The Journal of Biochemistry
Online ISSN : 1756-2651
Print ISSN : 0021-924X
Volume 83, Issue 2
Displaying 1-38 of 38 articles from this issue
  • Osamu TARUTANI, Toshihiko KONDO, Daniel J. SMITH, Sidney SHULMAN
    1978 Volume 83 Issue 2 Pages 333-340
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    A 4S protein which is thyroglobulin-related and which is immunologically distinguished from unusually iodinated albumin-like proteins or impaired substances of thyroglobulin synthesis, was found in the thyroid-soluble fraction even in the normal gland, though in minute amounts. In the present study, purification and characterization of the substance has been carried out using human and hog thyroid glands.
    This protein contains some of the immunospecific determinants of thyroglobulin, although the amino acid composition of the protein differs from that of thyroglobulin. The 4S protein has a molecular weight of about 58, 500 as determined by the sedimentation equilibrium method. Interestingly, the present 4S protein is eluted just in front of, or together with, thyroglobulin by gradient elution chromatography on a DEAE-cellulose column, and it is precipitated with thyroglobulin at 1.9M ammonium sulfate. Therefore, it was concluded that a molecular-sieving process is necessary for the purification of thyroglobulin, in addition to a complex procedure which consists of ammonium sulfate fractionation and DEAE-cellulose chromatography.
    Download PDF (1192K)
  • Jun-ichi HAYASHI, Tamio HIRABAYASHI
    1978 Volume 83 Issue 2 Pages 341-348
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    1. Tropomyosin, one of the regulatory proteins in muscle contraction, was prepared from chickens, rabbits, frogs, shrimps, and shellfish, and conserved characteristics were studied using an enzymological technique.
    2. All tropomyosins tested, irrespective of their sources, were found to have the ability to mediate the inhibitory activity of rabbit troponin toward rabbit Mg2+-activated actomyosin ATPase (Mg2+-ATPase) activity in the absence of Ca2+ ions.
    3. The effect of tropomyosin on the Mg2+-ATPase activity in the presence of Ca2+ ions varied, depending on the source, and this variation appeared to reflect the evolutionary course of this protein.
    4. Tropomyosin from shellfish adductor muscle had the ability to bind to rabbit skeletal muscle troponin and actin. This ability is also considered to be a basic characteristic of tropomyosin which has been conserved during evolution.
    Download PDF (2553K)
  • Kiyokazu MORIOKA, Tetsuo ONO
    1978 Volume 83 Issue 2 Pages 349-356
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    A novel component which specifically binds butyrate was found in rat and mouse liver. This component, termed butyrate binding protein (BBP), is localized in the cytosolic fraction and exhibits protein characteristics, such as heat- and protease-sensitivity. The size of BBP was found to be 7.6S, while it was converted to subunits of 45, 000-48, 000 dalton by treatment with sodium dodecyl sulfate. The dissociation constant of the binding of BBP with butyrate was 2.22 × 10-6 M in the standard assay. 30-Fold purification of BBP was achieved by batchwise adsorption and elution from CM-cellulose and hydroxylapatite column chromatography. BBP is clearly distinguishable from the fatty acid-binding protein found previously on the basis of its size and binding specificity.
    Download PDF (562K)
  • Masateru SHIN, Reiko OSHINO
    1978 Volume 83 Issue 2 Pages 357-361
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Ferredoxin immobilized on Sepharose 4B was prepared by reaction of CNBr-Sepharose 4B with spinach ferredoxin. The ferredoxin-Sepharose 4B conjugated ferredoxin-NADP+ reductase (NADPH: ferredoxin oxidoreductase, [EC 1. 6. 7. 1]) in dilute buffer solution and released it in high salt concentrations. A novel method of preparation for the reductase was established by a combination of affinity adsorption on the ferredoxin-Sepharose 4B column with usual purification procedures.
    It was found using the new method, that there are two forms of ferredoxin-NADP+ reductase, FNR I and FNR II, in spinach. Comparative studies of the two components suggest that FNR I may be a dimer of FNR II.
    Download PDF (377K)
  • Takashi MURAMATSU, Norio KOIDE, Ken-ichi MAEYAMA
    1978 Volume 83 Issue 2 Pages 363-370
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The purification procedure for endo-β-N-acetylglucosaminidase D was improved to yield an enzyme preparation which was homogeneous upon gel electrophoresis. The molecular weight of the enzyme as estimated by Sephadex G-200 column chromatography was 280, 000, while SDS-gel electrophoresis after reduction with 2-mercaptoethanol gave a value of 150, 000. The purified enzyme did not show any chitinase, hyaluronidase or lysozyme activity. In the presence of exoglycosidases removing peripheral sugars, the endoglycosidase acted on serum glycoproteins such as transferrin and fetuin. The enzyme also hydrolyzed an oligosaccharide, (Man)5(GlcNAc)2, indicating that the peptide portion of substrates does not have much effect on susceptibility to the enzyme.
    Download PDF (1246K)
  • Kenji TAKASE, Ryoya NIKI, Shunrokuro ARIMA
    1978 Volume 83 Issue 2 Pages 371-378
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The environment of tryptophan residues in α-lactalbumin was determined by comparative study of bovine, human, and guinea-pig α-lactalbumins in terms of solvent perturbation and denaturation. The results were interpreted on the basis of the fact that Trp-26 and Trp-60 of bovine α-lactalbumin are replaced by other amino acids in human and guinea-pig α-lact-albumins, respectively, and on the assumption that the three α-lactalbumins have similar conformations. Solvent perturbation measurements using ethylene glycol, glycerol, and dimethylsulfoxide as perturbants showed almost the same degree of exposure of tryptophan residues among the three α-lactalbumins at pH 7, and 25°C and 2°C, suggesting that Trp-26 is buried in the interior of the molecule and is inaccessible to solvent, and that Trp-60 is almost buried. A cationic detergent, cetyldimethylbenzylammonium chloride (CDBA) caused an abnormal difference spectrum of α-lactalbumin, which has a major positive peak at about 302 nm characteristic of the native structure of α-lactalbumin. The similarity of the spectrum of α-lactalbumin to that of lysozyme was considered to suggest that α-lactalbumin has a cleft-like region homologous with the active site cleft of lysozyme. By comparison of the ratio of the Δε value of the main peak to that for N-acetyl-L-tryptophan ethyl ester among the three α-lactalbumins, it was found that two tryptophan residues, Trp-60 and either Trp-104 or Trp-118, are located in this region.
