The Journal of Biochemistry
Online ISSN : 1756-2651
Print ISSN : 0021-924X
97 巻, 2 号
選択された号の論文の37件中1~37を表示しています
  • Reiko URADE, Yukako HAYASHI, Makoto KITO
    1985 年 97 巻 2 号 p. 391-398
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    When lipids were removed from the culture medium, growth of V 79 cells ceased. Supplementation with cis-octadecenoic acids satisfied the requirement for lipids by V 79 cells. After starvation of the exogenous lipids by the shift-down of the medium to lipid-free medium, the content of octadecenoic acid in phospholipids increased more slowly in V 79 cells than in V 79-R cells, which can grow in the lipid-starved medium. The incorporation of [14C]acetic acid into monoenoic fatty acids and phospholipid molecular species containing monoenoic fatty acids in V 79 cells was lower than that in V 79-R cells. The reduced formation of monoenoic fatty acids was shown to be due to deficiency in the stimulation of activity of stearoyl-CoA desaturase which is a key enzyme to convert saturated fatty acids to monoenoic fatty acids.
  • Ikuo MATSUI, Kunio OISHI, Koichi KANAYA, Norio BABA
    1985 年 97 巻 2 号 p. 399-408
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    The morphology of an L-fucose specific lectin, SEL 100-2, from a Streptomyces sp. was studied. Electron microscopic observation showed that purified SFL 100-2 preparation consisted of particles homogeneous in size. The diameter was 25 nm. The digitized images of these particles had 2-fold rotation symmetry. The sedimentation coefficient (s020, W) was determined to be 20.6 S. The particle weight and the Stokes radius were calculated to be 8.0×105 daltons and 94 Å, respectively, by three independent methods, i.e., gel filtration, sedimentation equilibrium and velocity measurements. The frictional ratio (ƒ/ƒmin) was estimated to be 1.53. These values are quite similar to those of human α2-macroglobulin. 125I-Labeled peptide mapping indicated that these particles were built up of about twelve identical subunits (Mr=68, 000). The size of SFL 100-2 in culture broth was found to be the same as that of the particles in the purified preparations. The shape and other properties of SFL 100-2 are discussed and compared with those of the tail of λ phage and type 1 pili of Escherichia coli, whose amino acid compositions were quite similar to that of SFL 100-2 and also those of L-fucose specific plant lectins.
  • Yoshiko BANNO, Yoshinori NOZAWA
    1985 年 97 巻 2 号 p. 409-418
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    A large amount of lysosomal acid hydrolases was released into the medium by Tetrahymena pyriformis strain W during growth. An extracellular lysosomal acid α-glucosidase has been purified 500-fold with a 41% yield to homogeneity, as judged by polyacrylamide gel electrophoresis. It was found to be a glycoprotein and to consist of a single 110, 000-dalton polypeptide chain. The carbohydrate content of the α-glucosidase was equivalent to 2.8% of the total protein content, and the oligosaccharide moiety was composed of mannose and N-acetylglucosamine in a molar ratio of 6.7:2. The optimal pHs for hydrolysis of maltose and p-nitrophenyl-α-glucopyranoside were both pH 4.0. The Km values determined with p-nitrophenyl-α-glucopyranoside, maltose, isomaltose, and glycogen were 1.1mM, 2.5mM, 33.0mM, and 18.5mg/ml, respectively. This purified enzyme appears to have α-1, 6-glucosidase as well as α-1, 4-glucosidase activity. Turanose has a noncompetitive inhibitory effect on the hydrolysis of maltose. The antibody raised against Tetrahymena acid α-glucosidase inhibited the hydrolysis of all substrates tested. These properties of Tetrahymena acid α-glucosidase were found to be similar to those of the human liver lysosomal α-glucosidase.
  • Yasunori KUSHI, Shizuo HANDA, Ineo ISHIZUKA
    1985 年 97 巻 2 号 p. 419-428
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    A series of underivatized sulfoglycolipids (SM4g, lyso-SM4g, SM4s, SM3, SM2, SB2, and SBla) from various tissues were analyzed by both positive (POS-SI-MS) and negative (NEG-SI-MS) secondary ion mass spectrometry. By POS-SI-MS were detected the molecular ions of sulfoglycolipids in the form with sodium or potassium together with some fragment ions useful for the carbohydrate sequence determination. The analysis of monosulfogangliotriaosyl- or monosulfogangliotetraosylceramide and bis-sulfoglycolipid was difficult due to noise in the high mass region.
    On the other hand, NEG-SI-MS of sulfoglycolipids gave more intense signals from molecular ion of (M-H)- for monosulfoglycolipids and ((M-H+Na)-H)- for bis-sulfoglycolipid. Many fragment ions useful for the elucidation of the carbohydrate sequences were also obtained with significant intensities. The fragmentation was assessed to occur at the glycosidic linkages to form ions of the oligosaccharides with or without ceramide. These ions were useful for sugar sequencing and also for distinguishing the differences in the position of the sulfate group. The intensities of saccharide ions without sulfate were lower than those with sulfates. In the case of SB2 and SB1a, containing 2 mol of sulfate ester groups, the molecular ion was detected as ((M-H+Na)-H)-. Also, fragment ions with 2 mol of sulfate were detected as the sodium-additive form. It was concluded that NEG-SI-MS is a very useful technique for the structural elucidation of higher sulfoglycolipids.
  • Toshio SHIMADA, Teruko SUGO, Hisao KATO, Kei-ichi YOSHIDA, Sadaaki IWA ...
