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.
Dihydrofolate reductase. The stereochemistry of inhibitor selectivity   总被引:7,自引:0,他引:7  
X-ray structural results are reported for 10 triazine and pyrimidine inhibitors of dihydrofolate reductase, each one studied as a ternary complex with NADPH and chicken dihydrofolate reductase. Analysis of these data and comparison with structural results from the preceding paper (Matthews, D.A., Bolin, J.T., Burridge, J.M., Filman, D.J., Volz, K.W., Kaufman, B. T., Beddell, C.R., Champness, J.N., Stammers, D.K., and Kraut, J. (1985) J. Biol. Chem. 260, 381-391) in which we contrasted binding of the antibiotic trimethoprim (TMP) to chicken dihydrofolate reductase on the one hand with its binding to Escherichia coli dihydrofolate reductase on the other, permit identification of differences that are important in accounting for TMP's selectivity. The crystallographic evidence strongly suggests that loss of a potential hydrogen bond between the 4-amino group of TMP and the backbone carbonyl of Val-115 when TMP binds to chicken dihydrofolate reductase but not when it binds to the E. coli reductase is the major factor responsible for this drug's more potent inhibition of bacterial dihydrofolate reductase. A key finding of the current study which is important in understanding why TMP binds differently to chicken and E. coli dihydrofolate reductases is that residues on opposite sides of the active-site cleft in chicken dihydrofolate reductase are about 1.5-2.0 A further apart than are structurally equivalent residues in the E. coli enzyme.  相似文献   

2.
The binding of trimethoprim to dihydrofolate reductase from L1210 mouse lymphoma cells has been studied by measuring the changes in chemical shift of nuclei of the ligand that accompanying binding. The 6- and 2',6'-proton chemical shifts of bound trimethoprim have been determined by transfer of saturation experiments, and the 2-carbon chemical shift has been determined by using [2-13C]trimethoprim. The changes in proton chemical shift are substantially smaller than those accompanying binding to bacterial dihydrofolate reductase [Cayley, P. J., Albrand, J. P., Feeney, J., Robert, G. C. K., Piper, E. A., & Burgen, A. S. V. (1979) Biochemistry 18, 3886]. It is shown that this difference arises largely from the fact that trimethoprim adopts different conformations when bound to mammalian and to bacterial dihydrofolate reductase. The proton chemical shifts are interpreted in terms of ring-current contributions from the two aromatic rings of trimethoprim itself and the nearby aromatic amino acid residues of the enzyme. The latter have been located by using the refined crystallographic coordinates of the Lactobacillus casei and Escherichia coli reductases in their complexes with methotrexate [Bolin, J. T., Filman, D. J., Matthews, D. A. & Kraut, J. (1982) J. Biol. Chem. 257, 13650], under the assumption that, as indicated by the 13C chemical shifts, the diaminopyrimidine ring of trimethoprim binds in the same way as does the corresponding part of methotrexate. With use of these assumptions, the conformation of trimethoprim bound to the dihydrofolate reductases from L. casei, E. coli, and L1210 cells has been calculated.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
C Bystroff  S J Oatley  J Kraut 《Biochemistry》1990,29(13):3263-3277
The crystal structure of dihydrofolate reductase (EC 1.5.1.3) from Escherichia coli has been solved as the binary complex with NADP+ (the holoenzyme) and as the ternary complex with NADP+ and folate. The Bragg law resolutions of the structures are 2.4 and 2.5 A, respectively. The new crystal forms are nonisomorphous with each other and with the methotrexate binary complex reported earlier [Bolin, J. T., Filman, D. J., Matthews, D. A., Hamlin, R. C., & Kraut, J. (1982) J. Biol. Chem. 257, 13650-13662]. In general, NADP+ and folate binding conform to predictions, but the nicotinamide moiety of NADP+ is disordered in the holoenzyme and ordered in the ternary complex. A mobile loop (residues 16-20) involved in binding the nicotinamide is also disordered in the holoenzyme. We report a detailed analysis of the binding interactions for both ligands, paying special attention to several apparently strained interactions that may favor the transition state for hydride transfer. Hypothetical models are presented for the binding of 7,8-dihydrofolate in the Michaelis complex and for the transition-state complex.  相似文献   

