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.
Reactive oxidative species (ROS) and S‐glutathionylation modulate the activity of plant cytosolic triosephosphate isomerases (cTPI). Arabidopsis thaliana cTPI (AtcTPI) is subject of redox regulation at two reactive cysteines that function as thiol switches. Here we investigate the role of these residues, AtcTPI‐Cys13 and At‐Cys218, by substituting them with aspartic acid that mimics the irreversible oxidation of cysteine to sulfinic acid and with amino acids that mimic thiol conjugation. Crystallographic studies show that mimicking AtcTPI‐Cys13 oxidation promotes the formation of inactive monomers by reposition residue Phe75 of the neighboring subunit, into a conformation that destabilizes the dimer interface. Mutations in residue AtcTPI‐Cys218 to Asp, Lys, or Tyr generate TPI variants with a decreased enzymatic activity by creating structural modifications in two loops (loop 7 and loop 6) whose integrity is necessary to assemble the active site. In contrast with mutations in residue AtcTPI‐Cys13, mutations in AtcTPI‐Cys218 do not alter the dimeric nature of AtcTPI. Therefore, modifications of residues AtcTPI‐Cys13 and AtcTPI‐Cys218 modulate AtcTPI activity by inducing the formation of inactive monomers and by altering the active site of the dimeric enzyme, respectively. The identity of residue AtcTPI‐Cys218 is conserved in the majority of plant cytosolic TPIs, this conservation and its solvent‐exposed localization make it the most probable target for TPI regulation upon oxidative damage by reactive oxygen species. Our data reveal the structural mechanisms by which S‐glutathionylation protects AtcTPI from irreversible chemical modifications and re‐routes carbon metabolism to the pentose phosphate pathway to decrease oxidative stress.  相似文献   

2.
Using model compounds of the melanic component of neuromelanin (NM) prepared by tyrosinase oxidation at various ratios of dopamine (DA) and cysteine (Cys) under physiological conditions, we examined a biosynthetic pathway to NM and its aging process by following the time course of oxidation to NM and the subsequent structural modification of NM under various heating conditions. Chemical degradation methods were applied to the synthetic NM. 4‐Amino‐3‐hydroxyphenylethylamine (4‐AHPEA) and thiazole‐2,4,5‐tricarboxylic acid (TTCA) were used as markers of benzothiazine and benzothiazole units, respectively. By following the time course of the biosynthetic pathway of synthetic NM, we found that neurotoxic molecules are trapped in NM. An aging simulation of synthetic NM showed that benzothiazine units in NM are gradually converted to benzothiazole during the aging process. Thus, natural NM was found to be similar to aged (heated) NM prepared from a 2:1 molar ratio of DA and Cys.  相似文献   

3.
Interaction of the cell‐penetrating peptide (CPP) cysteine‐transportan (Cys‐TP) with model lipid membranes was examined by spin‐label electron paramagnetic resonance (EPR). Membranes were labeled with lipophilic spin probes and the influence of Cys‐TP on membrane structure was studied. The influence of Cys‐TP on membrane permeability was monitored by the reduction of a liposome‐trapped water‐soluble spin probe. Cys‐TP caused lipid ordering in membranes prepared from pure dimyristoylphosphatidylcholine (DMPC) and in DMPC membranes with moderate cholesterol concentration. In addition, Cys‐TP caused a large increase in permeation of DMPC membranes. In contrast, with high cholesterol content, at which model lipid membranes are in the so‐called liquid‐ordered phase, no effect of Cys‐TP was observed, either on the membrane structure or on the membrane permeability. The interaction between Cys‐TP and the lipid membrane therefore depends on the lipid phase. This could be of great importance for understanding of the CPP–lipid interaction in laterally heterogeneous membranes, while it implies that the CPP–lipid interaction can be different at different points along the membrane. Copyright © 2008 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