    The denaturations induced by guanidine hydrochloride, acid, and heat were followed by difference spectroscopy. The magnitude of the denaturation blue shift at 293 nm of human α-lactalbumin was found to be the smallest, which indicates that Trp-26 contributes to the blue shift and is thus in a nonpolar environment in the native state. The positive peak at about 303 nm observed for bovine and human α-lactalbumins is greatly reduced for guinea-pig α-lactalbumin, and it was considered that Trp-60 is close to charged groups.
    These results are consistent with the proposed conformations of α-lactalbumin based on its similarity to hen egg-white lysozyme, and support the validity of the proposed molecular models.
    Download PDF (571K)
  • Kyuichiro OKUDA, Yohko ATSUTA
    1978 Volume 83 Issue 2 Pages 379-383
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    5β-[11, 12-3H] Cholestane-3α, 7α-diol was synthesized as follows. 5β-Cholestane-3α, 7α, 12α-triol 3, 7-diacetate was treated with phosphorus oxychloride in pyridine solution and then the product, 5β-cholest-11-ene-3α, 7α-diol diacetate, was hydrogenated in acetic acid solution using platinum oxide as a catalyst under an atmosphere of tritium gas.
    5β-[11, 12-3H] Cholestane-3α, 7α-diol thus obtained was readily hydroxylated at C-26 by mitochondria in the presence of isocitric acid, magnesium chloride and potassium cyanide.
    Download PDF (309K)
  • Ryuichi MATSUNO, Toshihiko SUGANUMA, Haruko FUJIMORI, Kazuhiro NAKANIS ...
    1978 Volume 83 Issue 2 Pages 385-394
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    In order to analyze complex experimental results of amylase-catalyzed reactions, which include transglycosylation and condensation as well as hydrolysis, with amylose and linear oligosaccharides as substrates, a general equation governing the concentrations of substrate and product species in a reaction mixture is derived based on the subsite theory. The equation is devised so as to follow the location in products of radioactive glucose which was originally located at a certain position in a substrate. The reaction scheme assumed is as follows: 1) Enzyme can bind any substrate reversibly to form an enzyme-substrate complex, ES. 2) From the ES complex, a glucosyl bond or C1-OH bond at the reducing end of the substrate is cleaved to form a reactive intermediate, releasing a product of shorter chain or H2O. 3) The reactive intermediate binds an acceptor molecule reversibly to form a ternary complex from which the product of transglycosylation or condensation is produced. The binding constants between enzyme and substrate, and between reactive intermediate and acceptor can be estimated from the subsite affinities. The rate constants of catalytic processes such as cleavage of a glucosyl bond and C1-OH bond, hydrolysis and transglycosylation are assumed to be independent of the degree of polymerization as well as of the binding mode of the substrate.
    The applicability of this scheme and rate equation was tested successfully with the Takaamylase A-catalyzed reaction of maltohexaose labeled with 14C at the reducing end.
    Download PDF (629K)
  • Michiko IIJIMA
    1978 Volume 83 Issue 2 Pages 395-402
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The effects of pyocin R1 on transport systems and syntheses of macromolecules in Pseudonwonas aeruginosa strain P14 were studied under two conditions, 0.4mM and 40mM magnesium ions. In the presence of 0.4mM magnesium ions, the incorporation of [14C]leucine into protein and the uptake of [14C]leucine into the leucine pool were completely inhibited by pyocin. Furthermore, [14C]leucine which had been taken up into the leucine pool was depleted by pyocin. Comparing the concentration of amino acid within cells with that in the medium, it was concluded that the active transport in P14 cells was inhibited completely by pyocin. Such an effect of pyocin R1 on transport systems was not specific to leucine but common to many amino acids and uridine. On the other hand, under the condition of 40mM magnesium ions, [14C]-leucine once accumulated was not depleted by pyocin, but uptake of [14C]leucine into the leucine pool stopped immediately while incorporation of [14C]leucine into protein continued at a considerably reduced rate; therefore the machinery for protein synthesis was not directly inhibited by pyocin. It is suggested that the loss from the amino acid pool in cells brings about the rapid cessation of protein synthesis under the condition of 0.4mM magnesium.
    Download PDF (615K)
  • F-Actin-Inactivated Hydrolysis Showing Initial Burst
    Hisao FUJISAKI, Hiroshi ASAI
    1978 Volume 83 Issue 2 Pages 403-407
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The initial burst of P1; liberation was found in the hydrolysis of β-naphthyl triphosphate (β-NapP3) by heavy meromyosin (HMM) in the presence of Mg ions as well as in the hydrolysis of ATP. However, unlike that of ATP, the steady-state hydrolysis of β-NapP3 by HMM was inhibited by the addition of F-actin to the reaction solution. Although the possession of an initial burst-like property during interaction of a substrate and myosin is believed by many investigators to be a key factor in F-actin activation of substrate hydrolysis in vitro and in the molecular mechanism of muscle contraction, the above results suggest that this is not generally true. β-NapP3 did not induce superprecipitation of actomyosin solution and suppressed ATP-induced superprecipitation.
