The Journal of Biochemistry
Online ISSN : 1756-2651
Print ISSN : 0021-924X
Volume 99, Issue 6
Displaying 1-33 of 33 articles from this issue
  • Shin-ichi YOKOTA, Shunji KAYA, Takashi KAWAMURA, Yoshio ARAKI, Eiji IT ...
    1986 Volume 99 Issue 6 Pages 1551-1561
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Structural studies were carried out on the O-polysaccharide fraction obtained from the lipopolysaccharide of Pseudomonas aeruginosa IID 1008 (ATCC 27584). The O-polysaccharide comprises L-rhamnose, N-acetyl-D-quinovosamine, N-acetyl-D-galactosaminuronic acid, and N-formyl-D-galactosaminuronic acid. The characterization of oligosaccharide fragments resulting from acid hydrolysis, Smith degradation and alkaline degradation of the O-polysaccharide, together with 1H-NMR and 13C-NMR spectroscopic data of the polysaccharide, led to the following structure for the repeating units:→3)Rha(α1→4)Ga1NAcA(α1→4)Ga1NFoA(α1→3)QuiNAc(α1→. Almost all of the carboxyl groups of the N-acetylgalactosaminuronic acid residues and about half of the same groups of the N-formylgalactosaminuronic acid residues were in an amide form.
    Download PDF (865K)
  • Yasuo BAI, Yasuteru MURAGAKI, Ken-ichi OBATA, Kazushi IWATA, Akira OOS ...
    1986 Volume 99 Issue 6 Pages 1563-1570
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Monoclonal antibodies to human (8 clones) and rat (12 clones) prolyl 4-hydroxylase [EC 1.14.11.2] were prepared and characterized as regards subclass, subunit specificity, inhibition and crossreactivity. Among the antibodies to the human enzyme, four clones showed the IgG1 subclass, two IgA, one IgG2b, and one IgM. Four clones reacted with the α subunit of the enzyme, while the others reacted with the β subunit. The enzymatic activity was inhibited by four clones. Five clones crossreacted with the rat enzyme. One clone inhibited the rat enzyme. Among the antibodies to the rat enzyme, seven clones showed the IgG1 subclass, four IgG2a and one IgG2b. Seven clones reacted with the α subunit, and four with the β subunit. One reacted with neither subunit. The enzymatic activity was inhibited by seven clones. Seven clones crossreacted with the human enzyme. Three clones inhibited the human enzyme.
    Download PDF (1069K)
  • Masayoshi TAKEUCHI, Naoki ASANO, Yukihiko KAMEDA, Katsuhiko MATSUI
    1986 Volume 99 Issue 6 Pages 1571-1577
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    3-Ketovalidoxylamine A C-N lyase of Flavobacterium saccharophilum is a monomeric protein with a molecular weight of 36, 000. Amino acid analysis revealed that the enzyme contains 5 histidine residues and no cysteine residue. The enzyme was inactivated by diethylpyrocarbonate (DEP) following pseudo-first order kinetics. Upon treatment of the inactivated enzyme with hydroxylamine, the enzyme activity was completely restored. The difference absorption spectrum of the modified versus native enzyme exhibited a prominent peak around 240 nm, but there was no absorbance change above 270 nm. The pH-dependence of inactivation suggested the involvement of an amino acid residue having a pKa of 6.8. These results indicate that the inactivation is due to the modification of histidine residues. Substrates of the lyase, p-nitrophenyl-3-ketovalidamine, p-nitrophenyl-α-D-3-ketoglucoside, and methyl-α-D-3-ketoglucoside, protected the enzyme against the inactivation, suggesting that the modification occurred at or near the active site. Although several histidine residues were modified by DEP, a plot of log (reciprocal of the half-time of inactivation) versus log (concentration of DEP) suggested that one histidine residue has an essential role in catalysis.
    Download PDF (531K)
  • Yoshihiro YAMAMOTO, Tomoko OGAWA, Hideo SHINAGAWA, Tatuo NAKAYAMA, His ...
    1986 Volume 99 Issue 6 Pages 1579-1590
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Prior to the analysis of transcription and translation, the nucleotide sequence of the uvrD gene and its neighboring regions was determined by the method of Maxam and Gilbert (Maxam & Gilbert (1980) Methods Enzymol. 65, 499-560). Disagreement in 14 positions between the nucleotide sequence determined by us and that reported previously (Finch & Emmerson (1984) Nucl. Acids Res. 11, 5789-5799) was found. We reexamined these disputed regions. The initiation site of transcription of the uvrD gene was determined by analyzing the transcripts synthesized in vitro. It was found that transcription of the uvrD gene starts from the A nucleotide, which is the first one of the SOS box of the uvrD. The amino terminal sequence and the amino acid composition of the purified UvrD protein (helicase II) were determined. It was found that translation starts from the first ATG codon, which lies 77 nucleotides downstream from the initiation site of transcription. The amino acid composition of the purified UvrD protein agreed well with that deduced from the nucleotide sequence.