    1985 年 97 巻 2 号 p. 429-439
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    The activation of Factor XII and prekallikrein by polysaccharide sulfates and sulfatides in the presence of high-molecular-weight (HMW) kininogen was studied, and compared with the kaolin-mediated activation reaction. Among a variety of artificially-sulfated polysaccharides and native polysaccharide sulfates, amylose sulfate (M. W.=380, 000 and sulfur content, 19.1%) and sulfatide were found to have the most efficient ability to trigger the activation of prekallikrein by Factor XII. The effects of these two kinds of negatively-charged surfaces on the following three activation reactions were compared; the activation of prekallikrein by Factor XII (reaction 1), the activation of Factor XII by kallikrein (reaction 2) and the activation of prekallikrein by Factor XIIa (reaction 3). All three reactions mediated by the selected surfaces were strongly accelerated by HMW kininogen and its derivatives, kinin-free protein and fragment 1•2-linked light chain, like the kaolin-mediated activation. However, this accelerating effect of HMW kininogen on the amylose sulfate- and sulfatide-mediated activations (reaction 1) was diminished after treatment with fluorescein iso-thiocyanate, whereas the effect on the kaolin-mediated activation was not influenced by fluorescein-labeling. In addition, reaction 2 mediated by amylose sulfate and sulfatide was extremely slow even in the presence of HMW kininogen, and the results also differed from those with kaolin. The sulfatide-mediated activation of reaction 1 was not inhibited by fragment 1•2 (His-rich fragment), which is released from HMW kininogen by the action of kallikrein, and is known to be a potent inhibitor of the kaolin-dependent activation.
    These results indicate that the mechanisms responsible for surface activation triggered by soluble amylose sulfate, sulfatide micelles and kaolin differ from each other as regards the molecular interaction with the contact factors.
  • Hiroaki HAYASHI, Yoshihide OHE, Tomoko HAYASHI, Koichi IWAI
    1985 年 97 巻 2 号 p. 441-448
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    For the sequence analysis of histones rich in lysine, we modified the subprograms for two reagents of a JEOL JAS-47 KS protein sequence analyzer. Together with this modification, the use of a synthetic carrier, Polybrene, the minimization of aldehyde contamination in Quadrol buffer, and the introduction of hydrophilic groups into ε-N-amino groups of lysine residues, markedly increased the repetitive yield of PTH-amino acids. Tetrahymena histones H3 and H4 were thus sequenced up to residues 104 and 92, respectively, in each consecutive analysis (Hayashi, T., Hayashi, H., Fusauchi, Y., & Iwai, K. (1984) J. Biochem. 95, 1741-1749; Hayashi, H., Nomoto, M., & Iwai, K. (1984) J. Biochem. 96, 1449-1456). The details for these improved procedures and results are described here.
  • Anthony P. CORFIELD, Roland SCHAUER, Lambertus DORLAND, Johannes F.G. ...
    1985 年 97 巻 2 号 p. 449-461
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    1. Investigation of the action of highly purified Clostridium perfringens sialidase on ganglioside II3Neu5Ac-Gg4Cer and its oligosaccharide II3Neu5Ac-Gg4, in the presence and absence of sodium cholate, extend earlier results obtained with impure enzyme fractions.
    2. Sialidase labeled with 125I was found to bind to various ganglioside substrate micelles, including II3Neu5Ac-Gg4Cer, and to mixed ganglioside-sodium cholate micelles.
    3. No binding occurred between the enzyme and the ganglioside-derived oligosaccharide II3Neu5Ac-Gg4, even when radioactive II3Neu5Ac-Gg4-[3H]ol was used.
    4. The binding of sialidase to micellar substrate is a condition for enzymic hydrolysis. Correspondingly, II3Neu5Ac-Gg4Cer and II3Neu5Ac-Gg4Cer-sodium cholate micelles were hydrolyzed by the enzyme but II3Neu5Ac-Gg4 was not.
    5. Ganglioside oligosaccharide analogues containing an amino function at the reducing terminus or between two oligosaccharide chains, II3Neu5Ac-Gg4-NH2, and (II3Neu5Ac-Gg4)2NH, were hydrolyzed in the absence of cholate. A synthetic analogue of II3Neu5Ac-Gg4Cer containing only the fatty acid moiety and not the sphingosine residue (I1-deoxy-I1-stearamido-II3-monosialo-gangliotetraitol) behaved as the ganglioside in the presence and absence of sodium cholate.
    6. The results are discussed in the light of 500 MHz 1H-NMR data which may suggest less conformational freedom of the internal galactose-linked sialic acid residue in II3Neu5Ac-Gg4 relative to the sialic acid in terminal, linear position of IV3Neu5Ac-, II3Neu5Ac-Gg4. This difference in conformation of the two sialic acid linkages may give an explanation for the resistance of the internal sialic acid towards the action of sialidase.
  • Karl Heinz SCHEIT, Sisinthy SHIVAJI, Pushpa M. BHARGAVA
    1985 年 97 巻 2 号 p. 463-471
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    Seminalplasmin, an antimicrobial protein present in bovine seminal plasma, is shown to inhibit the growth of, as well as nucleic acid and protein synthesis in, wild-type Saccharomyces cerevisiae SM202, when used at concentrations>200 μg/ml, in contrast to 20 μg/ml that is sufficient for Escherichia coli. An osmotically labile strain of S. cerevisiae VY1160 is 1-2 orders of magnitude more sensitive to seminalplasmin than the wild-type strain. RNA synthesis in protoplasts and nuclei of S. cerevisiae SM202 was also about as sensitive to seminalplasmin as in E. coli and S. cerevisiae VY1160. The RNA polymerases I and II from S. cerevisiae were strongly inhibited by seminalplasmin in vitro, while DNA and protein syntheses were not affected by seminalplasmin in cell-free systems, unlike in the whole cells. It is concluded that seminalplasmin acts in S. cerevisiae by entering the cells and inhibiting transcription.
  • Takayuki HARA, Misako TANIGUCHI
    1985 年 97 巻 2 号 p. 473-482
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    Weanling male rats were fed a riboflavin-deficient diet for 5-8 weeks, and the decrease in NADPH-cytochrome P-450 reductase (FpT) activity in the liver microsomes was compared with the contents of riboflavin derivatives. The decrease of FpT activity for the reduction of cytochrome c was greater than that for the reduction of ferricyanide.
