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 15 毫秒
1.
We constructed a single-chain antibody fragment (scFv) of murine monoclonal antibody, MABL, which specifically bound to human CD47 (hCD47) and induced apoptosis of the leukemic cells. The scFv of MABL antibody with a 15-residue linker (MABL scFv-15) formed both dimer (Mr 50 kDa) and monomer (Mr 25 kDa). Both MABL scFv-15 dimer and monomer had binding activity for hCD47. MABL scFv-15 dimer strongly induced apoptosis of hCD47-introduced mouse leukemic cells in vitro and exhibited anti-tumor effect in a myeloma transplanted mice model. However, MABL scFv-15 monomer scarcely exhibited these activities. These results strongly demonstrate that the ligation of CD47 antigen by two antigen-binding sites of MABL dimer is needed for inducing apoptosis. The parent MABL antibody caused hemagglutination due to the CD47 expressed on erythrocytes. Interestingly, MABL scFv-15 dimer did not cause hemagglutination. This apoptosis-inducing dimer appears to be a lead candidate for novel leukemic therapy.  相似文献   

2.
R Zhang  G H Snyder 《Biochemistry》1988,27(10):3785-3794
Rate constants have been determined in 3 M guanidine hydrochloride for disulfide exchange reactions between glutathione and two synthetic peptides containing a cysteine-valine-cysteine region. Equilibrium experiments demonstrate the absence of noncovalent peptide aggregation in this solvent. Procedures are given for separating seven different components in quenched reactions, including the fully reduced cysteine cluster, the monomeric disulfide loop, parallel and antiparallel dimer loops, and the three monomers containing one or two mixed disulfides with glutathione. Intramolecular rate constants for (1) formation of a sterically strained monomer loop, (2) transfer of glutathione between the two cysteines on the same peptide chain, and (3) formation of unstrained dimer loops correspond to a series of processes forming rings of increasing size. In one sequence, these rate constants are 3, 6, and about 21 s-1, respectively. The larger loops are formed more easily. In the other sequence, rate constants for formation and opening of monomer loops are accelerated 180- and 1300-fold, respectively, relative to analogous reactions in a peptide containing eight residues between the two cysteines. This gives a 7-fold smaller equilibrium constant for ring closure in the cysteine cluster. Dimer formation occurs by a mechanism utilizing the accelerated opening of monomer loops. Results provide information assisting efforts to develop strategies for directing disulfide pairing in novel protein structures. Results also help define factors contributing to formation of undesired oligomers during efforts to refold cysteine-containing proteins obtained by bacterial expression of mammalian genes.  相似文献   

3.
Diacylglycerol acyltransferases (DGATs) esterify sn-1,2-diacylglycerol with a long-chain fatty acyl-CoA, the last and rate-limiting step of triacylglycerol (TAG) biosynthesis in eukaryotic organisms. At least 74 DGAT2 sequences from 61 organisms have been identified, but the expression of any DGAT2 as a partial or full-length protein in Escherichia coli had not been reported. The main objective of this study was to express and purify recombinant DGAT2 (rDGAT2) from E. coli for antigen production with a minor objective to compare rDGAT2 expression in yeast. A plasmid was engineered to express tung tree DGAT2 fused to maltose binding protein and poly-histidine (His) affinity tags. Immunoblotting showed that rDGAT2 was detected in the soluble, insoluble, and membrane fractions. The rDGAT2 in the soluble fraction was partially purified by amylose resin, nickel-nitrilotriacetic agarose (Ni-NTA) beads, and tandem affinity chromatography. Multiple proteins co-purified with rDGAT2. Size exclusion chromatography estimated the size of the rDGAT2-enriched fraction to be approximately eight times the monomer size. Affinity-purified rDGAT2 fractions had a yellow tint and contained fatty acids. The rDGAT2 in the insoluble fraction was partially solubilized by seven detergents with SDS being the most effective. Recombinant DGAT2 was purified to near homogeneity by SDS solubilization and Ni-NTA affinity chromatography. Mass spectrometry identified rDGAT2 as a component in the bands corresponding to the monomer and dimer forms as observed by SDS-PAGE. Protein bands with monomer and dimer sizes were also observed in the microsomal membranes of Saccharomyces cerevisiae expressing hemagglutinin-tagged DGAT2. Nonradioactive assay showed TAG synthesis activity of DGAT2 from yeast but not E. coli. The results suggest that rDGAT2 is present as monomer and dimer forms on SDS-PAGE, associated with other proteins, lipids, and membranes, and that post-translational modification of rDGAT2 may be required for its enzymatic activity and/or the E. coli protein is misfolded.  相似文献   