    Download PDF (1944K)
  • Tadashi MATSUMOTO, Koichiro TAKESHIGE, Shigeki MINAKAMI
    1986 Volume 99 Issue 6 Pages 1591-1596
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    When guinea pig peritoneal neutrophils were suspended in the isotonic medium of potassium, rubidium, and cesium ions at 37°C, the cells released superoxide, while low activity was observed in the isotonic medium of sodium and lithium ions. The activity induced in the potassium medium was enhanced by potassium-ionophores, valinomycin, and gramicidin, and decreased by a potassium channel blocker, 4-aminopyridine. The superoxide-releasing activity was not affected by the presence or absence of extracellular calcium but was inhibited by an intracellular calcium antagonist-8-(N, N-diethylamino)-octyl-3, 4, 5-trimethoxybenzoate(TMB-8) with the half-inhibition concentration of 50μM. The release of granular enzymes, lysozyme and β-glucronidase, was also induced in the isotonic potassium medium in the absence of extracellular calcium and inhibited by TMB-8. A remarkable elevation of the intracellular free calcium concentration in neutrophils, which was monitored by quin-2 fluorescence, was found when the cells were added to the potassium medium without calcium. The elevation was inhibited by the addition of TMB-8. These observations suggest that calcium mobilization from intracellular storage sites, not an influx of calcium from the extracellular medium, causes the release of superoxide and the granular enzymes in isotonic potassium medium.
    Download PDF (462K)
  • Jun TAKAMATSU, Koichiro TAKESHIGE, Shosuke TAKAHASHI, Junichi YOSHITAK ...
    1986 Volume 99 Issue 6 Pages 1597-1604
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Phagocytic vesicles with superoxide-forming NADPH oxidase activity were obtained from human monocytes phagocytosing oil droplets. The superoxide-forming activity in the monocyte vesicles increased for the first 5 min during incubation with oil droplets and remained constant for 30 min. NADPH-dependent activities of 2, 6-dichlorophenol-indophenol (DCIP) reduction and ubiquinone-1 (Q1) reduction were found in the vesicles and the activities were closely associated with the superoxide-forming oxidase. The values of apparent Km for NADPH of these three activities were essentially the same and the activities were inhibited with a similar pattern by p-chloromercuribenzoate and a cationic detergent, cetyltrimethylammonium bromide. The activities were extremely labile and the DCIP reductase activity was most labile. The superoxide-forming oxidase and the Q1 reductase could be extracted with a mixture of deoxycholate and Tween-20. The extracted activities were not enhanced by the addition of FAD.
    Download PDF (639K)
  • Michio NOJIMA, Shoichi ISHIURA, Takeshi YAMAMOTO, Teruaki OKUYAMA, Hir ...
    1986 Volume 99 Issue 6 Pages 1605-1611
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    We purified a high-molecular-weight protease, ingensin, from extract of human placenta by successive DEAE-cellulose, hydroxyapatite, and high performance liquid chromatographies. The activity of ingensin was determined by using a synthetic substrate, succinyl-leucyl-leucyl-valyl-tyrosine-methylcoumarinamide (MCA). The purified ingensin, which gave a single band in 6.5% nondenaturing polyacrylamide gel electrophoresis, was activated by linoleic acid and sodium dodecyl sulfate (SDS). Maximum activity was observed at pH 9.5 in the presence of 0.06% SDS, but at pH 8.0 in the presence of linoleic acid. A subcellular fractionation study showed that a large amount of ingensin activity was present in the cytosol or microsome fraction rather than in the precipitate of low-speed centrifugation. The effect of protease inhibitors on the activated ingensin was also investigated.
    Download PDF (888K)
  • Hideaki NAGAMUNE, Osamu ORAYAMA, Yukichi HARA, Makoto NAKAO
    1986 Volume 99 Issue 6 Pages 1613-1624
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The ouabain-sensitive phosphatase activity of purified pig kidney Na, K-ATPase preparation in the absence of potassium ion ((-K)phosphatase) was examined precisely. During the preparation procedures, the (-K)3-O-methylfluoresceinphosphatase ((-K)3-OMFPase) activity or the (-K)p-nitrophenylphosphatase ((-K)pNPPase) activity appeared to be purified in parallel with the Na, K-ATPase activity. The (-K)phosphatase activity was competitively inhibited by ATP and by ADP, with the K1 values of 0.25μM and 1.4μM, respectively. These values are consistent with their Kd values for the high-affinity ATP binding site of the Na, K-ATPase (Hegyvary, C. & Post, R.L. (1971) J. Biol. Chem. 246, 5234-5240). The substrate, pNPP, apparently competed with covalently bound fluorescein-5'-isothiocyanate (FITC), which is known to bind in the neighborhood of the high-affinity ATP binding site of the Na, K-ATPase, in both the (-K)phosphatase and the (+K)phosphatase reactions.
    The FITC-fluorescence intensity of FITC-labeled enzyme at the maximal steadystate activity of the (-K)phosphatase reaction was at a similar level to that of the E2 species. However, the FITC-labeled enzyme in the presence of only magnesium ion or only pNPP gave a fluorescence level similar to that of the E1 species. Oligomycin inhibited the (-K)phosphatase activity by at most 46%.
    On the basis of these results, it is strongly suggested that the (-K)phosphatase reaction is catalyzed at the high-affinity ATP binding site of Na, K-ATPase, and the (-K)phosphatase reaction proceeds in a cyclic manner (E1→E2→E1).
