Network


Latest external collaboration on country level. Dive into details by clicking on the dots.

Hotspot


Dive into the research topics where Hirohide Toyama is active.

Publication


Featured researches published by Hirohide Toyama.


Advances in Microbial Physiology | 1994

Respiratory chains and bioenergetics of acetic acid bacteria.

Kazunobu Matsushita; Hirohide Toyama; Osao Adachi

Publisher Summary The chapter discusses the respiratory chains and bioenergetics of acetic acid bacteria. Acetic acid bacteria are obligate aerobes and well known as “vinegar producers.” They produce acetic acid from ethanol by two sequential catalytic reactions of membrane-bound alcohol dehydrogenase (ADH) and aldehyde dehydrogenase (ALDH). Besides alcohols and aldehydes, acetic acid bacteria are able to oxidize various sugars and sugar alcohols such as D-glucose, glycerol, and D-sorbitol. Such oxidation reactions are called “oxidative fermentations”, because they involve incomplete oxidation of alcohols or sugars accompanied by accumulation of the corresponding oxidation products in huge amounts in the growth medium. Bacteria capable of effecting oxidative fermentations are called “oxidative bacteria,” of which the most prominent are acetic acid bacteria. Acetic acid bacteria are also important for the fermentation industries to produce biomaterials such as vinegar and L-sorbose. Acetic acid bacteria are classified into two genera, Gluconobacter and Acetobacter of the family Acetobacteraceae. Gluconobacter species catalyze highly active oxidation reactions on ethanol or D-glucose— including also oxidative reactions on sugars such as D-gluconic acid, D-sorbitol, and glycerol. By contrast, Acetobacter species have a highly active ethanoloxidizing system but not enzymes for sugar oxidation. The respiratory chain in Acetobacter spp. has ubiquinone, cytochrome b , cytochrome c , and a terminal ubiquinol oxidase, which is either cytochrome a 1 or cytochrome o .


Applied Microbiology and Biotechnology | 2002

Quinoproteins: structure, function, and biotechnological applications.

Kazunobu Matsushita; Hirohide Toyama; Mamoru Yamada; Osao Adachi

Abstract. A new class of oxidoreductase containing an amino acid-derived o-quinone cofactor, of which the most typical is pyrroloquinoline quinone (PQQ), is called quinoproteins, and has been recognized as the third redox enzyme following pyridine nucleotide- and flavin-dependent dehydrogenases. Some quinoproteins include a heme c moiety in addition to the quinone cofactor in the molecule and are called quinohemoproteins. PQQ-containing quinoproteins and quinohemoproteins have a common structural basis, in which PQQ is deeply embedded in the center of the unique superbarrel structure. Increased evidence for the structure and function of quinoproteins has revealed their unique position within the redox enzymes with respect to catalytic and electron transfer properties, and also to physiological and energetic function. The peculiarities of the quinoproteins, together with their unique substrate specificity, have encouraged their biotechnological application in the fields of biosensing and bioconversion of useful compounds, and also to environmental treatment.


Applied Microbiology and Biotechnology | 2003

NEW DEVELOPMENTS IN OXIDATIVE FERMENTATION

Osao Adachi; Duangtip Moonmangmee; Hirohide Toyama; Mamoru Yamada; Emiko Shinagawa; Kazunobu Matsushita

