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 0 毫秒
1.
The three-dimensional structure of a bacterial superantigen, Staphylococcus aureus enterotoxin H (SEH), bound to human major histocompatibility complex (MHC) class II (HLA-DR1) has been determined by X-ray crystallography to 2.6 A resolution (1HXY). The superantigen binds on top of HLA-DR1 in a completely different way from earlier co-crystallized superantigens from S.aureus. SEH interacts with high affinity through a zinc ion with the beta1 chain of HLA-DR1 and also with the peptide presented by HLA-DR1. The structure suggests that all superantigens interacting with MHC class II in a zinc-dependent manner present the superantigen in a common way. This suggests a new model for ternary complex formation with the T-cell receptor (TCR), in which a contact between the TCR and the MHC class II is unlikely.  相似文献   

2.
Krahn JM  Jackson MR  DeRose EF  Howell EE  London RE 《Biochemistry》2007,46(51):14878-14888
Type II dihydrofolate reductase (DHFR) is a plasmid-encoded enzyme that confers resistance to bacterial DHFR-targeted antifolate drugs. It forms a symmetric homotetramer with a central pore which functions as the active site. Its unusual structure, which results in a promiscuous binding surface that accommodates either the dihydrofolate (DHF) substrate or the NADPH cofactor, has constituted a significant limitation to efforts to understand its substrate specificity and reaction mechanism. We describe here the first structure of a ternary R67 DHFR.DHF.NADP+ catalytic complex, resolved to 1.26 A. This structure provides the first clear picture of how this enzyme, which lacks the active site carboxyl residue that is ubiquitous in Type I DHFRs, is able to function. In the catalytic complex, the polar backbone atoms of two symmetry-related I68 residues provide recognition motifs that interact with the carboxamide on the nicotinamide ring, and the N3-O4 amide function on the pteridine ring. This set of interactions orients the aromatic rings of substrate and cofactor in a relative endo geometry in which the reactive centers are held in close proximity. Additionally, a central, hydrogen-bonded network consisting of two pairs of Y69-Q67-Q67'-Y69' residues provides an unusually tight interface, which appears to serve as a "molecular clamp" holding the substrates in place in an orientation conducive to hydride transfer. In addition to providing the first clear insight regarding how this extremely unusual enzyme is able to function, the structure of the ternary complex provides general insights into how a mutationally challenged enzyme, i.e., an enzyme whose evolution is restricted to four-residues-at-a-time active site mutations, overcomes this fundamental limitation.  相似文献   

3.
Four phosphoprotein phosphatases, with the ability to act upon hydroxymethylglutaryl (HMG)-CoA reductase, phosphorylase, and glycogen synthase have been purified from rat liver cytosol through a process that involves DEAE-cellulose, aminohexyl-Sepharose-4B, and Bio-Gel A 1.5 m chromatographies. Protein phosphatase II (Mr 180,000) was the major enzyme (68%) with a very broad substrate specificity, showing similar activity toward the three substrates. Phosphatases I1 (Mr 180,000) and I3 (Mr 250,000) accounted for only 12 and 15% of the total activity, respectively, and they were also able to dephosphorylate the three substrates. In contrast, phosphatase I2 (Mr 200,000) showed only phosphorylase phosphatase activity with insignificant dephosphorylating capacity toward HMG-CoA reductase and glycogen synthase. Upon ethanol treatment at room temperature, the Mr of all phosphatases changed; protein phosphatases I2, I3, and II were brought to an Mr of 35,000, while phosphatase I1 was reduced to an Mr of 69,000. Glycogen synthase phosphatase activity was decreased in all four phosphatases. There was also a decrease in phosphatase I1 activity toward HMG-CoA reductase and phosphorylase as substrates. The HMG-CoA reductase phosphatase and phosphorylase phosphatase activities of phosphatases I2, I3, and II were increased after ethanol treatment. Each protein phosphatase showed a different optimum pH, which changed depending on the substrate. The four phosphatases increased their activity in the presence of Mn2+ and Mg2+. In general, Mn2+ was a better activator than Mg2+, and phosphatase I1 showed a stronger dependency on these cations than any other phosphatase. Phosphorylase was a competitive substrate in the HMG-CoA reductase phosphatase and glycogen synthase phosphatase reactions of protein phosphatases I1, I3, and II. HMG-CoA reductase was also able to compete with phosphorylase and glycogen synthase for phosphatase activity. Glycogen synthase phosphatase activity presented less inhibition in the low-Mr forms. A comparison has been made with other protein phosphatases previously reported in the literature.  相似文献   

