The Journal of Biochemistry
Online ISSN : 1756-2651
Print ISSN : 0021-924X
Volume 96, Issue 6
Displaying 1-39 of 39 articles from this issue
  • Yousuke SEYAMA, Kaztunasa OHASHI, Etsuko YASUGI, Hideaki OTSUKA
    1984 Volume 96 Issue 6 Pages 1639-1643
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The fatty acid composition of cardiolipin from the Harderian gland of guinea pig was examined by gas chromatography and gas chromatography-mass spectrometry. At least 33 kinds of fatty acids were detected. Oleic acid was the most prominent component, accounting for 18.2 mol% of the total fatty acids. About 70.2 mol% of fatty acids had methyl branches. Ethyl branches were also detected (1.3 mol%). Straight chain saturated acids comprised only 10.3 mol% On the other hand, linoleic, linolenic, and arachidonic acids were not found in this lipid. The 2- (2'-) acyl moieties contained larger amounts of oleic acid and smaller amounts of branched chain acids than the 1- (1'-) acyl moieties, but the saturated straight chain acids showed even distribution between the 1- (1'-) and 2- (2'-) positions. The fatty acids of cardiolipin from the liver of the same animal were also examined. Linoleic acid was the most abundant component (66.9 mol%), and saturated straight chain acids occupied 21.9 mol%. Branched chain acids were detected but comprised only 11.2 mo1%.
    Download PDF (1750K)
  • Hiroshi HOMMA, Nobuyoshi CHIBA, Tetsuyuki KOBAYASHI, Ichiro KUDO, Keiz ...
    1984 Volume 96 Issue 6 Pages 1645-1653
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Detergent-resistant phospholipase A (DR-phospholipase A) of E. coli is a 28K-dalton protein and is exclusively located in the outer membrane. We cloned the p1dA gene of E. coli, which is responsible for the activity of DR-phospholipase A. Strains bearing the plasmid which contained the p1dA gene yielded a large amount of the outer membrane protein with a molecular weight of about 28K daltons and overproduced 20 to 65 times as much DR-phospholipase A activity as the wild type
    Download PDF (1236K)
  • Hiroshi HOMMA, Tetsuyuki KOBAYASHI, Nobuyoshi CHIBA, Ken KARASAWA, Hir ...
    1984 Volume 96 Issue 6 Pages 1655-1664
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The nucleotide sequence of the pldA gene of Escherichia coli K-12, which codes for detergent-resistant phospholipase A (DR-phospholipase A), located in the outer membrane, was determined and the amino acid sequence of DR-phospholipase A was deduced. DR-phospholipase A contains 269 amino acids, resulting in a protein with a molecular weight of 30, 809. It does not contain any cysteine residues and seems to be synthesized first as a precursor with a typical signal peptide composed of 20 amino acids. The NH2-terminus of the mature protein is glutamine, a polar amino acid, while other outer membrane proteins so far determined have a nonpolar amino acid there. The hydropathy profile of the deduced amino acid sequence of DR-phospholipase A was studied. Most of the region was rather hydrophilic and there were no stretches of hydrophobic amino acids. Computer analysis showed that there are no homologies between DR-phospholipase A and other extracellular phospholipases whose amino acid sequences are known. The candidates for the promoter region of the pldA gene, the 5'-flanking region, have a significantly high AT content, while the AT content of the coding region is about the same as the average AT content of the E. coli chromosome. A typical ρ-inde-pendent transcription termination site is also present at the 3'-flanking region. This is the first example of the primary structure of a membrane-bound phospholipase.
    Download PDF (704K)
  • Satoshi NAKANISHI, Tokuzo NISHINO, Yoshiyasu YABUSAKI, Shingo FUJISAKI ...
    1984 Volume 96 Issue 6 Pages 1665-1672
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The method for the preparation of zymosterol was improved (13mg of zymosterol/g dry cells) by the aerobic adaptation of the cells in the presence of 1mM DL-ethionine. Lanosterol was also found to accumulate (5.0mg/g dry cells) when the cells were adapted aerobically in the presence of 10-4 M buthiobate. Pure lanosterol could be obtained by separation of the unsaponifiable lipids on TLC. Pure [14C] lanosterol with a high specific radioactivity (56 Ci/mol) could be prepared by incubation of the desiccated cells with [14C] isopentenyl pyrophosphate, cofactors such as ATP and NADPH-generating system, and buthiobate in phosphate buffer. The method using desiccated cells may also be applicable to the preparation of other radioactive sterol intermediates.
    Download PDF (608K)
  • Mitsukuni YASUI, Toshiaki ARATA, Akio INOUE
    1984 Volume 96 Issue 6 Pages 1673-1680
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The rate constant for the binding of myosin subfragment-1 (S-1) with F-actin in the absence of nucleotide, k1, and that for dissociation of the F-actin-myosin subfragment-1 complex (acto-S-1), k-l, were measured independently. The rate of S-1 binding with F-actin was measured from the time course of the change in the light scattering intensity after mixing S-1 with various concentrations of F-actin and k1 was found to be 2.55×106M-1•s-1 at 20°C. The dissociation rate of acto-S-1 was determined using F-actin labeled with pyrenyl iodoacetamide (Pyr-FA). Pyr-FA, with its fluorescence decreased by binding with S-l, was mixed with acto-S-1 complex and the rate of displacement of F-actin by Pyr-FA was measured from the decrease in the Pyr-FA fluorescence intensity. The k-l value was calculated to be 8.5×10-3 s-1 (or 0.51min-1). The value of the dissociation constant of S-1 from acto-S-1 complex, Kd, was calculated from Kd=k-l/k1 to be 3.3×10-9 M at 20°C. Kd was also measured at various temperatures (0-30°C), and the thermodynamic parameters, ΔG°, ΔH°, and ΔS°, were estimated from the temperature dependence of Kd to be -11.3 kcal/mol, +2.5 kcal/mol, and +47 cal/deg•mol, respectively. Thus, the binding of the myosin head with F-actin was shown to be endothermic and entropy-driven.
    Download PDF (599K)
  • Yasutomo SUGIMURA, Ken HOSOYA, Fuminori YOSHIZAKI, Masami SHIMOKORIYAM ...