    Download PDF (366K)
  • Isamu SHIIO, Hachiro OZAKI
    1978 Volume 83 Issue 2 Pages 409-421
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The distribution of 5'-nucleotidases of Bacillus subtilis K No. 231, an inosine-producing bacterium, was investigated. It was found that this organism has an extracellular 5'-nucleotidase as well as various cellular 5'-nucleotidases, which include a major one bound to the cell wall and minor ones in the cytoplasm and on the cytoplasmic membrane. The cell wallbound 5'-nucleotidase was specifically solubilized by treatment of the sonic cell debris with lysozyme, but not by treatment with various other hydrolyzing enzymes, detergents or high concentrations of inorganic salts. The extracellular enzyme was more repressible by inorganic phosphate than the cellular ones. Among the three cellular 5'-nucleotidases the cytoplasmic enzyme was the most repressible by inorganic phosphate. The crude preparations of the extracellular, cell wall-bound and cytoplasmic 5'-nucleotidase showed single peaks of activity on gel filtration, at the elution volumes corresponding to molecular weights of 67, 000, 900, 000 and 600, 000, respectively. Upon DEAE-cellulose column chromatography of the peak fractions of the three preparations, two peaks (E1 and E2), three peaks (W1, W2, and W3) and two peaks (C1 and C2) were obtained, respectively. The molecular weights of E1, E2, W1, W2. W3, C1, and C2 were determined to be 27, 000, 46, 000, 37, 000, 21, 000, 640, 000, 38, 000, and 260, 000, respectively. All seven enzymes showed optimum pH between 7.0 and 8.0, and specifically hydrolyzed the phosphomonoester bonds of purine nucleoside-5'-monophosphates but not those of 5'-pyrimidine nucleotides. The cellular enzymes shoved slight activity toward UDP-glucose whereas the extracellular enzymes did not. None of the enzymes hydrolyzed 4-nitrophenylphosphate or bis(4-itrophenyl)phosphate. All the enzymes showed the same Km values, 1.5-1.9, μM for AMP. High concentrations (0.1M and 1M) of NaCl, KCl, and (NH4)2SO4 activated the cellular 5'-nucleotidases more strongly than the extracellular enzymes. Mg2+ and Mn2+ stimulated the activities of the cytoplasmic enzymes, C1 and C2, specifically. All the enzymes were inhibited by Zn2+, Cu2+, ATP, and PO42-. The PO42- inhibition was of mixed type with respect to AMP, and the inhibitor constant for PO42- was almost the same, 0.4-1.1mM, for all the enzymes.
    Download PDF (906K)
  • Formation of EF-1α•EF-1βγ and EF-α• EF-1β Complexes
    Shigekazu NAGATA, Kazuo MOTOYOSHI, Kentaro IWASAKI
    1978 Volume 83 Issue 2 Pages 423-429
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    In the preceding papers, we showed that one of the two complementar factors of polypeptide chain elongation factor 1 (EF-1) from pig liver, EF-1α, functionally corresponds to bacterial EF-Tu (Nagata, S., Iwasaki, K., and Kaziro, Y. (1976) Arch. Biochem. Biophys. 172, 168), while the other, EF-1βγ, as well as one of its subunits, EF-1β, corresponds to bacterial EF-Ts (Motoyoshi, K. and Iwasaki, K. (1977) J. Biochem. 82, 703). Therefore, the interaction between EF-1α and EF-1β, corresponds to bacterial EF-Ts (Motoyoshi, K. and Iwasaki, K. (1977) J. Biochem. 82, 703). Therefore, the interaction between EF-1α and EF-1βγ or EF-1β was examined and the following results were obtained. i) EF-1βγ catalytically promoted the exchange of [14C]GDP bound to EF-1α with exogenous [3H] GDP. ii) In the absence of the exogenous guanine nucleotide, EF-1βγ as well as β could displace GDP bound to EF-1αto form an EF-1α•EF-1βγ as well as an EF-1α•EF-1β complex. iii) The occurrence of α•EF-1βγ and EF-1α•EF-1β complexes was demonstrated by gel filtration on Sephadex G-150. These results strongly indicate that the mechanism of the action of EF-1βγ or EF-1β in converting EF-1α•GDP into EF-1α•GTP is analogous to bacterial EF-Ts, and the reaction is accomplished by the following reactions;
    EF-1α•GDP+FE-1βγ(or EF-1β)_??_EF-1α•EF-1βγ(or EF-1β)+GDP
    EF-1α•EF-1βγ(OR EF-1β)+GTP_??_EF-1α•GTP+EF-1βγ(or EF-1β).
    Download PDF (567K)
  • Yoichiro HIROKATA, Akio SHIGEMATSU, Tsuneo OMURA
    1978 Volume 83 Issue 2 Pages 431-440
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    NADH could support the lipid peroxidation of rat liver microsomes in the presence of ferric ions chelated by ADP (ADP-Fe). The reaction had a broad pH optimum (pH 5.8-7.4) and was more active in the acidic pH range.
    Antibodies to NADH-cytochrome b5 reductase [EC 1. 6. 2. 2] and cytochrome b5 inhibited NADH-dependent lipid peroxidation in the presence of ADP-Fe, whereas the antibody against NADPH-cytochrome c reductase [EC 1. 6. 2. 4] showed no inhibition. These observations suggest that the electron from NADH was supplied to the lipid peroxidation reaction via NADH-cytochrome b5 reductase and cytochrome b5. On the other hand, NADPH-supported
    lipid peroxidation was strongly inhibited by the antibody against NADPH-cytochrome c reductase, confirming the participation of this flavoprotein in the NADPH-dependent reaction.
    In the presence of both ADP-Fe and ferric ions chelated by EDTA (EDTA-Fe), NADH-de-pendent lipid peroxidation was highly stimulated up to the level of the NADPH-dependent reaction. In this case, the antibody against cytochrome b5 could not inhibit the reaction, while the antibody against NADH-cytochrome b5 reductase did inhibit it, suggesting the direct transfer of electrons from NADH-cytochrome b5 reductase to EDTA-Fe complex.
