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.
A series of new compounds, N,N'-bis(dialkylphosphoryl)diamines and S,S'-bis(dialkylphosphoryl)-1,3-propanedithiols were prepared by a Todd-Atherton like reaction of dialkylphosphites with symmetrical diamines and 1,3-propanedithiols in a biphasic system [F.R. Athertoon, H.T. Howard, A.R. Todd, J. Chem. Soc. (1948) 1106-1111; F.R. Athertoon, H.T. Openshaw, A.R. Todd, J. Chem. Soc. (1945) 660-663]. The structures were characterized by IR, 1H NMR, 13C NMR and mass spectrometry. Compounds with butoxy, isobutoxy and isopropoxy groups linked in the phosphorus atom showed the lowest LD50 values when tested against Musca domestica and Stomoxys calcitrans. The pharmacological and toxicological evaluation of N,N'-bis(diisobutylphosphoryl)-1,3-propylenediamine and S,S'-bis(diisobutylphosphoryl)-1,3-propanedithiol, which were very active against M. domestica and S. calcitrans, demonstrated that these compounds present no toxicological effects against mice in a concentration of 200mg/kg. An explanation for the observed activity profile is presented based on results obtained in a molecular modeling study with insect and mammalian acetylcholinesterase models.  相似文献   

2.
K Bruzik  M D Tsai 《Biochemistry》1984,23(8):1656-1661
Chirally labeled 1,2-dipalmitoyl-sn-glycero-3-phosphocholines (DPPC) with known configuration were synthesized by N-methylation of chirally labeled 1,2-dipalmitoyl-sn-glycero-3-phosphoethanolamine (DPPE). Transphosphatidylation of (RP)- and (SP)-[18O]DPPC catalyzed by phospholipase D from cabbage gave (RP)- and (SP)-[18O]DPPE, respectively, as indicated by 31P nuclear magnetic resonance (NMR) analysis of [18O]DPPE. Therefore, phospholipase D catalyzes transphosphatidylation with overall retention of configuration at phosphorus. The steric course of hydrolysis of DPPC catalyzed by the same enzyme was elucidated by the following procedures. Hydrolysis of (RP)-[17O, 18O]DPPC by phospholipase D gave 1,2-dipalmitoyl-sn-glycero-3-[ 16O , 17O, 18O]phosphate ( [ 16O , 17O, 18O] DPPA ) with unknown configuration. The latter compound was then converted to 1-[ 16O , 17O, 18O]phospho-(R)-propane-1,2-diol by a procedure involving no P-O bond cleavage [ Bruzik , K., & Tsai, M.-D. (1984) J. Am. Chem. Soc. 106, 747-754]. The configuration of the phosphopropane -1,2-diol was determined as RP by 31P NMR analysis following ring closure and methylation [ Buchwald , S. L., & Knowles, J. R. (1980) J. Am. Chem. Soc. 102, 6601-6603]. The results indicated that hydrolysis of DPPC catalyzed by phospholipase D also proceeds with retention of configuration at phosphorus. Our results therefore support a two-step mechanism involving a phosphatidyl-enzyme intermediate in the reactions catalyzed by phospholipase D from cabbage.  相似文献   