    1984 Volume 96 Issue 6 Pages 1681-1687
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Cytochrome c-552 was extracted from a red alga, Polysiphonia urceolata, by immersion of the frozen trichomes in deionized water. Purification was carried out by acrinol treatment, ammonium sulfate fractionation, DEAE-Sephacel chromatography, hydroxyapatite column chromatography, and Bio-Gel P-10 gel filtration. The ferrocytochrome c-552 has absorption maxima at 551.5 (α), 522.5 (β), 416.3 (γ), 318 (δ), 292, and 270 nm; those of the ferricytochrome are at 525, 408.5 (γ), and 358 nm. The pyridine ferrohemochrome showed absorption maxima at 550 (α), 520 (β), and 414 nm (γ). The α-band of the ferrocytochrome is symmetric without any shoulder at room temperature, and does not split even at liquid nitrogen temperature. The ferricytochrome showed a weak absorption shoulder at 695 nm, suggesting a methionine sulfur to be the sixth ligand of heme c iron. The cytochrome is oxidized by ferricyanide and reduced by ferrocyanide, cysteine, ascorbate, and hydrosulfite. Autoxidation was not observed. The midpoint potential (Em) of the cytochrome was determined by equilibration with the ferro- and ferricyanide system to be 0.333 volt at pH 7.0 and 25°C. The isoelectric points of the ferro- and ferricytochromes were determined to be at pH 3.85 and 4.02, respectively, by the isoelectric focusing method. The molecular weight was estimated to be about 12, 000 from the results of gel filtration and SDS polyacrylamide gel electrophoresis.
    Download PDF (1249K)
  • Fumitaka OYAMA, Shigeki MIZUNO, Kensuke SHIMURA
    1984 Volume 96 Issue 6 Pages 1689-1694
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Silk fibroin is composed of heavy (H-) and light (L-) chains linked with disulfide bond(s). This paper describes immunological properties of fibroin H- and L-chains. Antiserum against H-chain, L-chain or whole fibroin could be obtained, indicating that each of these polypeptides is antigenic in rabbits. The antiserum against whole fibroin reacted with H- and L-chains. The antiserum against H-chain did not react with L-chain. Similarly, the anti-L-chain serum did not cross-react with H-chain. These results demonstrated that H- and L-chains have different antigenic determinant groups. Sericin-free fibroin samples prepared by boiling the cocoon protein in acid or 1% Na-oleate reacted with both anti-L-chain and anti-H-chain sera, indicating that the subunit structure of fibroin is stable under the drastic conditions of desericinization.
    Download PDF (1210K)
  • Noboru OTOTANI, Chie KODAMA, Masaki KIKUCHI, Zensaku YOSIZAWA
    1984 Volume 96 Issue 6 Pages 1695-1703
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Whale heparin was partially digested with a purified heparinase and the oligosaccharide fractions with 8-20 monosaccharide units were isolated from the digest by gel filtration on Sephadex G-50, followed by affinity chromatography on a column of antithrombin III immobilized on Sepharose 4B. A marked difference in the inhibitory activity for thrombin in the presence of antithrombin III was observed between the high-affinity fractions for antithrombin III of octasaccharide_??_hex-adecasaccharide and those of octadecasaccharide_??_eicosasaccharide. The disaccharide compositions of these hexadeca-, octadeca-, and eicosasaccharides were analyzed by high-performance liquid chromatography after digestion with a mixture of purified heparitinases 1 and 2 and heparinase. The analytical data indicated that the proportions of trisulfated disaccharide (IdUA (2S) αl→4GlcNS (6S)) and disulfated disaccharide (UAl→4GIcNS (6S)) increased with the manifestation of high thrombin-inhibitory activity, while that of monosulfated disaccharide (UAl→ 4GIcNS) decreased. The present observations, together with those so far reported, suggest that the presence of the former structural elements, specifically IdUA (2S)αl →4GIcNS (6S), as well as the antithrombin III-binding pentasaccharide at the proper positions in the molecules of whale heparin oligosaccharides is essential for the manifestation of high inhibitory activity for thrombin in the presence of antithrombin III. The structural bases for the manifestation of the anticoagulant activity of whale and porcine heparins and their oligosaccharides are also discussed.
    Download PDF (727K)
  • Masakazu TAKAHASHI, Kenji SUGIYAMA, Seigo SHUMIYA, Sumi NAGASE
    1984 Volume 96 Issue 6 Pages 1705-1712
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    An albumin-deficient and jaundiced strain of rats (AJR) was established by crossing Nagase analbuminemic rats (NAR) with jaundiced Gunn rats. AJR have no serum albumin and die with kernicterus within 3 weeks after birth. Their serum bilirubin level was 25 per cent (20 μg/ml) of that of Gunn rats at 1-2 weeks of age, while their brain bilirubin content was 1.2-2.7 times (4-5 μg/g brain) that of Gunn rats.
    Binding of bilirubin to NAR plasma proteins was examined. The bilirubin binding protein in NAR plasma was found to be lipoprotein, showing an association constant of 6.7×106 M-1. Bilirubin-transport into the brain of postnatal NAR was investigated by intravenous infusion of bilirubin with taurocholate. The brain bilirubin level of NAR infused with free bilirubin was 1.6 times that of normal rats. In NAR, the brain level on infusion of lipoprotein-bound bilirubin was similar to that on infusion of free bilirubin, but albumin-bound bilirubin scarcely entered the brain. These findings suggest that lipoprotein-bound bilirubin can diffuse across the blood-brain barrier into the brain.
    Download PDF (636K)
  • Hidenori HAYASHI, Kenji FUKUI, Fumie YAMASAKI
    1984 Volume 96 Issue 6 Pages 1713-1719
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The association of liver peroxisomal fatty acyl-CoA β-oxidizing system (FAOS) with the synthesis of bile acids was investigated. When rats were given clofibrate, a peroxisome proliferator and stimulator of peroxisomal FAOS, the biosynthesis of bile acids was significantly increased. Di (2-ethylhexyl) phthalate, another peroxisome proliferator, also increased the biosynthesis of bile acids. On the other hand, administration of orotate, an inhibitor of mitochondrial FAOS activity, did not affect the biosynthesis.
    It is known that fatty acyl-CoA oxidase [EC 1. 3. 99. 3] in peroxisomal FAOS conjugates with catalase [EC 1. 11. 1. 6]. When the catalase activity of liver peroxisomes was irreversibly inhibited by administration of 3-amino-1, 2, 4-triazole (aminotriazole), the biosynthesis of bile acids was suppressed to about one-third, and the serum cholesterol level was increased. However, the bile acid components of the bile obtained from aminotriazole-treated rats were not essentially different from those of control rats, and no accumulation of intermediates of bile acid synthesis was found in this experiment. Peroxisomal FAOS activity of the liver from aminotriazole-treated rats was considerably lower than that of control liver.