    Download PDF (947K)
  • Tetsuo MURO, Yoshio TOMINAGA, Shigetaka OKADA
    1986 Volume 99 Issue 6 Pages 1625-1630
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The protease from Streptomyces cellulosae preferentially catalyzed the formation of (L-Leu-Gly)2 (P1) and (L-Leu-Gly)3 (P2) in highly concentrated solutions of L-Leu-Gly, although it weakly hydrolyzed the substrate at the same time. The formation of P1, P2, L-Leu, and Gly was studied at various pH values, temperatures, and substrate concentrations. The initial velocities (ν1, ν2, and vH) of formation of P1, P2, and L-Leu (or Gly) and the sum (ν1T) of ν1 and ν2, were determined.
    The effects of pH and temperature on ν1, ν2, νH, and ν1T were examined at a fixed substrate concentration. The optimum pH and optimum temperature for each of the processes forming P1, P2, L-Leu, and Gly were 8.0 and 65°C, respectively.
    In the study on the effect of substrate concentration, the plots of the initial velocities versus substrate concentrations were sigmoidal at lower substrate concentrations. The dependence of ν1T on the substrate concentration could be explained by a mechanism involving a single active center forming the peptide bonds and two substrate-binding sites located on the left sites (S1 and S2) and the right sites (S1' and S2') of the active center of this enzyme.
    Download PDF (361K)
  • Yukihiro ISODA, Yasunori NITTA
    1986 Volume 99 Issue 6 Pages 1631-1637
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The synthesized 2', 3'-epoxypropyl α-D-glucopyranoside (α-EPG) inactivated soybean β-amylase completely. The incorporation of α-EPG into the enzyme at 92% inactivation was 1.1 mol per mol of enzyme, as determined by using 14C-labeled α-EPG. The inactivation obeyed saturation kinetics of a two-step mechanism. The dissociation constant of α-EPG-enzyme complex and the rate constant of the irreversible inactivation step were estimated to be 119 mm and 1.14×10-3 s-1, respectively. α-Cyclodextrin, a competitive inhibitor of this enzyme, protected the enzyme against the inactivation by α-EPG in a competitive manner. This suggests that α-EPG binds to the active site of the enzyme. The above results indicate that α-EPG acts on soybean β-amylase as an affinity labeling reagent. It was also shown that an essential SH group near the active site, but not the catalytic one, scarcely participated in the inactivation by α-EPG.
    Download PDF (506K)
  • Hiroyasu SATOW, Senkiti SAKAI, Masuo OBINATA
    1986 Volume 99 Issue 6 Pages 1639-1643
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Stage-specific gene expression of 26 k casein, a major milk protein, was examined in mouse mammary glands. Mature mRNA (1.1 kb) with putative precursor RNA (5 kb) was detected in the lactating gland. In contrast, the putative precursor RNA (5 kb) was also detected in the mammary glands of virgin or weaned mice when milk was not produced, although mature mRNA was not detected. The RNA transcripts showing heterogeneous size distribution were abundant in the nonlactating stages. These RNAs are shown to be transcribed in the opposite direction to the mature mRNA. These results suggest the regulation at the post-transcriptional level in the synthesis of caseins in the mammary glands.
    Download PDF (1775K)
  • Junko AMANO, Akira KOBATA
    1986 Volume 99 Issue 6 Pages 1645-1654
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    An α-mannosidase differing from 1, 2-α-mannosidase was found to occur in Aspergillus saitoi. By a series of column chromatographies the enzyme was purified up to 1, 000fold, and its properties were studied in detail.
    The enzyme preparation, which was practically free from other exoglycosidases, showed a pH optimum of 5.0. In contrast to 1, 2-α-mannosidase, the enzyme was strongly activated by Ca2+ ions. p-Nitrophenyl α-mannopyranoside was not hydrolyzed by the enzyme. Accordingly, the substrate specificity of the new α-mannosidase was studied by using a variety of tritium-labeled oligosaccharides. Studies with linear oligosaccharides revealed that the enzyme cleaves the Manα1→3Man linkage more than 10 times faster than the Manα1→6Man and the Manα1→2Man linkages. Furthermore, it cleaves the Manα1→6Man linkage of the Manα1→6(Manα1→3)Manβ1→4GlcNAcβ1→4GlcNAcOT only after its Manα1→3 residue is removed. Because of this specificity, the enzyme can be used as an effective reagent to discriminate R→Manα1→6(Manα1→3)Manβ1→4G1cNAcβ1→4(±Fucα1→6)GlcNAcOT from its isomeric counterparts, Manα1→6(R→Manα1→3)Manβ1→4GlcNAcβ1→4(±Fucα1→6)GlcNAcOT, in which R represents sugars.
