The Journal of Biochemistry
Online ISSN : 1756-2651
Print ISSN : 0021-924X
Volume 106, Issue 3
Displaying 1-37 of 37 articles from this issue
  • Use of Two-Dimensional Homonuclear Hartmann-Hahn Spectroscopy
    Koichi Kato, Yoshifumi Nishimura, Markus Waelchli, Yoji Arata
    1989 Volume 106 Issue 3 Pages 361-364
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    A1H NMR study of a selectively deuterated mouse anti-dansyl monoclonal antibody is reported. Two-dimensional homonuclear Hartmann-Hahn (2D-HOHAHA) spectroscopy was found to be effective for establishing the connectivity between the C2-H and C4-H protons of His residues in the antibody molecule. It has been concluded that 1) even in the case of large proteins such as an antibody, HOHAHA peaks can be observed for amino acid residues that are located in a flexible environment, and 2) deuterium labeling is effective in reducing the efficiency of spin relaxation and makes it possible to increase the number of observed HOHAHA cross peaks. It was suggested that 2D-HOHAHA can also be used to obtain information concerning the flexible parts of antibody molecules.
    Download PDF (1140K)
  • Ping-Fan Rao, Toshio Takagi
    1989 Volume 106 Issue 3 Pages 365-371
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    A reassessment study was made on the viscosity behavior of SDS-protein polypeptide complexes, since information about the nature of the complexes is important in establishing the principles of SDS-polyacrylamide gel electrophoresis. Measurements were made under two conditions, i. e., in a buffer of low concentration, and in a buffer of high concentration (usually used in standard SDS-polyacrylamide gel electrophoresis). Results in the former case were not consistent with viscometric data previously reported and widely accepted [Reynolds, J. A. and Tanford, C.(1970) J. Biol. Chem. 245, 5161-5165], and indicated that the complexes did not behave as a series of pseudo-homopolymers under the former conditions. Results in the latter case indicated that the complexes behaved much more like homologous polymers, but their viscosity behavior can only be interpreted in terms of flexible chains rather than a series of rigid rods with a constant diameter and variable lengths depending on their molecular weights.
    Download PDF (2159K)
  • Inhibition by Hexokinase/Glucose or an ATP-Binding Site Blocker
    Kazuaki Tatei, Koichi Kimura, Yasumi Ohshima
    1989 Volume 106 Issue 3 Pages 372-375
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    Addition of yeast hexokinase and glucose at various time points of the pre-mRNA splicing reaction rapidly depleted ATP and inhibited further progress of the reaction, indicating that ATP is required for both the first and second steps of splicing. ATP analogues, p-fluorosulfonylbenzoy1-5'-adenosine (FSBA) and 7-chloro-4-nitrobenzo-2-oxa-1, 3-diazole (NBD-Cl), which can modify amino acids at the ATP-binding site of a protein, inactivated the splicing activity of the nuclear extract. While the inactivation by the former was irreversible, the splicing activity was complemented by a Micrococcal nucleasetreated extract. This ATP analogue (FSBA) may be a useful tool for identification of ATP-dependent splicing factors.
    Download PDF (1716K)
  • Hisashi Sasamoto, Kiyoshi Nakazawa, Kiyofumi Tsutsumi, Kenji Takase, K ...
    1989 Volume 106 Issue 3 Pages 376-382
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    Mature α-amylase of Bacillus subtilis is known to be formed from its precursor by the removal of the NH2-terminal 41 amino acid sequence (41 amino acid leader sequence). DNA fragments coding for short sequences consisting of 28 (Pro as the COOH terminus), 29 (Ala), 31 (Ala), and 33 (Ala) amino acids from the translation initiator, Met, in the leader sequence were prepared and fused in frame to the DNA encoding the mature α-amylase. The secretion activity of the 33 amino acid sequence was nearly twice as high as that of the parental 41 amino acid sequence, whereas the activity of the 31 amino acid sequence was 75%of that of the parent. In contrast, almost no secretion activity was observed with the 28 and 29 amino acid sequences. The signal peptide cleavage site of the precursor expressed from the plasmid encoding the 33 amino acid sequence was located between Ala and Leu at positions 33 and 34 and that from the 31 amino acid sequence between Thr and Ala at positions 33 and 34.The NH2-terminal amino acid from the latter corresponded to the 3rd amino acid of the mature enzyme. These results indicated that the functional signal peptide of the B. subtilis β-amylase consists of the first 33 amino acids from the initiator, Met.
    Download PDF (1935K)
  • Yuji Shimada, Akio Sugihara, Yoshio Tominaga, Taro Iizumi, Susumu Tsun ...
    1989 Volume 106 Issue 3 Pages 383-388
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    The cDNA clone of Geotrichum candidum (Geo.) lipase was isolated from the Geo. cDNA library by colony hybridization using 32P-labeled oligonucleotides corresponding to a partial amino acid sequence of this enzyme. The nucleotide sequence of the cDNA determined by the dideoxy chain terminating method included some partial amino acid sequences determined by Edman degradation, and the overall amino acid composition deduced from the cDNA coincided with that from amino acid analysis of this protein. The cloned cDNA coded a protein of 554 amino acids and a hydrophobic signal sequence of 19 amino acids. Geo. lipase contained the-Gly-X-Ser-X-Gly-sequence which is believed to form part of the interfacial lipid recognition site.