3.
The study of sulfur within the plasma cells of Ascidia ceratodes [Carlson, R. M. K. (1975) Proc. Natl. Acad. Sci. U.S.A. 72, 2217-2221; Frank, P., Carlson, R. M. K., & Hodgson, K. O. (1986) Inorg. Chem. 25, 470-478; Hedman, B., Frank, P., Penner-Hahn, J. E., Roe, A. L., Hodgson, K. O., Carlson, R. M. K., Brown, G., Cerino, J., Hettel, R., Troxel, T., Winick, H., & Yang, J. (1986) Nucl. Instrum. Methods Phys. Res., Sect. A 246, 797-800] has been extended with X-ray absorption near-edge structure (XANES) spectroscopy. An intense absorption feature at 2482.4 eV and a second feature at 2473.7 eV indicate a large endogenous sulfate concentration, as well as smaller though significant amounts of thiol or thioether sulfur, respectively. A strong shoulder was observed at 2481.7 eV on the low-energy side of the sulfate absorption edge, deriving from a novel type of sulfur having a slightly lower oxidation state than sulfate sulfur. The line width of the primary transition on the sulfur edge of a vanadium (III) sulfate solution was found to be broadened relative to that of sodium sulfate, possibly deriving from the formation of the VSO4+ complex ion [Britton, H. T. S., & Welford, G. (1940) J. Chem. Soc., 761-764; Duffy, J. A., & Macdonald, W. J. D. (1970) J. Chem. Soc., 977-980; Kimura, T., Morinaga, M., & Nakano, J. (1972) Nippon Kagaku Zaishi, 664-667]. Similar broadening appears to characterize the oxidized sulfur types in vanadocytes. A very good linear correlation between oxidation state and peak position (in electronvolts) was found for a series of related sulfur compounds. This correlation was used to determine a 5+ oxidation state for the additional sulfur type at 2481.7 eV. (ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
Hydroxymethylbilane synthase (porphobilinogen deaminase) from Escherichia coli uses a novel pyrromethane cofactor to bind the growing pyrrolic chain for hydroxymethylbilane biosynthesis [Hart, Miller, Leeper & Battersby (1987) J. Chem. Soc. Chem. Commun. 1762-1765]. We show that this cofactor is bound to the protein through the sulphur atom of a cysteine residue.  相似文献   

5.
J L Weaver  R W Williams 《Biochemistry》1988,27(25):8899-8903
Measurements of the colligative properties of nucleosides and their derivatives have shown that bases form transient aggregates in solution [Ts'o (1967) J. Am. Chem. Soc. 89, 3612-3622]. Aggregation of nucleotides cannot be measured by osmometry due to the presence of counterions. Sedimentation measurements are difficult to obtain and have been complicated by differences in pH [Ferguson et al. (1974) Biophys. Chem. 1, 325-337]. Raman studies of oligonucleotides have shown that the intensities due to base vibrational modes depend on the extent of base stacking, but this dependence has not been quantitated. We have measured this dependence by relating changes in the Raman spectra of nucleotides and nucleosides with previous measurements of colligative properties. Visible Raman spectra of ATP, AMP, and adenosine, taken over a range of concentrations from 1 to 1000 mM, show that the peak intensity ratio (I1305 + I1380)/I1340 varies linearly with the log of the concentration for all three bases. This concentration-dependent change correlates with published molal osmotic coefficient data for functionally similar bases with a correlation coefficient of 0.99. In contrast, UV resonance Raman spectra of the same bases show changes that vary linearly with concentration.  相似文献   