4.
Crystal structure of thrombin bound to heparin   总被引:5,自引:0,他引:5  
Thrombin is the final protease in the blood coagulation cascade and serves both pro- and anticoagulant functions through the cleavage of several targets. The ability of thrombin to specifically recognize a wide range of substrates derives from interactions that occur outside of the active site of thrombin. Thrombin possesses two anion binding exosites, which mediate many of its interactions with cofactors and substrates, and although many structures of thrombin have been solved, few such interactions have been described in molecular detail. Glycosaminoglycan binding to exosite II of thrombin plays a major role in switching off the procoagulant functions of thrombin by mediating its irreversible inhibition by circulating serpins and by its binding to the endothelial cell surface receptor thrombomodulin. Here we report the 1.85-A structure of human alpha-thrombin bound to a heparin fragment of eight monosaccharide units in length. The asymmetric unit is composed of two thrombin dimers, each sharing a single heparin octasaccharide chain. The observed interactions are fully consistent with previous mutagenesis studies and illustrate on a molecular level the cofactor interaction that is critical for the restriction of clotting to the site of blood vessel injury.  相似文献   

5.
The enzyme 3-hydroxy-3-methylglutaryl-coenzyme-A (HMG-CoA) reductase is known as the rate-limiting enzyme in early sterol biosynthesis in eukaryotic cells. To eliminate this regulation in the yeast Saccharomyces cerevisiae, a truncated HMG1 gene, producing a form of the enzyme that lacks the membrane-binding region (i.e. amino acids 1–552), was constructed and overexpressed in this yeast. The transformed strains accumulated large amounts of the sterol precursor squalene, while the levels of ergosterol and a number of other sterol compounds were only slightly elevated. These findings suggest that HMG-CoA reductase is not the only rate-limiting step in sterol synthesis and its overexpression cannot significantly influence this pathway beyond the sterol precursor squalene. Received: 9 June 1997 / Received revision: 1 September 1997 / Accepted: 19 September 1997  相似文献   

6.
Type II dihydrofolate reductases (DHFRs) encoded by the R67 and R388 plasmids are sequence and structurally different from known chromosomal DHFRs. These plasmid-derived DHFRs are responsible for confering trimethoprim resistance to the host strain. A derivative of R388 DHFR, RBG200, has been cloned and its physical properties have been characterized. This enzyme has been shown to transfer the pro-R hydrogen of NADPH to its substrate, dihydrofolate, making it a member of the A-stereospecific class of dehydrogenases [Brito, R. M. M., Reddick, R., Bennett, G. N., Rudolph, F. B., & Rosevear, P. R. (1990) Biochemistry 29,9825]. Two distinct binary RBG200.NADP+ complexes were detected. Addition of NADP+ to RBG200 DHFR results in formation of an initial binary complex, conformation I, which slowly interconverts to a second more stable binary complex, conformation II. The binding of NADP+ to RBG200 DHFR in the second binary complex was found to be weak, KD = 1.9 +/- 0.4 mM. Transferred NOEs were used to determine the conformation of NADP+ bound to RBG200 DHFR. The initial slope of the NOE buildup curves, measured from the intensity of the cross-peaks as a function of the mixing time in NOESY spectra, allowed interproton distances on enzyme-bound NADP+ to be estimated. The experimentally measured distances were used to define upper and lower bound distance constraints between proton pairs in distance geometry calculations. All NADP+ structures consistent with the experimental distance bounds were found to have a syn conformation about the nicotinamide-ribose (X = 94 +/- 26 degrees) and an anti conformation about the adenine-ribose (X = -92 +/- 32 degrees) glycosidic bonds.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
Regulation of hydroxymethylglutaryl-CoA reductase in rat leukocytes   总被引:4,自引:0,他引:4  
Methods were developed for the assay of hydroxymethylglutaryl-CoA reductase (NADPH) activity in microsomes from rat leukocytes. The activity in freshly isolated leukocytes is low compared to rat liver but can be assayed reliably. The patterns of response of leukocyte reductase in the assay to variation in substrate concentration, protein concentration, and time mimic those of rat liver reductase. Reductase activity in leukocyte microsomes, as in liver microsomes, is depressed by dietary cholesterol and by fasting and is elevated by dietary cholestyramine. Unlike liver reductase, leukocyte reductase activity does not exhibit a detectable diurnal rhythm. We conclude that the assay of reductase in freshly isolated leukocytes holds promise as a technique for detecting the effects of various factors on cholesterol synthesis in vivo.  相似文献   