    The above results indicate that liver peroxisomes play a role in the biosynthesis of bile acids in vivo.
    Download PDF (534K)
  • Hiroo ITOH, Hiroyoshi HIDAKA
    1984 Volume 96 Issue 6 Pages 1721-1726
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Trypsin-treated Ca2+/calmodulin-dependent phosphodiesterase (Ca2+-PDE), which had lost its sensitivity to Ca2+-calmodulin, was inhibited by various calmodulin antagonists, trifluoperazine, chlorpromazine, N- (6-aminohexyl) -5-chloro-1-naphth-alenesulfonamide (W-7) and aminoalkyl chain analogues of W-7 (A-3, A-4, A-5, I-240, A-6, A-7). These inhibitory effects were less than those on calmodulin-activated Ca2+-PDE. The ability of these compounds to inhibit trypsin-treated Ca2+-PDE correlated well with the inhibitory effect on calmodulin-activated Ca2+-PDE. W-7 inhibited trypsin-treated Ca2+-PDE in a competitive fashion with respect to cyclic GMP and the K1, value was 300 μM. The inhibition of trypsin-treated Ca2+-PDE by W-7 (300 μM) or A-7 (100 μM) was overcome by the addition of excess calmodulin. Trypsin-treated Ca2-PDE can bind to W-7-coupled cyanogen bromide-activated Sepharose 4B in the presence of 1mM EGTA. These results suggest that Ca2+-PDE possesses a binding site for calmodulin antagonists and that the binding site for these antagonists on this enzyme may be structurally similar to the binding site on calmodulin itself.
    Download PDF (442K)
  • Masayoshi IMAGAWA, Seiichi HASHIDA, Eiji ISHIKAWA, J. William FREYTAG
    1984 Volume 96 Issue 6 Pages 1727-1735
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    A method is described for the preparation of a monomeric Fab'-β-D-galactosidase conjugate, which is required for the development of a sensitive immunoenzymometric assay. Anti-human IgG F(ab')2 was labeled with 2, 4-dinitrophenyl groups, split into Fab' by reduction and reacted with excess maleimide groups which had been introduced into β-D-galactosidase through thiol groups using N, N'-o-phenyl-enedimaleimide. The monomeric 2, 4-dinitrophenyl Fab'-β-D-galactosidase conjugate was subsequently separated from unconjugated β-D-galactosidase by affinity chromatography on a column of (anti-2, 4-dinitrophenyl) IgG-Sepharose 4B. In the monomeric conjugate preparation, 98% of β-D-galactosidase activity was asso-ciated with Fab' and 90% was associated with specific (anti-human IgG) Fab'. This conjugate allowed the measurement of 0.1 fmol of human IgG by an immunoenzymometric assay technique.
    Download PDF (750K)
  • Shin YAZAWA, Ken FURUKAWA, Naohisa KOCHIBE
    1984 Volume 96 Issue 6 Pages 1737-1742
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Fucosyl glycoproteins were fractionated from a sialoglycoprotein preparation of human erythrocyte membrane by using Aleuria aurantia lectin (AAL) coupled to Sepharose 4B. The affinity eluates were characterized as having high fucose content and significant H activity as measured in terms of N-acetylgalactosamine (GaINAc) incorporation with A1-enzyme and hemagglutination inhibition assay with anti-H sera, and the unadsorbed fractions contained low levels of fucose and were devoid of apparent H activity. Neuraminidase treatment of the material improved the recovery of the affinity eluate. Thus, 66% of the applied asialoglycoprotein was recovered in the eluate, though only 10% of the untreated material was bound and eluted. Moreover, a fucose-rich and H-active fraction was obtained through the affinity chromatography of the previously unbound fraction after neuraminidase treatment. In sodium dodecyl sulfate-gel electrophoresis, the main component of both the unadsorbed and eluted fraction was revealed to be PAS-1 glycoprotein.
    These results indicate that AAL-Sepharose was effective for isolating fucose-containing compounds from the membrane glycoprotein especially after neuraminidase treatment. The reasons for the appearance of H activity in the affinity eluates are discussed.
    Download PDF (1003K)
  • Yasuhiro SAGARA, Akio ITO, Tsuneo OMURA
    1984 Volume 96 Issue 6 Pages 1743-1752
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    In vitro translation of bovine adrenal cortex RNA in rabbit reticulocyte lysate cellfree system produced the precursor form of adrenodoxin having a molecular weight of approximately 22, 000 daltons, which was about 10, 000 daltons larger than mature adrenodoxin. The precursor of adrenodoxin was efficiently imported into adrenal cortex mitochondria in vitro. The precursor was also imported into rat liver mitochondria, suggesting the lack of tissue specificity and species specificity of the import process.
    The enzyme which processed the precursor of adrenodoxin to the mature form was in the matrix fraction from bovine adrenal cortex mitochondria, and the processing protease was partially purified from the matrix fraction. The apparent molecular weight of the processing protease was about 60, 000 daltons as determined by Sephadex G-150 gel filtration, and its activity was optimal at pH 8.5. The processing protease was not inhibited by various bacterial protease inhibitors examined. Metal chelators (EGTA, GTP, 8-hydroxyquinoline, and Zincon) inhibited the processing, and EDTA and o-phenanthroline were more strongly inhibitory than other chelators. The processing protease was completely inactivated by incubation with 10 μM EDTA, and its activity was restored by addition of excess amounts of Mn2+, Fe2+, or Co2+. These results indicate that the maturation of the precursor of adrenodoxin is catalyzed by a soluble metalloprotease in the matrix.
    Download PDF (2989K)
  • Shoji TAJIMA, Shinji YOKOYAMA, Akira YAMAMOTO
    1984 Volume 96 Issue 6 Pages 1753-1767
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Triolein particles stabilized by a phosphatidylcholine monolayer were used to study the lipoprotein lipase (LpL) reaction. They were prepared in two different sizes and with triolein and phosphatidylcholine in the molar ratios of 0.9-1.2:1 (small particles) and 8-17:1 (large particles). The rate of hydrolysis by LpL of phosphatidylcholine on the surface of both lipid particles was only 1/20 as much as that of triolein, even if it was activated to the maximum by apolipoprotein C-II (apoC-II). Thus, the phospholipase activity of LpL was low enough to measure the initial rate of hydrolysis of triolein without causing a gross change of the surface of the lipid particle. When the hydrolysis of triolein by LpL was monitored, fatty acid was released at a constant rate until all of the triolein molecules were hydrolyzed. The enzyme required 220±17 and 66±9 nM apoC-II for its half-maximal activity (Km (apoC-II)) with small and large particles as a substrate (1.15mM triolein for small and 2.13mM triolein for large particles), respectively, using various concentrations of LpL.