6.
For gadolinium chelates, we determined that there is a linear correlation between calculated solvent-accessible surface area and q-value, the number of rapidly exchanging water molecules directly bound to the gadolinium ion. A calibration curve was developed to predict q-value based on the solvent-accessible surface area of gadolinium. This predictive method was validated with the following gadolinium crystal structures: (ethylenediaminetetraacetic acid)-gadolinium(III) [Gd(EDTA)] [Templeton, L. K., Templeton, D. H., Zalkin, A., and Ruben, H. W. (1982) Anomalous Scattering by Praseodymium, Samarium, and Gadolinium and Structures of their Thylenediaminetetraacetate (EDTA) Salts. Acta Crystallogr., Sect. B 38, 2155], (1,4,7,10-tetraazacyclododecane-N,N',N' ',N' "-tetraacetic acid)-gadolinium(III) [Gd(DOTA)] [Dubost, J.-P., Leger, J.-M., Langlois, M.-H., Meyer, D., and Schaefer, M. (1991) Structure of a Magnetic Resonance Imaging Agent - The Gadolinium-DOTA Complex C(16)H(24)N(4)O(8)NaGd, 5H(2)O. C. R. Acad. Sci., Ser. 2 312, 349], (diethylenetriaminepentaacetic acid)-gadolinium(III) [Gd(DTPA)] [Stezowski, J. J., and Hoard, J. L. (1984) Heavy Metal Ionophores - Correlations Among Structural Parameters of Complexed Nonpeptide Polyamino Acids. Isr. J. Chem. 24, 323], (diethylenepenta-acetato)-gadolinium(III) [Gd(DTPA-BEA)] [Smith, P. H., Brainard, J. R., Morris, D. E., Jarvinen, G. D., and Ryan, R. R. (1989) Solution and Solid-State Characterization of Europium and Gadolinium Schiff-Base Complexes and Assessment of their Potential as Contrast Agents in Magnetic Resonance Imaging. J. Am. Chem. Soc. 111, 7437], and (1,7,13-triaza-4,10, 16-trioxacyclo-octadecane-N,N',N' '-triacetato)-gadolinium(III) [Gd(TTTA)] [Chen, D., Squattrito, P. J., Martell, A. E., and Clearfield, A. (1990) Synthesis and Crystal Structure of a 9-Coordinate Gadolinium(III) Complex of 1,7,13-Triaza-4,10, 16-Trioxacyclooctadecane-N,N',N' '-Tri-Acetic Acid. Inorg. Chem. 29, 4366]. Predicted q-values were in complete agreement with experimentally determined q-values. A genetic algorithm-based conformational search method was developed to generate valid 3D models for gadolinium chelates. The method was successfully tested on the following gadolinium chelates: Gd(EDTA) (Templeton et al., 1982), Gd(DOTA) (Dubost et al., 1991), Gd(DTPA-BEA) (Smith et al., 1989), Gd(TTTA) (Chen et al., 1990), Gd(triethylene glycol) [Rogers, R. D., Voss, E. J., and Etzenhouser, R. D. (1988) F-Element Crown Ether Complexes. 17. Synthetic and Structural Survey of Lanthanide Chloride Tiethylene Glycol Complexes. Inorg. Chem. 27, 533], and Gd(tetraethylene glycol) [Rogers, R. D., Etzenhouser, R. D., Murdoch, J. S., and Reyes, E. (1991) Macrocycle Complexation Chemistry. 35. Survey of the Complexation of the Open-Chain 15-Crown-5 Analogue tetraethylene Glycol with the Lanthanide Chlorides. Inorg. Chem. 30, 1445].  相似文献   

7.
The molecular structure of the active site of myoglobin from the shark, Galeorhinus japonicus, has been studied by 1H-NMR. Some hyperfine-shifted amino acid proton resonances in the met-cyano form of G. japonicus myoglobin have been unambiguously assigned by the combined use of various two-dimensional NMR techniques; they were compared with the corresponding resonances in Physter catodon myoglobin. The orientations of ThrE10 and IleFG5 residues relative to the heme in G. japonicus met-cyano myoglobin were semiquantitatively estimated from the analysis of their shifts using the magnetic susceptibility tensor determined by a method called MATDUHM (magnetic anisotropy tensor determination utilizing heme methyls) [Yamamoto, Y., Nanai, N. & Ch?j?, R. (1990) J. Chem. Soc., Chem. Commun., 1556-1557] and the results were compared with the crystal structure of P. catodon carbonmonoxy myoglobin [Hanson, J. C. & Schoenborn, B. P. (1981) J. Mol. Biol. 153, 117-124]. In spite of a substantial difference in shift between the corresponding amino acid proton resonances for the two proteins, the orientations of these amino acid residues relative to the heme in the active site of both myoglobins were found to be highly alike.  相似文献   

8.
The interaction of formate and acetate ions with cobalt-substituted carbonic anhydrase (CA) has been investigated through 13C-NMR and one-dimensional and two-dimensional 1H-NMR spectroscopy. 13C data on formate are consistent with a regularly coordinated ligand, as previously proposed for the acetate anion [Bertini, I., Luchinat, C. & Scozzafava, A. (1977) J. Chem. Soc. Dalton Trans., 1962-1965]. 1H-NOE experiments on both anions give evidence of through-space interactions between ligand protons and protein protons. The latter are assigned to specific residues in the active cavity through nuclear Overhauser effect spectroscopy (NOESY) experiments. The 13C-derived and 1H-derived constrains allow reliable docking of these ligands in the active-site cavity. The resulting geometries are similar to one another and consistent with five-coordinated structures around the metal ion, as previously proposed from electronic spectroscopy [Bertini, I., Canti, G., Luchinat, C. & Scozzafava, A. (1978) J. Am. Chem. Soc. 100, 4873-4877]. The results are discussed in light of the current debate on anion binding to metal ions in carbonic anhydrase [Lindahl, M., Svensson, A. & Liljas, A. (1992) Proteins, in the press]; Bertini, I., Luchinat, C., Pierattelli, R. & Vila, A. J. (1992) Inorg. Chem., in the press; Banci, L. & Merz, K. (1992) unpublished results] and, in particular, of the proposed long Zn-O distance found in the recent X-ray results on the formate adduct [Hakanson, K., Carlsson, M., Svensson, A. & Liljas, A. (1992) J. Mol. Biol., in the press].  相似文献   