    Download PDF (1349K)
  • VIII. Purification and Characterization of Cathepsins D from Japanese Monkey Lung
    Akihiko MORIYAMA, Kenji TAKAHASHI
    1978 Volume 83 Issue 2 Pages 441-451
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Two kinds of cathepsin D were found in Japanese monkey lung and were named cathepsins D-I and D-II. Cathepsin D-I was partially purified by ammonium sulfate fractionation and DEAE-cellulose column chromatography. It had properties common to other ordinary cathepsins D in terms of the elution position from a DEAE-cellulose column at pH 8.0, the pH-dependence of activity toward acid-denatured hemoglobin, and the molecular weight of 35, 000 as determined by Sephadex G-100 gel filtration.
    On the other hand, cathepsin D-II was purified about 1, 000-fold by a combination of ammonium sulfate fractionation and column chromatographies on DEAE-cellulose and Sephadex G-100. It was a very acidic protein as judged from its elution position from a DEAE-cellulose column at pH 8.0, and the high mobility toward the anode on disc gel electrophoresis at pH 8.6. Its molecular weight was determined to be 35, 000 by Sephadex G-100 gel filtration and 39, 000 by SDS-polyacrylamide gel electrophoresis. It was optimally active at pH 2.8 against acid-denatured hemoglobin as a substrate, showing 80% of the optimal activity at pH 1.0, and almost no activity above pH 4.0. This pH-profile of activity was similar to that of monkey pepsin C (gastricsin). It did not hydrolyze N-acetyl-L-phenylalanyl-3, 5-diiodo-L-tyrosine, a synthetic substrate for pepsin, but was inhibited by a series of pepsin inhibitors such as pepstatin, 1, 2-epoxy-3-(p-nitrophenoxy)propane, p-bromophenacyl bromide, and diazoacetyl-DL-norleucine methyl ester, although the diazo reagent was a rather weak inhibitor of the enzyme. The amino acid composition of cathepsin D-II was found to be fairly different from those of other cathepsins D. However, it showed a striking resemblance to that of Japanese monkey pepsinogen C, suggesting some evolutionary relationship between them.
    Download PDF (1152K)
  • Atsushi MIYAJIMA, Yoshito KAZIRO
    1978 Volume 83 Issue 2 Pages 453-462
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The amounts of the polypeptide chain elongation factors Tu, 1s, and G, and ribosomal protein S1 were assessed under various growth conditions using three independent procedures: (a) Immunoprecipitation and gel electrophoresis, (b) radioimmune assay, and (c) activity measurements. It was demonstrated that, during balanced growth of E. coli, the intracellular levels of these proteins increased in proportion to the growth rate, and the ratio of EF-Tu: EF-Ts: EF-G: protein SI was 4-5: 1:1:1, at all growth rates.
    The effects of isoleucine starvation on the rates of synthesis of these proteins were examined using a pair of isogenic stringent and relaxed strains. The syntheses of all these proteins were found to be under the influence of stringent control.
    These results indicate that in E. coli the syntheses of the above four proteins are regulated in a coordinated manner and are subject to stringent control.
    Download PDF (797K)
  • Kei-ichi UEMURA, Matsuko YUZAWA, Tamotsu TAKETOMI
    1978 Volume 83 Issue 2 Pages 463-471
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Several neutral glycolipids and gangliosides were isolated from bovine erythrocyte stroma. Their structures were determined by partial acid hydrolysis, methylation analysis, periodate oxidation and CrO3 oxidation. Two major neutral glycolipids were characterized as lactosylceramide and galactosyl (α1-3) galactosyl (β1-4) N-acetylglucosaminyl (β1-3) galactosyl (β1-4) glucosyl (β1-1) ceramide. Two major gangliosides were N-glycolylneuraminosyl (2-3) galactosyl (β1-4) glucosyl (β1-1) ceramide, and N-glycolylneuraminosyl (2-3) galactosyl (β1-4) N-acetylglucosaminyl (βl-3) galactosyl (β1-4) glucosyl (β1-1) ceramide. Minor glycolipids were glucosyl- and galactosylceramide, glucosamine-containing tri- and tetraglycosylceramide, glucosamine-containing disialosylhexaglycosylceramide, and gangliosides containing N-acetylneuraminic acid. The ceramide moiety of each glycolipid contained predominantly d18:1 sphingosine, and normal fatty acids of C16:0, C22:0, C24:0, and C24:1.
    Download PDF (2277K)
  • Nobuyuki SUETSUGU, Ken'ichi TAKEO, Yutaka SANAI, Takashi KUGE
    1978 Volume 83 Issue 2 Pages 473-478
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Sweet potato β-amylase [EC 3. 2. 1. 2, α-1, 4-D-glucan maltohydrolase]-catalyzed hydrolyses of aryl β-maltotriosides with substituents, NO2-, Cl-, and Br- at the o-, m-, and p-positions in the phenyl ring were studied at pH 4.8 and 25°C. The hydrolyses of a few of the maltotriosides by soybean *-amylase [EC 3. 2. 1. 2, α-1, 4-D-glucan maltohydrolase] were also studied at pH 5.4 and 25°C. It was found that the aryl β-maltotriosides were preferentially hydrolyzed into maltose and aryl β-D-glucosides by both β-amylases.
    The Michaelis constant Km and the molecular activity k0 were determined for the hydrolyses of these maltotriosides and compared with those of maltotriose and maltotetraose. Aryl β-maltotriosides were more rapidly hydrolyzed than maltotriose by a factor of 30-80, and more slowly hydrolyzed than maltotetraose by a factor of 10-30, depending on the kinds of substituents.
    The rapid hydrolysis of aryl, β-maltotrioside as compared with maltotriose may be due to the interaction of an aryl group with the subsite of β-amylase. This is in contrast with glucoamylase [EC 3. 2. 1. 3, α-1, 4-D-glucan glucohydrolase] of Rhizopus niveus-catalyzed hydrolysis of phenyl β-maltoside, whose phenyl group does not interact so much with the subsite of the enzyme.
    Download PDF (456K)
  • Further Purification, Characterization, and Analysis of the Activation and Inactivation
    Kenji YAMAGUCHI, Yu HOSOKAWA, Noriyuki KOHASHI, Yoshihiko KORI, Shigek ...