    The Km (apoC-II) values for these two substrates became similar when LpL activity was analyzed with respect to the density of apoC-II on the phosphatidyl-choline monolayer at the surface of the particles (bound apoC-II/phosphatidyl-choline). The concentration of substrate particles did not affect the Km(apoC-II) values. The presence of an adequate amount of apoC-II increased the maximal activity of LpL (Vmax (triolein)) from 0.48±0.21 to 6.81 ±0.45 and from 0.32±0.04 to 7.13 ±0.64 mmol/h/mg with a slight decrease in the apparent Michaelis constant (Km (triolein)) for small (from 90 to 54 μM triolein) and large (from 1.00 to 0.65mM triolein) particles, respectively.
    Although the apparent Km for triolein in large particles was about ten times greater than that in small particles, the values became similar when they were corrected for the concentration of phosphatidylcholine (50-100 μM phosphatidylcholine), which corresponded to the surface area of the substrate particles. It was suggested that bound apoC-II molecules were transferred relatively slowly to other lipid particles while LpL molecules moved rapidly among the lipid particles.
    These results indicate that i) LpL activity is modulated by the concentration of apoC-II which is bound to the surface of the substrate particles, ii) LpL attacks triolein in or near the phosphatidylcholine layer at the surface of lipid particles, and iii) apoC-II activates LpL by increasing the catalytic rate with little change in the affinity of triolein for the enzyme.
    Download PDF (2308K)
  • Kazuo NAGASAKI, Michiki KASAI
    1984 Volume 96 Issue 6 Pages 1769-1775
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Mg2+ and Ca2+ efflux rates from sarcoplasmic reticulum vesicles were measured by using a fluorescence chelate probe, chlortetracycline. Mg2+ efflux rate was almost the same as Ca2+ efflux rate under the same medium conditions. Both effluxes were activated by micromolar concentrations of extravesicular Ca2+ and blocked by millimolar concentrations of extravesicular Ca2+. Both effluxes were also activated by caffeine and inhibited by procaine. This result shows that Mg2+ can permeate through the Ca2+-induced Ca2+ release channel at the same rate as Ca2+: namely this channel has no selectivity between Ca2+ and Mg2+. Gating specificity of the channel was studied by measuring Mg2+ efflux. It was shown that this channel could be opened by Sr2+ as well as Ca2+, but not by Ba2+ or Mg2+. The apparent dissociation constant of the activation site for Sr2+ was slightly higher than that for Ca2+ but the apparent dissociation constant of the inhibition site for Sr2+ was almost the same as that for Ca2+.
    Download PDF (426K)
  • Yasuhiro HASHIMOTO, Tamio YAMAKAWA, Yuichi TANABE
    1984 Volume 96 Issue 6 Pages 1777-1782
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Genetic polymorphism was observed in the sialic acid species constituting the terminal sugar residues of hematosides from dog erythrocytes. One was N-acetyl-neuraminic acid and the other phenotype was N-glycolylneuraminic acid, regulated by an autosomal dominant allele (Yasue, S., Handa, S., Miyagawa, S., Inoue, J., Hasegawa, A., & Yamakawa, T. (1978) J. Biochem. 83, 1101-1107). In this study we analyzed blood samples from 1, 591 dogs of 36 breeds and demonstrated that the expression of N-glycolylneuraminic acid was limited to several breeds of oriental dogs in spite of its dominant nature. Moreover, the incidence of N-glycolylneu-raminic acid was higher in native breeds of northern China, Korea and the southern part of Japan than in other oriental breeds. On the other hand, the Hokkaido-dog is unique in not expressing N-glycolylneuraminic acid. These results suggest that the native breeds in the southern part of Japan came from northern China via the Korean peninsula in contrast with indigenous breeds of the northern part of Japan.
    Download PDF (326K)
  • Shigeyuki FUKUI, Yoshito NUMATA, Ikuo YAMASHINA
    1984 Volume 96 Issue 6 Pages 1783-1788
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Protein sulfation in baby hamster kidney cells (BHK) and their polyoma virus transformants (PY-BHK) was studied comparatively. On in vivo labeling, [35S] -sulfate was incorporated into the 50K protein and proteins in the 100-180K range, represented by the 155K protein. The incorporation into both the 50K and 155K protein was elevated 2-3 fold in PY-BHK cells compared to in BHK cells. Tyrosine-O-sulfate was the only identifiable sulfated amino acid in both proteins. On in vitro labeling with [35S] 3'-phosphoadenosine 5'-phosphosulfate (PAPS), at least 6 radioactive protein bands were discernible on gel electrophoresis. Of these, sulfation of the 57K and 60K proteins was elevated in PY-BHK cells compared to in BHK cells, whereas sulfation of the 39K protein was depressed in PY-BHK cells. Tyrosine-O-sulfate was the only identifiable sulfated amino acid in these proteins.
    Download PDF (1759K)
  • Hiroshi TSUCHIHASHI, Toshiro YADOMAE, Toshio MIYAZAKI
    1984 Volume 96 Issue 6 Pages 1789-1797
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The occurrence of three forms, I, II, and III, of exo-β-D-glucuronidase of the fungus Kobayasia nipponica was demonstrated. These three forms were purified 1, 905-fold, 857-fold, and 357-fold, respectively. Forms I, II, and III of exo-β-D-glucuronidase behaved differently on heparin-Sepharose chromatography, and differed in optimum pH (3.5, 3.2, and 2.6, respectively), pH-stability, Km (0.22, 0.16, and 0.13mM, respectively), and Vmax (V) values. Their molecular weights, as estimated by gel filtration through Sephacryl S-200, were all 70, 000; polyacrylamide gel electrophoresis in the presence of sodium dodecyl sulfate gave a value of 72, 000. These three forms were very active towards 1-4 linked β-D-glucuronans, D-GlcUAi D-GlcUA and D-GlcUA L-IdUA D-GlcUA, but weak or inactive towards protuberic acid and several glycosaminoglycans.
