphilic form. Several bands corresponding to the catalytic subunits of bovine DBH were observed in sodium dodecyl sulfate-polyacrylamide gel electrophoresis, but this multiplicity was not simply correlated with the amphiphilic character of the enzyme. In the case of human DBH, we observed two bands of 78 and 84 kDa. As previously reported by others, the presence of the heavy subunit characterizes the amphiphilic forms of the enzyme. We discuss the nature of the hydrophobic domain, which could be an uncleaved signal peptide, and the organization of the different amphiphilic and nonamphiphilic DBH forms. We present two models in which dimers may possess either one hydrophobic domain or two domains belonging to each subunit; in both cases, a single detergent micelle would be bound per dimer.  相似文献   

15.
Catalase (hydroperoxidase) HPII of Escherichia coli is the largest catalase so far characterized, existing as a homotetramer of 84 kDa subunits. Each subunit has a core structure that closely resembles small subunit catalases, supplemented with an extended N-terminal sequence and compact flavodoxin-like C-terminal domain. Treatment of HPII with trypsin, chymotrypsin, or proteinase K, under conditions of limited digestion, resulted in cleavage of 72-74 residues from the N-terminus of each subunit that created a homotetramer of 76 kDa subunits with 80% of wild-type activity. Longer treatment with proteinase K removed the C-terminal domain, producing a transient 59 kDa subunit which was subsequently cleaved into two fragments, 26 and 32 kDa. The tetrameric structure was retained despite this fragmentation, with four intermediates being observed between the 336 kDa native form and the 236 kDa fully truncated form corresponding to tetramers with a decreasing complement of C-termini (4, 3, 2, and 1). The truncated tetramers retained 80% of wild-type activity. The T(m) for loss of activity during heating was decreased from 85 to 77 degrees C by removal of the N-terminal sequence and to 59 degrees C by removal of the C-terminal domain, revealing the importance of the C-terminal domain in enzyme stability. The sites of cleavage were determined by N- and C-terminal sequencing, and two were located on the surface of the tetramer with a third being exposed by removal of the C-terminal domain.  相似文献   

16.
Type II restriction endonuclease BsaWI recognizes a degenerated sequence 5′-W/CCGGW-3′ (W stands for A or T, ‘/’ denotes the cleavage site). It belongs to a large family of restriction enzymes that contain a conserved CCGG tetranucleotide in their target sites. These enzymes are arranged as dimers or tetramers, and require binding of one, two or three DNA targets for their optimal catalytic activity. Here, we present a crystal structure and biochemical characterization of the restriction endonuclease BsaWI. BsaWI is arranged as an ‘open’ configuration dimer and binds a single DNA copy through a minor groove contacts. In the crystal primary BsaWI dimers form an indefinite linear chain via the C-terminal domain contacts implying possible higher order aggregates. We show that in solution BsaWI protein exists in a dimer-tetramer-oligomer equilibrium, but in the presence of specific DNA forms a tetramer bound to two target sites. Site-directed mutagenesis and kinetic experiments show that BsaWI is active as a tetramer and requires two target sites for optimal activity. We propose BsaWI mechanism that shares common features both with dimeric Ecl18kI/SgrAI and bona fide tetrameric NgoMIV/SfiI enzymes.  相似文献   

17.
Bovine IF(1), a basic protein of 84 amino acids, is involved in the regulation of the catalytic activity of the F(1) domain of ATP synthase. At pH 6.5, but not at basic pH values, it inhibits the ATP hydrolase activity of the enzyme. The oligomeric state of bovine IF(1) has been investigated at various pH values by sedimentation equilibrium analytical ultracentrifugation and by covalent cross-linking. Both techniques confirm that the protein forms a tetramer at pH 8, and below pH 6.5, the protein is predominantly dimeric. By covalent cross-linking, it has been found that at pH 8.0 the fragment of IF(1) consisting of residues 44-84 forms a dimer, whereas the fragment from residues 32-84 is tetrameric. Therefore, some or all of the residues between positions 32 and 43 are necessary for tetramer formation and are involved in the pH-sensitive interconversion between dimer and tetramer. One important residue in the interconversion is histidine 49. Mutation of this residue to lysine abolishes the pH-dependent activation-inactivation, and the mutant protein is active and dimeric at all pH values investigated. It is likely from NMR studies that the inhibitor protein dimerizes by forming an antiparallel alpha-helical coiled-coil over its C-terminal region and that at high pH values, where the protein is tetrameric, the inhibitory regions are masked. The mutation of histidine 49 to lysine is predicted to abolish coiled-coil formation over residues 32-43 preventing interaction between two dimers, forcing the equilibrium toward the dimeric state, thereby freeing the N-terminal inhibitory regions and allowing them to interact with F(1).  相似文献   