4.
5.
Dihydrofolate reductases from different species contain several highly conserved arginines, some of which have been shown by x-ray crystallography to have their guanido groups near the p-aminobenzoyl glutamate moiety of enzyme-bound methotrexate. The orientation of one of these (Arg-52) appears to be completely reversed in comparing the crystal structures of Escherichia coli with Lactobacillus casei enzyme (Bolin, J. T., Filman, D. J., Matthews, D. A., Hamlin, R. C., and Kraut, J. (1982). J. Biol. Chem. 257, 13650-13662). We synthesized a novel antifolate containing a glyoxal group designed to react specifically with active-site guanido groups which are able to approach the p-aminobenzoyl carbonyl of methotrexate. The binding of this compound to the enzyme was competitive with dihydrofolate (DHF) in ordinary buffers. In borate buffer at pH 8.0 it inactivated dihydrofolate reductases from both E. coli and L. casei at similar maximum rates, while the chicken liver enzyme was more slowly inactivated. The inactivation was stoichiometric, paralleled the loss of the glyoxal chromophore, and showed saturation kinetics. Inhibitor binding and thus inactivation was enhanced by NADPH, while DHF protected the enzyme. This allowed calculation of the Kd for DHF which was found to be identical with its Km. The stoichiometrically inactivated enzyme displayed the 340-nm chromophore characteristic of 4-aminopteridines bound to dihydrofolate reductase confirming active-site labeling with normal orientation of the ligand. The ligand remained covalently bound to inactivated enzyme upon denaturation at low pH but dissociated at neutral pH. Computer graphic modeling of the crystal structures predicted reaction of Arg-31 but not Arg-52 in L. casei dihydrofolate reductase and of only Arg-52 in the E. coli enzyme. Purification of the CNBr fragments from the inactivated enzymes gave a single labeled peptide for each species. The particular peptide tagged in each case was unaffected by the presence of NADPH and was in excellent agreement with the crystallographic predictions.  相似文献   

6.
The 2.2-A crystal structure of chicken liver dihydrofolate reductase (EC 1.5.1.3, DHFR) has been solved as a ternary complex with NADP+ and biopterin (a poor substrate). The space group and unit cell are isomorphous with the previously reported structure of chicken liver DHFR complexed with NADPH and phenyltriazine [Volz, K. W., Matthews, D. A., Alden, R. A., Freer, S. T., Hansch, C., Kaufman, B. T., & Kraut, J. (1982) J. Biol. Chem. 257, 2528-2536]. The structure contains an ordered water molecule hydrogen-bonded to both hydroxyls of the biopterin dihydroxypropyl group as well as to O4 and N5 of the biopterin pteridine ring. This water molecule, not observed in previously determined DHFR structures, is positioned to complete a proposed route for proton transfer from the side-chain carboxylate of E30 to N5 of the pteridine ring. Protonation of N5 is believed to occur during the reduction of dihydropteridine substrates. The positions of the NADP+ nicotinamide and biopterin pteridine rings are quite similar to the nicotinamide and pteridine ring positions in the Escherichia coli DHFR.NADP+.folate complex [Bystroff, C., Oatley, S. J., & Kraut, J. (1990) Biochemistry 29, 3263-3277], suggesting that the reduction of biopterin and the reduction of folate occur via similar mechanisms, that the binding geometry of the nicotinamide and pteridine rings is conserved between DHFR species, and that the p-aminobenzoylglutamate moiety of folate is not required for correct positioning of the pteridine ring in ground-state ternary complexes. Instead, binding of the p-aminobenzoylglutamate moiety of folate may induce the side chain of residue 31 (tyrosine or phenylalanine) in vertebrate DHFRs to adopt a conformation in which the opening to the pteridine binding site is too narrow to allow the substrate to diffuse away rapidly. A reverse conformational change of residue 31 is proposed to be required for tetrahydrofolate release.  相似文献   