    Download PDF (959K)
  • Hiroshi TSUCHIHASHI, Toshiro YADOMAE, Toshio MIYAZAKI
    1984 Volume 96 Issue 6 Pages 1799-1805
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    An endo-β-D-glucuronidase was isolated and characterized from Kobayasia nipponica. The enzyme was purified by ammonium sulfate fractionation, CM-Sephadex chromatography, gel filtration with Sephacryl S-200, and heparin-Sepharose chromatography. The enzyme shows the following properties: optimum pH 5.0, thermal stability below 37°C, pH stability 5-6, optimum temperature 45-55°C, and Km 0.12% for L-idurono-D-glucuronan (protuberic acid (PA), L-IdUA:D-GlcUA=1:2) from Kobayasia nipponica, 0.19% for PF (L-IdUA:D-GlcUA=1:3) from Pseudocolus fusiformis, and 0.23% for (1-4)-β-D-glucuronan (mucoric acid) from Mucor mucedo as determined from Hofstee plots. The molecular weight values estimated by gel filtration through Sephacryl S-200 and Sephadex G-50 and by polyacrylamide gel electrophoresis in the presence of sodium dodecyl sulfate were 10, 500 and 10, 200, respectively. The endo-β-D-glucuronidase was inactive towards several glycosaminoglycans.
    Download PDF (451K)
  • Yoshifumi NISHIMURA, Masamichi TSUBOI
    1984 Volume 96 Issue 6 Pages 1807-1811
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The cutting modes of restriction endonucleases which recognize tetradeoxyribonu-cleotide sequences are classified into two groups. d (GGCC) and d (CGCG), for example, are cut to produce blunt ends, while d (CCGG) and d (GCGC) are cut to produce two-base-long cohesive ends. A conformational analysis by the Calladine-Dickerson method indicates that d (GGCC) and d (CGCG) should have a roll angle of successive base-pairs open towards the major groove at the central (second) base-pair step. On the other hand, d (CCGG) and d (GCGC) have such open roll angles at the first and third base-pair steps. It is suggested that, in general, the cutting mode of a tetramer-specific enzyme depends primarily upon the substrate conformation, rather than upon the enzyme. Similar correlations between the mode of action and substrate conformation are also suggested for hexamer-specific enzymes.
    Download PDF (310K)
  • Kiyoshi TAKENAKA, Teruhiro TAKABE
    1984 Volume 96 Issue 6 Pages 1813-1821
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    To study the relationship between the electron transfer rate and the net or local charge of protein, chemically modified cytochrome f, in which positively charged amino groups are replaced with negatively charged carboxyl groups, has been prepared by using an arylating reagent 4-chloro-3, 5-dinitrobenzoic acid. Four distinct species of chemically modified cytochrome f, having 1 to 4 mol of modified amino residues per mol of cytochrome f, were separated by preparative polyacrylamide gel electrophoresis. The rate of electron transfer from the reduced singly substituted cytochrome f to the oxidized spinach plastocyanin was only about 50% of that of the native unmodified cytochrome f. The reaction rate further decreased about 50% upon the modification of each amino residue. The biphasic oxidation of cytochrome f by plastocyanin was observed when more than 2 mol of amino residues were modified. The rate of the second phase also decreased with an increasing number of modified amino residues. On the other hand, the oxidation of chemically modified cytochrome f by potassium ferricyanide was clearly monotonic. The rate decreased about 30% upon the modification of each amino residue. The midpoint potentials of chemically modified cytochrome f were almost the same as that of the native protein. These results clearly indicate the importance of local positive charges on cytochrome f, since the overall net charge of cytochrome f is negative at neutral pH. The theory of electrostatic corrected outer-sphere electron transfer of Marcus explained the effect of charge on cytochrome f for the reaction with the small molecule of ferricyanide well, but not the reaction with the protein of plastocyanin.
    Download PDF (1984K)
  • Koh IBA, Ken-ichiro TAKAMIYA, Hiroyuki ARATA, Yoshihiro TOH, Mitsuo NI ...
    1984 Volume 96 Issue 6 Pages 1823-1830
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The photochemical reaction centers from Rhodopseudomonas sphaeroides were reconstituted with soybean phospholipids into liposomes by the cholate-dialysis method. The transmembrane orientation of the reaction centers in the proteoliposomes and the morphology of the vesicles were investigated. The orientation was determined by the reduction of externally added cytochrome c after its photooxidation by a flash. The structure of the vesicles was examined by electron microscope.
    Discontinuous sucrose density gradient centrifugation yielded several proteoliposome fractions with different vesicular sizes and reaction-center orientations. The proportion of the reaction centers that exposed their cytochrome c reacting sites to the outside of the vesicles increased from 45 to 85% with an increase of the vesicular size. The proportion also depended on the ionic composition of the dialysis buffer. The optimal ionic environment during the dialysis (100mM NaCl or 2.5mM MgSO4) gave a liposome yield of 25-30% with a highly asymmetric orientation (>60%). Entrapping of cytochrome c molecules into the phospholipid vesicles had little effect on the orientation of the reaction centers.
    Download PDF (1374K)
  • III. Basic Structure of the Photosynthetic Unit and Its Relation to Other Bacteriochlorophyll Forms
    Shinya NAKAMOTO, Mikio KATAOKA, Tatzuo UEKI
    1984 Volume 96 Issue 6 Pages 1831-1839
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    We have performed X-ray diffraction studies on photosynthetic units of Rhodospirillum rubrum and solubilized *B800+B890 complex from chromatophores of Chromatium vinosum, to investigate the homology of their molecular structures. The native chromatophores of Chromatium vinosum, which contain other bacterio-chlorophyll forms, were examined by an X-ray diffraction technique, in order to assess the interactions between the complexes as well as the molecular structures of the bacteriochlorophyll forms.
    1. The subchromatophore particles, solubilized by Triton X-100 from cells of Chromatium vinosum, exhibit a major absorption maximum at 881 nm and a minor one at 804 nm, consisting of bacteriochlorophyll form *B800+B890. The near-IR absorption spectrum of the particle is very similar to that of chromatophores of Rhodospirillum rubrum although the major absorption maximum is shifted slightly.
    2. The X-ray diffraction pattern of the subchromatophore particles is very similar to that of chromatophores of Rhodospirillum rubrum. Thus, the subchromatophore particles are considered to be the “photoreaction unit” of Rhodospirillum rubrum. Since the bacteriochlorophyll form, *B800+B890, is common in the purple bacteria, it is strongly suggested that the photoreaction unit is the basic and common structure existing in the photosynthetic units of purple bacteria.