4.
Binding of the 58 kDa monomer and 44 kDa alpha beta dimer forms of terminal deoxynucleotidyl transferase to double stranded DNA was demonstrated by gel retardation and tryptophan fluorescence quenching. The dissociation constants and cooperativity parameters were similar to those that have been determined for binding of these two forms of terminal transferase to single stranded DNA. However, the double stranded DNA binding site size of 10 nucleotides was half the size expected. The efficacy of blunt ended DNA as an initiator in the polymerization reaction catalyzed by terminal transferase was demonstrated by radiometric assays and product analyses on agarose gels. The initial reaction kinetics indicated that dGTP but not dATP was added efficiently to a blunt double stranded DNA 3' end. These results are correlated with current models for in vivo terminal transferase function.  相似文献   

5.
A naturally occurring tandem duplication of the 7-kDa type III antifreeze protein from Antarctic eel pout (Lycodichthys dearborni) is twice as active as the monomer in depressing the freezing point of a solution. We have investigated the basis for this enhanced activity by producing recombinant analogues of the linked dimer that assess the effects of protein size and the number and area of the ice-binding site(s). The recombinant dimer connected by a peptide linker had twice the activity of the monomer. When one of the two ice-binding sites was inactivated by site-directed mutagenesis, the linked dimer was only 1.2 times more effective than the monomer. When the two monomers were linked through a C-terminal disulfide bond in such a way that their two ice-binding sites were opposite each other and unable to engage the same ice surface simultaneously, the dimer was again only 1.2 times as active as the monomer. We conclude from these analyses that the enhanced activity of the dimer stems from the two ice-binding sites being able to engage to ice at the same time, effectively doubling the area of the ice-binding site.  相似文献   

6.
A number of different monomer and dimer derivatives of protein L7/L12 has been studied in EF-G-dependent reactions on the ribosome. It has been shown that only dimer derivatives of protein L7/L12 are able to interact with the ribosome. This means that it is the dimer forms of protein L7/L12 that are present in the functionally active ribosome. It is likely that the N-terminal sequence of protein L7/L12 is responsible for dimerization of the protein in solution and at the same time contributes mainly to the interaction of the protein L7/L12 dimer with the ribosome. The results obtained suggest that there are four copies of protein L7/L12 in the translating ribosome.  相似文献   

7.
In our isolation procedure, the surface antigens of the variants AnTat 1.1 and 1.10 (Trypanosoma brucei brucei) are essentially obtained as a disulfide-linked dimer while the AnTat 1.8 surface antigen is found as a mixture of monomer and disulfide-linked dimer. This observation may be related to the localization of the cysteine residues in the protein sequences. In the purification procedure using concanavalin-A Sepharose chromatography, besides the VSG elution by methyl-alpha-D-mannopyranoside, a quantitative elution of still bound VSG may be obtained by the addition of beta-mercaptoethanol to methyl-alpha-D-mannopyrannoside in the elution buffer. The surface antigen of the variant AnTat 1.1 was examined for molecular form at several different times during the release procedure. The disulfide-linked dimer could be observed within 30 min of the surface coat release, indicating its presence within the parasite.  相似文献   

8.
The ATM protein kinase, is a serine/threonine protein kinase that is recruited and activated by DNA double-strand breaks, mediates responses to ionizing radiation in mammalian cells. Here we show that ATM is held inactive in unirradiated cells as a dimer and phosphorylates the opposite strand of the dimer in response to DNA damage. Cellular irradiation induces rapid intermolecular autophosphorylation of serine 1981 that causes dimer dissociation and initiates cellular ATM kinase activity. ATM cannot phosphorylate the substrates when it could not undergo dimer monomer transition. After DNA repair, the active monomer will undergo dephosphorylation to form dimer again and dephosphorylation is critical for dimer re-formation. Our work reveals novel function of ATM dimer monomer transition and explains why ATM dimer monomer transition plays such important role for ATM cellular activity during DNA repair.  相似文献   