    The FpT's of riboflavin-deficient and control rats were indistinguishable in the Ouchterlony immunodiffusion test against anti-FpT, and were shown to have the same molecular weight of 78, 000 by SDS-polyacrylamide slab gel electrophoresis. However, the purified FpT of the riboflavin-deficient rats contained 14.2, 4.9, and 1.9 nmol of FAD, FMN, and riboflavin per mg of protein, respectively, while that of the control rats contained 10.6 and 9.5 nmol of FAD and FMN per mg of protein, respectively.
    After riboflavin injection into the riboflavin-deficient rats, NADPH-cytochrome c reductase activity and FMN content of the FpT were restored to the control levels in 36 h, NADPH-ferricyanide reductase activity recovered in 18 h, and riboflavin content diminished in 18 h. On incubation of the purified FpT of the riboflavindeficient rats with FMN, NADPH-cytochrome c reductase activity and FMN content were restored to those of control rats. These results indicated that a part of FMN in the FpT of the riboflavin-deficient rats was replaced with FAD and riboflavin.
  • Mamoru NAKANISHI, Sayori TSUNEDA, Seizo TAKAHASHI, Masamichi TSUBOI
    1985 年 97 巻 2 号 p. 483-491
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    The kinetics of the hydrogen-deuterium exchange reactions of deoxyguanosine (dG), deoxycytidine (dC), double-helical poly[d(G-C)]•poly[d(G-C], and double-helical poly(dG)•poly(dC) have been examined at 20°C, pH 7.0, and in low-salt (0.15M NaCl) medium by stopped-flow ultraviolet spectrophotometry, in the spectral region of 260 to 320 nm. The rate constant was found to be 78.9 s-1 for dG-NH, 2.2 s-1 for dG-NH2, 39.3 s-1 for dC-NH2, 2.4 s-1 (fast) and 0.94 s-1 (slow) for poly[d(G-C)]•poly[d(G-C)], and 2.2 s-1 (fast) and 0.92 s-1 (slow) for poly(dG)•poly(dC). From these values, the probability of base-pair opening of the G•C containing B-form double helix is estimated to be (3±1)×10-3. This is much greater than what is expected from an extrapolation of the van't Hoff plot at the helix-coil transition region, i.e. at about 110°C. The mechanism of these base-pair openings at 20°C (as well as the mechanism of base-pair reformation) is suggested to be totally different from those in the melting temperature range.
  • Shigeharu NAGASAWA, Chiharu KOBAYASHI, Tomoko MAKI-SUZUKI, Noriaki YAM ...
    1985 年 97 巻 2 号 p. 493-499
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    The C3 convertase of the classical pathway of the complement system is a labile complex, C-4b, 2a, and is activated by limited proteolysis of two components, C4 and C2, by C-ls. By utilizing iodine-treated C2 and size exclusion high-performance liquid chromatography (HPLC), we have succeeded in isolating for the first time the classical pathway C3 convertase. Size exclusion HPLC demonstrated that the apparent molecular mass of the C3 convertase was 280 K daltons. The C3 convertase decay-dissociates spontaneously into C4b and C2a. The decay-dissociation is a temperature-dependent reaction and the half-lives of the C3 convertase at 24, 30, and 37°C were estimated to be 400, 180, and 60min, respectively. The decaydissociation was also dependent on pH and was accelerated by increasing pH. In addition, the decay-dissociation of the C3 convertase was accelerated by C2b. This result suggests that C2b acts as a feedback inhibitor on the activation of the classical pathway of complement system.
  • Kazue NAGATA, Mitsuko OKADA
    1985 年 97 巻 2 号 p. 501-507
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    The properties of crude and purified mitochondrial aspartate aminotransferase preparations from pyridoxine-deficient and control rat livers were compared. The preparations from the two sources showed very similar behaviors on heat treatment, electrophoresis and chromatofocusing, and had similar molecular weights, but their visible absorption spectra and circular dichroism properties were different. These results suggest that mitochondrial aspartate aminotransferase from pyridoxinedeficient and control rat livers have very similar properties, but differ somewhat in conformation in the region of the pyridoxal phosphate binding site.
  • Minoru SUZUKI, Akemi SUZUKI, Tamio YAMAKAWA, Ei MATSUNAGA
    1985 年 97 巻 2 号 p. 509-515
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    A molecular species of sialic acid was isolated in a free form from cerumen of the wet type, but not of the dry type, by an ion-exchange column chromatography and preparative high-performance liquid chromatography. Structural analysis of this sialic acid was performed by gas-liquid chromatography/mass spectrometry with chemical ionization (CI) and electron ionization (EI). In the CI mass spectra, the protonated molecular ion of the trimethylsilyl derivative was observed at m/z 580 and that of the methyl ester-trimethylsilyl derivative was at m/z 522. In the EI mass spectrum, the methyl ester-trimethylsilyl derivative gave characteristic ions at m/z 506, 462, 418, 416, 328, 316, 238, 228, 205, 186, and 173. This mass spectrum was identical with that of 2, 7-anhydro-N-acetylneuraminic acid, which was reported by Lifely and Cottee (Carbohydr. Res. 107, 187-197, 1982) as the mass spectrum of a by-product prepared from N-acetylneuraminic acid by methanolysis. These results indicate that the compound in the wet cerumen is 2, 7-anhydro-N-acetylneu-raminic acid. Since this sialic acid species could not be detected in cerumens of the dry type, its formation in the wet type may be controlled by an autosomal dominant gene.
  • Yutaka ASHIKARI, Yoji ARATA, Kozo HAMAGUCHI
    1985 年 97 巻 2 号 p. 517-528
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    Circular dichroism and proton nuclear magnetic resonance measurements were made to clarify how reduction and reduction plus alkylation of the intrachain disulfide bond affect the conformation and stability of the constant (CL) fragment of the immunoglobulin light chain. The pH titration behavior of the two histidine residues in the intact CL, reduced CL, and reduced and alkylated CL fragments were followed by proton nuclear magnetic resonance spectroscopy. It was shown that reduction of the intrachain disulfide bond does not affect the solution conformation of the immunoglobulin fold, whereas reduction and alkylation results in extensive unfolding of the protein molecule. These results are consistent with the previous results obtained by Goto and Hamaguchi ((1979) J. Biochem. 86, 1433-1441) using circular dichroism, fluorescence, and titration of SH groups.