4.
Fn14 is the smallest member of the tumor necrosis factor (TNF) receptor superfamily, and specifically binds to its ligand, TWEAK (TNF‐like weak inducer of apoptosis), which is a member of the TNF superfamily. The receptor‐ligand recognition between Fn14 and TWEAK induces a variety of cellular processes for tissue remodeling and is also involved in the pathogenesis of some human diseases, such as cancer, chronic autoimmune diseases, and acute ischaemic stroke. The extracellular ligand‐binding region of Fn14 is composed of 53 amino acid residues and forms a single, cysteine‐rich domain (CRD). In this study, we determined the solution structure of the Fn14 CRD (Glu28‐Ala70) by heteronuclear NMR, with a 13C‐/15N‐labeled sample. The tertiary structure of the CRD comprises a β‐sheet with two strands, followed by a 310 helix and a C‐terminal α‐helix, and is stabilized by three disulfide bonds connecting Cys36‐Cys49, Cys52‐Cys67, and Cys55‐Cys64. Comparison of the disulfide bond connectivities and the tertiary structures with those of other CRDs revealed that the Fn14 CRD is similar to the fourth CRD of TNF receptor 1 (A1‐C2 module type), but not to the CRD of B‐cell maturation antigen and the second CRD of transmembrane activator and CAML (calcium modulator and cyclophilin ligand) interactor (A1‐D2 module type). This is the first structural report about the A1‐C2 type CRD that could bind to the known target.  相似文献   

5.
All five L‐amino acids tested (L‐serine, L‐lysine, L‐leucine, L‐cysteine and L‐arginine) were used by Chlamydomonas reinhardtii as sole nitrogen source. Among these, L‐Cys was special as it has not been reported before. While these amino acids could be used in the dark only in the presence of acetic acid, in conditions of light they could support the growth of C. reinhardtii without the supplementation of acetic acid. When cultured in the TAP‐N medium, the chlorophyll content was found to be lower in the dark, but higher in the light for the cells grown with L‐Arg than with other four amino acids. Exogenously supplied L‐Ser and L‐Lys did not accumulate in the cells, demonstrating that they were used by supplying ammonium to the cells from the activity of an extracellular deaminase. Further results showed that the induction of the extracellular deaminase activity required a period of nitrogen starvation, regardless of the medium containing acetic acid or not. Results also showed that the uptake of L‐Cys was similar to L‐Leu, most likely via passive diffusion. When L‐Cys and L‐Leu were supplied together to the nitrogen‐starved cells, the absorption of L‐Cys did not affect the uptake of L‐Leu.  相似文献   

6.
Selenocysteine (Sec) is the 21st amino acid in the genetic code and it is present in a small number of proteins where it replaces the much more commonly used amino acid cysteine (Cys). It is both more complicated and bioenergetically costly to insert Sec into a protein in comparison to Cys, and this cost is most likely compensated by a gain of function to the enzyme/protein in which it is incorporated. Here we investigate one such gain of function, the enhancement of one‐electron transfer catalysis. We compared the ability of Sec‐containing mouse mitochondrial thioredoxin reductase (mTrxR2) to catalyze the reduction of bovine cytochrome c, ascorbyl radical, and dehydroascorbate in comparison to Cys‐containing thioredoxin reductases from D. melanogaster (DmTrxR) and P. falciparum (PfTrxR). The Sec‐containing mTrxR2 was able to reduce all three substrates, while the Cys‐containing enzymes had little or no activity. In addition, we constructed Cys?Sec mutants of DmTrxR and PfTrxR and found that this substitution resulted in a gain of function, as these mutant enzymes were now able to catalyze the reduction of these substrates. We also found that in the case of PfTrxR, reduction of cytochrome c was enhanced five‐fold in a truncated PfTrxR in which the C‐terminal redox center was removed. This shows that some of the ability of thioredoxin reductase to reduce this substrate comes from the flavin coenzyme. We also discuss a possible mechanism by which Sec‐containing thioredoxin reductase reduces dehydroascorbate to ascorbate by two sequential, one‐electron reductions, in part catalyzed by Sec.  相似文献   

7.
Pain is a major symptom of many diseases and results in enormous pressures on human body or society. Currently, clinically used analgesic drugs, including opioids and nonsteroidal anti‐inflammatory drugs, have adverse reactions, and thus, the development of new types of analgesic drug candidates is urgently needed. Animal venom peptides have proven to have potential as new types of analgesic medicine. In this research, we describe the isolation and characterization of an analgesic peptide from the crude venom of centipede, Scolopendra subspinipes mutilans. The amino acid sequence of this peptide was identical with SsmTX‐I that was previously reported as a specific Kv2.1 ion channel blocker. Our results revealed that SsmTX‐I was produced by posttranslational processing of a 73‐residue prepropeptide. The intramolecular disulfide bridge motifs of SsmTX‐I was Cys1–Cys3 and Cys2–Cys4. Functional assay revealed that SsmTX‐I showed potential analgesic activities in formalin‐induced paw licking, thermal pain, and acetic acid‐induced abdominal writhing mice models. Our research provides the first report of cDNA sequences, disulfide motif, successful synthesis, and analgesic potential of SsmTX‐I for the development of pain‐killing drugs. It indicates that centipede peptide toxins could be a treasure trove for the search of novel analgesic drug candidates. Copyright © 2017 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