    1978 Volume 83 Issue 2 Pages 479-491
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Rat liver cysteine dioxygenase has been purified to homogeneity. It is a single subunit protein having a molecular weight of 22, 500+1, 000, with a pI of 5.5. The enzyme purified was catalytically inactive and activated by anaerobic incubation with either L-cysteine or its analogues such as carboxymethyl-L-cysteine, carboxyethyl-L-cysteine, S-methyl-L-cysteine, D-cysteine, cysteamine, N-acetyl-L-cysteine, and DL-homocysteine. The enzyme thus activated with L-cysteine was rapidly inactivated under aerobic condition. This rapid inactivation was observed at 0°C where no formation of either the reaction product cysteine sulfinate or the autoxidation product of cysteine, cystine, was detected. Further analysis shows that the inactivation of the activated enzyme was due to oxygen but unrelated to either the presence of substrate, enzyme turnover or accumulation of inhibitor produced during assay. A distinct rat liver cytoplasmic protein, called protein-A, could completely prevented the enzyme from the aerobic inactivation. The loss of activity during assay in the absence of protein-A was shown to be a first order decay process. From the plots of log(Δproduct/min) versus time, the initial velocity (V0) and the velocity at 7 min (V7) were obtained. The apparent Km value for L-cysteine in the absence of protein-A was calculated from the initial velocity as 4.5×10-4M. Protein-A did not alter the apparent Km value for L-cysteine. The chelating agents such as o-phenanthroline, α, α'-dipyridyl, bathophenanthroline, 8-hydroxyquinoline, EGTA, and EDTA strongly inhibited the enzyme activity when these chelating agents were added before preactivation. The purified cysteine dioxygenase contains 1 atom of iron per mol of enzyme protein. By the activation procedure, the enzyme became less susceptible to the heat denaturation, the inhibitory effects of chelating agents and the tryptic digestion.
    Download PDF (977K)
  • Susumu UJITA
    1978 Volume 83 Issue 2 Pages 493-502
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Fructose 1, 6-bisphosphate aldolase [EC 4. 1. 2. 13] from spores of Bacillus subtilis PCI 219 was purified to a nearly homogeneous state; the corresponding enzyme from vegetative cells was also partially purified, and their properties were compared.
    The molecular weight of the aldolase from spores was similar to that of the enzyme from vegetative cells; they were found to be 150, 000 ±20, 000 by Sephadex G-200 gel filtration, 130, 000 ±5, 000 by disc gel electrophoresis, and 120, 000 ±5, 000 by sucrose density gradient centrifugation. The molecular weights of subunits were estimated to be 30, 000 ±2, 000 by SDS-gel electrophoresis for both vegetative cell and spore aldolases. The enzyme purified from vegetative cells in the presence of diisopropylfluorophosphate also gave the same results.
    The enzymes from vegetative cells and spores were both inhibited by metal chelators but not by treatment with sodium borohydride in the presence of the substrate. For these enzymes, the pH optima were 7. 5, the Km values were 2mM for fructose 1, 6-bisphosphate, and fructose 1-phosphate was not utilized as a substrate. These results suggest that Bacillus subtilis PCI 219 aldolases can be classified as Class II aldolases. No differences in properties between these vegetative cell and spore enzymes were detected as regards inhibition by calcium ions and p-chloromercuribenzoate, or activation by sulfhydryl compounds.
    During the course of growth and sporulation, a constant level of the aldolase was detected, and a single active band was obtained by gel electrophoresis.
    Download PDF (2513K)
  • Susumu UJITA, Teruaki SHIROZA, Kinuko KIMURA
    1978 Volume 83 Issue 2 Pages 503-510
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Vegetative cell and spore fructose 1, 6-bisphosphate aldolases [EC 4. 1. 2. 13] were partially purified from Bacillus subtilis 168, Bacillus licheniformis and Bacillus cereus IFO 3466, and their properties were compared with those of the aldolases previously purified from Bacillus subtilis PCI 219.
    The molecular weights of aldolases from B. subtilis 168 and B. licheniformis were similar to those of the aldolases from B. subtilis PCI 219 and were found to be 150, 000±20, 000 by gel filtration, but the molecular weights of B. cereus aldolases were smaller than those of B. subtilis PCI 219 aldolases, being 60, 000±3, 000. However, no differences between the vegetative cell aldolase and the spore aldolases were detected even in the case of B. cereus IFO 3466. On the other hand, the subunit sizes of their aldolases were estimated by SDS-gel electrophoresis and all of the subunit sizes of vegetative cell and spore aldolases from the bacilli were found to be similar, 30, 000±2, 000.
    Differences between the aldolases of B. subtilis PCI 219 and B. cereus IFO 3466 were found in electrophoretic mobilities, but all of the aldolases resembled each other in such properties as the inhibition by metal chelators and a sulfhydryl reagent. No differences between vegetative cell and spore aldolases from each bacilli were observed except for the thermal stability of B. cereus IFO 3466 aldolases.
    The molecular weights of aldolases from B. subtilis PCI 219 and B. cereus IFO 3466 in the presence of substrate were also estimated by means of sucrose density gradient centrifugation. The molecular weight of each aldolase did not change; the same values as in the absence of substrate were obtained.