    Download PDF (815K)
  • Yasuyuki SUZUKI, Hyogo SINOHARA
    1986 Volume 99 Issue 6 Pages 1655-1665
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Guinea pig α-macroglobulin was purified to apparent homogeneity by sequential chromatography on Sephacryl S-300, DEAE-cellulose, and hydroxyapatite. A molecular weight of 780, 000 was obtained by equilibrium sedimentation. The preparation migrated as a single band of Mr=180, 000 in sodium dodecyl sulfatepolyacrylamide gel electrophoresis under reducing conditions. Rabbit antiserum raised against the final preparation partially cross-reacted with human and rat α-2-macroglobulins but not with rat α-l-macroglobulin. Guinea pig α-macroglobulin stimulated the amidolytic activity of trypsin towards a small substrate, but inhibited the proteolytic activity of trypsin towards remazol brilliant blue hide powder. When treated with trypsin or methylamine, four thiol groups per molecule were newly generated. The reaction with trypsin proceeded with at least at two different rates: half of the thiol groups were generated in a fast reaction and the remaining half in a slower reaction. On the other hand, such a two-step reaction was not detected in the reaction with methylamine. The methylamine-treated a-macroglobulin retained half the capacity to bind trypsin and its mobility in polyacrylamide gel under nondenaturing conditions remained virtually unchanged. These properties are in marked contrast to those reported for human α-2-macroglobulin, but resemble those of rat α-2- and mouse a-macroglobulins. The amidase activity of trypsin bound to guinea pig α-macroglobulin was impaired by soybean trypsin inhibitor to a much greater degree than that of trypsin bound to human or rat α-2-macroglobulin.
    Download PDF (2270K)
  • Shinji IIJIMA, Man-Jin OH, Takashi SAIKI, Teruhiko BEPPU
    1986 Volume 99 Issue 6 Pages 1667-1672
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Heat-stable malate dehydrogenase isolated from Thermus flavus AT62 was completely inactivated by treatment with diethylpyrocarbonate. The inactivation was accompanied by the loss of 1.2 histidine residues per subunit of the enzyme. The enzyme was protected from inactivation by NADH.
    The enzyme was also inactivated by dye-sensitized photooxidation. Methionine residues, in addition to histidine residues, were destroyed in the inactivated enzyme. Kinetic analyses of the inactivation indicated that the pK value of the residue involved in the inactivation was 8.20 at 25.0°C and 7.52 at 60.0°C. From the pK values and the heat of ionization calculated from the van't Hoff plot of pKs, a histidine residue was identified to be primarily involved in the inactivation. The effect of temperature on the pK value of the essential group in this enzyme from a thermophilic organism is discussed.
    Download PDF (477K)
  • Rika SUZUKI, Shigeo MASAKI, Osamu KOIWAI, Shonen YOSHIDA
    1986 Volume 99 Issue 6 Pages 1673-1679
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    It has been shown that DNA primase activity is tightly associated with 10S DNA polymerase α from calf thymus (Yoshida, S. et al. (1983) Biochim. Biophys. Acta 741, 348-357). In the present study, the primase activity was separated from DNA polymerase α by treating purified lOS DNA polymerase α with 3.4 M urea followed by a fast column chromatography (Pharmacia FPLC, Mono Q column equilibrated with 2 M urea). Ten to 20% of the primase activity was separated from 10S DNA polymerase α by this procedure but 80-90% remained in the complex. The separated primase activity sedimented at 5.6S through a gradient of glycerol. The separated primase was strongly inhibited by araATP (K1=10μM) and was also sensitive to salts such as KCl (50% inhibition at 30 mm). The primase used poly(dT) or poly(dC) as templates efficiently, but showed little activity with poly(dA) or poly(dI). These properties agree well with those of the primase activity in the DNA polymerase α-primase complex (10S DNA polymerase α). These results indicate that the calf thymus primase may be a part of the 10S DNA polymerase α and its enzymological characters are preserved after separation from the complex.
    Download PDF (1133K)
  • Keizo MIYATA, Yoko YAMAMOTO, Masamichi UEDA, Yoshimi KAWADE, Kayoko MA ...
    1986 Volume 99 Issue 6 Pages 1681-1688
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    A simple procedure for purifying human interferon-γ from leukocytes was established, based on monoclonal antibody affinity chromatography. The recovery of interferon activity was essentially quantitative, and the specific activity of the product was (4-12)×107 international units/mg protein. SDS-polyacrylamide gel electrophoresis reproducibly revealed four components associated with interferon activity (and no other proteins): two major ones with molecular weights (MW) of 24, 000-25, 000 (25K) and 19, 000-20, 000 (20K), a minor one with MW 14, 000-15, 000 (15K) (these three bands were doublets), and a still less prominent one(s) with MW 40, 000-48, 000. Gel filtration in neutral solution indicated that all the 25K, 20K, and 15K species exist as oligomers, probably dimers. By means of experiments using a cleavable crosslinking reagent, the dimers were shown to comprise both homo- and heterodimers. Gel filtration in alkali (the condition used during purification) indicated that the molecules are largely in a monomeric state. Thus, the molecules once dissociated in alkali appear to reassociate at random upon neutralization; this process takes place without being accompanied by inactivation.
    Download PDF (1865K)
  • Kazuki SAITO, Atsuko SHINOHARA, Tetsuya KAMATAKI, Ryuichi KATO
    1986 Volume 99 Issue 6 Pages 1689-1697
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    N-Hydroxyarylamine O-acetyltransferase, arylhydroxamic acid N, O-acetyltransferase, and arylamine N-acetyltransferase in hamster liver cytosol were co-purified almost to electrophoretical homogeneity by ion exchange chromatography on DEAF-cellulose, gel filtration on Cellulofine GCL-2000-sf and high-performance KB-hydroxyapatite chromatography. The molecular weight of the acetyltransferase was estimated to be 33, 000 by gel filtration and SDS-polyacrylamide gel electrophoresis. The three acetyltransferase activities were inhibited by iodoacetamide, pentachlorophenol, and 1-nitro-2-naphthol. Furthermore, 2-aminofluorene, a substrate for arylamine N-acetyltransferase, inhibited the reactions of N-hydroxyarylamine O-acetyl transfer and arylhydroxamic acid N, O-acetyl transfer. These results suggest that the same enzyme catalyzes the three types of acetyl transfer reactions. The acetyltransferase could activate N-hydroxyarylamines, such as 2-hydroxyamino-6-methyldipyrido[1, 2-a: 3', 2'-d]imidazole, 3-hydroxyamino-l-methyl-5H-pyrido[4, 3-b]indole, and N-hydroxy-2-aminofluorene, to the corresponding N-acetoxyarylamines, which are capable of binding to nucleic acid. Polyguanylic acid was most efficiently modified by the N-acetoxyarylamines formed by the acetyltransferase.