    Download PDF (1933K)
  • Shonen Yoshida, Katsuyuki Tamai, Hayato Umekawa, Motoshi Suzuki, Kiyoh ...
    1989 Volume 106 Issue 3 Pages 389-395
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    A DNA polymerase α-primase complex, which had been purified by means of immunoaffinity column chromatography, showed little activity in a reaction mixture composed of Tris-HCl buffer, but showed full activity in potassium phosphate buffer. It was found that potassium ion is required for the reaction by the immunoaffinity-purified enzyme. On the other hand, the DNA polymerase α purified by the orthodox biochemical method showed full activity in both buffer systems. A protein factor, which could restore the activity of immunoaffinity-purified DNA polymerase α-primase complex in the potassium-free reaction mixture, was separated from biochemically purified DNA polymerase α. The factor, designated as factor T, was stable to heat up to 70°C, but was sensitive to trypsin. It sedimented at about 4S through a glycerol gradient. SDS-polyacrylamide gel electrophoresis revealed two polypeptide bands at 56 and 54 kDa. By immunoprecipitation, the factor T was shown to be physically associated with DNA polymerase α-primase complex. The stimulation was also observed with poly [d (A-T)], primed M13 DNA, and heat-denatured DNA.
    Download PDF (2286K)
  • A Proton Nuclear Magnetic Resonance Study
    Toshiya Endo, Ichio Shimada, David Roise, Fuyuhiko Inagaki
    1989 Volume 106 Issue 3 Pages 396-400
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    Two-dimensional proton nuclear magnetic resonance (NMR) spectra of a synthetic peptide (p25) corresponding to the amino-terminus of the yeast mitochondrial cytochrome oxidase subunit IV precursor protein have been analyzed. Sequence-specific resonance assignments of the peptide have been made in the presence of micelles of a phospholipid analog, perdeuterated dodecylphosphocholine (DPC), with the aid of such techniques as HOHAHA, DQF-COSY, and NOESY. The interresidue nuclear Overhauser effects (NOEs) indicate that the N-terminal half of p25 (53-F11) takes a helical structure while the C-terminal half does not take a regular secondary structure. Addition of DPC to the solution of p25 induced chemical shift changes only of the resonances from the residues in the N-terminal half, suggesting that the N-terminal half of p25 is directly involved in binding to DPC. The induced helical structure in the N-terminal half at a lipid-water interface may be important in the ability of this presequence to direct a “passenger” protein into mitochondria.
    Download PDF (996K)
  • Takashi Kawasaki, Michiki Kasai
    1989 Volume 106 Issue 3 Pages 401-405
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    Ca2+ channels of isolated sarcoplasmic reticulum were incorporated into a planar lipid bilayer and their pharmacological properties were studied. The results show that the channel is a Ca2+-induced Ca2+ release channel like that observed in skinned muscle fibers and isolated vesicles.(i) The open channel probability was increased by the addition of micromolar amounts of Ca2+ to the cis (myoplasmic) side and further increased by millimolar ATP.(ii) The channel was closed by millimolar Mg2+ and micromolar ruthenium red. We found that two disulfonic stilbene derivatives, 4, 4'-diisothiocyanostilbene-2, 2'-disulfonic acid (DIDS) and 4-acetoamido-4'-isothiocyanostilbene-2, 2'-disulfonic acid (SITS), when added to the cis side open the channel and lock it irreversibly at open without changing the single channel conductance. Ca2+ efflux from SR vesicles was also enhanced by SITS and DIDS, as monitored by a tracer assay. Further, Ag+ activated the channel transiently. These results suggest that certain amino and SH residues play important roles in gating the Ca2+ channel.
    Download PDF (1027K)
  • Comparative Physiology of Hemoglobin
    Antonio Tsuneshige, Kiyohiro Imai, Hiroshi Hori, Itiro Tyuma, Toshio G ...
    1989 Volume 106 Issue 3 Pages 406-417
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    The physicochemical properties of giant hemoglobin (Hb) of the marine polychaete Perinereis aibuhitensis were extensively studied and the following results were obtained.(1) Light absorption spectra of the oxy, deoxy, CO, met, and cyanomet derivatives were similar to those for human Hb, except for a somewhat peculiar shape and pH-dependence of the met derivative, and high absorbance values around 277 nm for all these derivatives of Perinereis Hb. Abnormal pH dependence for the met derivative was confirmed by powder electron paramagnetic resonance (EPR) spectroscopy, which revealed that a water molecule does not coordinate to the heme iron as a sixth ligand. The high absorption around 277 nm is indicative of the existence of some non-heme polypeptide chains and/or a high content of aromatic residues in the molecule.(2) UV difference and derivative spectra revealed oxygenation-induced conformational changes in the protein moiety that are related to the degree of cooperativity.(3) The EPR spectrum for the nitrosyl derivative showed wellresolved triplet-triplet splittings due to 14N, indicating that the proximal residue is probably a histidine.(4) The oxygen affinity and cooperativity of this Hb were pHdependent. Mg2+ markedly increased the oxygen affinity, the Bohr effect, and the cooperativity, which was maximal at physiological pH. CO2 and anions such as 2, 3-diphosphoglycerate and inositol hexaphosphate had no effect on the oxygenation properties. Thus, different from vertebrate Hb, the oxygen-binding properties of this Hb are regulated by divalent cations which bind preferentially to the oxy form. The low temperature-dependence of oxygen affinity observed for this Hb is a sign of adaptation to the environment by this poikilothermic organism.(5) By using a graphic method, the minimal functional unit that preserves the full cooperativity (allosteric unit) was inferred to be the one containing 6 heme groups and its significance is discussed in connection with the structural hierarchy of the molecule.