    3. Chromatium vinosum cells exhibit different near-IR absorption spectra, depending on the culture media and also on the intensity of the illumination during culture. The chromatophores from these cells give different equatorial X-ray diffraction patterns. These patterns are much broader than that of solubilized subchromatophore particles, though they have common features. Thus, the molecular structures in the photosynthetic units are different, depending on their constituent bacterio-chlorophyll forms.
    4. The rather broad X-ray diffraction pattern from the chromatophores and the distinct pattern from solubilized subchromatophore particles of Chromatium vinosum reflect the interaction between bacteriochlorophyll forms in the photosynthetic unit. Namely, the rather broad pattern can be understood in terms of the sum of the patterns from each form, indicating that the interaction between the forms is weak. Further, the bacteriochlorophyll forms, B800+B820 and B800+B850, must have well-defined molecular structures rather than liquid-like, disordered structures.
    Download PDF (1342K)
  • Eiki KOMINAMI, Yoshiaki BANDO, Kunio II, Kazuo HIZAWA, Nobuhiko KATUNU ...
    1984 Volume 96 Issue 6 Pages 1841-1848
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    In dystrophic hamsters, increases in the levels of cathepsin B plus L and thiol proteinase inhibitor were marked in skeletal muscle, but only slight in heart muscle. The lysosomal hydrolases did not increase in parallel in dystrophic muscle: cathepsin B plus L and β-glucuronidase increased, but cathepsin C and acid phosphatase did not. In immunohistochemical studies with antibodies against rat liver cathepsin B and thiol proteinase inhibitor, the proteinase and inhibitor were both stained in phagocytes, chiefly macrophages, but not in muscle cells, indicating that the increases in cathepsin B plus L and thiol proteinase inhibitor in dystrophic muscle are due to their presence in invading phagocytes. The levels of cathepsin B plus L, β-glucuronidase and thiol proteinase inhibitor in isolated peritoneal macrophages were 50 to 180 times higher than those in skeletal muscle, but the levels of acid phosphatase and cathepsin C were only about 10 to 30 times those in skeletal muscle. Plots of the cathepsin B plus L activities versus the level of thiol proteinase inhibitor in homogenates of tissues of various animals showed an exponential rather than a linear relation between the two activities, suggesting that the syntheses of the proteinases are higher than that of the inhibitor in phagocytes invading dystrophic muscle.
    Download PDF (1212K)
  • Kunio YAMANE, Yu-ichi HIRATA, Takashi FURUSATO, Hisato YAMAZAKI, Akira ...
    1984 Volume 96 Issue 6 Pages 1849-1858
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The regulatory gene, amyR2, and the structural gene, amyEn+, coding for N-type α-amylase from Bacillus subtilis N7 have been cloned in the B. subtilis plasmid pUB110. The complete nucleotide sequence of amyR2 and amyEn+ has been determined. Starting from an ATG initiator codon, there was an open reading frame comprising 477 amino acids (1, 431 bp), giving a molecular weight of 52, 678. The NH2-terminal portion of amyEn+ encoded a 41-amino acid-long signal sequence. The DNA nucleotide sequence was compared with the sequences of amyEm+ coding for M-type α-amylase from B. subtilis NA64 (Yamazaki et al. (1983) J. Bacteriol. 156, 327-337) and another B. subtilis α-amylase gene (Yang et al. (1983) Nucl. Acids Res. 11, 237-249). Almost all the sequences were identical in the three genes. However in the sequence of amyEn+ 32 bp of the other two α-amylase genes were deleted in the region from nucleotide 1, 406 to 1, 437. This deletion region was included in the direct repeat structure of the two genes. The reading frame downstream of the deletion region of amyEn+ shifted and a new termination codon (TGA) appeared at 26 bp downstream. Thus, the differences of the M-type and N-type α-amylases from amyEm+ and amyEn+ seemed to be caused by the occurrence of translation termination at different sites of the α-amylase gene.
    Download PDF (1259K)
  • Kazuhisa SEKIMIZU, Masami HORIKOSHI, Shunji NATORI
    1984 Volume 96 Issue 6 Pages 1859-1865
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    By use of purified RNA polymerase II, it was demonstrated that S-II, a stimulatory protein of RNA polymerase II, enhanced the frequency of initiation of transcription from discrete sites on the promoter region of the silk fibroin gene integrated in a supercoiled plasmid DNA in the presence of manganese. In the absence of S-II, RNA polymerase II preferentially initiated RNA synthesis from site+25, 25 bases downstream from the cap site. Of these initiation sites, the initiation from site+25 was not affected by S-II, suggesting that site+25 is structurally different from other initiation sites, including the cap site, and that S-II modifies the latter sites to the same structure as that of site+25. Direct interaction between S-II and DNA in the initiation complex was shown by demonstrating a conformational change of S-II on its interaction with DNA; namely, S-II in the initiation complex was as sensitive to chymotryptic digestion as S-II interacting with DNA, whereas free S-II was completely insensitive to chymotryptic digestion.
    Download PDF (3166K)
  • Koji NISHIO, Makoto KAWAKAMI
    1984 Volume 96 Issue 6 Pages 1867-1874
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Alanyl-tRNA synthetase was purified from the posterior silk glands of Bombyx mori by ammonium sulfate fractionation and chromatography on DEAE-Sephacel and hydroxyapatite columns. The yield was about 100mg of the enzyme per 1kg of the glands. The enzyme required both L-alanine and alanine tRNA for pyrophosphate formation from ATP. The PP1 formation was observed even after tRNA was fully aminoacylated. The enzyme was found to be a monomer of 115K daltons by SDS-polyacrylamide gel electrophoresis, gel filtration and suberimidate cross-linking experiments. The monomeric enzyme did not dimerize in the presence of the alanine tRNA. The enzyme and the tRNA formed a 1:1 complex. The results indicate that Bombyx mori alanyl-tRNA synthetase functions in a monomeric state.