8.
Yersinia enterocolitica is a Gram‐negative bacterium that causes yersiniosis, a zoonotic disease affecting the gastrointestinal tract of humans, cattle, and pigs, among others. The lipopolysaccharide of Y. enterocolitica O:8 contains an unusual sugar, 6‐deoxy‐d ‐gulose, which requires four enzymes for its biosynthesis. Here, we describe a combined structural and functional investigation of WbcA, which catalyzes the third step in the pathway, namely an epimerization about the C‐3′ carbon of a CDP‐linked sugar. The structure of WbcA was determined to 1.75‐Å resolution, and the model was refined to an overall R‐factor of 19.5%. The fold of WbcA places it into the well‐defined cupin superfamily of sugar epimerases. Typically, these enzymes contain both a conserved histidine and a tyrosine residue that play key roles in catalysis. On the basis of amino acid sequence alignments, it was anticipated that the “conserved” tyrosine had been replaced with a cysteine residue in WbcA (Cys 133), and indeed this was the case. However, what was not anticipated was the fact that another tyrosine residue (Tyr 50) situated on a neighboring β‐strand moved into the active site. Site‐directed mutant proteins were subsequently constructed and their kinetic properties analyzed to address the roles of Cys 133 and Tyr 50 in WbcA catalysis. This study emphasizes the continuing need to experimentally verify assumptions that are based solely on bioinformatics approaches.  相似文献   

9.
The MarR/DUF24‐type repressor YodB controls the azoreductase AzoR1, the nitroreductase YodC and the redox‐sensing regulator Spx in response to quinones and diamide in Bacillus subtilis. Previously, we showed using a yodBCys6‐Ala mutant that the conserved Cys6 apparently contributes to the DNA‐binding activity of YodB in vivo. Here, we present data that mutation of Cys6 to Ser led to a form of the protein that was reduced in redox‐sensing in response to diamide and 2‐methylhydroquinone (MHQ) in vivo. DNA‐binding experiments indicate that YodB is regulated by a reversible thiol‐modification in response to diamide and MHQ in vitro. Redox‐regulation of YodB involves Cys6‐Cys101' intermolecular disulfide formation by diamide and quinones in vitro. Diagonal Western blot analyses confirm the formation of intersubunit disulfides in YodB in vivo that require the conserved Cys6 and either of the C‐terminal Cys101' or Cys108' residues. This study reveals a thiol‐disulfide switch model of redox‐regulation for the YodB repressor to sense electrophilic compounds in vivo.  相似文献   

10.
Kringle 5(K5) is the fifth kringle domain of human plasminogen and its anti‐angiogenic activity is more potent than angiostatin that includes the first four kringle fragment of plasminogen. Our recent study demonstrated that K5 suppressed hepatocarcinoma growth by anti‐angiogenesis. To find high efficacy and minimal peptide sequence required for the anti‐angiogenic and anti‐tumour activities of K5, two deletion mutants of K5 were generated. The amino acid residues outside kringle domain of intact K5 (Pro452‐Ala542) were deleted to form K5mut1(Cys462‐Cys541). The residue Cys462 was deleted again to form K5mut2(Met463‐Cys541). K5mut1 specifically inhibited proliferation, migration and induced apoptosis of endothelial cells, with an apparent two‐fold enhanced activity than K5. Intraperitoneal injection of K5mut1 resulted in more potent tumour growth inhibition and microvessel density reduction than K5 both in HepA‐grafted and Bel7402‐xenografted hepatocarcinoma mouse models. These results suggested that K5mut1 has more potent anti‐angiogenic activity than intact K5. K5mut2, which lacks only the amino terminal cysteine of K5mut1, completely lost the activity, suggesting that the kringle domain is essential for the activity of K5. The activity was enhanced to K5mut1 level when five acidic amino acids of K5 in NH2 terminal outside kringle domain were replaced by five serine residues (K5mut3). The shielding effect of acidic amino acids may explain why K5mut1 has higher activity. K5, K5mut1 and K5mut3 held characteristic β‐sheet spectrum while K5mut2 adopted random coil structure. These results suggest that K5mut1 with high efficacy is the minimal active peptide sequence of K5 and may have therapeutic potential in liver cancer.  相似文献   