    Download PDF (1252K)
  • Takashi SASAKI, Takashi MORITA, Sadaaki IWANAGA
    1986 Volume 99 Issue 6 Pages 1699-1705
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    A plasminogen-binding site of human α2-plasmin inhibitor was studied. The chromatogram of digest from the amidinated α2-plasmin inhibitor (67K-daltons, plasminogen-binding form) with trypsin was almost identical with that obtained from the 65K-daltons derivative (non-plasminogen-binding form) treated with the same procedure, except for the three tryptic peptides. One of the three peptides, the deamidinated peptide T-11, was found to have a strong ability to inhibit the interaction of α2-plasmin inhibitor with human plasmin. Moreover, the dissociation constant Kd for interaction between the peptide T-11 and plasmin was estimated to be 5.5μM, indicating that Kd is about 10-fold lower than that of ε-aminocaproic acid. The sequence of the peptide T-11 was determined by the Edman method as follows: NH2-G-D-K-L-F-G-P-D-L-K-L-V-P-P-M-E-E-D-Y-P-Q-F-G-S-P-K-COOH. α2-Plasmin inhibitor and its 65K-daltons derivative were found to have the same NH2-terminal sequence of Asn(Asp)-Gln-Glu-Gln-. These results indicated that the plasminogen-binding site(s) of α2 plasmin inhibitor could be located in the COOH-terminal region of its molecule and that some of ε-NH2-groups in the deamidinated peptide T-11 may be involved in the lysine-binding site(s) of plasmin(ogen).
    Download PDF (482K)
  • Yoshiharu SHIRASU, Hideaki YOSHIDA, Toshifumi MIKAYAMA, Shigeru MATSUK ...
    1986 Volume 99 Issue 6 Pages 1707-1712
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    We have cloned a DNA that is complementary to the messenger RNA that encodes porcine pancreatic elastase 1 from pancreas using rat pancreatic elastase 1 cDNA as a probe. This complementary DNA contains the entire protein coding region of 798 nucleotides which encodes an elastase of 266 amino acids, and 22 and 136 nucleotides of the 5' and 3'-untranslated sequences. When this deduced amino acid sequence was compared with known amino acid sequences, a carboxy-terminal 240 amino acids long peptide was found to be identical with a mature form of porcine pancreatic elastase 1, except for two amino acids. The porcine enzyme contains the same number of amino acid residues as the rat enzyme, and their amino acid sequences are 85% homologous. Taking the above findings together, we conclude that the cloned cDNA encodes a mature enzyme of 240 amino acids including a leader and activation peptide of 26 amino acids. We expressed the cloned porcine pancreatic elastase 1 cDNA in E. coli as a lac-fused protein. The resulting fused protein showed enzymatic activity and immunoreactivity toward anti-elastase serum.
    Download PDF (1096K)
  • Yasuyuki ITO, Yoshihiro OH-HASHI, Harumi OKUYAMA
    1986 Volume 99 Issue 6 Pages 1713-1718
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    A partially purified fatty acid synthetase from Candida utilis synthesized fatty acids with varying chain lengths that depended on the assay temperature; the stearate/palmitate ratio decreased with decreasing temperature. This temperature-dependency was also observed in vivo for the newly synthesized fatty acids in cells incubated at various temperatures, although to a lesser extent than that observed in vitro. The difference in the temperature-dependencies observed in vivo and in vitro appeared to be due to the difference in the acceptors used in in vitro assays; a temperature-dependency comparable to that observed in vivo was reproduced in vitro on using microsomes rather than bovine serum albumin as the acceptor of the fatty acid synthetase products. Thus, the fatty acid synthetase was identified as a thermo-receptor in Candida cells known to possess a temperature-dependent, inducible desaturase system.
    Download PDF (493K)
  • Hiraku ITIKAWA, Hiroshi FUJITA, Morimasa WADA
    1986 Volume 99 Issue 6 Pages 1719-1724
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The cellular concentrations of ppGpp in the dnaK(Ts) and dnaJ(Ts) mutants of Escherichia coli were examined, since the thermosensitive RNA synthesis of these mutants is relaxed by an additional mutation in the relA gene. The results showed that ppGpp accumulated extensively in the dnaK(Ts) and dnaJ(Ts) mutants after a temperature shift up, reaching levels of 5mm and 0.5mm, respectively. This unusual accumulation of ppGpp was suppressed by the relA1 mutation, implying that it results from induction of a stringent response in these mutants at a nonpermissive temperature.