    Download PDF (1135K)
  • III. Chemical Composition of Microsomal and Lysosomal β-Glucuronidases Purified from Rat Liver
    Masaru HIMENO, Yukio NISHIMURA, Keikichi TAKAHASHI, Keitaro KATO
    1978 Volume 83 Issue 2 Pages 511-518
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The microsomal (pI 6.9-7.6) and lysosomal (pI 5.8-6.6) β-glucuronidases [EC 3. 2. 1. 31] purified from rat liver by the method of Himeno et al. (I) were subjected to chemical analyses. The enzymes had very similar amino acid compositions and did not contain sulfur-containing amino acids. The microsomal β-glucuronidase contained 6.5% carbohydrate, consisting of mannose, glucose, fucose, galactose, and glucosamine in a ratio of 69; 12; 3; 3; 25. Sialic acid was not detected in this enzyme. The lysosomal enzyme contained 5.9% carbohydrate, consisting of mannose, glucose, fucose, galactose, and glucosamine in a ratio of 57; 14; 3; 4; 23, with a trace of sialic acid. To confirm the presence of sialic acid in the lysosomal enzyme, more acidic forms of lysosomal β-glucuronidase with PI values ranging from pH 5.4 to 6.1 were also purified. They contained 0.11% sialic acid and were neuraminidase-sensitive. This suggests that a very small part of the lysosomal enzyme may be a sialoglycoprotein. Since it has been suggested that sialic acid in the lysosomal glycoproteins is cleaved rapidly by autolytic degradation (Goldstone, A. & Koenig, H. (1974) Biochem. J. 141, 527-535), the sugar acceptor activity of the purified lysosomal β-glucuronidase was determined using sialyltransferase in the Golgi fraction of rat liver. As the lysosomal β-glucuronidase was not a substrate for the sialyltransferase, the possibility of cleavage of terminal sialic acid by autolytic degradation is probably ruled out. The microsomal β-glucuronidase was also not a substrate for the sialyltransferase. This suggests that sialylation may occur very seldom in β-glucuronidase in the Golgi region. The relationship between the carbohydrate moieties of the microsomal and lysosomal βglucuronidases is discussed in connection with intracellular transport of β-glucuronidase.
    Download PDF (1121K)
  • Koichi OGAWA, Akira ICHIHARA
    1978 Volume 83 Issue 2 Pages 519-525
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    A synthetic polypeptide, poly-L-ornithine (pLo; average molecular weight, 13, 000), inhibited protein synthesis in mouse L cells in suspension culture. The inhibition was dependent upon both the concentration of pLo and the cell density, and the dose for half-inhibition was correlated with the cell density. Cell viability remained high in the concentration range of PLo causing inhibition of protein synthesis. This inhibition was prevented by addition of calf serum with pLo, but not by washing the treated cells with serum or a high concentration of salt.
    PLo had no effect on protein synthesis in a cell-free system prepared from L cells or rabbit reticulocytes. The polysome profiles of cells treated with pLo were similar to those of control cells. Experiments on the effect of interchange of the ribosomes and supernatants of control and pLo-treated cells showed that the ribosomes from pLo-treated cells were inactive. The inactivation of these ribosomes was partially prevented by the presence of 120mM K+ in the medium during pLo treatment. The inhibition of protein synthesis by pLo, therefore, may result from the binding of pLo to the cell membranes, causing leakage of intracellular K+ and thus inactivating ribosomes.
    Download PDF (513K)
  • Michio OGURA, Midori GOTO, Yoshikazu AYAKI
    1978 Volume 83 Issue 2 Pages 527-535
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The formation of bile acids was studied in isolated rat livers perfused with Krebs-Ringer bicarbonate buffer (pH 7.4)-5.5mM glucose oxygenated with O2/CO2 (95:5).
    When livers were perfused with 150ml of the medium in a recirculating system, bile was secreted at a rate of around 30mg/g liver/h during a 3-h period. Gas chromatographic analysis of biliary bile acid components gave the following results. As regards cholic acid, after that present in the hepatic pool was washed out during the initial 0.5 h, secretion of de novo cholic acid at a level of about 3nmol/g liver/h lasted for at least 2h. The pattern of bile acid was essentially similar to that in bile-fistula rats. However, secretion of β-muricholic acid relative to cholic acid increased more than 60% in the perfusion period of 1.5-2.5h compared with that in the period of 0.5-1.0h.
    In tracer experiments with [24-*C]chenodeoxycholic acid, no accelerated incorporation into &- and β-muricholic acids was observed in the perfusion period of 1.5-2.5h, as compared with that of 0.5-1.5h. Specific radioactivity of muricholic acids, especially of β-muricholic acid, was far lower than that of chenodeoxycholic acid.
    These data suggest the existence of a compartmentalized pool of endogenous chenodeoxycholic acid.
    Download PDF (682K)
  • Reiko YAMAMOTO, Hiroshi MUNAKATA, Masahiko YAMAMOTO, Zensaku YOSIZAWA
    1978 Volume 83 Issue 2 Pages 537-542
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    A crude microsomal fraction (M-Fr) was separated from the endometrial scrapings of uteri of ovariectomized rabbits with or without hormonal treatment. The effects of estrogen and progesterone on the incorporation into M-Fr of L-[U-14C]-fucose and N-acetyl-D-[6-3H]-glucosamine from their nucleotides were investigated. Estrogen increased the incorporation of these sugars, whereas progesterone suppressed this effect. The results of fractionation on a DEAE-Sephadex A-25 (Cl- form) column of the isotope-labelled complex saccharide mixtures, obtained by pronase digestion of the incubation mixtures, indicated that biosynthesis of sulfated glycoprotein was most sensitive to the hormones among the complex saccharides in M-Fr. Thus, a hormonal effects on the biosynthesis of sulfated glycoprotein in the endometrium of ovariectomized rabbit has been unambiguously confirmed at the microsomal level.
    Download PDF (402K)
  • Tadashi NOTO, Tetsunori TANAKA, Teruo NAKAJIMA
    1978 Volume 83 Issue 2 Pages 543-552
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Urinary metabolites of polyamines in rats were studied systematically by the intraperitoneal injection of radioactive polyamines. Urinary metabolites were fractionated into 4 fractions containing non-polar and acidic compounds, acidic and neutral ampholytes, basic ampholytes and polyamines. A large amount of radioactivity was detected in the fractions containing non-polar and acidic compounds and polyamines of urine of rats injected with radioactive putrescine, while in the case of the injection of radioactive spermidine or spermine a relatively large amount of radioactivity was found in the basic ampholyte fraction as well as in the polyamine fraction.