    The pH-induced unfolding of the reduced CL fragment was compared with that of the intact CL fragment. The stability of the intact CL fragment to acid was much greater than that of the reduced CL fragment. The equilibria of the unfolding by acid could be explained by assuming that the same ionizable groups participate in the unfolding for the reduced CL fragment as for the intact CL fragment, and that the stability of the intact CL fragment is 100 times greater than that of the reduced CL fragment. In the alkaline pH region, the reduced CL fragment was unfolded above pH 8.5, while the intact CL fragment was not unfolded until pH 11.6. The kinetics of the unfolding by alkali of the reduced CL fragment was also studied. The unfolding by alkali of the reduced CL fragment could be explained by assuming that the reduced CL molecule can no longer adopt the folded conformation when either or both SH groups ionize.
  • Hiromichi KUMAGAI, Masaoki IMAZAWA, Kanji MIYAMOTO
    1985 年 97 巻 2 号 p. 529-532
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    Microtubule proteins were purified from chick brains at various developmental stages from the 12-day embryo to adult. Three species of microtubule-associated protein-1 (MAP-1) and 5-7 molecular components of tau proteins were observed by SDS-polyacrylamide gel electrophoresis. The molecular compositions were observed to change during development of the chick brain.
  • Tsutomu KODAKI, Fumiaki KATAGIRI, Masahide ASANO, Katsura IZUI, Hirohi ...
    1985 年 97 巻 2 号 p. 533-539
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    The phosphoenolpyruvate carboxylase gene (ppc) from Anacystis nidulans, a cyanobacterium (blue-green alga), was cloned in Escherichia coli. Chromosomal DNA of A. nidulans was partially digested with Sau3AI, and the obtained DNA fragments were ligated in the BamHI site of pBR322. The hybrid plasmids were first transformed into E. coli K802 (hsdR-, hsdM+) to obtain the gene bank of A. nidulans. The bank consisted of about 12, 000 clones. These hybrid plasmids were then transformed into E. coli PCR1 (ppc2-, recAl-, hsdR+, hsdM+), and the transformants were selected by complementation of the ppc mutation (phenotype of glutamate requirement). In the cell-free extracts of E. coli strains having the cloned ppc gene, PEPCase activities were detected, but their properties were different from those of the E. coli enzyme. Analysis by subcloning showed that the ppc gene was included in a DNA fragment 3, 500 base pairs long and the maxicell method revealed that the molecular weight of the gene product was about 108, 000. It is suggested that the ppc gene is expressed in E. coli mainly by read-through transcription, being initiated by the promoter of tetracycline-resistance gene of pBR322, but the significant expression in reversed orientation of the cloned ppc gene indicates that the gene includes a promoter capable of functioning in E. coil cells.
  • Takayuki MIYANISHI, Tetsuo MAITA, Fumi MORITA, Shuhei KONDO, Genji MAT ...
    1985 年 97 巻 2 号 p. 541-551
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    Smooth muscle myosin from scallop (Patinopecten yessoensis) adductor muscle contains two kinds of regulatory light chains (regulatory light chains a and b), and myosin having regulatory light chain a is suggested to be suitable for inducing “catch contraction” rather than myosin having regulatory light chain b (Kondo, S. & Morita, F. (1981) J. Biochem. 90, 673-681). The amino acid sequences of these two light chains were determined and compared. Regulatory light chain a consists of 161 amino acid residues, while regulatory light chain b consists of 156 amino acid residues. Amino acid substitutions and insertions were found only in the N-terminal regions of the sequences. The structural difference between the two light chains may contribute to the functional difference between myosins having regulatory light chains a and b.
  • Fumi MORITA, Shuhei KONDO, Ken TOMARI, Osamu MINOWA, Mitsuhiko IKURA, ...
    1985 年 97 巻 2 号 p. 553-561
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    Calcium binding was studied with two regulatory light chains (RLC-a and RLC-b) of smooth muscle myosin of scallop. With the equilibrium dialysis method, the binding of 0.98 mol Ca2+ per mol of RLC-b was observed with a dissociation constant of 2.3×10-5M. Similar values for RLC-b, 1.9×10-5M, and RLC-a, 1.5×10-5M, were obtained by measuring the difference absorption spectrum induced by Ca2+. The difference molar absorption coefficient at 288 nm was 159 and 209 M-1•cm-1 for RLC-a and RLC-b, respectively, while it was - 34M-1•cm-1 for the regulatory light chain of striated muscle myosin of scallop (RLC-st). Proton NMR spectra of the three light chains were very similar to each other and were broader than those of other Ca2+ binding proteins, parvalbumin and calmodulin. The regulatory light chains may be more rigid than in these Ca2+ binding proteins. CD spectra were measured for the three light chains, and the estimated helix contents were 27, 29, and 24%, respectively, for RLC-a, RLC-b, and RLC-st. All these results in comparison with the primary structures led us to suppose that the polypeptide of regulatory light chains is folded in such a way that domain 4 becomes near to the calcium binding site of domain 1.
    The decrease in intact light chains on trypsin digestion was determined for the gel electrophoretic patterns. RLC-a was 6 times more susceptible to the tryptic digestion than RLC-b. The susceptible residues may be located in the N-terminal region which had a different amino acid sequence from that of RLC-b. RLC-b was protected completely by the binding to myosin, but RLC-a was not at all. The regulatory light chains seem to bind to myosin so that the N-terminal region is oriented toward the outside, and that of RLC-a in myosin is exposed so as to be accessible to trypsin.