9.
Eason PD  Imperiali B 《Biochemistry》1999,38(17):5430-5437
Recent work has resulted in the development of potent inhibitors of oligosaccharyl transferase (OT), the enzyme that catalyzes the cotranslational glycosylation of asparagine [Hendrickson, T. L., Spencer, J. R., Kato, M., and Imperiali, B. (1996) J. Am. Chem. Soc. 118, 7636-7637; Kellenberger, C., Hendrickson, T. L., and Imperiali, B. (1997) Biochemistry 36, 12554-12559]. However, no specific OT inhibitors that function in the cellular environment have yet been reported. The peptide cyclo(hex-Amb-Cys)-Thr-Val-Thr-Nph-NH2 was previously shown to exhibit nanomolar inhibition (Ki = 37 nM) through slow tight binding kinetics [Hendrickson, T. L., Spencer, J. R., Kato, M., and Imperiali, B. (1996) J. Am. Chem. Soc. 118, 7636-7637]. Included herein is the redesign of this prototype inhibitor for achieving both passive and active translocation into model membrane systems representing the endoplasmic reticulum (ER). The strategy for passive transport involved the incorporation of a membrane permeable import function previously shown to carry various peptides across the outer as well as the interior cellular membranes [Rojas, M., Donahue, J. P., Tan, Z., and Lin, Y.-Z. (1998) Nat. Biotechnol. 16, 370-375]. Assessment of function in intact ER membranes revealed that the inhibitor targeted toward passive diffusion demonstrated concentration-dependent inhibition of two different glycosylation substrates. Thus, this modified inhibitor achieved potent inhibition of glycosylation after being successfully transported through the ER membrane. In the active translocation approach, the lead OT inhibitor and a corresponding substrate were redesigned to include features recognized by the transporter associated with antigen processing (TAP). This protein translocates peptides into the lumen of the ER [Heemels, M.-T., Schumacher, T. N. M., Wonigeit, K., and Ploegh, H. L. (1993) Science 262, 2059-2063]. However, although acceptance of the cyclized substrate by the TAP receptor was demonstrated via efficient transport and glycosylation, the modified inhibitor was not translocated by TAP machinery, and therefore, active translocation was achieved for the modified substrate only. Both of these ER transport methods afforded redesigned OT inhibitors that retained their inhibitor properties in vitro, regardless of the extensions to the carboxy-terminus of the root inhibitor. The above family of redesigned inhibitors provides a template for generating a transcellular pathway and represents the first step toward OT inhibition in intact cells.  相似文献   

10.
B Imperiali  K L Shannon 《Biochemistry》1991,30(18):4374-4380
A series of tripeptides that satisfy the -Asn-Xaa-Thr/Ser- primary sequence requirement [Marshall, R. D. (1972) Annu. Rev. Biochem. 41, 673-702] for N-glycosylation have been synthesized and examined as potential acceptors in an oligosaccharyltransferase assay. Of these, six (Ac-Asn-Ala-Thr-NH2, Ac-Asn-Leu-Thr-NH2, Ac-Asn-Asp-Thr-NH2, Ac-Asn-D-Ala-Thr-NH2, Ac-Asn-Pro-Thr-NH2, and Ac-Asn-AIB-Thr-NH2) were examined for solution conformational properties in dimethyl sulfoxide with use of amide proton temperature coefficients, 3JHN alpha analysis [Pardi, A., et al. (1984) J. Mol. Biol. 180, 741-751], and 2-D ROESY experiments [Bothner-By, A. A., et al. (1984) J. Am. Chem. Soc. 106, 811-813]. The analysis reveals that the peptides that serve as acceptors in the transferase assay demonstrate similar conformational properties in solution. These are highlighted by a secondary structural motif that involves the interaction between the asparagine side-chain carboxamide and the backbone amide of the threonine. The peptides that show very poor acceptor, or even nonacceptor, properties in the oligosaccharyltransferase assay demonstrate different conformational features in solution. These observations may explain the distinct biological activity observed for these peptides.  相似文献   