    Download PDF (3000K)
  • Nobuo Makino
    1989 Volume 106 Issue 3 Pages 418-422
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    Hemocyanin was prepared from an Asian horseshoe crab, Tachypleus gigas. The hemocyanin was found to be similar to Limulus hemocyanin in the size of native molecules (48-mer) and dissociation under nonphysiological conditions. It also showed the reverse Bohr effect. The 02 affinity of the dissociated monomer was higher than that of the native molecule. Equilibrium O2 binding to T. gigas hemocyanin was studied with special attention to the effect of inorganic ions. Neutral salts decreased the O2 affinity of the associated hemocyanin. In the presence of CaCl2 the strength of the effect was in the order of Na+>Cs+≈K+ for the series of chlorides, and Br-≈Cl-> SO42- for the series of Na+ salts. A high concentration of CaCl2 (50-500mM) considerably increased the Hill coefficient. The O2 binding data obtained under various ionic conditions were analyzed by model fitting. The two-state concerted model could be fitted to the data, if the ligand affinity of the states was allowed to vary. Statistical tests of the fitting showed that the hexameric structure can be regarded as the functional unit under physiological conditions.
    Download PDF (1154K)
  • Nobuo Makino
    1989 Volume 106 Issue 3 Pages 423-429
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    Six subunits (I to VI) were isolated from hemocyanin of an Asian horseshoe crab, Tachypleus gigas, by anion exchange chromatography of the dissociated hemocyanin. The subunit preparations were nearly homogeneous as judged by alkaline electrophoresis, but they still showed the presence of isoproteins in isoelectric focusing. The subunits were reassembled (in 10mM CaCl2 at pH 7.5) and tested for restoration of the cooperativity in O2 binding. The reassembly of the subunits gave equilibrium mixtures of the monomer and hexamer with small amounts of larger molecules. Homogeneous and heterogeneous hexamers were prepared by reassembling a single kind or two kinds of subunits, followed by isolation of the hexamer fraction by gel filtration. Among the homohexamers, only the subunit V hexamer showed cooperativity in O2 binding with the Hill coefficient of 1.6. Among the heterohexamers the subunit I/V hybrid was most noteworthy, showing a Hill coefficient (1.7) higher than that of any other heterohexamer examined. It was concluded that there are specific interactions between the subunits I and V. It is suggested that their interactions are important for the cooperativity in the native hemocyanin.
    Download PDF (2391K)
  • Yasuyuki Ishizuka, Shinji Iizumi, Fumiaki Akiyama, Hitoshi Hori, Roger ...
    1989 Volume 106 Issue 3 Pages 430-435
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    Monoclonal antibodies were raised against a synthetic peptide (43 amino acid residues) that corresponds to the complete profragment of human prorenin. Seven monoclonal antibodies were chosen for further characterization. Two antibodies, 2-X-C1 and 4-X-E1, reacted with the middle region and C-terminus of the profragment and were isotyped IgG1. The affinity constants of these antibodies against the human profragment were 7.6×108 and 3.0×107 M-1, respectively. Immunoaffinity columns containing the antibodies 2-X-C1 and 4-X-E1, respectively, were used for the characterization of active prorenin in human plasma. This active prorenin strongly bound to the 4-X-E1 column and eluted as two separate peaks which corresponded to fully and partially active prorenin, respectively. The partially active prorenin had higher activity with a small substrate, tridecapeptide, than with a large one, angiotensinogen, although the fully active prorenin had the same renin activity irrespective of the size of the substrate. These data suggest that new forms of prorenin, active prorenin, exist in human plasma and that their active sites are completely or partially exposed to the substrates. Moreover, the active prorenin in plasma was found not only in human but also in all tested mammalians. Cross-reactivity among the profragments of mammalian plasma prorenins can be explained by conservation of the amino acid sequence (epitope) of the combining site.
    Download PDF (1907K)
  • Ryuji Furuta, Junichi Yamagishi, Hirotada Kotani, Fumiko Sakamoto, Tos ...
    1989 Volume 106 Issue 3 Pages 436-441
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    A putative mature human neutrophil chemotactic factor (NCF) corresponding to the C-terminal 72 amino acids of its precursor was directly produced in Escherichia coli by recombinant DNA technology. Human NCF was present in both the soluble and insoluble protein fractions of the homogenate of host cells, and it was partially purified as a water-soluble polypeptide from both fractions, separately. The partially purified NCF preparation was highly purified to an endotoxin-free homogeneous polypeptide by means of CM-Sepharose CL-6B column chromatography and gel filtration on Toyopearl HW-55. No difference between the human NCF preparations purified from both starting materials could be found concerning purity, primary structure, solubility, molecular weight, and chemotactic activity for human neutrophils. The amino acid sequence of recombinant human NCF was identical to the sequence deduced from the cDNA sequence. A methionine residue due to the translation initiation codon was removed. Recombinant human NCF was found to be biologically active and to exhibit chemotactic activity for human neutrophils in vitro and cause a neutrophil infiltration in vivo in mice.