  • Eiro MUNEYUKI, Eisuke NISHIDA, Kazuo SUTOH, Hikoichi SAKAI
    1985 年 97 巻 2 号 p. 563-568
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    Cofilin, a 21, 000 molecular weight protein originally purified from porcine brain that is capable of binding to actin filaments in a molar ratio of the protein to actin monomer of 1:1 in the filament (Nishida et al. (1984) Biochemistry 23, 5307-5313), was purified from porcine kidney in the present study. The two cofilins from brain and kidney were indistinguishable from each other with respect to the mobility on polyacrylamide gel electrophoresis in the presence of sodium dodecyl sulfate, the one-dimensional peptide map, and the mode of interaction with actin. Treatment of the actin-cofilin complex with a zero-length cross-linker, 1-ethyl-3-[3-(dimethylamino)propyl]carbodiimide (EDC), generated a cross-linked product with an apparent molecular weight of 63, 000. Analysis of this product by peptide mapping (Sutoh (1982) Biochemistry 21, 3654-3661) showed that cofilin was cross-linked with the N-terminal segment of actin containing residues 1-12.
  • Eisuke HANADA, Kunio KONNO
    1985 年 97 巻 2 号 p. 569-577
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    A disialosylganglioside was isolated from adult bovine nasal cartilage, and its structure was determined by analysis of sugar composition, permethylation analysis, exoglycosidase treatment, and mild acid hydrolysis. The structure of this ganglioside was identified as disialo-lacto-N-norhexaosyl ceramide, NeuNAc(α2-8)NeuNAc(α2-3)Gal(β1-4)G1cNAc(β1-3)Gal(β1-4)G1cNAc(β1-3)Gal(1-4)Glc(1-1)Cer. Furthermore, we also isolated from this cartilage gangliosides whose structures were presumed to be monosialo-lacto-N-norhexaosyl ceramide, and mono- and disialo-lacto-N-neotetraosyl ceramide. The major fatty acids of the four gangliosides isolated were palmitic, stearic, behenic and lignoceric acids. The predominant long chain bases were sphingenine, heptadecasphingenine and hexadecasphingenine.
  • Hiroshi AKANUMA, Toshikazu YAMANOUCHI, Hideki ONO, Keiichi NOMURA, Yas ...
    1985 年 97 巻 2 号 p. 579-588
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    This report describes an application of liquid chromatography to the determination of sorbitol in red blood cells. The chromatograph employed in the present study was made up of sub- and main-separation systems and a detector portion. The sub-separation system was for concentration of polyols and involved two small columns, each containing the same anion exchange resin. The first was a tiny column which, in borate form, served as the concentrator of polyols and sugars charged in a large volume, while the second, in acetate form, separated the carbohydrates from the borate. The main system was for fine separation of each carbohydrate and employed cation exchange columns. The detector part utilized a flow fluorometric method comprising two successive reactions: periodate oxidation followed by the Hantzsch reaction. The resulting whole chromatographic system was applied to the determination of sorbitol in red blood cells obtained from normal rats and rats made diabetic by the administration of streptozotocin; a part of the latter group had also received an aldose reductase inhibitor. Our results supported the concepts that a prolonged duration of high blood glucose level induces an elevated level of sorbitol inside red blood cells and that aldose reductase inhibitors are effective in reducing this level.
  • Fusahiro OGATA, Satoru MAKISUMI
    1985 年 97 巻 2 号 p. 589-597
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    Several trypsin inhibitors with different mobilities on polyacrylamide gel electrophoresis occur in the tubers of taro (Colocasia antiquorum), and they each have a dimeric molecular weight of 40, 000. Of all the constituent subunits, molecular weight 20, 000, of the taro trypsin inhibitor (TTI), three major subunit components were separated by chromatography on SP-Sephadex C-25 in 8M urea, and they were named protomers α, β, and γ in the order of their elution from the SP-Sephadex column. After removal or dilution of the urea, the three protomers could be either reassociated individually or hybridized with each other to form dimeric inhibitors. All of the reassociated dimers were powerful inhibitors of trypsin. Among them, each dimer derived from protomers α and γ was a weak inhibitor of chymotrypsin, whereas the dimer of protomer β did not inhibit the enzyme. Therefore TTI is presumed to be a mixture of heterogeneous and homogenous dimers whose properties reflect those of their constituent protomers. It was also proved that the major three trypsin inhibitors (TTI-I, TTI-II, and TTI-III) previously isolated from taro tubers are composed of protomers a and γ, i.e., TTI-II is a heterogeneous dimer of protomers α and γ, and TTI-I and TTI-III are homogeneous dimers of protomers α and γ, respectively. The molecular weight of a trypsin-TTI complex saturated with trypsin was found to be 79, 000, suggesting the formation of a tetrameric complex.
  • Koshi SAITO, Kunihiro YAMAMOTO, Takaji TAKAI, Takeshi KUZUYA
    1985 年 97 巻 2 号 p. 599-604
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    Efflux of preloaded I- from the thyroid induced by externally added I- was studied using a biological model of the thyroid I- transport system. Phospholipid vesicles (P-vesicles) made from thyroid plasma membranes and soybean phospholipids were capable of accumulating I- in the presence of external Na+. P-vesicles incubated in 136mM Na+ containing 0.9 μM I- with 125I- for 2min accumulated I- so that the I- concentration in the vesicles became about 2 μM. Addition of 5-20 μM stable I- to the incubation mixture at 2min incubation resulted in a dose-dependent decrease in previously loaded 125I- in the vesicles. In other words, a dose-dependent increase in efflux of preloaded 121I- was observed. While the efflux occurred, Na+ -dependent I- influx into P-vesicles was preserved. When 2mM CIO4-, a specific inhibitor of Na+ -dependent I- influx, was added together with 10 μM I-, the external I- failed to diminish preloaded 125I- in P-vesicles. The 125I- efflux did not occur when a large amount of stable I- entered P-vesicles independently of Na+ in the presence of ClO4-. Similar 125I- efflux induced by externally added 5 μM SCN- was also blocked by simultaneously added ClO4-. These observations suggest that such I- efflux from the thyroid is a certain type of uphill I- transport which is closely related to Na+ -dependent I- transport and that ClO4- and SCN- act on a common site of the I- transport system.