11.
J J Led  E Neesgaard 《Biochemistry》1987,26(1):183-192
A detailed analysis of the stability and activity of Mn(II) human carbonic anhydrase I and the kinetics and mechanism of its catalysis of the HCO3-/CO2 exchange have been performed at pH 8.5. The analysis was based on the paramagnetic relaxation rates R1p and R2p of the 13C atom of HCO3- in the Mn2+/apoenzyme/HCO3-/CO2 system and the HCO3(-)----CO2 interconversion rate obtained by the magnetization-transfer technique. The R1p and R2p rates were measured as functions of the temperature, magnetic field strength, and substrate and apoenzyme concentrations and were interpreted on the basis of the Solomon-Bloembergen-Morgan theories and general equations for the ligand exchange [Led, J. J., & Grant, D. M. (1977) J. Am. Chem. Soc. 99, 5845-5858]. From the analysis of the data, a formation constant for the Mn(II) enzyme of log KMAM = 5.8 +/- 0.4 was obtained while the activity of the Mn(II) enzyme, measured as the HCO3-/CO2 interconversion rate at [HCO3-] = 0.100 M and pH 8.5, was found to be about 4% of that of the native Zn(II) enzyme. However, an effective dissociation constant KeffHCO3- less than or approximately 12 mM and a maximal exchange rate constant kcatexch approximately equal to 400 s-1, also derived by the analysis, result in an apparent second-order rate constant kcatexch/KeffHCO3- only a factor of 4 smaller than the corresponding rate constant for the native Zn(II) isoenzyme I. Most conspicuously, the resulting distance of only 2.71 +/- 0.03 A between the Mn2+ ion of the enzyme and the 13C atom of HCO3- in the enzyme-bicarbonate complex indicates that the bicarbonate is bound to the metal ion by two of its oxygen atoms in the central catalytic step, thereby supporting the modified Zn(II)-OH mechanism [Lindskog, S., Engberg, P., Forsman, C., Ibrahim, S. A., Jonsson, B.-H., Simonsson, I., & Tibell, L. (1984) Ann. N.Y. Acad. Sci. 429, 61-75 (and references cited therein)]. In contrast, this binding mode differs from the structure of the complexes suggested in the rapid-equilibrium kinetic model [Pocker, Y., & Deits, T. L. (1983) J. Am. Chem. Soc. 105, 980-986; Pocker, Y., & Deits, T. L. (1984) Ann. N.Y. Acad. Sci. 429, 76-83].  相似文献   

12.
Molecular dynamics simulations were performed on models of the dodecamer DNA double-stranded segment, [d(CGCGAATTCGCG)](2), in which each of the adenine residues, individually or jointly, was replaced by the water-mimicking analog 2'-deoxy-7-(hydroxy-methyl)-7-deazaadenosine (hm(7)c(7)dA) [Rockhill, J.K., Wilson,S.R. and Gumport,R.I. (1996) J. Am. Chem. Soc.,118, 10065-10068]. The simulations, when compared with those of the dodecamer itself, show that incorporation of the analog affects neither the overall DNA structure nor its hydrogen-bonding and stacking interactions when it replaces a single individual base. Furthermore, the water molecules near the bases in the singly-substituted oligonucleotides are similarly unaffected. Double substitutions lead to differences in all the aforementioned parameters with respect to the reference sequence. The results suggest that the analog provides a good mimic of specific 'ordered' water molecules observed in contact with DNA itself and at the interface between protein and DNA in specific complexes.  相似文献   