    Download PDF (1697K)
  • Sumihiro HASE, Toshiaki SUGIMOTO, Hiroshi TAKEMOTO, Tokuji IKENAKA, Ka ...
    1986 Volume 99 Issue 6 Pages 1725-1733
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    While the structure of the major oligosaccharide of Japanese quail ovomucoid was reported earlier (Hase, S. et al. (1982) J. Biochem. 91, 735-737), the structures of the minor oligosaccharide units were investigated for the first time in the present studies. For this purpose, the glycans of the protein were liberated from the polypeptide chain by hydrazinolysis. After N-acetylation, the reducing ends of the oligosaccharides obtained were coupled with 2-aminopyridine, and then the resulting fluorescent derivatives were purified by Bio-Gel P-2 column chromatography and reversedphase HPLC. The chemical structures of two minor oligosaccharide units were determined with the aid of exoglycosidases, and by methylation analysis and Smith degradation. The results demonstrated that the ovomucoid contains the following two monoantennary glycans: Mana1-6(Galβ1-4GlcNAcβ1-2Manα1-3)Manβ1-4GlcNAcβ1-4GlcNAc and Galβ1-4GlcNAcβ1-2Manα1-6(Manα1-3)Manβ1-4Glc-NAcβ1-4GlcNAc. The latter structure was not predicted by the classical metabolic pathway for the N-glycans to be formed. The structures of three additional minor heterosaccharides were deduced from their elution positions on HPLC together with the results of determination of their molecular sizes and the HPLC elution positions of their enzymatic degradation products. It is noteworthy that for the latter procedure for the estimation of the structures of oligosaccharides only minute quantities of glycans (several hundreds pmol) are required.
    Download PDF (662K)
  • Akihiko KAWAGUCHI, Nobuo UEMURA, Shigenobu OKUDA
    1986 Volume 99 Issue 6 Pages 1735-1742
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Curtobacterium pusillum contains 11-cyclohexylundecanoic acid as a major component of cellular fatty acids. A trace amount of 13-cyclohexyltridecanoic acid is also present. Fatty acids other than ω-cyclohexyl fatty acids present are 13-methyltetradecanoic, 12-methyltetradecanoic, n-pentadecanoic, 14-methylpentadecanoic, 13-methylpentadecanoic, n-hexadecanoic, 15-methylhexadecanoie, 14-methylhexadecanoic, and n-heptadecanoic acids. The fatty acid synthetase system of this bacterium was studied. Various 14C-labeled precursors were added to the growth medium and the incorporation of radioactivity into cellular fatty acids was analyzed. Sodium [14C]acetate and [14C]glucose were incorporated into almost all species of cellular fatty acids, the incorporation into 11-cyclohexyhmdecanoic acid being predominant. [14C]Isoleucine was incorporated into 12-methyltetradecanoic and 14-methylhexadecanoic acids; [14C]leucine into 13-methyltetradecanoic and 15-methylhexadecanoic acids; and [14C]valine into 14-methylpentadecanoic acid. [14C]-Shikimic acid was incorporated almost exclusively into ω-cyclohexyl fatty acids. The fatty acid synthetase activity of the crude enzyme preparation of C. pusillum was reconstituted on the addition of acyl carrier protein. This synthetase system required NADPH and preferentially utilized cyclohexanecarbonyl-CoA as a primer. The system was also able to use branched- and straight-chain acyl-CoAs with 4 to 6 carbon atoms effectively as primers but was unable to use acetyl-CoA. However, if acetyl acyl carrier protein was used as the priming substrate, the system produced straight-chain fatty acids. The results imply that the specificity of the initial acyl-CoA: acyl carrier protein acyltransferase dictates the structure of fatty acids syn-thesized and that the enzymes catalyzing the subsequent chain-elongation reactions do not have the same specificity restriction.
    Download PDF (1173K)
  • Isao HAMAMOTO, Atsuo HIWATASHI, Yoshiyuki ICHIKAWA
    1986 Volume 99 Issue 6 Pages 1743-1748
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The zona glomerulosa, zona fasciculata, zona reticularis, and medulla were separated from bovine adrenal glands and cytochromes P-450 and related enzymes in each zone were investigated immunochemically by Western blotting using antisera from chickens or rabbits against cytochromes P-450scc, P-45011β, P-450s21, and b5, NADH-cytochrome b5 reductase, NADPH-cytochrome P-450 reductase, NADPH-adrenodoxin reductase, and adrenodoxin.
    Concentrations of cytochrome P-45011β, NADPH-cytochrome P-450 reductase, and cytochrome b5 per milligram of protein of homogenate were higher in the zona glomerulosa than in the other zones; the levels of the other components were higher in the zona fasciculata. The total enzyme content of all components was the highest in the zona fasciculata. The amount of adrenodoxin was about 10 times that of NADPH-adrenodoxin reductase in each zone.
    Download PDF (1117K)
  • Sohsuke SHIDARA, Hidekazu IWASAKI, Tetsuhiko YOSHIMURA, Shinnichiro SU ...
    1986 Volume 99 Issue 6 Pages 1749-1752
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Cytochrome c' was crystallized from Achromobacter xylosoxidans GIFU 543. The cytochrome was a basic protein and its molecular weight was 28, 000. The pyridine ferrohemochrome showed absorption peaks at 415, 521, and 551nm. The absorption spectra of the oxidized and reduced forms at neutral pH were almost the same as those of other cytochromes c' reported already. The reduced cytochrome c' reacted with CO and NO, and the NO complex showed a characteristic absorption spectrum. The midpoint redox potential of the hemoprotein was measured to be +110 mV at pH 7.2.