  • Kuniaki TAKAGI, Yasuo ICHIKAWA, Kazuhiro NAGATA
    1985 年 97 巻 2 号 p. 605-616
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    A homodimer protein consisting of two 38, 000 dalton peptides was isolated from a murine leukemia cell line (Ml). The binding molar ratio of the 38K-dimer protein to purified skeletal muscle actin was saturated at 1:3, and when the 38K-dimer/actin ratio exceeded 1:12, gelation occurred. This gelation was completely inhibited by the presence of either 10mM KCl or 20mM NaCl. The protein induced actin filament bundling, which required a higher 38K-dimer/actin ratio and was not affected by the presence of monovalent cations. During the differentiation of Ml cells, the sensitivity of the 38K protein to monovalent cations was decreased; that is 20mM KCl or 50mM NaCl was required to inhibit the gelation by the 38K protein isolated from differentiated cells. On the other hand, the intracellular K+ content of Ml cells decreased from 70±5mM to 18±3mM, and Na+ increased from 10±5mM to 40±10mM during the differentiation. These findings suggest that the differentiation brought about conditions favourable for the 38K protein to induce actin gelation, and in turn, the locomotive and phagocytic activities which were induced only after differentiation in this cell line.
  • Hideaki TSUNEMATSU, Hiroaki NISHIMURA, Koichi MIZUSAKI, Satoru MAKISUM ...
    1985 年 97 巻 2 号 p. 617-623
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    The rates of hydrolysis of Nα-benzoyl-p-guanidine-L-phenylalaninamide (Bz-GPA-NH2) and Nα-substituted p-nitroanilides (pNA) of GPA (benzyloxycarbonyl(Z)-GPA-pNA, benzoyl(Bz)-GPA-pNA and acetyl(Ac)-GPA-pNA) by bovine and porcine trypsins were compared with those of arginine (Arg) substrates. The amide type substrates of GPA were hydrolyzed as fast as those of Arg by the two enzymes with much the same kcat/Km values, though significant differences were found between the kcat and Km values of GPA derivatives and those of Arg derivatives. The kinetic behavior of porcine trypsin toward GPA substrates was almost the same as that of the bovine enzyme. The ratio of the kcat value for Bz-GPA-OEt to that for Bz-GPA-NH2 was much larger than that for the ester to amide substrates of arginine, suggesting that the conformational change of the active site of trypsin induced by a benzene ring in the side chain of Bz-GPA-OEt specifically increases the velocity of the deacylation process of the ester substrate. Remarkably low values of both kcat and Km were found for the tryptic hydrolysis of Z-GPA-pNA and Ac-GPA-pNA, as well as on that of Bz-GPA-pNA (Tsunematsu, H., et al. (1983) J. Biochem. 94, 123-128). Z-GPA-pNA is the best substrate for the two trypsins among the three Nα-substituted anilide substrates of GPA. Substrate activation was observed with bovine trypsin in the hydrolysis of the three anilide substrates of GPA in a substrate concentration range higher than about 5.0×10-4M, but it was found with the porcine enzyme only in the hydrolysis of Z-GPA-pNA. In contrast, no activation by the amide substrate of GPA could be detected with either enzyme.
  • Shooju KAMEYAMA, Hisashi ICHIKAWA, Yayoi SUNAGA, Sumio NAKATA, Yukio S ...
    1985 年 97 巻 2 号 p. 625-632
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    For both cardiac and skeletal myosin, the Ca-ATPase activity of myosin at acidic pH was shown to be different from that at alkaline pH, in the susceptibility to heatinactivation, the effects of organic solvents, and the effect of trinitrophenylation of the myosin. It is therefore suggested that there are two different types of Ca-ATPase of both cardiac and skeletal myosin.
    Differences in the Ca-ATPase activity were also found between cardiac and skeletal myosins. (a) The Ca-ATPase activity of cardiac myosin was more susceptible to heat-inactivation at alkaline pH than at acidic pH. In contrast, the activity of skeletal myosin was more susceptible to heat-inactivation at acidic pH than at alkaline pH. (b) Dioxane weakly stimulated the activity of cardiac myosin at acidic pH, but strongly activated that of skeletal myosin at acidic pH. Acetone very strongly inhibited the activity of cardiac myosin at alkaline pH, but not so strongly that of skeletal myosin at alkaline pH. (c) Trinitrophenylation of the myosin resulted in loss of the activity optimum at acidic pH with skeletal myosin but not with cardiac myosin.
    As reported by Srivastava et al. (J. Biochem. 86, 725-731, 1979), 1 mol of lysine residue per mol of cardiac myosin quickly reacted with 2, 4, 6-trinitrobenzene sulfonate (TNBS) either in the absence or presence of inorganic pyrophosphate (PPS). However, trinitrophenyl (TNP) groups bound to cardiac myosin in the presence of PP1 were significantly different, in the pH dependence of the absorption spectrum, from those bound (to cardiac myosin) in the absence of PP1. Moreover, the difference was found to be very similar to that observed between TNP groups bound to skeletal myosin in the presence of PP1 and those bound (to skeletal myosin) in the absence of PP1. It is therefore suggested that the sites on cardiac myosin for trinitrophenylation in the presence of PP1 are different from the sites for trinitrophenylation in the absence of PP1, as much as the sites on skeletal myosin are.
  • Takashi KOYAMA, Norio INOKUCHI, Masanori IWAMA, Masachika IRIE
    1985 年 97 巻 2 号 p. 633-641
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    1. In order to elucidate the structure-function relationship of glucoamylases [EC 3. 2. 1. 3, α-D-(1-4)-glucan glucohydrolase] from Aspergillus saitoi, the reaction of a minor component, Gluc M2 with 1-cyclohexyl-3-(2-morpholinyl-(4)-ethyl)carbodiimide metho p-toluenesulfonate (CMC) was studied at pH 4.5.
    2. Inactivation of Gluc M2 with [14C]CMC proceeded with the incorporation of about 5 CMC moieties. From the results of analyses of amino acid and sulfhydryl contents of CMC-modified Gluc M2 and the hydroxylamine treatment of the CMC-modified Gluc M2 at pH 7.0, it was concluded that the sites of CMC-modification were carboxylic acids of Gluc M2.