    Download PDF (2263K)
  • Properties and N-Terminal Amino Acid Sequence
    Koichiro Takeuchi, Takao Shimizu, Nobuya Ohishi, Yousuke Seyama, Fumim ...
    1989 Volume 106 Issue 3 Pages 442-445
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    Angiotensin-converting enzyme from the human lung was purified to apparent homogeneity, using high-performance liquid chromatography following trypsin treatment of the detergent-extract. A 1, 750-fold purification was achieved with a 26% yield. The specific activity of the enzyme was 105 units per mg protein with the substrate hippuryl-L-histidyl-L-leucine (HHL) at 37°C, and the Km value for HHL was 1.9mM. The molecular weight was estimate to be 170, 000 by sodium dodecyl sulfate gel electrophoresis, and the isoelectric point was about 4.8, by chromatofocusing. The N-terminal amino acid sequence was (NH2)-X-X-Pro-Gly-Leu-Glu-Pro-Gly-X-Phe-Ser-Ala-Arg-Glu-Ala-Gly-Ala. This is highly homologous to the corresponding sequences of the enzymes from bovine and rabbit lung and from pig, bovine, and mouse kidney, but significantly different from that of the human kidney enzyme.
    Download PDF (1085K)
  • Yohtalou Tashima, Miyuki Terui, Hideaki Itoh, Hideo Mizunuma, Ryoji Ko ...
    1989 Volume 106 Issue 3 Pages 446-454
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    The [3H] corticosterone binders from rat brain and kidney were characterized by binding affinity and chromatographies, and compared with the binders for [3H] aldosterone and [3H] triamcinolone acetonide. Corticosterone-binding globulin-like molecules at very high concentrations in crude extracts were completely eliminated by a DEAE-gel adsorption procedure.[3H] Aldosterone binder in the renal, DEAE-treated fraction was recovered in a single peak by gcl-filtration chromatography and by ultracentrifugation in linear sucrose gradients, independent of hormone-binding and tungstate, a stabilizer of the binder. The Stokes' radius and sedimentation coefficient of the renal aldosterone binder were 6.6 nm and 9.3S, respectively, indicating an apparent molecular weight of 263, 000. Corticosteronepreferring binder also existed in the DEAE-treated fraction. Both aldosterone and corticosterone binders were found in the brain and kidney preparations. Comparison among the binders showed identical values of Stokes' radius and elution pattern from DEAE Toyopearl in a linear salt gradient regardless of the organ and the hormones. Scatchard analyses of [3H] aldosterone and [3H] corticosterone binding showed for each ligand only one group of high-affinity sites with the equivalent dissociation constants, 4-7 nM. The orders of steroids in competing for the two high-affinity sites were equivalent: corticosterone≥aldosterone>>triamcinolone acetonide, and that for the triamcinolone acetonide binding was triamcinolone acetonide>>aldosterone≥corticosterone. Hydroxyapatite column chromatography separated the aldosterone and corticosterone binders from the triamcinolone acetonide binder, but not the aldosterone binder from the corticosterone binder. It is concluded that aldosterone and corticosterone binders distinct from triamcinolone acetonide binder exist in rat brain and kidney. The corticosterone binder is very similar or identical to the aldosterone binder.
    Download PDF (2282K)
  • Minoru Hoshimaru, Yasushi Fujio, Kenji Sobue, Tetsuo Sugimoto, Shigeta ...
    1989 Volume 106 Issue 3 Pages 455-459
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    In a previous study, we identified a new mammalian myosin heavy chain, termed myosin I heavy chain-like protein (MIHC), by molecular cloning of a bovine intestinal cDNA clone. In this investigation, we examined the relationship between MIHC and the 110-kDa intestinal brush-border protein, which possesses a myosin-like ATPase activity. We raised antibodies against a chemically synthesized oligopeptide representing a part of the MIHC sequence. These antibodies reacted specifically in immunoblots with the 110-kDa protein in both purified 110-kDa protein-calmodulin complex and crude microvillar protein extracts. Staining of tissue sections with these antibodies was specifically localized to the brushborder microvilli of small intestines, indicating an identical cellular localization for both MIHC and the 110-kDa protein. Furthermore, analysis of the MIHC sequence revealed two putative calmodulin-binding sites, which is consistent with the fact that the 110-kDa protein forms a complex with calmodulin. These results strongly support the conclusion that MIHC is identical to the 110-kDa protein and suggest that not only the conventional myosin system but also the MIHC (110-kDa protein)-calmodulin complex may play an important role in ATP-dependent and Ca2+-induced brush-border contraction.
    Download PDF (2866K)
  • Pyruvate Aminotransferase of Rat Liver Mitochondria Expressed in Escherichia coli
    Toshiaki Oda, Hiroaki Miyajima, Yoshie Suzuki, Takeshi Ito, Sadaki Yok ...