    Download PDF (912K)
  • Shin-ichi HAYASHI, Mitsuhide NOSHIRO, Kyuichiro OKUDA
    1986 Volume 99 Issue 6 Pages 1753-1763
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    A molecular species of cytochrome P-450 that catalyzes the 25-hydroxylation of cholecalciferol (P-450cc25) was purified from rat liver microsomes on the basis of its catalytic activity. The purification procedure consisted of polyethylene glycol fractionation, and column chromatographies on octylamino Sepharose 4B, hydroxylapatite, DEAE-Sepharose CL-6B, and CM-Sepharose CL-6B. The specific cytochrome P-450 content of the final preparation was 17.0 nmol/mg of protein. The enzymatic activity was reconstituted with the purified cytochrome P-450, NADPH-cytochrome P-450 reductase, an NADPH-generating system, and dilauroylglyceryl-3-phosphorylcholine, the specific activity obtained being 3.7 nmol/min/mg of protein, which was 4, 000 times as high as that in microsomes. The apparent molecular weight of the P-450cc25 was 50, 000, based on the results of sodium dodecyl sulfate polyacrylamide gel electrophoresis. The absorption spectra of the oxidized form of the enzyme showed a Soret band at 416 nm, which is typical of the low spin state of cytochrome P-450, and α and β bands at 570 and 536 nm, respectively. The Soret peak of the reduced cytochrome P-450-CO complex was at 450 nm. The purified enzyme not only catalyzed the 25-hydroxylation of cholecalciferol but also showed hydroxylation activity toward a variety of substrates, i.e. 1α-hydroxycholecalciferol (at 25), testosterone (at 2α and 16α) and dehydroepiandrosterone (at 16α). Amino terminal sequence of the purified cytochrome P-450 was determined by the manual sequence method to be H2N-Met-Asp-Pro-Val-Leu-Val-Leu-Val-. The antibody elicited against the purified enzyme in a rabbit inhibited the cholecalciferol 25-hydroxylation activity by more than 90% with a concentration of 2mg of immunoglubulin per nmol of cytochrome P-450.
    Download PDF (1363K)
  • Shozo IIDA, James D. POTTER
    1986 Volume 99 Issue 6 Pages 1765-1772
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The effects of Mg2+ ion, pH, and KCl concentration on Ca2+ binding to calmodulin were studied by using a Ca2+ ion-sensitive electrode. The Ca2+ ion affinity of calmodulin increased with increasing pH or decreasing KCl concentration. Cooperativity between the Ca2+-binding sites was observed, and increased with decreasing pH or increasing KCI concentration. Free Ca2+ ion concentration was decreased by adding MgCl2 ion at low Mg2+ concentration and increased at higher concentrations in the presence of small amounts of Ca2+ ion. The decrease of free Ca2+ ion concentration by Mg2+ ion strongly suggests cooperativity between the Ca2+-binding sites, and it is difficult to explain the decrease in terms of the ordered binding models previously proposed. These results can be explained by a simple model which has four equivalent binding sites that bind Ca2+ and Mg2+ competitively, and showing cooperativity when either Ca2+ or Mg2+ is bound. Mg2+ ion binding to calmodulin was measured in the presence or absence of Ca2+ to confirm the validity of this model, and no Mg2+-specific site was observed.
    Download PDF (655K)
  • Isamu YAMAMOTO, Norimichi OKUBO, Makoto ISHIMOTO
    1986 Volume 99 Issue 6 Pages 1773-1779
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Escherichia coli trimethylamine N-oxide (TMAO) reductase I, the major enzyme among inducible TMAO reductases, was purified to homogeneity by an improved method including heat treatment, ammonium sulfate precipitation, and chromatographies on Bio-Gel A-1.5m, DEAE-cellulose, and Reactive blue-agarose. The molecular weight was estimated by gel filtration to be approximately 200, 000. A single subunit peptide with a molecular weight of 95, 000 was found by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. This enzyme contained 1.96 atoms of molybdenum, 0.96 atoms of iron, 1.52 atoms of zinc, and less than 0.4 atoms of acid-labile sulfur per molecular weight of 200, 000. The absorption spectrum of the enzyme showed a peak at 278nm and a shoulder at 288 nm, but no characteristic absorption was found from 350 to 700nm. A fluorescent derivative of molybdenum cofactor was found when the enzyme was boiled with iodine in acidic solution; its fluorescence spectra were almost the same as those of the form A derivative of molybdopterin found in sulfite oxidase. The molybdenum cofactor released from heated TMAO reductase I reconstituted nitrate reductase in the extracts of Neurospora crassa mutant strain nit-1 lacking molybdenum cofactor. Thus, TMAO reductase I contains molybdopterin, which is a common constituent of some molybdenum-containing enzymes. Some kinetic properties were also determined.