7.
The effects of pH upon the C-2 resonances of the 5 histidine residues of Escherichia coli MB 1428 dihydrofolate reductase in binary complexes with methotrexate, aminopterin, folate, methopterin, and trimethoprim were studied by 300-MHz 1H nmr spectroscopy. Three of the five histidine residues, labeled 1, 2, and 3, exhibited similar pK' values and chemical shifts for their C-2 protons in the five binary complexes. One histidine, 4, was quite different in the folate complex and the last histidine, 5 was quite different in the trimethoprim complex. For all five binary complexes, each histidine had a pK' which was significantly different from the other 4 histidines of that complex. Titration of the binary methotrexate complex of a 5,5'-dithiobis(2-nitrobenzoate)-modified enzyme showed that 2 histidines were not perturbed by this modification of Cys 152, and that the alkaline form of histidine 2, the acid form of histidine 4, and, to a lesser extent, the acid form of histidine 3 were slightly perturbed. Titration of the binary methotrexate complex of a N-bromosuccinimide-modified enzyme demonstrated that this modification slightly affected all of the histidines and drastically affected histidine 5. Histidines 3 and 5 of the binary methotrexate complex reacted rapidly with the histidine-specific reagent, ethoxyformic anhydride, while histidines 2 and 4 reacted at a moderate rate and histidine 1 reacted slowly if at all. The local electrostatic environments of the 5 histidine residues as deduced from the crystal structure of the binary complex of the enzyme with methotrexate (Matthews, D.A., Alden, R.A., Bolin, J.T., Freer, S.T., Hamlin, R., Xuong, N., Kraut, J., Poe, M., Williams, M.N., and Hoogsteen, K. (1977) Science 197, 594-597) were used as the basis for proposed assignments of the five histidine C-2 nmr resonances. The assignments were: 1, pK' 7.9 to 8.2, His 124; 2, pK' 7.2 to 7.4, His 141; 3, pK' 6.5 to 6.7, His 149; 4, pK' 5.7 to 6.3, His 114; and 5, pK' 5.2 to 5.9, His 45. The effect of the chemical modifications upon the enzyme's histidine residues were consistent with the assignments, but no direct chemical evidence in support of the assignments was obtained. It was proposed that, since the crystallographic data provided consistent assignments of the histidine nmr data for both native and chemically modified enzyme, the local environment of each of the 5 histidine residues was similar in the crystal and in solution.  相似文献   

8.
We have explored the substrate protonation mechanism of Escherichia coli dihydrofolate reductase by changing the location of the proton donor. A double mutant was constructed in which the proton donor of the wild-type enzyme, aspartic acid-27, has been changed to serine and simultaneously an alternative proton donor, glutamic acid, has replaced threonine at position 113. The active site of the resulting variant enzyme molecule should therefore somewhat resemble that proposed for the R67 plasmid-encoded dihydrofolate reductase [Matthews, D. A., Smith, S. L., Baccanari, D. P., Burchall, J. J., Oatley, S. J., & Kraut, J. (1986) Biochemistry 25, 4194]. At pH 7, the double-mutant enzyme has a 3-fold greater kcat and an unchanged Km(dihydrofolate) as compared with the single-mutant Asp-27----Ser enzyme described previously [Howell, E. E., Villafranca, J. E., Warren, M. S., Oatley, S. J., & Kraut, J. (1986) Science (Washington, D.C.) 231, 1123]. Additionally, its activity vs pH profiles together with observed deuterium isotope effects, suggest that catalysis depends on an acidic group with a pKa of 8. It is concluded that the dihydropteridine ring of a bound substrate molecule can indeed be protonated by a glutamic acid side chain at position 113 (instead of an aspartic acid side chain at position 27), but with greatly decreased efficiency: at pH 7, the double mutant still has a 25-fold lower kcat (1.2 s-1) and a 2900-fold lower kcat/km(dihydrofolate) (8.6 X 10(3) s-1 M-1) than the wild-type enzyme.  相似文献   