9.
The finding that exchange of tubulin subunits between tubulin dimers (alpha-beta + alpha'beta' <--> alpha'beta + alphabeta') does not occur in the absence of protein cofactors and GTP hydrolysis conflicts with the assumption that pure tubulin dimer and monomer are in rapid equilibrium. This assumption underlies the many physical chemical measurements of the K(d) for dimer dissociation. To resolve this discrepancy we used surface plasmon resonance to determine the rate constant for dimer dissociation. The half-time for dissociation was approximately 9.6 h with tubulin-GTP, 2.4 h with tubulin-GDP, and 1.3 h in the absence of nucleotide. A Kd equal to 10(-11) M was calculated from the measured rate for dissociation and an estimated rate for association. Dimer dissociation was found to be reversible, and dimer formation does not require GTP hydrolysis or folding information from protein cofactors, because 0.2 microM tubulin-GDP incubated for 20 h was eluted as dimer when analyzed by size exclusion chromatography. Because 20 h corresponds to eight half-times for dissociation, only monomer would be present if dissociation were an irreversible reaction and if dimer formation required GTP or protein cofactors. Additional evidence for a 10(-11) M K(d) was obtained from gel exclusion chromatography studies of 0.02-2 nM tubulin-GDP. The slow dissociation of the tubulin dimer suggests that protein tubulin cofactors function to catalyze dimer dissociation, rather than dimer assembly. Assuming N-site-GTP dissociation is from monomer, our results agree with the 16-h half-time for N-site GTP in vitro and 33 h half-life for tubulin N-site-GTP in CHO cells.  相似文献   

10.
DPP-IV is a prolyl dipeptidase, cleaving the peptide bond after the penultimate proline residue. It is an important drug target for the treatment of type II diabetes. DPP-IV is active as a dimer, and monomeric DPP-IV has been speculated to be inactive. In this study, we have identified the C-terminal loop of DPP-IV, highly conserved among prolyl dipeptidases, as essential for dimer formation and optimal catalysis. The conserved residue His750 on the loop contributes significantly for dimer stability. We have determined the quaternary structures of the wild type, H750A, and H750E mutant enzymes by several independent methods including chemical cross-linking, gel electrophoresis, size exclusion chromatography, and analytical ultracentrifugation. Wild-type DPP-IV exists as dimers both in the intact cell and in vitro after purification from human semen or insect cells. The H750A mutation results in a mixture of DPP-IV dimer and monomer. H750A dimer has the same kinetic constants as those of the wild type, whereas the H750A monomer has a 60-fold decrease in kcat. Replacement of His750 with a negatively charged Glu (H750E) results in nearly exclusive monomers with a 300-fold decrease in catalytic activity. Interestingly, there is no dynamic equilibrium between the dimer and the monomer for all forms of DPP-IVs studied here. This is the first study of the function of the C-terminal loop as well as monomeric mutant DPP-IVs with respect to their enzymatic activities. The study has important implications for the discovery of drugs targeted to the dimer interface.  相似文献   

11.
It was previously shown that the Cro repressor from phage lambda, which is a dimer, can be converted into a stable monomer by a five-amino acid insertion. Phe58 is the key residue involved in this transition, switching from interactions which stabilize the dimer to those which stabilize the monomer. Structural studies, however, suggested that Phe58 did not penetrate into the core of the monomer as well as it did into the native dimer. This was strongly supported by the finding that certain core-repacking mutations, including in particular, Phe58-->Trp, increased the stability of the monomer. Unexpectedly, the same substitution also increased the stability of the native dimer. At the same time it decreased the affinity of the dimer for operator DNA. Here we describe the crystal structures of the Cro F58W mutant, both as the monomer and as the dimer. The F58W monomer crystallized in a form different from that of the original monomer. In contrast to that structure, which resembled the DNA-bound form of Cro, the F58W monomer is closer in structure to wild-type (i.e. non-bound) Cro. The F58W dimer also crystallizes in a form different from the native dimer but has a remarkably similar overall structure which tends to confirm the large changes in conformation of Cro on binding DNA. Introduction of Trp58 perturbs the position occupied by the side-chain of Arg38, a DNA-contact residue, providing a structural explanation for the reduction in DNA-binding affinity.The improved thermal stability is seen to be due to the enhanced solvent transfer free energy of Trp58 relative to Phe58, supplemented in the dimer structure, although not the monomer, by a reduction in volume of internal cavities.  相似文献   