11.
In order to synthesize interferon‐γ glycoform having an oligosaccharide at the 97 position by a semi‐synthetic method, interferon‐γ‐polypeptide‐(1–94)‐α‐hydrazide was prepared by the specific Cys‐cyanylation of polypeptide‐(1–94)‐Cys‐His6 expressed from E. coli and subsequent hydrazinolysis in 22% yield (two steps). This polypeptide‐α‐hydrazide was then converted into corresponding polypeptide‐α‐thioester under NaNO2/acid conditions followed by thiolysis in 83% yield. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

12.
The interactions of Met and Cys with other amino acid side chains have received little attention, in contrast to aromatic–aromatic, aromatic–aliphatic or/and aliphatic–aliphatic interactions. Precisely, these are the only amino acids that contain a sulfur atom, which is highly polarizable and, thus, likely to participate in strong Van der Waals interactions. Analysis of the interactions present in membrane protein crystal structures, together with the characterization of their strength in small‐molecule model systems at the ab‐initio level, predicts that Met–Met interactions are stronger than Met–Cys ≈ Met–Phe ≈ Cys–Phe interactions, stronger than Phe–Phe ≈ Phe–Leu interactions, stronger than the Met–Leu interaction, and stronger than Leu–Leu ≈ Cys–Leu interactions. These results show that sulfur‐containing amino acids form stronger interactions than aromatic or aliphatic amino acids. Thus, these amino acids may provide additional driving forces for maintaining the 3D structure of membrane proteins and may provide functional specificity.  相似文献   

13.
Phosphonium and uronium salt‐based reagents enable efficient and effective coupling reactions and are indispensable in peptide chemistry, especially in machine‐assisted SPPS. However, after the activating and coupling steps with these reagents in the presence of tertiary amines, Fmoc derivatives of Cys are known to be considerably racemized during their incorporation. To avoid this side reaction, a coupling method mediated by phosphonium/uronium reagents with a weaker base, such as 2,4,6‐trimethylpyridine, than the ordinarily used DIEA or that by carbodiimide has been recommended. However, these methods are appreciably inferior to the standard protocol applied for SPPS, that is, a 1 min preactivation procedure of coupling with phosphonium or uronium reagents/DIEA in DMF, in terms of coupling efficiency, and also the former method cannot reduce racemization of Cys(Trt) to an acceptable level (<1.0%) even when the preactivation procedure is omitted. Here, the 4,4′‐dimethoxydiphenylmethyl and 4‐methoxybenzyloxymethyl groups were demonstrated to be acid‐labile S‐protecting groups that can suppress racemization of Cys to an acceptable level (<1.0%) when the respective Fmoc derivatives are incorporated via the standard SPPS protocol of phosphonium or uronium reagents with the aid of DIEA in DMF. Furthermore, these protecting groups significantly reduced the rate of racemization compared to the Trt group even in the case of microwave‐assisted SPPS performed at a high temperature. © 2013 The Authors. European Peptide Society published by John Wiley & Sons, Ltd.  相似文献   

14.
CrtJ from Rhodobacter capsulatus is a regulator of genes involved in the biosynthesis of haem, bacteriochlorophyll, carotenoids as well as structural proteins of the light harvesting‐II complex. Fluorescence anisotropy‐based DNA‐binding analysis demonstrates that oxidized CrtJ exhibits ~ 20‐fold increase in binding affinity over that of reduced CrtJ. Liquid chromatography electrospray tandem ionization mass spectrometric analysis using DAz‐2, a sulfenic acid (–SOH)‐specific probe, demonstrates that exposure of CrtJ to oxygen or to hydrogen peroxide leads to significant accumulation of a sulfenic acid derivative of Cys420 which is located in the helix–turn–helix (HTH) motif. In vivo labelling with 4‐(3‐azidopropyl)cyclohexane‐1,3‐dione (DAz‐2) shows that Cys420 also forms a sulfenic acid modification in vivo when cells are exposed to oxygen. Moreover, a Cys420 to Ala mutation leads to a ~ 60‐fold reduction of DNA binding activity while a Cys to Ser substitution at position 420 that mimics a cysteine sulfenic acid results in a ~ 4‐fold increase in DNA binding activity. These results provide the first example where sulfenic acid oxidation of a cysteine in a HTH‐motif leads to differential effects on gene expression.  相似文献   