13.
M D Davies  S G Sligar 《Biochemistry》1992,31(46):11383-11389
Camphor is hydroxylated in Pseudomonas putida by a three-component system comprised of an oxidase, cytochrome P-450cam, and a two-protein electron-transfer chain, putidaredoxin and putidaredoxin reductase [Tyson et al. (1972) J. Biol. Chem. 274, 5777-5784]. The enzymatic removal of putidaredoxin's C-terminal tryptophan is known to cause a much reduced rate of enzymatic activity in the reconstituted camphor hydroxylase system [Sligar et al. (1974) Proc. Natl. Acad. Sci. U.S.A. 71, 3906-3910]. To further study the role of tryptophan in the association and/or electron-transfer reactions of putidaredoxin, the gene coding for the iron-sulfur protein was altered so that the tryptophan codon was either deleted or replaced by Phe, Tyr, Asp, Leu, Val, or Lys. Although the initial evaluation of these variant proteins [Davies et al. (1990) J. Am. Chem. Soc. 112, 7396-7398] showed much reduced velocities of electron transfer between P-450cam and the nonaromatic C-terminal proteins, the relative contributions of the binding specificity and intracomplex electron-transfer rates were not addressed. We report here a complete kinetic characterization of these proteins where the dependence of the rate constant on the putidaredoxin concentration was used to determine the intracomplex electron-transfer rate constants and the association energies for all the putidaredoxins in both oxidation states. The sum of forward and reverse intracomplex electron-transfer rate constants varies from 4.90 s-1 for the Lys C-terminal variant to 172 s-1 for the native protein.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
The conformation of Forssman glycolipid, GalNAc alpha 1-3GalNAc beta 1-3Gal alpha 1-4Gal beta 1-4Glc beta 1-1ceramide, was analysed with the aid of the rotating frame NOE and Hartmann-Hahn spectroscopy. NOE contacts between C-, O-, and N-linked protons were used for distance mapping. The glycosidic bonds that are common to globotriaosylceramide and globoside showed a similar flexibility as found for these compounds [Poppe et al., (1990) Eur. J. Biochem. 189, 313-325; J. Am. Chem. Soc. 112, 7762-7771]. In contrast, the conformational mobility of the terminal GalNAc alpha 1-3GalNAc beta linkage appears to be restrained. A new approach, based on 2D exchange spectroscopy, was proposed for revealing of spatial proximities between exchangeable protons in Me2SO solution.  相似文献   

15.
16.
17.
Glycogen is a storage form of glucose utilized as an energy reserve by many organisms. Glycogen synthase, which is essential for synthesizing this glucose polymer, is regulated by both covalent phosphorylation and the concentration of glucose-6-P. With the yeast glycogen synthase Gsy2p, we recently identified two mutants, R579A/R580A/R582A [corrected] and R586A/R588A/R591A, in which multiple arginine residues were mutated to alanine that were completely insensitive to activation by glucose-6-P in vitro (Pederson, B. A., Cheng, C., Wilson, W. A., and Roach, P. J. (2000) J. Biol. Chem. 275, 27753-27761). We report here the expression of these mutants in Saccharomyces cerevisiae and, as expected from our findings in vitro, they were not activated by glucose-6-P. The R579A/R580A/R582A [corrected] mutant, which is also resistant to inhibition by phosphorylation, caused hyperaccumulation of glycogen. In contrast, the mutant R586A/R588A/R591A, which retains the ability to be inactivated by phosphorylation, resulted in lower glycogen accumulation when compared with wild-type cells. When intracellular glucose-6-P levels were increased by mutating the PFK2 gene, glycogen storage due to the wild-type enzyme was increased, whereas that associated with R579A/R580A/R582A [corrected] was not greatly changed. This is the first direct demonstration that activation of glycogen synthase by glucose-6-P in vivo is necessary for normal glycogen accumulation.  相似文献   