12.
13.
Six pRNAs (p for packaging) of bacterial virus phi29 form a hexamer complex that is an essential component of the viral DNA translocating motor. Dimers, the building block of pRNA hexamer, assemble in the order of dimer --> tetramer --> hexamer. The two-dimensional structure of the pRNA monomer has been investigated extensively; however, the three-dimensional structure concerning the distance constraints of the three stems and loops are unknown. In this report, we probed the three-dimensional structure of pRNA monomer and dimer by photo affinity cross-linking with azidophenacyl. Bases 75-81 of the left stem were found to be oriented toward the head loop and proximate to bases 26-31 in a parallel orientation. Chemical modification interference indicates the involvement of bases 45-71 and 82-91 in dimer formation. Dimer was formed via hand-in-hand contact, a novel RNA dimerization that in some aspects is similar to the kissing loops of the human immunodeficiency virus. The covalently linked dimers were found to be biologically active. Both the native dimer and the covalently linked dimer were found by cryo-atomic force microscopy to be similar in global conformation and size.  相似文献   

14.
Enzyme I of the bacterial phosphotransferase system can exist in a monomer/dimer equilibrium which may have functional significance. Each monomer contains two tryptophan residues. It is demonstrated that the decay of both the monomer and the dimer can be described by a biexponential. The decay times depend on the temperature and at 6 degrees C the decay times are tau 1 = 0.4 ns and tau 2 = 3.2 ns for the monomer and tau 3 = 3.2 ns and tau 4 = 7.2 ns for the dimer form of the enzyme. The changes in the fluorescence decay parameters can be utilized to measure the equilibrium constant for the monomer/dimer transition.  相似文献   

15.
Perfusion of isolated small intestine with the procyanidin dimers B2 and B5 extracted from cocoa indicated that both forms of dimer are transferred to the serosal side of enterocytes but only to a very small extent (<1% of the total transferred flavanol-like compounds). However, perfusion of dimer mainly resulted in large amounts of unmetabolised/unconjugated epicatechin monomer being detected on the serosal side (95.8%). The cleavage of dimer during transfer seemed to be energy-dependent, requiring an intact cell system, as incubation with jejunal homogenates failed to yield epicatechin. Low levels methylated dimer were also detected (3.2%), but no conjugates and metabolites of epicatechin indicating that metabolism of monomer and dimer is limited during dimer cleavage/translocation. The methylation of dimer may be by catechol-O-methyltransferase, however, at high concentrations of dimer COMT activity is reduced leading to an inhibition of both monomer and dimer O-methylation.  相似文献   

16.
目的进行重组鼠疫耶尔森菌LcrV抗原原液二聚体含量及性质研究,确定LcrV原液的相关质控标准。方法在不同缓冲体系(0.85%NaCl(NS)、20 mmol/L PBS),不同蛋白浓度(2.0、1.5、1.0、0.5、0.1 mg/mL)及不同保存温度(4℃、-20℃、-70℃)条件下保存LcrV抗原,采用SDS-PAGE和HPSEC方法定期检测LcrV二聚体含量及纯度。将连续三批检定合格的LcrV抗原原液进行质谱相对分子质量测定、等电点测定、N末端氨基酸序列测定、圆二色(CD)谱、HPLC肽谱及氨基酸组成分析,研究LcrV抗原的相关性质。结果随着保存时间的延长LcrV抗原二聚体含量增加,低温保存时二聚体不易大量形成。在-20℃和-70℃条件下,NS保存的LcrV抗原比PBS体系保存稳定,而在4℃条件下NS保存的LcrV抗原容易降解。LcrV抗原高浓度保存容易发生聚合。LcrV抗原在低质量浓度(0.1 mg/mL)保存时免疫学活性明显下降。质谱检测到LcrV单体和二聚体共同存在,且与理论相对分子质量一致。LcrV原液检测等电点范围为4.6~6.3。N末端测序、CD谱、HPLC肽谱图及氨基酸组成分析与理论结果一致。结论 LcrV抗原原液保存条件确定为:NS体系,蛋白质量浓度1.0~2.0 mg/mL,-20℃以下冻存。制备的LcrV抗原各项检测结果与理论结果一致,抗原性质稳定。  相似文献   