    Download PDF (1253K)
  • Koji NISHIO, Makoto KAWAKAMI
    1984 Volume 96 Issue 6 Pages 1875-1881
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Alanyl-tRNA synthetase of 115 K daltons from Bombyx mori was cleaved into two fragments of 62 K and 47 K daltons by trypsin. The 47 K fragment was active in aminoacylation of tRNA, whereas the 62K fragment was inactive. The 47 K and 62 K fragments were found to be located at the N- and C-terminal ends, respectively, in the intact enzyme. The intact enzyme was protected from trypsin-attack by the cognate tRNA. The Km value of the 47 K fragment for tRNA was 22 μM which is about 16-fold higher than that for the intact enzyme (1.4 μM). The molecular activities of the fragment and the intact enzyme were 2.2 s-1 and 16.8 s-1, respectively. This indicates that the 62 K domain enhances affinity for tRNA and it is responsible for the full activity of tRNA aminoacylation. These results do not support the “covalently linked dimer” hypothesis, but indicate that the alanyl-tRNA synthetase is a functional monomer consisting two large domains.
    Download PDF (2938K)
  • Masatoshi KAKU, Kosuke ICHIHARA, Emi KUSUNOSE, Kiyokazu OGITA, Satoru ...
    1984 Volume 96 Issue 6 Pages 1883-1891
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Earlier studies (Kusunose, E., Kaku, M., Ichihara, K., Yamamoto, S., Yano, I., & Kusunose, M. (1984) J. Biochem. 95, 1733-1739) showed that a form of cytochrome P-450 isolated from microsomes of rabbit small intestinal mucosa had the highest prostaglandin A1 (PGA1) hydroxylase activity so far reported among cytochrome P-450s. The present paper describes the procedure for the purification and further characterization of this cytochrome (designated as cytochrome P-450ia). Cytochrome P-450ia, had a monomeric molecular weight of 53, 000. The CO-difference spectra of its reduced form showed a maximal absorption at 451 nm, and the absolute spectra of its oxidized form indicated that cytochrome P-450ia was present largely in the low-spin state, and partially in the high-spin state. The cytochrome efficiently catalyzed the hydroxylation of fatty acids as well as prostaglandins in a reconstituted system containing cytochrome P-450, NADPH-cytochrome P-450 reductase, phospholipid, and cytochrome b5. PGA1 was the most efficient substrate, followed by myristate, laurate, palmitate, caprate, and PGE1 or PGE2. Among phospholipids, didecanoyl- and dilauroylphosphatidyl-cholines had the most stimulatory effect for both activities. 20-Hydroxy PGA1 was identified as the hydroxylation product of PGA1 by gas chromatography-mass spectrometry and mass fragmentography; the possibility of 19-hydroxy PGA1 being the product was excluded. In contrast, both ω- and (ω-1) -hydroxy fatty acids were identified as hydroxylation products of fatty acids. Cytochrome P-450ia had no detectable activity toward aminopyrine, benzphetamine, p-nitroanisole, 7-ethoxycoumarin, benzo (a) pyrene, or hexadecane.
    Download PDF (1320K)
  • Nobuaki WATANABE, Masashi KOBAYASHI, Hiroshi MAEGAWA, Osamu ISHIBASHI, ...
    1984 Volume 96 Issue 6 Pages 1893-1902
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The in vitro effect of glucocorticoid on insulin binding and glucose transport was studied with rat adipocytes. Isolated rat adipocytes were incubated with or without 0.70 μg/ml (1.9μmol) of hydrocortisone in TCM 199 medium at 37°C, 5% CO2/95%, air (v/v), pH 7.4, for 2, 4, and 8 h, and then fat cell insulin binding and insulinstimulated 3-O-methylglucose transport were measured. Hydrocortisone did not affect insulin binding in terms of affinity or receptor number. Glucose transport in the absence of insulin was significantly decreased at the incubation time of 2 h and continued to decrease up to 8 h of incubation with hydrocortisone. Decreased insulin sensitivity of glucose transport (i.e., a right-ward shift of the dose response curve) was also demonstrated after 2 h incubation with hydrocortisone, and the ED50 of insulin was maximally increased at 4 h of incubation (0.53 ng/ml for treated vs. 0.22 ng/ml for control cells). Maximal insulin responsiveness was also significantly decreased in treated cells after 8 h incubation with hydrocortisone. When percent maximum glucose transport was expressed relative to receptor-bound insulin, the ED50 values of treated and control cells were 10.5 and 7.2 pg of bound insulin, per 2×103 cells, respectively. Thus, it was evident that glucocorticoid induced a postreceptor coupling defect in the signal transmission of insulin-receptor complex.
    Download PDF (791K)
  • Keizo TESHIMA, Yuji SAMEJIMA, Saju KAWAUCHI, Kiyoshi IKEDA, Kyozo HAYA ...
    1984 Volume 96 Issue 6 Pages 1903-1913
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The pH dependence of the chemical reaction rate of p-bromophenacyl bromide (BPB) with His 48 of cobra (Naja naja atra) venom phospholipase A2, in which the α-NH2 group had been selectively modified to an α-keto group, was studied at 25°C and ionic strength 0.1 in the absence of Ca2+ The pH-dependence curve was monophasic with a midpoint at pH 7.9, which corresponds to the pK value of His 48 of the α-NH2 modified enzyme, whereas the curve for the intact enzyme was biphasic, indicating participation of two ionizable groups with pK values of 7.3 and 8.55 (Teshima et al. (1982) J. Biochem. 91, 1778-1788). These two groups were thus identified as His 48 and the α-NH2 group, respectively.
    The pH dependence of the binding constant of Ca2+ to the α-NH2 modified enzyme was studied at 25°C and ionic strength 0.1 by measuring the tryptophyl fluorescence changes. The pH-dependence curve was very similar to that for the intact enzyme (Teshima et al. (1981) J. Biochem. 89, 13-20), and it was interpreted in terms of participation of His 48 and Asp 49 (pK 5.4). The absence of participation of the α-NH2 group in the Ca2+ binding was thus confirmed.
    Bindings of monodispersed n-dodecylphosphorylcholine (n-C12YPC) and micellar n-hexadecylphosphorylcholine (n-C16PC) to the α-NH2-modified enzyme were studied at 25°C and ionic strength 0.1 by the aromatic circular dichroism (CD) and tryptophyl fluorescence methods, respectively. The binding constant of the monodispersed substrate was very similar to that for the intact enzyme (Teshima et al. (1981) J. Biochem. 89, 1163-1174). The binding constant of the micellar substrate to the modified enzyme in the presence of Ca2+ was also very similar to that for the intact enzyme-Ca2+ complex (Teshima et al. (1983) J. Biochem. 94, 223-232), and the pH-dependence curve was interpreted in terms of participation of His 48. On the other hand, the binding constant of the micellar substrate to the modified apoenzyme was much smaller than that for the intact apoenzyme. Nevertheless, the pH-dependence curve could be interpreted in terms of participation of His 48 and Asp 49. From these findings, it was concluded that the ionization state of the α-NH2 group of cobra venom phospholipase A2 is essentially irrelevant to the bindings of Ca2+ and also of the monodispersed and micellar substrates.