15.
The ubiquitously distributed peroxiredoxins (Prxs) have been shown to have diverse functions in cellular defense‐signaling pathways. They have been largely classified into three Prx classes, 2‐Cys Prx, atypical 2‐Cys Prx and 1‐Cys Prx, which can be distinguished by how many Cys residues they possess and by their catalytic mechanisms. Proteins belonging to the typical 2‐Cys Prx group containing the N‐terminal peroxidatic Cys residue undergo a cycle of peroxide‐dependent oxidation to sulfenic acid and thiol‐dependent reduction during H2O2 catalysis. However, in the presence of high concentrations of H2O2 and catalytic components, including thioredoxin (Trx), Trx reductase and NADPH, the sulfenic acid can be hyperoxidized to cysteine sulfinic acid. The overoxidized 2‐Cys Prxs are slowly reduced by the action of the adenosine 5′‐triphosphate‐dependent enzyme, sulfiredoxin. Upon exposure of cells to strong oxidative or heat‐shock stress conditions, 2‐Cys Prxs change their protein structures from low‐molecular weight to high‐molecular weight complexes, which trigger their functional switching from peroxidases to molecular chaperones. The C‐terminal region of 2‐Cys Prx also plays an essential role in this structural conversion. Thus, proteins with truncated C‐termini are resistant to overoxidation and cannot regulate their structures or functions. These reactions are primarily guided by the active site peroxidatic Cys residue, which serves as an ‘H2O2‐sensor’ in cells. The reversible structural and functional switching of 2‐Cys Prxs provides cells with a means to adapt to external stresses by presumably activating intracellular defense‐signaling systems. In particular, plant 2‐Cys Prxs localized in chloroplasts have dynamic protein structures that undergo major conformational changes during catalysis, forming super‐complexes and reversibly attaching to thylakoid membranes in a redox‐dependent manner.  相似文献   

16.
The side‐chain dihedral angle distributions of all amino acids have been measured from myriad high‐resolution protein crystal structures. However, we do not yet know the dominant interactions that determine these distributions. Here, we explore to what extent the defining features of the side‐chain dihedral angle distributions of different amino acids can be captured by a simple physical model. We find that a hard‐sphere model for a dipeptide mimetic that includes only steric interactions plus stereochemical constraints is able to recapitulate the key features of the back‐bone dependent observed amino acid side‐chain dihedral angle distributions of Ser, Cys, Thr, Val, Ile, Leu, Phe, Tyr, and Trp. We find that for certain amino acids, performing the calculations with the amino acid of interest in the central position of a short α‐helical segment improves the match between the predicted and observed distributions. We also identify the atomic interactions that give rise to the differences between the predicted distributions for the hard‐sphere model of the dipeptide and that of the α‐helical segment. Finally, we point out a case where the hard‐sphere plus stereochemical constraint model is insufficient to recapitulate the observed side‐chain dihedral angle distribution, namely the distribution P3) for Met. Proteins 2014; 82:2574–2584. © 2014 Wiley Periodicals, Inc.  相似文献   

17.
Human protein C (hPC) is glycosylated at three Asn‐X‐Ser/Thr and one atypical Asn‐X‐Cys sequons. We have characterized the micro‐ and macro‐heterogeneity of plasma‐derived hPC and compared the glycosylation features with recombinant protein C (tg‐PC) produced in a transgenic pig bioreactor from two animals having approximately tenfold different expression levels. The N‐glycans of hPC are complex di‐ and tri‐sialylated structures, and we measured 78% site occupancy at Asn‐329 (the Asn‐X‐Cys sequon). The N‐glycans of tg‐PC are complex sialylated structures, but less branched and partially sialylated. The porcine mammary epithelial cells glycosylate the Asn‐X‐Cys sequon with a similar efficiency as human hepatocytes even at these high expression levels, and site occupancy at this sequon was not affected by expression level. A distinct bias for particular structures was present at each of the four glycosylation sites for both hPC and tg‐PC. Interestingly, glycans with GalNAc in the antennae were predominant at the Asn‐329 site. The N‐glycan structures found for tg‐PC are very similar to those reported for a recombinant Factor IX produced in transgenic pig milk, and similar to the endogenous milk protein lactoferrin, which may indicate that N‐glycan processing in the porcine mammary epithelial cells is more uniform than in other tissues.  相似文献   