    Analysis of these fractions indicated that γ-aminobutyric acid, N-monoacetylputrescine, 2 (3)-hydroxyputrescine, putreanine, N-(3-aminopropyl)-4-aminobutyric acid (isoputreanine), spermic acid, N-(3-aminopropyl), N'-(2-carboxyethyl)-1, 4-diaminobutane, and N-mono-acetylspermidine A and B were excreted as urinary metabolites of the polyamines in addition to putrescine, spermidine and spermine.
    Download PDF (709K)
  • An Application of the Thiobarbituric Acid Method to the Assay of D-Gluco-4-Enepyranosyluronic Acid-Containing Disaccharides
    Masai KOSEKI, Atsushi KIMURA, Koichi TSURUMI
    1978 Volume 83 Issue 2 Pages 553-558
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    An improved method is described for the micro determination of acidic glycosaminoglycans after digestion with chondroitinase-ABC and -AC.
    The determination is based on the color production of D-gluco-4-enepyranosyluronic acidcontaining disaccharides produced by the action of chondroitinase-ABC and -AC (acidic glycosaminoglycans-endoeliminase) when the periodate-thiobarbituric acid method is applied to the α, β-unsaturated disaccharides. Suitable conditions for the quantitative assay are described.
    Download PDF (445K)
  • Takashi MORITA, Sadaaki IWANAGA
    1978 Volume 83 Issue 2 Pages 559-570
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    A prothrombin activator from Echis carinatus venom was highly purified by gel-filtration on Sephadex G-150 followed by chromatographies on DEAE-Sephadex A-25 and DEAF-cellulose, gel-filtration on Sepharose 6B, and rechromatography on DEAE-cellulose. Through these procedures, about 2mg of the purified material was obtained from one gram of the lyophilized venom and about 57-fold purification was achieved. The purified preparation was found to give a single band on polyacrylamide-gel disc electrophoresis in the presence and absence of sodium dodecyl sulfate (SDS). Moreover, no additional protein band was observed on the reduced SDS-gel, suggesting that the protein consists of a single polypeptide chain.
    The molecular weight of prothrombin activator was estimated to be approximately 56, 000 by SDS-gel electrophoresis and by molecular sieving on Sepharose 6B in the presence of 6M guanidine. The activator appeared to be a glycoprotein, as judged by a positive periodic acid-Schiff reaction against the protein band on SDS-gel and from the amino acid analysis. The isoelectric point was determined to be 4.5 by isoelectric focusing.
    The purified activator showed strict specificity towards prothrombin and its derivative; it did not activate factor X, factor IX, prekallikrein, plasminogen, or trypsinogen from a bovine source. The Km values towards bovine prothrombin and prethrombin I were estimated to be 1.03×10-6M and 1.05×10-6M, respectively. Further, the activator did not hydrolyze any of the synthetic ester and anilide substrates widely used for α-thrombin, factor Xa, and kallikrein; neither caseinolytic nor fibrinogenolytic activities were detected in it. Chelating agents including EDTA, o -phenanthroline, and cysteine strongly inhibited the enzyme activity, while diisopropyl phosphorofluoridate, antithrombin III, and naturally occurring proteinase inhibitors, which inhibit serine proteinases, had no effect. These results indicate that the venom prothrombin activator is not a serine proteinase like coagulation factors found in mammalian plasma, but is one of the metal proteinases. Thus, the enzymatic properties differ remarkably from those of factor Xa.
    Download PDF (1442K)
  • Evidence for Two Kinds of Phosphorylated Intermediate with and without Bound p-Nitrophenol
    Yoichi NAKAMURA, Yuji TONOMURA
    1978 Volume 83 Issue 2 Pages 571-583
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The reaction mechanism of p-nitrophenylphosphatase (PNPPase) of SR was elucidated by analyzing its partial reaction using radioactive PNPP as substrate. The following results were obtained:
    1. Phosphoprotein was produced by the reaction of SR with PNPP at an alkaline pH and in the presence of a high concentration of Ca 2+ ions, where it decomposed very slowly. The amount of phosphoprotein reached 0.2mol/105g in the presence of 10mM CaCl2 and 1mM MgCl2 at pH 8.75 and 15°C. The rate of phosphoprotein formation of PNPPase was much lower than that of ATPase. The amount of phosphoprotein decreased, while the rate of PNP liberation increased, immediately after the pH and/or free Ca2+ concentration was reduced.
    2. When ADP reacted with 32P-labelled phosphoprotein produced from [32P]PNPP, the amount of 32P-labelled phosphoprotein decreased rapidly and [γ-32P] ATP was formed almost stoichiometrically. The pH stability and hydroxylamine sensitivity of phosphoprotein of PNPPase denatured with TCA were the same as those of phosphoprotein of ATPase.
    3. The phosphorylation reaction was started by adding [32P]PNPP and stopped by dilution with a 40-fold volume of the reaction mixture containing unlabelled PNPP, then the time course of the subsequent decrease in the amount of phosphoprotein was measured. The time course was biphasic, indicating the existence of two kinds of phosphoprotein. In the presence of 5mM CaCl2, and 5mM MgCl2 at pH 8.0, the decay constants were about 2min-1 and 0.1min-1, respectively. The ratio of the amounts of fast and slowly decomposing components of the phosphoprotein was about 1:4. The decay rates and the ratio of the two components were unaffected by preincubation of SR with PNP. The decay constant of the slow phase increased when the Mg2+ concentration increased or the Ca2+ concentration
    Download PDF (849K)
  • Isamu TAKAGAHARA, Yasuo SUZUKI, Tuyosi FUJITA, Juniti YAMAUTI, Katsumi ...
    1978 Volume 83 Issue 2 Pages 585-597
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Glucose-6-phosphate dehydrogenase, 6-phosphogluconate dehydrogenase, glutathione reductase and pyruvate kinase of Candida utilis and baker's yeast, when in anionic form, were adsorbed on a cation exchanger, P-cellulose, due to affinities similar to those for the phosphoric groups of their respective substrates; thus, glucose-6-phosphate dehydrogenase was readily eluted by either NADP+ or NADPH, glutathione reductase by NADPH, 6-phosphogluconate dehydrogenase by 6-phosphogluconate, and pyruvate kinase by either ATP or ADP. This type of chromatography may be called “affinity-adsorption-elution chromatography”; the main principle is different from that of so-called affinity-elution chromatography.