    1989 Volume 106 Issue 3 Pages 460-467
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    In the previous study (Oda, T., et al.(1985) Eur. J. Biochem. 150, 415-421), we isolated a cDNA clone which expressed in Escherichia coli a specific size of product having the activity of rat liver serine: pyruvate aminotransferase (SPTm). This specific product (SPT10) was purified to homogeneity through three different column chromatographies. The amino acid composition and N-terminal amino acid sequence of the purified enzyme agreed with those predicted from the nucleotide sequence of cDNA and showed that SPT10 consists of the whole amino acid sequence of mature SPTm and several extra amino acid residues at the N-terminus. The catalytic and physical properties of SPT10, such as substrate specificity, Km for α-keto acids, electric charge, and quaternary structure, were all very similar to those of SPTm. Using several cDNA clones which lack a 5'-terminal sequence corresponding to a portion of the N-terminal amino acid sequence of SPTm, we examined the expression profile of the specific product in bacteria transformed with each cDNA clone. The products encoded by these cDNAs were segregated into inclusion bodies and were neither catalytically active nor easily solubilized by sonication. In contrast, the inclusion bodies were not formed in the bacteria transformed with the cDNA clone for SPT10.
    Download PDF (2041K)
  • Yasushi Kikuta, Emi Kusunose, Shyuichiro Matsubara, Yoshihiko Funae, S ...
    1989 Volume 106 Issue 3 Pages 468-473
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    Prostaglandin ω-hydroxylase, designated as cytochrome P-450LPGω (P-450LPGω), has been purified, to a specific content of 15 nmol of cytochrome P-450/mg of protein, from liver microsomes of pregnant rabbits. The purified P-450LPGω was found to be homogeneous on sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) and to have an apprent molecular weight of 52, 000. The enzyme showed a maximum at 450 nm in the carbon monoxide (CO)-difference spectrum for its reduced form. This cytochrome P-450 efficiently catalyzed the ω-hydroxylation of prostaglandin E1 (PGE1), prostaglandin E2 (PGE2), prostaglandin D2 (PGD2), prostaglandin F (PGF), prostaglandin A1 (PGA1), and prostaglandin A2 (PGA2), as well as the ω-and (ω-1)-hydroxylation of myristate and palmitate, in a reconstituted system containing cytochrome P-450, NADPH-cytochrome P-450 reductase, phospholipid, and cytochrome b5. Various monovalent and divalent cations further stimulated these reactions in the presence of cytochrome b5. In addition, the reactions were also markedly enhanced by various organic solvents, such as ethanol and acetone. This cytochrome P-450 showed no detectable activity toward several xenobiotics tested. P-450LPGω was very similar or identical to the pulmonary prostaglandin whydroxylase (P-450p-2)(Yamamoto, S., Kusunose, E., Ogita, K., Kaku, M., Ichihara, K., & Kusunose, M.(1984) J. Biochem. 96, 593-603) in its molecular weight, absorption spectra, catalytic activity, peptide mapping pattern, and N-terminal amino acid sequence. However, P-450LPGω was more unstable than P-450P-2 on storage. In sharp contrast to P-450p-2, P-450LPGω was not induced by progesterone.
    Download PDF (2668K)
  • Purification and Characterization
    Tatsuo Kurihara, Mitsuyoshi Ueda, Atsuo Tanaka
    1989 Volume 106 Issue 3 Pages 474-478
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    Acetoacetyl-CoA thiolase (Thiolase I) and 3-ketoacyl-CoA thiolase (Thiolase III) found in peroxisomes of an n-alkane-utilizing yeast, Candida tropicalis pK 233, were each purified to homogeneity by successive column chromatographies. Thiolase I was composed of six identical subunits whose molecular masses were 41, 000 Da, and Thiolase III was a homodimer composed of 43, 000 Da subunits. The results of limited proteolysis of the respective thiolases indicated that they were quite different in peptide components. Furthermore, these enzymes were immunochemically distinguishable. The kinetic studies showed that the substrates with long chains were degraded exclusively by Thiolase III, while acetoacetyl-CoA was degraded preferentially by Thiolase I. Thus, in the yeast, the complete degradation of fatty acids is suggested to be carried out efficiently in peroxisomes.
    Download PDF (2241K)
  • Kuniaki Mukai, Sadao Wakabayashi, Hiroshi Matsubara
    1989 Volume 106 Issue 3 Pages 479-482
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    The amino acid sequence of the mature protein of Euglena gracilis cytochrome c1 was determined by sequencing of its cDNA. A cDNA expression library was constructed from Euglena poly (A)+ RNA in phage λgtll and screened with an antiserum raised against cytochrome c polypeptide isolated from purified E. gracilis complex III. An isolated cDNA clone consisted of 872 base pairs and encoded the mature protein with 243 amino acids. The deduced amino acid sequence contained the unusual heme binding sequence-Phe-Ala-Pro-Cys-His-(Mukai, K. et al.(1989) Eur. J. Biochem. 178, 649-656) instead of the typical sequence, -Cys-X-Y-Cys-His-, commonly found in C-type cytochromes. Comparison of the sequence with those of several other cytochromes c1 revealed that Euglena cytochrome c1 conserved the residues probably ligating heme-iron, those supposed to interact with cytochrome c and regions anchoring the mitochondrial inner membrane.