    3. In the presence of maltose, when Gluc M2 was treated with [14C]CMC, ca. 4 CMC moieties were incorporated with a simultaneous decrease in activity (30%). The Gluc M2 modified in the presence of maltose was re-modified with CMC after elimination of maltose. The CMC-modified Gluc M2 (70% activity) was inactivated completely with the further incorporation of ca. 2 CMC moieties.
    4. The logarithm of the half-life of the inactivation of Gluc M2 by CMC was a linear function of log[CMC] indicating that one carboxyl group among the modified ones was crucial for the inactivation of Gluc M2.
    5. From the results of these modification reactions, it was concluded that one or two carboxylic acids in Gluc M2 were crucial for the catalysis of glucoamylase from A. saitoi.
    6. Based on the analysis of the pH-profile of CMC inactivation of Gluc M2, the participation of a carboxylic acid having pKa 5.7 in the active site is proposed.
  • Takashi DAIHO, Haruhiko TAKISAWA, Taibo YAMAMOTO
    1985 年 97 巻 2 号 p. 643-653
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    The effects of intra- and extravesicular calcium and magnesium ions on the hydrolysis of the phosphoenzyme (EP) intermediate formed in the reaction of Ca2+, Mg2+-dependent ATPase of the sarcoplasmic reticulum were investigated. The rate constants of EP hydrolysis were measured under conditions that allowed a single turnover of ATP hydrolysis to minimize the increase in calcium concentration inside the vesicles.
    The EP formed during a single turnover was hydrolyzed biphasically and could be resolved into fast- and slow-decomposing components. When free Mg2+ outside the vesicles was chelated by adding excess EDTA, EP could also be kinetically resolved into two components; EDTA-sensitive EP, which could be quickly decomposed by adding EDTA, and EDTA-insensitive EP, which could be prevented from decomposing by adding EDTA. The amount of EDTA-sensitive EP decreased rapidly during the initial phase of the reaction, while that of EDTA-insensitive EP decreased slowly with the same rate constant as that of the slow-decomposing EP. These results showed that the biphasic time course of EP hydrolysis was caused by the formation of EDTA-sensitive and -insensitive EP during the reaction. The time course of EP hydrolysis could be quantitatively analyzed in terms of the following reaction mechanism.
    EDTA-sensitive EP_??_EDTA-insensitive EP
    E+P1
    The decomposition of EDTA-insensitive EP required Mg2+ outside the vesicles and was competitively inhibited by extravesicular Ca2+. The decomposition of EDTA-sensitive EP was inhibited by Ca2+ inside the vesicles but not by external Ca2+. The linear relationships between the inverse of the rate constants of EP decomposition during the initial phase and the intravesicular CaCl2 concentrations suggested that decomposition of EDTA-sensitive EP was inhibited by the binding of 1 mol of intravesicular Ca2+ to 1 mol of EP. Furthermore, Mg2+ inside the vesicles scarcely affected the inhibition of EP hydrolysis by intravesicular Ca2+. These results suggested that magnesium ions are not counter-transported during the active transport of calcium by SR vesicles.
  • Kiyoshi KAWAKAMI, Kiyohisa MIZUMOTO, Akira ISHIHAMA, Kazuko SHINOZAKI- ...
    1985 年 97 巻 2 号 p. 655-661
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    During the generation of primer for transcription initiation by endonucleolytic cleavage of capped RNA, the influenza virus-associated RNA-dependent RNA polymerase recognizes three signals of capped RNA, i.e., cap-1 structure at 5' termini, RNA sequences at the cleavage sites, and distance between these two signals (Plotch, S. J., Bouloy, M., & Krug, R. M. (1979) Proc. Natl. Acad. Sci. U. S. 76, 1618-1622; Kawakami, K., Mizumoto, K., & Ishihama, A. (1983) Nucl. Acids Res. 11, 3637-3649). The free cap-1 structure, i.e., m7GpppNm not associated with polynucleotides, stimulates the RNA synthesis as well, irrespective of the presence or absence of dinucleotide primers. Since the stimulation by m7GpppNm and ApG is additive and the free cap-1 structure is not incorporated into 5' termini of product RNA, we propose that the cap-1 structure stimulates the RNA polymerase by allosteric modulation rather than by priming the transcription initiation.
  • Masatoshi KAKU, Emi KUSUNOSE, Satoru YAMAMOTO, Kosuke ICHIHARA, Masami ...
    1985 年 97 巻 2 号 p. 663-670
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    Three cytochrome P-450 preparations, designated as cytochrome P-450ca, cytochrome P-450cb, and cytochrome P-448c fraction, were separated and purified about 23-, 50-, and 29-fold, respectively, from the cholate extracts of rabbit colon mucosa microsomes. Their specific contents were 1.2, 2.6, and 1.5 nmol of cytochrome P-450 per mg of protein, respectively. Cytochrome P-450ca and cytochrome P-450cb migrated as heme-containing polypeptide bands with molecular weights of about 53, 000 and 57, 000, respectively, on SDS-polyacrylamide gel electrophoresis. The CO-reduced difference spectra of cytochrome P-450ca, cytochrome P-450cb, and cytochrome P-448c fraction showed maxima at 451, 450, and 449 nm, respectively. Cytochrome P-450ca efficiently catalyzed the ω-hydroxylation of prostaglandin A, (PGA1) and the ω- and (ω-1)-hydroxylation of caprate, laurate, and myristate in the reconstituted system containing cytochrome P-450ca, NADPH-cytochrome P-450 reductase, cytochrome b5, and phosphatidylcholine. In contrast, cytochrome P-450cb and cytochrome P-448c fraction had no detectable activity toward PGA1 and fatty acids. Both catalyzed aminopyrine and benzphetamine N-demethylation. Cytochrome P-448c fraction also hydroxylated benzo(a)pyrene, and phosphatidylinositol or phosphatidylserine exhibited a stimulatory effect on this activity. The results show that rabbit colon microsomes contain catalytically different cytochrome P-450, one of which is specialized for the ω-oxidation of prostaglandins, the others being involved in the metabolism of exogenous compounds such as drugs and polycyclic hydrocarbons.