18.
The formation mechanism of Maillard peptides was explored in Maillard reaction through diglycine/glutathione(GSH)/(Cys‐Glu‐Lys‐His‐Ile‐Met)–xlyose systems by heating at 120 °C for 30–120 min. Maximum fluorescence intensity of Maillard reaction products (MRPs) with an emission wavelength of 420~430 nm in all systems was observed, and the intensity values were proportional to the heating time. Taken diglycine/GSH–[13C5]xylose systems as a control, it was proposed that the compounds with high m/z values of 379 and 616 have the high molecular weight (HMW) products formed by cross‐linking of peptides and sugar. In (Cys‐Glu‐Lys‐His‐Ile‐Met)–xylose system, the m/z value of HMW MRPs was not observed, which might be due to the weak signals of these products. According to the results of gel permeation chromatography, HMW MRPs were formed by Maillard reaction, especially in (Cys‐Glu‐Lys‐His‐Ile‐Met)–xylose system, the percentage of Maillard peptides reached 52.90%. It was concluded that Maillard peptides can be prepared through the cross‐linking of sugar and small peptides with a certain MW range. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
Peroxiredoxins are ubiquitous thioredoxin‐dependent peroxidases presumed to display, upon environmental constraints, a chaperone function resulting from a redox‐dependent conformational switch. In this work, using biochemical and genetic approaches, we aimed to unravel the factors regulating the redox status and the conformation of the plastidial 2‐Cys peroxiredoxin (2‐Cys PRX) in plants. In Arabidopsis, we show that in optimal growth conditions, the overoxidation level mainly depends on the availability of thioredoxin‐related electron donors, but not on sulfiredoxin, the enzyme reducing the 2‐Cys PRX overoxidized form. We also observed that upon various physiological temperature, osmotic and light stress conditions, the overoxidation level and oligomerization status of 2‐Cys PRX can moderately vary depending on the constraint type. Further, no major change was noticed regarding protein conformation in water‐stressed Arabidopsis, barley and potato plants, whereas species‐dependent up‐ and down‐variations in overoxidation were observed. In contrast, both 2‐Cys PRX overoxidation and oligomerization were strongly induced during a severe oxidative stress generated by methyl viologen. From these data, revealing that the oligomerization status of plant 2‐Cys PRX does not exhibit important variation and is not tightly linked to the protein redox status upon physiologically relevant environmental constraints, the possible in planta functions of 2‐Cys PRX are discussed.  相似文献   

20.
The diuretic drug ethacrynic acid (EA), both an inhibitor and substrate of pi class glutathione S‐transferase (GST P1‐1), has been tested in clinical trials as an adjuvant in chemotherapy. We recently studied the role of the active site residue Tyr‐108 in binding EA to the enzyme and found that the analysis was complicated by covalent binding of this drug to the highly reactive Cys‐47. Previous attempts to eliminate this binding by chemical modification yielded ambiguous results and therefore we decided here to produce a double mutant C47S/Y108V by site directed mutagenesis and further expression in Escherichia coli and the interaction of EA and its GSH conjugate (EASG) examined by calorimetric studies and X‐ray diffraction. Surprisingly, in the absence of Cys‐47, Cys‐101 (located at the dimer interface) becomes a target for modification by EA, albeit at a lower conjugation rate than Cys‐47. The Cys‐47 → Ser mutation in the double mutant enzyme induces a positive cooperativity between the two subunits when ligands with affinity to G‐site bind to enzyme. However, this mutation does not seem to affect the thermodynamic properties of ligand binding to the electrophilic binding site (H‐site) and the thermal or chemical stability of this double mutant does not significantly affect the unfolding mechanism in either the absence or presence of ligand. Crystal structures of apo and an EASG complex are essentially identical with a few exceptions in the H‐site and in the water network at the dimer interface. Copyright © 2010 John Wiley & Sons, Ltd.  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