    Download PDF (996K)
  • Nobuko Kawasaki, Toshisuke Kawasaki, Ikuo Yamashina
    1989 Volume 106 Issue 3 Pages 483-489
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    Serum mannan-binding protein (MBP), which is a lectin specific for mannose and N-acetylglucosamine and is known to activate complement via the classical pathway, has been revealed to have a complement-dependent bactericidal activity, as tested on rough strains of Escherichia coli, K-12 and B. The bacteria, which had been sensitized with purified human serum MBP in the presence of Ca2+, followed by incubation with guinea pig complement, showed a marked decrease of colony forming ability compared with those not sensitized with the lectin. The bactericidal effect depended on the concentrations of the lectin and complement. The C4-dependency of the reaction indicated that the complementdependent bactericidal action by MBP is expressed through the classical pathway. The bacteria were aggregated by the lectin. Scatchard plot analysis of 125I-labeled MBP binding to the bacteria showed that the dissociation constant (Kd) and the maximum binding capacity were 6×10-9M and 30, 000 molecules of MBP per cell, respectively. The binding was inhibited by mannose, N-acetylglucosamine, N-acetylmannosamine, L-fucose, mannoheptulose, and sedoheptulose, suggesting that MBP recognized L-glycero-D-manno-heptose and N-acetylglucosamine constituting the core oligosaccharide of the E. coli K-12 cell wall, and L-glycero-D-manno-heptose for E. coli B. These findings suggest the physiological significance of the serum lectin in host defense, being consistent with the avirulence of E. coli rough strains in mammals.
    Download PDF (2535K)
  • Identification of the Protein as Type IV Collagen
    Akira Murakami, Hiroshi Konomi, Naoya Itokazu, Masataka Arima, Norio S ...
    1989 Volume 106 Issue 3 Pages 490-494
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    Production of an unusual collagenous protein was observed in culture of dermal fibroblasts from four patients with Marfan syndrome. The apparent molecular weight of the protein was about 185kDa after reduction with 2-mercaptoethanol and 175kDa after limited pepsin treatment. The 185kDa protein was susceptible to the bacterial collagenase but resistant to the animal collagenase. Immunoprecipitation revealed the specific interaction of the pepsin-treated 175kDa collagenous protein with monoclonal and polyclonal antibodies to human type IV collagen. From the patterns of CNBr peptide mapping the 185kDa band was identified as αl (IV) chain. Type IV collagen in the skin is generally considered to be of non-fibroblastic origin. However, in “diseased” condition, dermal fibroblasts might produce type IV collagen. The clinical manifestation in relation to production of type IV collagen by cultured skin fibroblasts from Marfan patients is discussed.
    Download PDF (3256K)
  • Keiji Tanaka, Tetsuro Yoshimura, Akira Ichihara
    1989 Volume 106 Issue 3 Pages 495-500
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    Previously, we reported that proteasomes (large multi-protease complexes) are present in a latent state in a variety of eukaryotic cells, and can be activated by treatment with various compounds such as sodium dodecyl sulfate (SDS) or poly-lysine (Tanaka et al.(1988) J. Biol. Chem. 263, 16209-16217). In the present study, the mechanism of activation of latent proteasomes by SDS was examined. Latent proteasomes were greatly activated by addition of low concentrations of 0.04 to 0.08% SDS in the presence of substrate. This activation appeared to be reversible, because SDS-activated proteasomes returned to a latent state when the concentration of SDS was reduced by dilution. In contrast, in the absence of substrate, latent proteasomes lost their activity almost completely in an irreversible fashion within a few minutes during treatment with SDS at either 0 or 37°C. Interestingly, SDS-treated proteasomes were markedly protected against this rapid inactivation by either a peptide or protein substrate. Moreover, removal of the substrate after activation of proteasomes caused their rapid irreversible inactivation. These results indicate that the substrate is necessary for reversible activation of latent proteasomes by SDS. This effect of substrate is presumably important in regulation of intracellular protein breakdown by activated proteasomes in eukaryotic cells.
    Download PDF (1380K)
  • Mizuho Une, Nobuko Kisaka, Michiko Yoshii, Takahiko Hoshita
    1989 Volume 106 Issue 3 Pages 501-504
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    In order to confirm the occurrence of 3α, 6α, 7α, 12α-tetrahydroxy-5β-cholestanoic acid in Zellweger's syndrome, the nature of tetrahydroxycholestanoic acids present in a patient with this disease was studied. Urinary bile acids were extracted with a Sep-pak C18 cartridge and methylated after alkaline hydrolysis. The methyl esters were purified by silica gel column chromatography, and the methyl tetrahydroxycholestanoate fraction was analyzed by gas liquid chromatography-mass spectrometry. Along with already known side chain hydroxylated derivatives of 3α, 7α, 12α-trihydroxy-5β-cholestanoic acid, 3α, 7α, 12α, 24- and 3α, 7α, 12α, 26-tetrahydroxy-5β-cholestanoic acids, three nuclear hydroxylated derivatives of 3α, 7α, 12α-trihydroxy-5β-cholestanoic acid were found. One of them was identified as 3α, 6α, 7α, 12α-tetrahydroxy-5β-cholestanoic acid by direct comparison with the authentic standard which was chemically synthesized from 3α, 6α, 7α, 12α-tetrahydroxy-5β-cholanoic acid by side chain elongation.