    Based on these findings, a large-scale procedure suitable for successive purification of several enzymes having affinities for the phosphoric groups of their substrates was devised. As an example, glucose-6-phosphate dehydrogenase was highly purified from baker's yeast and crystallized.
    Download PDF (930K)
  • Kazuhiko NAGATA, Koichi SUZUKI, Kazutomo IMAHORI
    1978 Volume 83 Issue 2 Pages 599-606
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    A reactive ATP analog, N6-(6-bromoacetamidohexyl)-AMP•PCP, was synthesized in an attempt to covalently label the binding sites for adenine nucleotides, especially ATP, of various enzymes which utilize adenine nucleotides as substrates, cofactors, inhibitors or allosteric effectors. This reagent rapidly inactivated rabbit muscle glyceraldehyde 3-phosphate dehydrogenase (GPD), myokinase (MK), and creatine kinase (CK) under very mild conditions. Adenine nucleotide substrates prevented the inactivation. In the case of GPD, complete inactivation was observed when 1 mol of the reagent per mol of enzyme subunit was incorporated into the enzyme. These results indicate that the present ATP analog may be useful as an affinity labeling reagent for various adenine nucleotide-dependent enzymes.
    Download PDF (627K)
  • Akira OMORI, Mariko ISHIZUKA-KOBAYASHI, Tsuneko UCHIDA
    1978 Volume 83 Issue 2 Pages 607-615
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    1. Four DNases were found in the dried liver extract of a top shell, Turbo cornutus. The major one was purified 120-fold by phosphocellulose column chromatography, sulfoethylcellulose column chromatography and gel-filtration on Sephadex G-150. The yield was 2.7%.
    2. The enzyme activity was not affected by Mg2+ (10-3-10-2M), EDTA (10-3-10-2M), or NaCl (10-1M). It showed a pH optimum of 4.7-4.8. Ionic strength was found to be critical for the maximal activity. The isoelectric point was 8.5-9.0. On heating at 50°C for 5min the enzymic activity fell to half the initial value.
    3. The enzyme preparation degraded native as well as heat-denatured DNA, but not RNA. It degraded heat-denatured DNA endonucleolytically to give oligonucleotides with 3'-phosphates.
    4. The 3'-phosphate and 5'-hydroxy termini of oligonucleotides were investigated. At both the 3'- and 5' -terminal positions, purine nucleotides were predominant.
    Download PDF (673K)
  • Oyin SOMORIN, Jiri SKOREPA
    1978 Volume 83 Issue 2 Pages 617-623
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    An enzyme with characteristics typical of aliesterase has been found in human blood serum using a gas solid chromatographic assay technique. This conflicts with the findings of several authors that aliesterase is absent in the human blood. Another aliesterase is released into the blood stream after intravenous administration of heparin. Partial purification of the aliesterase in normal (preheparin) and postheparin sera was effected by column chromatography using CM- and DEAE-Sephadex. The preheparin aliesterase and postheparin aliesterase have different pH optima of 7.0 and 8.5 respectively. The preheparin aliesterase activity was very sensitive to sodium fluoride and insensitive to a negatively charged detergent, sodium lauryl sulfate, unlike the postheparin esterase which was highly sensitive to sodium lauryl sulfate and comparatively less sensitive to sodium fluoride.
    Download PDF (534K)
  • Nobuhiko KATUNUMA, Nobumasa YASOGAWA, Kiyoaki KITO, Yukihiro SANADA, H ...
    1978 Volume 83 Issue 2 Pages 625-628
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    The activities of serine protease in muscles from normal persons and from patients with progressive muscular and neuromuscular diseases have been determined. A significant increase in the level of serine protease was found in muscle of patients with Duchenne-type muscular dystrophy and with Becker-type muscular dystrophy, but the activity was not increased in muscle of a patient with amyotrophic lateral sclerosis.
    Download PDF (244K)
  • Koyu HON-NAMI, Tairo OSHIMA
    1978 Volume 83 Issue 2 Pages 629-631
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Two cytochromes of the C-type, c-554 and c-549, were isolated from the soluble fraction of an extreme thermophile, Thermus thermophilus HB8. Highly purified cytochrome c-554 had absorption maxima at 554, 522, and 417 nm in the reduced state, and at 410 nm in the oxidized state. The α-band of the reduced state resembled that of “split-α” cytochromes. The isoelectric point was at pH 4.9, and the molecular weight was about 29, 000. Cytochrome c-549, partially purified, had absorption maxima at 549, 520, and 416 not in the reduced form, and at 408 nm in the oxidized form. The molecular weight was about 25, 000. Both were slowly auto-oxidizable, and did not combine with CO.
    Download PDF (154K)
  • Showbu SATO
    1978 Volume 83 Issue 2 Pages 633-635
    Published: February 25, 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    A site specific endonuclease from Thermus aquaticus, Taq I, cleaves Simian virus 40 (SV40) DNA at a single site. The cleavage site was localized on the physical map by double digestions, using the previously characterized fragments produced by digestion with HaeII, HaeIII, AluI, HhaI, HinfI, or BstI. The TaqI site is located at the position that is 56.5% of the unit length from the Eco RI site.
    Download PDF (780K)
  • 1978 Volume 83 Issue 2 Pages 637a
    Published: 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Download PDF (26K)
  • 1978 Volume 83 Issue 2 Pages 637b
    Published: 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Download PDF (26K)
  • 1978 Volume 83 Issue 2 Pages 637c
    Published: 1978
    Released on J-STAGE: November 18, 2008
    JOURNAL FREE ACCESS
    Download PDF (26K)
feedback
Top