  • Saori TAKAHASHI, Retsu MIURA, Yoshihiro MIYAKE
    1985 年 97 巻 2 号 p. 671-677
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    We purified, from human kidney, a protein that reacts with rabbit anti-porcine kidney renin binding protein (RnBP) antiserum by trapping with porcine kidney renin. The purified preparation showed a single protein peak on gel filtration by high performance liquid chromatography (HPLC) and two protein bands on sodium dodecyl sulfate polyacrylamide gel electrophoresis (SDS-PAGE). The latter two kinds of protein were identified as the porcine renin and human kidney protein from their electrophoretic mobilities and reactivity toward rabbit anti-porcine kidney renin and RnBP antisera. The molecular weights of the purified preparation and the human kidney protein were estimated to be 56, 000 by HPLC and 43, 000 by SDS-PAGE, respectively. The specific activity of porcine renin in the purified preparation was 8.6mg angiotensin I per mg of protein per h at 37°C and pH 6.5. This specific activity was about one-fifth that of free porcine renin. Therefore, it is suggested from the reactivity toward the anti-porcine RnBP antiserum and inhibitory action toward porcine renin that the human kidney protein is RnBP and that the human RnBP is purified as a complex with porcine renin.
  • Atsunori KASHIWAGI, Yutaka HARANO, Keisuke KOSUGI, Takamitsu NAKANO, H ...
    1985 年 97 巻 2 号 p. 679-684
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    The specific [125I]insulin binding to primary cultured hepatocytes was significantly greater than that to freshly isolated hepatocytes. Low affinity insulin binding sites in cultured cells were 6-fold greater in number than those of freshly isolated cells without a significant change in high affinity sites. However, both sensitivity (insulin concentration for half maximum stimulation) and responsiveness (% of increase above the basal level) to insulin for the stimulation of ODC activity were similar for isolated and cultured cells indicating an important role of high affinity sites in the insulin action.
    On the other hand, the specific [125I]glucagon binding to cultured cells was significantly decreased. Low affinity glucagon binding sites in cultured cells decreased by about 50% in cultured cells without a significant change in high affinity sites. Both sensitivity and responsiveness to glucagon for the stimulation of ketogenesis from palmitate also decreased as compared with those of isolated cells, indicating an important role of low affinity sites in the glucagon action.
    These results indicate that insulin and glucagon receptors were reciprocally changed in cultured cells, as compared with isolated cells.
  • Nobuko TAKEI, Tokuhiko HIGASHI
    1985 年 97 巻 2 号 p. 685-692
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    Rat liver catalase was found to interact with deoxycholate (DOC). When purified, the peroxisomal catalase was precipitated at pH 6 in the presence of DOC, whereas in the peroxisomal extract (with DOC) it was unsedimentable at pH 6. The membrane fraction in the extract interacted with the catalase instead of DOC, and prevented the precipitation of catalase with DOC at pH 6. The peroxisomal catalase seemed to be easily modified by lysosomal protease during manipulation, and this proteolytic cleavage rendered the molecule able to interact with the membrane.
    On the other hand, the cytosolic catalase, both in the cytosol fraction and in the purified preparation, sedimented at pH 6 in the presence of DOC. The cytosolic catalase was far more resistant to proteolytic modification than the peroxisomal catalase. The molecule of peroxisomal catalase is assumed to have a site for recognizing the membrane, whereas such a structure may be absent in the cytosolic catalase or may not be easily exposed by proteolytic cleavage.
  • Setsuro EBASHI
    1985 年 97 巻 2 号 p. 693-695
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    A simple method for preparing actin from chicken gizzard was described. This method takes advantage of a property of gizzard tropomyosin, that is, that it does not form Mg paracrystals readily.
  • Kenji IKEHARA, Eriko KAMITANI, Chika KOARATA, Akemi OGURA
    1985 年 97 巻 2 号 p. 697-700
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    A stringent response was induced in Bacillus subtilis vegetative cells by streptomycin. This was confirmed as follows: In B. subtilis stringent cells (BR 16 S), stable RNA synthesis was repressed, and pppGpp and ppGpp were transiently synthesized in the presence of required amino acids and streptomycin. However, these phenomena were not observed in the isogenic relaxed strain (BR 16 R) under the same conditions. On the other hand, tetracyclines did not induce the response, and, moreover, the stringent response to streptomycin upon pretreatment of the stringent cells with the antibiotics was released.
  • Susumu TSUNASAWA, Jun KONDO, Fumio SAKIYAMA
    1985 年 97 巻 2 号 p. 701-704
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    The phenylthiohydantoin (PTH) derivatives of protein amino acids have been separated by reverse-phase high performance liquid chromatography (HPLC) on a fully end-capped C18 column using an isocratic solvent system. The developing solvent was 0.01M sodium acetate buffer (pH 4.5) containing 39.5% acetonitrile and 0.02% sodium dodecylsulfate (SDS). With an automated liquid chromatography equipped with a dual-channel detector, operating at 254 and 313 nm, the present isocratic separation system was quite useful for routine microanalysis of PTH-amino acids released with a “gas-phase” sequencer. The time for one run was _??_23min and the limit of analysis _??_2.5 pmol of a PTH-amino acid.
  • Shin NAKAMURA
    1985 年 97 巻 2 号 p. 705-707
    発行日: 1985年
    公開日: 2008/11/18
    ジャーナル フリー
    A sensitive and quantitative amidolytic assay for thromboplastin (tissue factor) coupled to thrombin formation was established. A fluogenic peptide substrate, Boc-Val-Pro-Arg-MCA, was found to be suitable for the coupled amidolytic assay. The amidolytic assay was applied to measure TF activity of endotoxin-stimulated mononuclear leukocytes and monocytes. The amidolytic assay showed good corre-lation of 0.97 with the currently used clotting assay upon measuring TF activity of the cellular samples.
feedback
Top