9.
Crystal structure of human dihydrofolate reductase complexed with folate   总被引:8,自引:0,他引:8  
The crystal structure of recombinant human dihydrofolate reductase with folate bound in the active site has been determined and the structural model refined at 0.2-nm resolution. Preliminary studies of the binding of the inhibitors methotrexate and trimethoprim to the human apoenzyme have been performed at 0.35-nm resolution. The conformations of the chemically very similar ligands folate and methotrexate, one a substrate the other a potent inhibitor, differ substantially in that their pteridine rings are in inverse orientations relative to their p-aminobenzoyl-L-glutamate moieties. Methotrexate binding is similar to that previously observed in two bacterial enzymes but is quite different from that observed in the enzyme from a mouse lymphoma cell line [Stammers et al. (1987) FEBS Lett. 218, 178-184]. The geometry of the polypeptide chain around the folate binding site in the human enzyme is not consistent with conclusions previously drawn with regard to the species selectivity of the inhibitor trimethoprim [Matthews et al. (1985) J. Biol. Chem. 260, 392-399].  相似文献   

10.
T G Consler  S H Woodard  J C Lee 《Biochemistry》1989,28(22):8756-8764
Pyruvate kinase is an important glycolytic enzyme which is expressed differentially as four distinct isozymes whose catalytic activity is regulated in a tissue-specific manner. The kidney isozyme is known to exhibit sigmoidal kinetics, whereas the muscle isozyme exhibits hyperbolic kinetic properties. By integration of the crystallographic [Stuart, D. I., Levine, M., Muirhead, H., & Stammers, D.K. (1979) J. Mol. Biol. 134, 109-142] and primary sequence data [Noguchi, T., Inoue, H., & Tanaka, T. (1986) J. Biol. Chem. 261, 13807], it was shown that the primary sequence for the C alpha 1 and C alpha 2 regions may constitute the allosteric switching site. To provide insights into the effects of the localized sequence change on the global structural and functional behavior of the enzyme, kinetic studies under a wide spectrum of conditions were conducted for both the muscle and kidney isozymes. These conditions include measurements of enzyme activity as a function of substrate concentrations with different concentrations of allosteric inhibitors or activators. These results showed that both isozymes exhibit the same regulatory properties although quantitatively the distribution of active and inactive forms and the various dissociation constants which govern the binding of substrate and allosteric effectors with the enzyme are different. For such a majority of equilibrium constants to be altered, the localized primary sequence change must confer global perturbations which are manifested as differences in the various equilibrium constants. Structural information about these two isozymes was provided by phase-modulation measurement of the fluorescence lifetime of tryptophan residues under a variety of experimental conditions.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
The Escherichia coli lit gene product is required for the multiplication of bacteriophage T4 at temperatures below 34 degrees C. After infection of a lit mutant host, early gene product synthesis is normal, as is T4 DNA replication; however, the late gene products never appear, and early gene product synthesis eventually ceases. Consequently, at late times, there is no protein synthesis of any kind (W. Cooley, K. Sirotkin, R. Green, and L. Snyder, J. Bacteriol. 140:83-91, 1979; W. Champness and L. Snyder, J. Mol. Biol. 155:395-407, 1982), and no phage are produced. We have isolated T4 mutants which can multiply in lit mutant hosts. The responsible T4 mutations (called gol mutations) completely overcome the block to T4 gene expression (Cooley et al., J. Bacteriol. 140:83-91). We have proposed that gol mutations alter a cis-acting regulatory site on T4 DNA rather than a diffusible gene product and that the wild-type form of the gol site (gol+) somehow interferes with gene expression late in infection (Champness and Snyder, J. Mol. Biol. 155:395-409). In this communication, we report the sequence of the gol region of the T4 genome from five different gol mutants. The gol mutations are all single-base-pair transitions within 40 base pairs of DNA. Therefore, the gol site is at least 40 base pairs long. The sequence data confirm that the gol phenotype is not due to an altered protein. We also report that the gol+ site in plasmids prevents transformation of Lit- but not Lit+ E. coli. Thus, the gol site is at least partially active in the absence of the T4 genome.  相似文献   