18.
Structural properties of two similar beta-galactosidase fragments were investigated to determine how they influence the fragments' degradation rate in Escherichia coli. Both fragments resulting from a C-terminal nonsense mutation in lacZ, the CSH11 polypeptide and its 90 kDa degradative intermediate, exist predominantly as monomer subunits instead of in the tetrameric form characteristic of the native enzyme. However, both fragments appear to produce trace amounts of dimers and tetramers. The tetramer and higher molecular weight aggregates formed by the wild-type subunit confer greater protection for the enzyme's N-terminal auto-alpha polypeptide than does the monomer state of the beta-galactosidase fragments. The thermally induced aggregation of both beta-galactosidase fragments correlates with their sensitivity to alpha-chymotrypsin. The relatively low thermal stability of the 90 kDa degradative intermediate appears to be the cause of the significant increase in its proteolytic susceptibility at moderately high temperatures.  相似文献   

19.
HIV-1 Nef modulates disease progression through interactions with over 30 host proteins. Individual chains fold into membrane-interacting N-terminal and C-terminal core (Nefcore) domains respectively. Nef exists as small oligomers near membranes and associates into higher oligomers such as tetramers or hexadecamers in the cytoplasm. Earlier structures of the Nefcore in apo and complexed forms with the Fyn-kinase SH3 domain revealed dimeric association details and the role of the conserved PXXP recognition motif (residues 72–78) of Nef in SH3-domain interactions. The crystal structure of the tetrameric Nef reported here corresponds to the elusive cytoplasmic stage. Comparative analyses show that subunits of Nefcore dimers (open conformation) swing out with a relative displacement of ∼22 Å and rotation of ∼174° to form the ‘closed’ tetrameric structure. The changes to the association are around Asp125, a conserved residue important for viral replication and the important XR motif (residues 107–108). The tetramer associates through C4 symmetry instead of the 222 symmetry expected when two dimers associate together. This novel dimer-tetramer transition agrees with earlier solution studies including small angle X-ray scattering, analytical ultracentrifugation, dynamic laser light scattering and our glutaraldehyde cross-linking experiments. Comparisons with the Nefcore—Fyn-SH3 domain complexes reveal that the PXXP motif that interacts with the SH3-domain in the dimeric form is sterically occluded in the tetramer. However the 151–180 loop that is distal to the PXXP motif and contains several protein interaction motifs remains accessible. The results suggest how changes to the oligomeric state of Nef can help it distinguish between protein partners.  相似文献   

20.
Abstract: The hydrophilic, salt-soluble (SS) form of acetylcholinesterase (AChE) from bovine brain caudate nucleus exists mainly as a tetramer sedimenting at 10.3S (∼40%), and a monomer sedimenting at 3.4S (∼60%). The enzyme is N -glycosylated and contains similar HNK-1 carbohydrates as detergent-soluble (DS) AChE. No O-linked carbohydrates could be detected. Amino acid sequencing showed that the N terminus of SS-AChE is identical to that of DS-AChE. In tetrameric SS-AChE, two pairs of disulfide-linked dimers are associated by hydrophobic forces located in the C terminus. Antibodies were raised against a peptide identical to the last 10 amino acid residues of bovine brain DS-AChE. The peptide included the sequence of residues 574–583 (H-Tyr-Ser-Lys-Gln-Asp-Arg-Cys-Ser-Asp-Leu-OH) of the enzyme. The antibodies cross-reacted with tetrameric, but not with monomeric, SS-AChE, showing that in the latter form, the C terminus is truncated. Limited proteolysis of tetrameric SS-AChE at the C terminus led to the formation of an enzymatically active monomer, which did not react with anti-C-terminal antibody. Although the DS form of AChE contains a structural subunit that serves as membrane anchor, no anchor was detected in SS-AChE. Enzyme antigen immunoassays showed that SS-AChE reacted with all monoclonal antibodies directed against the catalytic subunit of DS-AChE, but not with monoclonal antibodies targeting the membrane-anchored subunits. From our results, we conclude that SS-AChE utilizes the same alternative splicing pattern as DS-AChE, leading to tetrameric SS-AChE devoid of the membrane anchor. The active monomer of SS-AChE is most likely derived from tetrameric forms by limited postsynthetic proteolysis.  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