8.
Guaiacol is a universal substrate for all peroxidases, and its use in a simple colorimetric assay has wide applications. However, its exact binding location has never been defined. Here we report the crystal structures of guaiacol bound to cytochrome c peroxidase (CcP). A related structure with phenol bound is also presented. The CcP-guaiacol and CcP-phenol crystal structures show that both guaiacol and phenol bind at sites distinct from the cytochrome c binding site and from the δ-heme edge, which is known to be the binding site for other substrates. Although neither guaiacol nor phenol is seen bound at the δ-heme edge in the crystal structures, inhibition data and mutagenesis strongly suggest that the catalytic binding site for aromatic compounds is the δ-heme edge in CcP. The functional implications of these observations are discussed in terms of our existing understanding of substrate binding in peroxidases [Gumiero A et al. (2010) Arch Biochem Biophys 500, 13-20].  相似文献   

9.
10.
Aptamers, an emerging class of therapeutics, are DNA or RNA molecules that are selected to bind molecular targets that range from small organic compounds to large proteins. All of the determined structures of aptamers in complex with small molecule targets show that aptamers cage such ligands. In structures of aptamers in complex with proteins that naturally bind nucleic acid, the aptamers occupy the nucleic acid binding site and often mimic the natural interactions. Here we present a crystal structure of an RNA aptamer bound to human thrombin, a protein that does not naturally bind nucleic acid, at 1.9 A resolution. The aptamer, which adheres to thrombin at the binding site for heparin, presents an extended molecular surface that is complementary to the protein. Protein recognition involves the stacking of single-stranded adenine bases at the core of the tertiary fold with arginine side chains. These results exemplify how RNA aptamers can fold into intricate conformations that allow them to interact closely with extended surfaces on non-RNA binding proteins.  相似文献   

11.
We report a crystal structure that shows an antibiotic that extracts a nucleobase from a DNA molecule ‘caught in the act’ after forming a covalent bond but before departing with the base. The structure of trioxacarcin A covalently bound to double-stranded d(AACCGGTT) was determined to 1.78 Å resolution by MAD phasing employing brominated oligonucleotides. The DNA–drug complex has a unique structure that combines alkylation (at the N7 position of a guanine), intercalation (on the 3′-side of the alkylated guanine), and base flip-out. An antibiotic-induced flipping-out of a single, nonterminal nucleobase from a DNA duplex was observed for the first time in a crystal structure.  相似文献   

12.
The regulation of hydroxymethylglutaryl-CoA reductase in cultured cells   总被引:4,自引:0,他引:4  
Growth-stimulated synchronized cells exhibit a rapid increase in 3-hydroxy-3-methylglutaryl coenzyme A reductase (HMG-CoA reductase, EC 1.1.1.88) activity prior to the onset of DNA synthesis. Under normal culture conditions, HMG-CoA reductase activity exhibits wide variations among experiments. To determine whether this phenomenon is dependent on cell replication, we used J774 macrophage-like cells to compare changes in reductase activity in cells synchronized by serum deprivation and then growth-stimulated by fresh media containing serum to unsynchronized cells treated with fresh media and serum. Under these conditions, no increase in [3H]thymidine incorporation into cell DNA was seen in unsynchronized cells, but a large increase was observed in synchronized cells 10-12 h after media change. Although the growth characteristics differed between the cells, reductase activity was low at the time of media change and increased 10 to 20-fold 5-10 h after media change, returning to basal levels by 24 h in both synchronized and unsynchronized cells. This pattern of reductase activity was observed in unsynchronized cells from a variety of cell lineages, although the magnitude of the changes varied. Fluctuations of [14C]acetate incorporation into cholesterol were observed in parallel to alterations in reductase activity. LDL receptor expression also paralleled the changes in reductase activity, but scavenger receptor expression was not affected. Addition of lipoproteins at the time of media change inhibited the rise in reductase activity by 80-90%. The increase in reductase activity was not due to a stimulation of cholesterol efflux into the medium, but evidence for the secretion into the media of an inhibitory factor was obtained. These results suggest that cell requirements for cholesterol are not always directly related to replication, and that standard culture conditions induce transient fluctuations in reductase activity and lipoprotein receptor expression.  相似文献   