12.
The sequence of more than 1,000 amino acid residues, derived from two different isoforms, has been determined from peptides generated from purified human erythrocyte membrane Ca2(+)-ATPase (hPMCA). Several of these peptide sequences correspond to the previously reported, cDNA deduced sequence of the "teratoma" Ca2+ pump isoform hPMCA1 (Verma, A. K., Filoteo, A. G., Stanford, D. R., Wieben, E. D., Penniston, J. T., Strehler, E. E., Fischer, R., Heim, R., Vogel, G., Matthews, S., Strehler-Page, M.-A., James, P., Vorherr, T., Krebs, J., and Carafoli, E. (1988) J. Biol. Chem. 263, 14152-14159). The complete primary structure of a novel isoform (hPMCA3) has been determined by molecular cloning and nucleotide sequencing of its corresponding cDNA. This new member of the plasma membrane Ca2+ pump family consists of 1,205 amino acid residues with a calculated Mr of 133,930, and it shows 88% similarity (75% identity) with the previously sequenced pump isoform. Specific probes detect major mRNA species of 5.6 kilobases for hPMCA1, and of 7.5 kilobases for hPMCA3, on Northern blots of human K562 erythroleukemic cell RNA. A large number of peptide sequences match perfectly with only one or the other of these isoforms and all peptides (with 6 exceptions corresponding to a contaminant protein or to a third minor Ca2+ pump isoform) are found in either only one or in both of the isoforms. The two erythrocyte Ca2+ pumps display high sequence divergence in a few localized regions that may determine isoform-specific functional specializations; for example, the putative extracellular loop separating transmembrane domains 1 and 2, the highly negatively charged region previously suggested to be involved in Ca2+ binding, and the site of cAMP-dependent protein kinase phosphorylation.  相似文献   

13.
ATP-dependent proteases are responsible for most energy-dependent protein degradation across all species. Proteases initially bind an unstructured region on a substrate and then translocate along the polypeptide chain, unfolding and degrading protein domains as they are encountered. Although this process is normally processive, resulting in the complete degradation of substrate proteins to small peptides, some substrates are released prematurely. Regions of low sequence complexity within the substrate such as the glycine-rich region (GRR) from p105 or glycine-alanine repeats (GAr) from the EBNA1 (Epstein-Barr virus nuclear antigen-1) protein, can trigger partial degradation and fragment release. Loss of processivity could be due to inability to hold on to the substrate (faster release) or inability to unfold and degrade a substrate domain (slower unfolding). I previously showed that the GRR slows domain unfolding by the proteasome (Kraut, D. A., Israeli, E., Schrader, E. K., Patil, A., Nakai, K., Nanavati, D., Inobe, T., and Matouschek, A. (2012) ACS Chem. Biol. 7, 1444–1453). In contrast, a recently published study concluded that GArs increase the rate of substrate release from ClpXP, a bacterial ATP-dependent protease (Too, P. H., Erales, J., Simen, J. D., Marjanovic, A., and Coffino, P. (2013) J. Biol. Chem. 288, 13243–13257). Here, I show that these apparently contradictory results can be reconciled through a reanalysis of the ClpXP GAr data. This reanalysis shows that, as with the proteasome, low complexity sequences in substrates slow their unfolding and degradation by ClpXP, with little effect on release rates. Thus, despite their evolutionary distance and limited sequence identity, both ClpXP and the proteasome share a common mechanism by which substrate sequences regulate the processivity of degradation.  相似文献   

14.
Fifteen single-site charge-reversal mutations of yeast cytochrome c peroxidase (CcP) have been constructed to determine the effect of localized charge on the catalytic properties of the enzyme. The mutations are located on the front face of CcP, near the cytochrome c binding site identified in the crystallographic structure of the yeast cytochrome c-CcP complex [Pelletier, H., and Kraut, J. (1992) Science 258, 1748-1755]. The mutants are characterized by absorption spectroscopy and hydrogen peroxide reactivity at both pH 6.0 and 7.5 and by steady-state kinetic studies using recombinant yeast iso-1-ferrocytochrome c(C102T) as a substrate at pH 7.5. Some of the charge-reversal mutations cause detectable changes in the absorption spectrum, especially at pH 7.5, reflecting changes in the equilibrium between penta- and hexacoordinate heme species in the enzyme. An increase in the amount of hexacoordinate heme in the mutant enzymes correlates with an increase in the fraction of enzyme that does not react with hydrogen peroxide. Steady-state velocity measurements indicate that five of the 15 mutations cause large increases in the Michaelis constant (R31E, D34K, D37K, E118K, and E290K). These data support the hypothesis that the cytochrome c-CcP complex observed in the crystal is the dominant catalytically active complex in solution.  相似文献   