    Download PDF (992K)
  • Masakazu Menju, Shoji Tajima, Akira Yamamoto
    1989 Volume 106 Issue 3 Pages 505-510
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    The human monocyte-like cell line, THP-1, differentiated into macrophage-like cells on the addition of a phorbol ester, 12-O-tetradecanoyl-phorbol-13-acetate. During the course of differentiation of THP-1 cells, the level of transcripts of the apolipoprotein E gene increased. Apolipoprotein E mRNA increased by more than a hundred times compared to the level prior to differentiation. The apolipoprotein E mRNA reached the maximal level on day 2 after the addition of the phorbol ester and then gradually decreased. After the level had decreased to half the maximal value on day 4 it remained constant. The time course of apolipoprotein E secretion, which showed a peak on day 2, was parallel to that of apolipoprotein E protein synthesis. Furthermore, the time course of apolipoprotein E protein synthesis showed a similar profile to that of the apolipoprotein E transcript level. This indicates that the induction of apolipoprotein E expression by the phorbol ester is due mainly to the increase in the number of transcripts. The synthesis of apolipoprotein E protein was reduced by about 60% on treatment of the differentiated THP-1 cells with 5μg/ml of lipopolysaccharide. The presence of 5μg/ml of lipopolysaccharide in the medium reduced the level of apolipoprotein E mRNA by about 50%. Thus the reduction in protein synthesis was mainly explained by the decrease in the level of apolipoprotein E transcripts. This reduction in the mRNA level caused by lipopolysaccharide was not mediated by the tumor necrosis factor or interleukin 1, which are known to reduce the transcriptional and post-transcriptional activity of lipoprotein lipase in adipocytes, respectively.
    Download PDF (3258K)
  • Hiroshi Tokumitsu, Masatoshi Hagiwara, Koji Onoda, Hiroyoshi Hidaka
    1989 Volume 106 Issue 3 Pages 511-514
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    We prepared monoclonal antibodies directed against chicken gizzard myosin light chain kinase (MLCK) and used them to study the contractile system of aortic smooth muscle. One monoclonal antibody, MM13, dose dependently inhibited actomyosin superprecipitation of bovine aortic smooth muscle, in accord with the suppression of 20 kDa myosin light chain phosphorylation by endogenous kinase. Immunoblotting analysis demonstrated that MM13 cross-reacted with the 150, 000 Mr peptide of bovine aortic actomyosin preparation. The bovine aortic MLCK was purified approximately 2, 400-fold to apparent homogeneity by three steps of column chromatography. The purified enzyme has a molecular weight of 150, 000 and a slower mobility than chicken gizzard MLCK (130, 000 Mr), as determined by SDS-polyacrylamide gel electrophoresis. MM13 also cross-reacted with purified bovine aortic MLCK and inhibited the kinase activity, in vitro. We interpret these findings to mean that binding of the anti-gizzard MLCK monoclonal antibody directly to aortic smooth muscle MLCK (150, 000 Mr) decreases the phosphorylation of the 20 kDa myosin light chain, thus suppressing the aortic smooth muscle myosin-actin interaction.
    Download PDF (2149K)
  • Nobuyoshi Nakajima, Kaoru Nakamura, Nobuyoshi Esaki, Hidehiko Tanaka, ...
    1989 Volume 106 Issue 3 Pages 515-517
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    We have established a simple procedure for the in situ analysis of stereospecificity of an NAD (P)-dependent dehydrogenase for C-4 hydrogen transfer of NAD (P) H by means of glutamate racemase [EC 5. 1. 13] and glutamate dehydrogenase [EC 1. 4. 1. 3]. Glutamate racemase inherently catalyzes the exchange of α-H of glutamate with 2H during racemization in 2H2O. When the reactions of glutamate racemase and glutamate dehydrogenase, which is pro-S specific for the C4-H transfer of NAD (P) H, are coupled in 2H2O, [4S-2H]-NAD (P) H is exclusively produced. Therefore, if 1H is fully retained at C-4 of NAD (P) +after incubation of a reaction mixture containing both the enzymes and a dehydrogenase to betested, the stereospecificity of the dehydrogenase is the same as that of glutamate dehydrogenase. When the C4-H of NAD (P) +is exchanged with 2H, the enzyme to be examined is different from glutamate dehydrogenase in stereospecificity. Thus, we can readily determine the stereospecificity by 1H-NMR measurement of NAD (P) +without isolation of the coenzymes and products.
    Download PDF (406K)
  • Keizo Teshima, Yutaka Kitagawa, Yuji Samejima, Saju Kawauchi, Shinobu ...