    Download PDF (1183K)
  • Hideo MIZUNUMA, Yohtalou TASHIMA
    1986 Volume 99 Issue 6 Pages 1781-1788
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Fructose 1, 6-bisphosphatase has been purified from rat muscle. Although the specific activity of the enzyme in the crude extract of rat muscle was extremely low, purification by the present procedure is highly reproducible. The purified enzyme showed a single band in SDS-polyacrylamide gel electrophoresis. The subunit molecular weight of the muscle enzyme was 37, 500 in contrast to 43, 000 in the case of the liver enzyme. Immunoreactivity of the muscle enzyme to anti-muscle and anti-liver fructose 1, 6-bisphosphatase sera was clearly distinct from that of the liver enzyme. All one-dimensional peptide mappings of the muscle enzyme with staphylococcal V8 protease, chymotrypsin, and papain showed different patterns from those of the liver enzyme. When incubated with subtilisin, the extent of activation of muscle fructose 1, 6-bisphosphatase at pH 9.1 was smaller than that of the liver enzyme. The subtilisin digestion pattern of the muscle enzyme on SDS-polyacrylamide gel electrophoresis was distinct from that of the liver enzyme. The AMP-concentration giving 50%. inhibition of the muscle enzyme was 0.54μM, whereas that of the liver enzyme was 85μm. The concentrations of fructose 2, 6-bisphosphate that gave 50% inhibition of rat muscle and liver enzymes were 6.3 and 1.5μM, respectively. Fructose 1, 6-bisphosphatase protein was not detected in soleus muscle by immunoelectroblotting with anti-muscle fructose 1, 6-bisphosphatase serum.
    Download PDF (3454K)
  • Shuzo OTANI, Isao MATSUI-YUASA, Yasuko MIMURA, Seiji MORISAWA
    1986 Volume 99 Issue 6 Pages 1789-1797
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Retinoic acid potentiated the increases in ornithine decarboxylase (L-ornithine carboxy-lyase [EC 4.1.1.17]) activity, [3H]difluoromethylornithine binding to ornithine decarboxylase, intracellular levels of polyamines and DNA synthesis in guinea pig lymphocytes stimulated with phytohemagglutinin. The stimulatory effect on the ornithine decarboxylase induction was dependent on the dose of retinoic acid and on the time of addition of the drug. Retinoic acid had to be added not later than 2h after phytohemagglutinin to elicit the potentiation. Retinyl acetate also potentiated ornithine decarboxylase induction caused by phytohemagglutinin. Both of these retinoids augmented ornithine decarboxylase induction caused by phorbol 12-myristate 13-acetate. The half-life of ornithine decarboxylase activity estimated after addition of actinomycin D was longer in cells treated with phytohemagglutinin or phorbol 12-myristate 13-acetate together with retinoic acid than in cells treated with the mitogen alone. The half-life after addition of cycloheximide was not affected by retionic acid. These results suggest that the retinoids are stimulators rather than inhibitors of ornithine decarboxylase induction caused by phytohemagglutinin or phorbol 12-myristate 13-acetate in guinea pig lymphocytes and that retinoic acid potentiates the enzyme activity at the transcriptional or posttranscriptional, but not at the post-translational stage.
    Download PDF (717K)
  • Shigenori ITO, Kotoyo YAMAZAKI, Tokiko MIYAZAKI, Hideo MATSUMOTO
    1986 Volume 99 Issue 6 Pages 1799-1802
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    An allotypic G3m(g) marker-specific substitution was studied by sequence analysis of glycopeptides derived from myeloma proteins Ba (G3m(g+)) and Bu (G3m(g-)). The experimental results indicate that glutamic acid at position 295 is responsible for the specificity. Based on the results of chemical modification (Arg, Tyr, and Glu), this antigenic epitope is presumed to involve five sequential residues from Arg-292 to Tyr-296.
    Download PDF (285K)
  • Keiko TAMIYA-KOIZUMI, Kiyohide KOJIMA
    1986 Volume 99 Issue 6 Pages 1803-1806
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The plasma membrane isolated from rat ascites hepatoma, AH 7974 cells was treated with 1% Triton X-100, which resulted in a more than 80 % reduction in the phospholipid content of the plasma membrane. The delipidized plasma membrane showed only 18% of the activity of the magnesium-dependent, neutral sphingomyelinase in the untreated plasma membrane. On the addition of acidic phospholipids, especially phosphatidylserine, however, the enzyme activity in the delipidized membrane was markedly restored up to 77% of that in the untreated membrane. It was suggested that, considering the phospholipid composition of the AH 7974 plasma membrane (Koizumi, K. et al. (1977) Cell Struct. Func. 2, 145-153), phosphatidylserine may be a natural activator of neutral sphingomyelinase.
    Download PDF (256K)
  • Yasuo IHARA, Nobuyuki NUKINA, Reiko MIURA, Midori OGAWARA
    1986 Volume 99 Issue 6 Pages 1807-1810
    Published: 1986
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Antisera to paired helical filaments (PHF) were found to contain a significant amount of tau antibodies specific for a phosphorylated form, but only a negligible amount of those specific for a non-phosphorylated form.
    Also, the phosphorylated tau-specific antibodies, but not the non-phosphorylated tau-specific ones, labeled neurofibrillary tangles isolated in the presence of sodium dodecyl sulfate (SDS) and stained both tangles and senile plaque neurites in fixed tissue sections in a very similar way to as the whole antiserum did. Taken together, these results strongly suggest that a major antigenic determinant of PHF is phosphorylated tau itself.
    Download PDF (1047K)
feedback
Top