17.
The structure of Drosophila LC8 pH-induced monomer has been determined by NMR spectroscopy using the program AutoStructure. The structure at pH 3 and 30 degrees C is similar to the individual subunits of mammalian LC8 dimer with the exception that a beta strand, which crosses between monomers to form an intersubunit beta-sheet in the dimer, is a flexible loop with turnlike conformations in the monomer. Increased flexibility in the interface region relative to the rest of the protein is confirmed by dynamic measurements based on (15)N relaxation. Comparison of the monomer and dimer structures indicates that LC8 is not a domain swapped dimer.  相似文献   

18.
The DNA in intranuclear yeast chromatin is protected from rapid staphylococcal nuclease degradation so as to yield an oligomeric series of DNA sizes. The course of production and disappearance of the various oligomers agrees quantitatively with a theory of random cleavage by the enzyme at uniformly susceptible sites. The sizes of the oligomers are integral repeats of a basic size, about 160 base pairs, and 80-90% of the yeast genome is involved in this repeating structure. Within this repeat there exists a 140 base pair core of more nuclease-resistant DNA. During the course of digestion, the sizes of the oligomers decrease continuously. The widths of the distribution of DNA sizes increase in order: monomer (1 X repeat size, half width = 5-7 base pairs) less than dimer (2 X repeat size, half width = 30 base pairs) less than trimer (3 X repeat size, half width = 40-45 base pairs). The yeast genome thus seems to have variable spacing of the nucleaseresistant cores, to produce the average repeat size of about 160 base pairs. Also, the presence of more than one species of monomer and dimer at certain times of digestion suggests a possible heterogeneity in the subunit structure.  相似文献   

19.
Stefin A folds as a monomer under strongly native conditions. We have observed that under partially denaturing conditions in the temperature range from 74 to 93 degrees C it folds into a dimer, while it is monomeric above the melting temperature of 95 degrees C. Below 74 degrees C the dimer is trapped and it does not dissociate. The dimer is a folded and structured protein as judged by CD and NMR, nevertheless it is no more functional as an inhibitor of cysteine proteases. The monomer-dimer transition proceeds at a slow rate and the activation energy of dimerization at 99 kcal/mol is comparable to the unfolding enthalpy. A large and negative dimerization enthalpy of -111(+/- 8) kcal/mol was calculated from the temperature dependence of the dissociation constant. An irreversible pretransition at 10-15 deg. below the global unfolding temperature has been observed previously by DSC and can now be assigned to the monomer-dimer transition. Backbone resonances of all the dimer residues were assigned using 15N isotopically enriched protein. The dimer is symmetric and the chemical shift differences between the monomer and dimer are localized around the tripartite hydrophobic wedge, which otherwise interacts with cysteine proteases. Hydrogen exchange protection factors of the residues affected by dimer formation are higher in the dimer than in the monomer. The monomer to dimer transition is accompanied by a rapid exchange of all of the amide protons which are protected in the dimer, indicating that the transition state is unfolded to a large extent. Our results demonstrate that the native monomeric state of stefin A is actually metastable but is favored by the kinetics of folding. The substantial energy barrier which separates the monomer from the more stable dimer traps each state under native conditions.  相似文献   

20.
Chemokine CXCL8 and its receptor CXCR1 are key mediators in combating infection and have also been implicated in the pathophysiology of various diseases including chronic obstructive pulmonary disease (COPD) and cancer. CXCL8 exists as monomers and dimers but monomer alone binds CXCR1 with high affinity. CXCL8 function involves binding two distinct CXCR1 sites – the N‐terminal domain (Site‐I) and the extracellular/transmembrane domain (Site‐II). Therefore, higher monomer affinity could be due to stronger binding at Site‐I or Site‐II or both. We have now characterized the binding of a human CXCR1 N‐terminal domain peptide (hCXCR1Ndp) to WT CXCL8 under conditions where it exists as both monomers and dimers. We show that the WT monomer binds the CXCR1 N‐domain with much higher affinity and that binding is coupled to dimer dissociation. We also characterized the binding of two CXCL8 monomer variants and a trapped dimer to two different hCXCR1Ndp constructs, and observe that the monomer binds with ~10‐ to 100‐fold higher affinity than the dimer. Our studies also show that the binding constants of monomer and dimer to the receptor peptides, and the dimer dissociation constant, can vary significantly as a function of pH and buffer, and so the ability to observe WT monomer peaks is critically dependent on NMR experimental conditions. We conclude that the monomer is the high affinity CXCR1 agonist, that Site‐I interactions play a dominant role in determining monomer vs. dimer affinity, and that the dimer plays an indirect role in regulating monomer function.  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