    1989 Volume 106 Issue 3 Pages 518-527
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    Phospholipases A2 are classsfied into two groups, I and II, according to differences in the polypeptide-chain length and the intramolecular-disulfide bondings. The effects of Ca2+on the kinetic parameters for the hydrolysis of monodispersed and micellar phosphatidylcholines, catalyzed by a cobra (Naja naja atra) enzyme (Group I) and by mamushi (Agkistrodon halys blomhoffii) and habu (Trimeresurus flavoviridis) enzymes (Group II), were studied by the pH-stat assay method at 25°C, pH 8.0-8.2, and ionic strength 0.1-0.2. The results were compared with those reported for other Group I and II enzymes. The Ca2+binding was clearly shown to be essential for the catalysis of all the phospholipases A2. However, the substrate binding to Group I enzymes was found to be independent of the Ca2+binding. On the other hand, the substrate binding to Group II enzymes was facilitated more than 10 times by the binding of Ca2+to the enzymes. This was interpreted in terms of conformation changes of the peptide loop of residues 26 to 44 accompanying the Ca2+binding. The latter result, but not the former, seems compatible with the hypothesis for interpreting the catalytic mechanism of phospholipases A2 that an intermediate complex should be stabilized by the coordination of the bound Ca2+ ion with the phosphoryl group and the carbonyl oxygen atom of the ester bond at the sn-2 position of the bound substrate molecule [Verheij et al.(1980) Biochemistry 19, 743-750 and (1981) Rev. Physiol. Biochem. Pharmacol. 91, 91-203]. According to the similarity in the primary and tertiary structures of the active sites of both types of enzymes [Renetseder et al.(1985) J. Biol Chem. 260, 11627-11634], it is supposed that similar intermediate complexes may occur even for Group I enzymes, at least in the transition state of the productive complexes.
    Download PDF (2422K)
  • Purification and Characterization
    Kenichi Nakano, Yasuko Omura, Mitsao Tagaya, Toshio Fukui
    1989 Volume 106 Issue 3 Pages 528-532
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    UDP-glucose pyrophosphorylase from potato tuber was purified 243-fold to a nearly homogeneous state with a recovery of 30%. The purified enzyme utilized UDP-glucose, but not ADP-glucose, as the substrate, and was not activated by 3-phosphoglyceric acid. Product inhibition studies revealed the sequential binding of UDP-glucose and MgPP1 and the sequential release of glucose-1-phosphate and MgUTP, in this order. Analyses of the effects of Mg2+ on the enzyme activity suggest that the MgPP1 and MgUTP complexes are the actual substrates for the enzyme reaction, and that free UTP acts as an inhibitor. The enzyme exists probably as the monomer of an approximately 50-kDa polypeptide with a blocked amino terminus. For structural comparison, 29 peptides isolated from a tryptic digest of the S-carboxymethylated enzyme were sequenced. The results show that the potato tuber enzyme is homologous to UDP-glucose pyrophosphorylase from slime mold, but not to ADP-glucose pyrophosphorylase from Escherichia coli, and provide structural evidence that UDP-glucose and ADP-glucose pyrophosphorylase are two different protein entities.
    Download PDF (1834K)
  • Toshirou Nishida, Toru Inoue, Wataru Kamiike, Yasunaru Kawashima, Kuni ...
    1989 Volume 106 Issue 3 Pages 533-538
    Published: September 01, 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    During anoxic incubation, depletion of mitochondrial ATP was followed by release of Ca2+with concomitant increase in the rate of state 4 respiration due to disruption of the diffusion barrier against protons. The external addition of ATP and its non-metabolizable analog, β, γ-methylene adenosine 5'-triphosphate, prevented both the release of Ca2+ and increase in the rate of state 4 respiration. Addition of EGTA, which did not prevent release of the ion, resulted in little increase in the respiration rate. Addition of an inhibitor of mitochondrial phospholipase A2, such as quinacrine, dibucaine, or chlorpromazine, also prevented increase in the respiration rate without affecting Ca2+ release from mitochondria during anoxic incubation. Non-esterified polyunsaturated fatty acids were also found to be liberated from anoxic mitochondria. External addition of the ATP-analog, EGTA, and inhibitors of phospholipase A2 suppressed the liberation of non-esterified polyunsaturated fatty acids. Melittin and Ca2+, which activate phospholipase A2 increased the rate of state 4 respiration and the liberation of fatty acids. These findings support the hypothesis proposed previously that the following sequence changes occurs in mitochondria during anoxia; depletion of ATP, liberation of free calcium from mitochondria, and disruption of the diffusion barrier against H+ of the inner membrane. The results also indicate another event; activation of phospholipase A2 by release Ca2+ which results in H+ leakiness of the inner membrane.
    Download PDF (1805K)
  • 1989 Volume 106 Issue 3 Pages 539a
    Published: 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    Download PDF (73K)
  • 1989 Volume 106 Issue 3 Pages 539b
    Published: 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    Download PDF (73K)
  • 1989 Volume 106 Issue 3 Pages 539c
    Published: 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    Download PDF (73K)
  • 1989 Volume 106 Issue 3 Pages 539d
    Published: 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    Download PDF (73K)
  • 1989 Volume 106 Issue 3 Pages 539e
    Published: 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    Download PDF (73K)
  • 1989 Volume 106 Issue 3 Pages 539f
    Published: 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    Download PDF (73K)
  • 1989 Volume 106 Issue 3 Pages 539g
    Published: 1989
    Released on J-STAGE: January 25, 2011
    JOURNAL FREE ACCESS
    Download PDF (73K)
feedback
Top