    Download PDF (814K)
  • Takeshi OHNO, Midori AOYAGI, Yuji YAMANASHI, Hiraku SAITO, Shuntaro IK ...
    1984 Volume 96 Issue 6 Pages 1915-1923
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The sequence of about 4, 500 nucleotides of the internal part of tobacco mosaic virus (TMV) -tomato strain (L) RNA has been newly determined using cloned cDNAs. Together with the previously determined partial sequences at both ends, the entire sequence of the 6, 384 nucleotide genome has been completed. The 130K (1, 115 amino acids), 180K (1, 615 amino acids), 30K (263 amino acids) and coat protein (158 amino acids) cistrons are located at residues 72-3442, 72-4922, 4906-5700, and 5703-6182 on the genome, respectively. Sequence polymorphism was not observed except for heterogeneity in the length of the A cluster near the 3' end. The homology of the nucleotide sequences of TMV-L and TMV-vulgare, a common strain, is about 80% on average. Remarkable differences between them were found in a part of the N-terminal portion of the 130K/180K protein and the C-terminal portion of the 30K protein. A new method for cDNA cloning was developed by which the cDNA of the 5'-terminus of viral RNA can be cloned efficiently.
    Download PDF (643K)
  • Ikuo IMAMURA, Takehiko WATANABE, Yutaka HASE, Yoshiteru SAKAMOTO, Yuko ...
    1984 Volume 96 Issue 6 Pages 1925-1929
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Histamine metabolism in histidinemic patients was studied by measuring the urinary levels of histamine and its metabolites. The urinary excretions of histamine, Nτ-methyl histamine, imidazole acetic acid, and its conjugate (s) were higher in patients with histidinemia than in controls, and these levels of excretion were correlated with the plasma histidine level. The urinary histamine levels of patients with eczema-like dermatitis were twice that of those without dermatitis. The urinary excretion of 3-methylhistidine showed a close correlation with the urinary histidine excretion. Thus, it was concluded that histamine metabolism is higher in histidinemic patients than in normal controls.
    Download PDF (348K)
  • Ikuo IMAMURA, Takehiko WATANABE, Kazutaka MAEYAMA, Akio KUBOTA, Akira ...
    1984 Volume 96 Issue 6 Pages 1931-1937
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The urinary excretions by young healthy men of histamine and its metabolites, Nτ-methylhistamine, imidazole acetic acid, and imidazole acetic acid conjugate (s), increased 1-3 h after food intake. The increase was seen even after the intake of konnyaku (mannan) as a protein-deficient food, suggesting that physical stimulation of the gastric mucosa by food is the main cause of histamine release. This suggestion was confirmed by the following findings in patients and mice. In patients with stomach diseases, gastrectomy resulted in decreases in the excretion of histamine and its metabolites in the urine, and patients subjected to intravenous hyperalimentation excreted less histamine and its metabolites in the urine than normal subjects. In mice, a correlation of histamine excretion with food intake was demonstrated experimentally. Namely, mice fed only during the night (21:00-0:00) showed increased excretions of histamine and its metabolites at 23:00-3:00, whereas those fed in the morning (9:00-12:00) showed increased excretions of those compounds at 11:00-15:00. All these results are consistent with the idea that urinary histamine and its metabolites mainly originate from the stomach.
    Download PDF (457K)
  • Tetsuya KAMATAKI, Kaori MAEDA, Miki SHIMADA, Tsuneji NAGAI, Ryuichi KA ...
    1984 Volume 96 Issue 6 Pages 1939-1942
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The effects of neonatal castration and treatment with testosterone on sex-specific forms of cytochrome P-450, namely P-450-male and P-450-female, were studied. Neonatal castration of male rats resulted in a change in the population of forms of cytochrome P-450. Castration 1 day after birth abolished the synthesis of P-450-male and stimulated the synthesis of P-450-female. The decrease in the amount of P-450-male as well as the activities of drug metabolizing enzymes was partially reversed by administration of testosterone after castration.
    Download PDF (271K)
  • Laura RIBONI, Anna MALESCI, Sergio Maria GAINI, Sandro SONNINO, Riccar ...
    1984 Volume 96 Issue 6 Pages 1943-1946
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The ganglioside pattern from normal human temporal cortex and cerebellum has been studied in fresh specimens obtained at surgery.
    The analyses have been performed by two-dimensional thin layer chromatography with an intermediate ammonia treatment which is a methodology particularly suitable for resolving alkali labile gangliosides.
    Alkali labile gangliosides were detected in all the analyzed specimens and their content contributed to 23% and 11% of total lipid bound sialic acid, in temporal cortex and cerebellum, respectively.
    Download PDF (1215K)
  • Sumiko KIMURA, Hideaki YOSHIDOMI, Koscak MARUYAMA
    1984 Volume 96 Issue 6 Pages 1947-1950
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Proteolytic fragments of 400 kD isolated from chymotrypsin-treated connectin, a muscle elastic protein, still retained the ability to cause aggregation of myosin filaments but lost the actin-bundling action. Tryptic digests of connectin showed similar effects. However, when connectin was hydrolyzed by pepsin to peptides smaller than approximately 40 kD, no such action was seen for both myosin and actin filaments. It is suggested that the actin bundling action of connectin filaments is due to topological restrictions.
    A modified reproducible procedure for the preparation of native connectin from chicken breast muscle is described in detail.
    Download PDF (766K)
  • Kazuko ÔBA, Atsuko NAKAMURA, Yukio IWAIKAWA
    1984 Volume 96 Issue 6 Pages 1951-1954
    Published: 1984
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Furanoterpene-containing particles were isolated by centrifugation on a discontinuous Ficoll density gradient from a homogenate of the non-infected tissue adjacent to the infected region of Ceratocystis fimbriata-infected sweet potato root tissue. The particles were recovered at a relatively high ratio in the 2% Ficoll fraction, in which there was no contamination by mitochondria and only little by endoplasmic reticulum judging from the distribution of the activities of their marker enzymes and electron micrographs. Each particle was enveloped in a single membrane, 7-10 nm thick.
    Download PDF (1036K)
feedback
Top