18.
Near-UV circular dichroic (CD) spectra of three colchicine analogues that differ at the C-10 position have been obtained in the presence and absence of tubulin. All three colchicine analogues show dramatic alterations in the low-energy near-UV CD band upon tubulin binding that cannot be mimicked by solvent, but in no event does the rotational strength of the CD band decrease to nearly zero as in the case of colchicine [Detrich, H. W., III, Williams, R. C., Jr., Macdonald, T. L., & Puett, D. (1981) Biochemistry 20, 5999-6005]. The effect of self-association of colchicine and one of the C-10 analogues, thiocolchicine, on the near-UV CD band was also investigated. A qualitative similarity was seen between the near-UV CD spectra of colchicine and thiocolchicine dimers and the spectra of these molecules bound to tubulin. These observations support the previous suggestion that ligands bound to the colchicine site on tubulin may be interacting with an aromatic amino acid in the colchicine binding site [Hastie, S. B., & Rava, R. P. (1989) J. Am. Chem. Soc. 110, 6993-7001].  相似文献   

19.
Katz et al. [Katz, J., Golden, S. & Wals, P.A. (1976) Proc. Natl Acad. Sci. USA 73, 3433-3437] were the first to report that in hepatocytes isolated from fasted rats and incubated with either dihydroxyacetone, glucose or other sugars, glycogen synthesis was greatly accelerated by addition of amino acids. We have looked for possible mediators responsible for this effect and have tested the effect of alanine, proline, asparagine, glutamine or a combination of ammonia with either pyruvate or lactate in activating glycogen synthesis from dihydroxyacetone. The following observations were made. 1. Stimulation of glycogen synthesis by alanine, proline or asparagine does not require production of glutamine since the effect also occurs in periportal hepatocytes which lack glutamine synthetase. 2. Under various conditions, stimulation of glycogen synthesis by added amino acids directly correlated with increases in the intracellular content of amino acids, expressed in osmotic equivalents. 3. 3-Mercaptopicolinic acid, the inhibitor of phosphoenolpyruvate carboxykinase, further enhances stimulation of glycogen synthesis by amino acids because it increases the intracellular accumulation of aspartate and glutamate. 4. The previously reported enhancement by leucine of the stimulation of glycogen synthesis by glutamine [Chen. K. S. & Lardy, H. A. (1985) J. Biol. Chem. 260, 14683-14688] can be ascribed to inhibition of urea synthesis by leucine which results in accumulation of glutamate and of ammonia, the essential activator of glutaminase. It is concluded that activation of glycogen synthesis by added amino acids is due to an increase in intracellular osmolarity following their uptake and the accumulation of intracellular catabolites. This results in an increase in hepatic volume which stimulates glycogen synthesis [Baquet, A., Hue, L., Meijer, A. J., van Woerkom, G. M. & Plomp, P. J. A. M. (1990) J. Biol. Chem. 265, 955-959].  相似文献   

20.
P Stover  V Schirch 《Biochemistry》1992,31(7):2148-2155
Solutions of 5,10-methenyltetrahydropteroylglutamate can be converted to a stable hydrated adduct by heating solutions at 50 degrees C at pH values of 3-5 for several hours. The adduct is stable at pH values from 4 to 9 for hours, but at pH values below 2 it is converted to 5,10-methenyltetrahydropteroylglutamate and at pH values above 8 it is converted to 5-formyltetrahydropteroylglutamate. Arguments are presented that the adduct is (11R)-5,10-hydroxymethylenetetrahydropteroylglutamate formed from (11S)-5,10-hydroxymethylenetetrahydropteroylglutamate by formation of an ylide at C-11 which undergoes inversion of the electron pair to form the (11R) isomer. The (11R) hydrated adducted is believed to be the isomer of 5,10-methenyltetrahydropteroylglutamate referred to as anhydroleucovorin B by Cosulich et al. [Cosulich, D. C., Roth, B., Smith, J. M., Hultquist, M. E., & Parker, R. P. (1952) J. Am. Chem. Soc. 74, 3252-3263]. In addition, a new mechanism for the formation of 5-formyltetrahydropteroylglutamate from either 5,10-methenyltetrahydropteroylglutamate or 10-formyltetrahydropteroylglutamate via (11R)-5,10-hydroxymethylenetetrahydropteroylglutamate is proposed. A requirement for this pathway is that the formyl proton of 10-formyltetrahydropteroylglutamate exchange with solvent protons. The exchange of this formyl proton was observed at all pH values from 5.5 to 11.5 at a rate which exceeded by more than an order of magnitude the rate of formation of 5-formyltetrahydropteroylglutamate.  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