15.
The present studies were performed to attempt to elucidate the basis for the discrepancy between results of Kondo and Schulz (1976, Biochim. Biophys Acta 419, 76–92), who found that cholecystokinin and cholinergic agents increase uptake of 45Ca by dispersed acinar cells from rat pancreas, and results of others (Matthews, E.K., Petersen, O.H. and Williams, J.A. (1973) J. Physiol. 234, 689–701; Chandler, D.E. and Williams, J.A. (1974) J. Physiol. 243, 831–846; Case, R.M. and Clausen, T. (1973) J. Physiol. 235, 75–102; Gardner, J.D., Conlon, T.P., Klaeveman, H.L., Adams, T.D. and Ondetti, M.A. (1975) J. Clin. Invest. 56, 366–375; Christophe, J.P., Frandsen, E.K., Conlon, T.P., Krishna, G. and Gardner, J.D. (1976) J. Biol. Chem. 251, 4640–4645; Shelby, H.T., Gross, L.P., Lichty, P. and Gardner, J.D. (1976) J. Clin. Invest. 58, 1482–1493 and Deschodt-Lanckman, M., Robberecht, P., de Neef, P., Lammens, M. and Christophe, J. (1976) J. Clin. Invest. 58, 891–898). They have reported that cholecystokinin and cholinergic agents do not alter or cause a slight decrease in uptake of 45Ca by pancreatic acinar cells. Our present results indicate that increased uptake of 45Ca by acinar cells incubated with cholecystokinin occurs only in cells washed with iced, 160 mM choline chloride and reflects increased cellular uptake of radioactivity from the wash solution but not from the incubation medium. We detected no effect of cholecystokinin on uptake of 45Ca by cells washed with 160 mM choline chloride containing 5 mM ethylenediaminetetraacetate or by cells washed with Krebs-Ringer bicarbonate. Furthermore, cells washed with 160 mM choline chloride accumulated a substantial amount of 45Ca from the wash solution and this accumulation was increased in cells that had been preincubated with cholecystokinin. Cells washed with Krebs-Ringer bicarbonate did not take up 45Ca from the wash solution.  相似文献   

16.
A C Anusiem  M Kelleher 《Biopolymers》1984,23(7):1147-1167
Interest in the thermodynamics of the iron-binding site in hemoproteins has increased in recent years due to refinements in x-ray crystallographic studies of hemoproteins [see Deathage, J. F., Lee, R. S., Anderson, C. M. & Moffat, K. (1976) J. Mol. Biol. 104 , 687–706; Heidner, E. J., Ladner, R. C. & Perutz, M. F. (1976) J. Mol. Biol. 104 , 707–722; Deathage, J. F., Lee, R. S. & Moffat, K. (1976) J. Mol. Biol. 104 , 723–728; Ladner, R. C., Heidner, E. J. & Perutz, M. F. (1976) J. Mol. Biol. 114 , 385–414; Fermi, G. & Perutz, M. F. (1977) J. Mol. Biol. 114 , 421–431; Takano, T. (1977) J. Mol. Biol. 110 , 537–568 and 569–589], the synthesis and x-ray analysis of model heme compounds [see Scheidt, W. R. (1977) Acc. Chem. Res. 10 , 339–345; Kastner, M. E., Scheidt, W. R., Mashino, T. & Reed, C. A. (1978) J. Am. Chem. Soc. 100 , 666–667; Mashiko, T., Kastner, M. E., Spartalian, K., Scheidt, W. R. & Reed, C. A. (1978) J. Am. Chem. Soc. 100 , 6354–6362; Hill, H. A. O., Skite, P. P., Buchler, J. W., Luchr, H., Tonn, M., Gregson, A. K. & Pellizer, G. (1979) Chem. Commun. 4 , 151–152; and Scheidt, W. R., Cohen, I. A. & Kastner, M. E. (1979) Biochemistry 18 , 3546–3556], and the numerous data on heme–protein interactions that account for the differences observed in ligand binding between the various species of animals. Numerous probes have been used and provide information about the structure and thermodynamics of the binding site, but no single probe can provide the complete picture [see Iizuka, T. & Yonetani, T. (1970) Adv. Biophys. 1 , 157–182; Smith, D. W. & Williams, R. J. P. (1970) Struct. Bond. 7 , 1–45; and Spiro, T. G. (1975) Biochim. Biophys. Acta 416 , 169–189].  相似文献   