13.
14.
Biliverdin reductase (BVR) is a soluble cytoplasmic enzyme that catalyzes the conversion of biliverdin to bilirubin using NADH or NADPH as electron donor. Bilirubin is a significant biological antioxidant, but it is also neurotoxic and the cause of kernicterus. In this study, we have determined the crystal structure of rat BVR at 1.4 A resolution. The structure contains two domains: an N-terminal domain characteristic of a dinucleotide binding fold (Rossmann fold) and a C-terminal domain that is predominantly an antiparallel six-stranded beta-sheet. Based on this structure, we propose modes of binding for NAD(P)H and biliverdin, and a possible mechanism for the enzyme.  相似文献   

15.
This study investigates a proposed design of a peptide sequence that is based on a bioactive conformation of statins that act as the competitive inhibitors of HMG-CoA for HMGR. To bridge these heterogeneous organic compounds, a conformational aspect relating to an analysis of the flexibility of the peptide molecules and their occupied volumes was applied to the peptide design. The design criterion was formulated in terms of a proximity parameter (Pr), reflecting the probability of an active peptide conformation to approximate the statin. Through a structure-functional analysis of previously synthesized peptides and statin molecules, nine peptides were selected for the peptide library. Comparing the calculated proximity parameters, four peptides (IAVE, YAVE, IVAE, and YVAE) from the library were selected and synthesized. In vitro assays elucidated the inhibition properties for HMGR that are exhibited by these peptides. Among all peptides, YVAE showed the highest ability to inhibit HMGR. A kinetic analysis revealed that this peptide is a competitive inhibitor of HMG-CoA with an equilibrium constant of inhibitor binding (K(i)) of 15.2 +/- 1.4 microM. The calculated coefficient correlation (R) between log (IC(50)) and the inverse value of proximity parameter (1/Pr) was found to be 0.99, indicating a high degree of correlation and efficacy of the given approach in the peptide sequence design.  相似文献   

16.
Biliverdin reductase (BVR) catalyzes the last step in heme degradation by reducing the gamma-methene bridge of the open tetrapyrrole, biliverdin IXalpha, to bilirubin with the concomitant oxidation of a beta-nicotinamide adenine dinucleotide (NADH) or beta-nicotinamide adenine dinucleotide phosphate (NADPH) cofactor. Bilirubin is the major bile pigment in mammals and has antioxidant and anticompliment activity. We have determined X-ray crystal structures of apo rat BVR and its complex with NADH at 1.2 A and 1.5 A resolution, respectively. In agreement with an independent structure determination of the apo-enzyme, BVR consists of an N-terminal dinucleotide-binding domain (Rossmann-fold) and a C-terminal domain that contains a six-stranded beta-sheet that is flanked on one face by several alpha-helices. The C-terminal and N-terminal domains interact extensively, forming the active site cleft at their interface. The cofactor complex structure reported here reveals that the cofactor nicotinamide ring extends into the active site cleft, where it is adjacent to conserved amino acid residues and, consistent with the known stereochemistry of the reaction catalyzed by BVR, the si face of the ring is accessible for hydride transfer. The only titratable side-chain that appears to be suitably positioned to function as a general acid in catalysis is Tyr97. This residue, however, is not essential for catalysis, since the Tyr97Phe mutant protein retains 50% activity. This finding suggests that the dominant role in catalysis may be performed by hydride transfer from the cofactor, a process that may be promoted by proximity of the invariant residues Glu96, Glu123, and Glu126, to the nicotinamide ring.  相似文献   

17.
Delta-crystallin, the major soluble protein component of avian and reptilian eye lenses, is highly homologous to the urea cycle enzyme, argininosuccinate lyase (ASL). In duck lenses, there are two highly homologous delta crystallins, delta I and delta II, that are 94% identical in amino acid sequence. While delta II crystallin has been shown to exhibit ASL activity in vitro, delta I is enzymatically inactive. The X-ray structure of a His to Asn mutant of duck delta II crystallin (H162N) with bound argininosuccinate has been determined to 2.3 A resolution using the molecular replacement technique. The overall fold of the protein is similar to other members of the superfamily to which this protein belongs, with the active site located in a cleft formed by three different monomers in the tetramer. The active site of the H162N mutant structure reveals that the side chain of Glu 296 has a different orientation relative to the homologous residue in the H91N mutant structure [Abu-Abed et al. (1997) Biochemistry 36, 14012-14022]. This shift results in the loss of the hydrogen bond between His 162 and Glu 296 seen in the H91N and turkey delta I crystallin structures; this H-bond is believed to be crucial for the catalytic mechanism of ASL/delta II crystallin. Argininosuccinate was found to be bound to residues in each of the three monomers that form the active site. The fumarate moiety is oriented toward active site residues His 162 and Glu 296 and other residues that are part of two of the three highly conserved regions of amino acid sequence in the superfamily, while the arginine moiety of the substrate is oriented toward residues which belong to either domain 1 or domain 2. The analysis of the structure reveals that significant conformational changes occur on substrate binding. The comparison of this structure with the inactive turkey delta I crystallin reveals that the conformation of domain 1 is crucial for substrate affinity and that the delta I protein is almost certainly inactive because it can no longer bind the substrate.  相似文献   

18.
Ye Q  Hyndman D  Li X  Flynn TG  Jia Z 《Proteins》2000,38(1):41-48
Chinese hamster ovary (CHO) reductase is an enzyme belonging to the aldo-keto reductase (AKR) superfamily that is induced by the aldehyde-containing protease inhibitor ALLN (Inoue, Sharma, Schimke, et al., J Biol Chem 1993;268: 5894). It shows 70% sequence identity to human aldose reductase (Hyndman, Takenoshita, Vera, et al., J Biol Chem 1997;272:13286), which is a target for drug design because of its implication in diabetic complications. We have determined the crystal structure of CHO reductase complexed with nicotinamide adenine dinucleotide phosphate (NADP)+ to 2.4 A resolution. Similar to aldose reductase and other AKRs, CHO reductase is an alpha/beta TIM barrel enzyme with cofactor bound in an extended conformation. All key residues involved in cofactor binding are conserved with respect to other AKR members. CHO reductase shows a high degree of sequence identity (91%) with another AKR member, FR-1 (mouse fibroblast growth factor-regulated protein), especially around the variable C-terminal end of the protein and has a similar substrate binding pocket that is larger than that of aldose reductase. However, there are distinct differences that can account for differences in substrate specificity. Trp111, which lies horizontal to the substrate pocket in all other AKR members is perpendicular in CHO reductase and is accompanied by movement of Leu300. This coupled with movement of loops A, B, and C away from the active site region accounts for the ability of CHO reductase to bind larger substrates. The position of Trp219 is significantly altered with respect to aldose reductase and appears to release Cys298 from steric constraints. These studies show that AKRs such as CHO reductase are excellent models for examining the effects of subtle changes in amino acid sequence and alignment on binding and catalysis.  相似文献   

19.
Uracil-DNA glycosylase (UDG) removes uracil generated by the deamination of cytosine or misincorporation of deoxyuridine monophosphate. Within the UDG superfamily, a fifth UDG family lacks a polar residue in the active-site motif, which mediates the hydrolysis of the glycosidic bond by activation of a water molecule in UDG families 1-4. We have determined the crystal structure of a novel family 5 UDG from Thermus thermophilus HB8 complexed with DNA containing an abasic site. The active-site structure suggests this enzyme uses both steric force and water activation for its excision reaction. A conserved asparagine residue acts as a ligand to the catalytic water molecule. The structure also implies that another water molecule acts as a barrier during substrate recognition. Based on no significant open-closed conformational change upon binding to DNA, we propose a "slide-in" mechanism for initial damage recognition.  相似文献   

20.
R67 is a Type II dihydrofolate reductase (DHFR) that catalyzes the reduction of dihydrofolate (DHF) to tetrahydrofolate by facilitating the addition of a proton to N5 of DHF and the transfer of a hydride ion from NADPH to C6. Because this enzyme is a plasmid-encoded DHFR from trimethoprim-resistant bacteria, extensive studies on R67 with various methods have been performed to elucidate its reaction mechanism. Here, Raman difference measurements, conducted on the ternary complex of R67.NADP(+).DHF believed to be an accurate mimic of the productive DHFR.NADPH.DHF complex, show that the pK(a) of N5 in the complex is less than 4. This is in clear contrast to the behavior observed in Escherichia coli DHFR, a substantially more efficient enzyme, where the pK(a) of bound DHF at N5 is increased to 6.5 compared with its solution value of 2.6. A comparison of the ternary complexes in R67 and E. coli DHFRs suggests that enzymic raising of the pK(a) at N5 can significantly increase the catalytic efficiency of the hydride transfer step. However, R67 shows that even without such a strategy an effective DHFR can still be designed.  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