17.
The primary site of mercury-induced injury is the kidney due to uptake of the reactive Hg(2+)-conjugated organic anions in the proximal tubule. Here, we investigated the in vivo role of Oat1 (organic anion transporter 1; originally NKT (Lopez-Nieto, C. E., You, G., Bush, K. T., Barros, E. J., Beier, D. R., and Nigam, S. K. (1997) J. Biol. Chem. 272, 6471-6478)) in handling of known nephrotoxic doses of HgCl(2). Oat1 (Slc22a6) is a multispecific organic anion drug transporter that is expressed on the basolateral aspects of renal proximal tubule cells and that mediates the initial steps of elimination of a broad range of endogenous metabolites and commonly prescribed pharmaceuticals. Mercury-induced nephrotoxicity was observed in a wild-type model. We then used the Oat1 knock-out to determine in vivo whether the renal injury effects of mercury are mediated by Oat1. Most of the renal injury (both histologically and biochemically as measured by blood urea nitrogen and creatinine) was abolished following HgCl(2) treatment of Oat1 knock-outs. Thus, acute kidney injury by HgCl(2) was found to be mediated mainly by Oat1. Our findings raise the possibility that pharmacological modulation of the expression and/or function of Oat1 might be an effective therapeutic strategy for reducing renal injury by mercury. This is one of the most striking phenotypes so far identified in the Oat1 knock-out. (Eraly, S. A., Vallon, V., Vaughn, D. A., Gangoiti, J. A., Richter, K., Nagle, M., Monte, J. C., Rieg, T., Truong, D. M., Long, J. M., Barshop, B. A., Kaler, G., and Nigam, S. K. (2006) J. Biol. Chem. 281, 5072-5083).  相似文献   

18.
The plasma membrane calcium/calmodulin-dependent calcium ATPase (PMCA) (Shull, G.E., and J. Greeb. 1988. J. Biol. Chem. 263:8646-8657; Verma, A.K., A.G. Filoteo, D.R. Stanford, E.D. Wieben, J.T. Penniston, E.E. Strehler, R. Fischer, R. Heim, G. Vogel, S. Mathews, et al. 1988. J. Biol. Chem. 263:14152-14159; Carafoli, E. 1997. Basic Res. Cardiol. 92:59-61) has been proposed to be a regulator of calcium homeostasis and signal transduction networks of the cell. However, little is known about its precise mechanisms of action. Knock-out of (mainly neuronal) isoform 2 of the enzyme resulted in hearing loss and balance deficits due to severe inner ear defects, affecting formation and maintenance of otoconia (Kozel, P.J., R.A. Friedman, L.C. Erway, E.N. Yamoah, L.H. Liu, T. Riddle, J.J. Duffy, T. Doetschman, M.L. Miller, E.L. Cardell, and G.E. Shull. 1998. J. Biol. Chem. 273:18693-18696). Here we demonstrate that PMCA 4b is a negative regulator of nitric oxide synthase I (NOS-I, nNOS) in HEK293 embryonic kidney and neuro-2a neuroblastoma cell models. Binding of PMCA 4b to NOS-I was mediated by interaction of the COOH-terminal amino acids of PMCA 4b and the PDZ domain of NOS-I (PDZ: PSD 95/Dlg/ZO-1 protein domain). Increasing expression of wild-type PMCA 4b (but not PMCA mutants unable to bind PDZ domains or devoid of Ca2+-transporting activity) dramatically downregulated NO synthesis from wild-type NOS-I. A NOS-I mutant lacking the PDZ domain was not regulated by PMCA, demonstrating the specific nature of the PMCA-NOS-I interaction. Elucidation of PMCA as an interaction partner and major regulator of NOS-I provides evidence for a new dimension of integration between calcium and NO signaling pathways.  相似文献   

19.
A method has been used to quantitate the reaction between eukaryotic type I DNA topoisomerase and topological forms of DNA. This procedure (Trask, D.K., DiDonato, J.D. and Muller, M.T. (1984) Eur. Mol. Biol. Organ. J. 3, 671-676) measures the efficiency of DNA cleavage and concurrent formation of a covalent enzyme/DNA complex. Eukaryotic type I topoisomerases react preferentially by 5-10-fold with supercoiled DNA. The effect of supercoiling is clearly evident in that both the initial rate and final extent of the reaction is elevated. Because the dissociation rate is much lower than the association rate, it is possible to isolate native topoisomerase/DNA complexes. These complexes are comprised of enzyme molecules which are catalytically active when challenged with a second supercoiled DNA substrate. Collectively, the data support the conclusion that a functional intermediate in the reaction sequence is being detected and that the avian topoisomerase I preferentially cleaves supercoiled DNA.  相似文献   

20.
The 2.2-A X-ray structure for CCP(MI), a plasmid-encoded form of Saccharomyces cerevisiae cytochrome c peroxidase (CCP) expressed in Escherichia coli [Fishel, L.A., Villafranca, J. E., Mauro, J. M., & Kraut, J. (1987) Biochemistry 26, 351-360], has been solved, together with the structures of three specifically designed single-site heme-cleft mutants. The structure of CCP(MI) was solved by using molecular replacement methods, since its crystals grow differently from the crystals of CCP isolated from bakers' yeast used previously for structural solution. Small distal-side differences between CCP(MI) and bakers' yeast CCP are observed, presumably due to a strain-specific Thr-53----Ile substitution in CCP(MI). A Trp-51----Phe mutant remains pentacoordinated and exhibits only minor distal structural adjustments. The observation of a vacant sixth coordination site in this structure differs from the results of solution resonance Raman studies, which predict hexacoordinated high-spin iron [Smulevich, G., Mauro, J.M., Fishel, L. A., English, A. M., Kraut, J., & Spiro, T. G. (1988) Biochemistry 27, 5477-5485]. The coordination behavior of this W51F mutant is apparently altered in the presence of a precipitating agent, 30% 2-methyl-2,4-pentanediol. A proximal Trp-191----Phe mutant that has substantially diminished enzyme activity and altered magnetic properties [Mauro, J. M., Fishel, L. F., Hazzard, J. T., Meyer, T. E., Tollin, G., Cusanovich, M. A., & Kraut, J. (1988) Biochemistry 27, 6243-6256] accommodates the substitution by allowing the side chain of Phe-191, together with the segment of backbone to which it is attached, to move toward the heme. This relatively large (ca. 1 A) local perturbation is accompanied by numerous small adjustments resulting in a slight overall compression of the enzyme's proximal domain; however, the iron coordination sphere is essentially unchanged. This structure rules out a major alteration in protein conformation as a reason for the dramatically decreased activity of the W191F mutant. Changing proximal Asp-235 to Asn results in two significant localized structural changes. First, the heme iron moves toward the porphyrin plane, and distal water 595 now clearly resides in the iron coordination sphere at a distance of 2.0 A. The observation of hexacoordinated iron for the D235N mutant is in accord with previous resonance Raman results. Second, the indole side chain of Trp-191 has flipped over as a result of the mutation; the tryptophan N epsilon takes part in a new hydrogen bond with the backbone carbonyl oxygen of Leu-177.(ABSTRACT TRUNCATED AT 400 WORDS)  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