Abstract. Oxidative fermentations have been well established for a long time, especially in vinegar and in L-sorbose production. Recently, information on the enzyme systems involved in these oxidative fermentations has accumulated and new developments are possible based on these findings. We have recently isolated several thermotolerant acetic acid bacteria, which also seem to be useful for new developments in oxidative fermentation. Two different types of membrane-bound enzymes, quinoproteins and flavoproteins, are involved in oxidative fermentation, and sometimes work with the same substrate but produce different oxidation products. Recently, there have been new developments in two different oxidative fermentations, D-gluconate and D-sorbitol oxidations. Flavoproteins, D-gluconate dehydrogenase, and D-sorbitol dehydrogenase were isolated almost 2 decades ago, while the enzyme involved in the same oxidation reaction for D-gluconate and D-sorbitol has been recently isolated and shown to be a quinoprotein. Thus, these flavoproteins and a quinoprotein have been re-assessed for the oxidation reaction. Flavoprotein D-gluconate dehydrogenase and D-sorbitol dehydrogenase were shown to produce 2-keto-D-gluconate and D-fructose, respectively, whereas the quinoprotein was shown to produce 5-keto-D-gluconate and L-sorbose from D-gluconate and D-sorbitol, respectively. In addition to the quinoproteins described above, a new quinoprotein for quinate oxidation has been recently isolated from Gluconobacter strains. The quinate dehydrogenase is also a membrane-bound quinoprotein that produces 3-dehydroquinate. This enzyme can be useful for the production of shikimate, which is a convenient salvage synthesis system for many antibiotics, herbicides, and aromatic amino acids synthesis. In order to reduce energy costs of oxidative fermentation in industry, several thermotolerant acetic acid bacteria that can grow up to 40°C have been isolated. Of such isolated strains, some thermotolerant Acetobacter species were found to be useful for vinegar fermentation at a high temperature such 38–40°C, where mesophilic strains showed no growth. They oxidized higher concentrations of ethanol up to 9% without any appreciable lag time, while alcohol oxidation with mesophilic strains was delayed or became almost impossible under such conditions. Several useful Gluconobacter species of thermotolerant acetic acid bacteria are also found, especially L-erythrulose-producing strains and cyclic alcohol-oxidizing strains. Gluconobacter frateurii CHM 43 is able to rapidly oxidize meso-erythritol at 37°C leading to the accumulation of L-erythrulose, which may replace dihydroxyacetone in cosmetics. G. frateurii CHM 9 is able to oxidize cyclic alcohols to their corresponding cyclic ketones or aliphatic ketones, which are known to be useful for preparing many different physiologically active compounds such as oxidized steroids or oxidized bicyclic ketones. The enzymes involved in these meso-erythritol and cyclic alcohol oxidations have been purified and shown to be a similar type of membrane-bound quinoproteins, consisting of a high molecular weight single peptide. This is completely different from another quinoprotein, alcohol dehydrogenase of acetic acid bacteria, which consists of three subunits including hemoproteins.


Applied and Environmental Microbiology | 2003

5-Keto-d-Gluconate Production Is Catalyzed by a Quinoprotein Glycerol Dehydrogenase, Major Polyol Dehydrogenase, in Gluconobacter Species

Kazunobu Matsushita; Yoshikazu Fujii; Yoshitaka Ano; Hirohide Toyama; Masako Shinjoh; Noribumi Tomiyama; Taro Miyazaki; Teruhide Sugisawa; Tatsuo Hoshino; Osao Adachi

ABSTRACT Acetic acid bacteria, especially Gluconobacter species, have been known to catalyze the extensive oxidation of sugar alcohols (polyols) such as d-mannitol, glycerol, d-sorbitol, and so on. Gluconobacter species also oxidize sugars and sugar acids and uniquely accumulate two different keto-d-gluconates, 2-keto-d-gluconate and 5-keto-d-gluconate, in the culture medium by the oxidation of d-gluconate. However, there are still many controversies regarding their enzyme systems, especially on d-sorbitol and also d-gluconate oxidations. Recently, pyrroloquinoline quinone-dependent quinoprotein d-arabitol dehydrogenase and d-sorbitol dehydrogenase have been purified from G. suboxydans, both of which have similar and broad substrate specificity towards several different polyols. In this study, both quinoproteins were shown to be identical based on their immuno-cross-reactivity and also on gene disruption and were suggested to be the same as the previously isolated glycerol dehydrogenase (EC 1.1.99.22). Thus, glycerol dehydrogenase is the major polyol dehydrogenase involved in the oxidation of almost all sugar alcohols in Gluconobacter sp. In addition, the so-called quinoprotein glycerol dehydrogenase was also uniquely shown to oxidize d-gluconate, which was completely different from flavoprotein d-gluconate dehydrogenase (EC 1.1.99.3), which is involved in the production of 2-keto-d-gluconate. The gene disruption experiment and the reconstitution system of the purified enzyme in this study clearly showed that the production of 5-keto-d-gluconate in G. suboxydans is solely dependent on the quinoprotein glycerol dehydrogenase.


Applied Microbiology and Biotechnology | 2006

Correlation between acetic acid resistance and characteristics of PQQ-dependent ADH in acetic acid bacteria

Janja Trček; Hirohide Toyama; Jerzy Czuba; Anna Misiewicz; Kazunobu Matsushita

In this study, we compared the growth properties and molecular characteristics of pyrroloquinoline quinone (PQQ)-dependent alcohol dehydrogenase (ADH) among highly acetic acid-resistant strains of acetic acid bacteria. Ga. europaeus exhibited the highest resistance to acetic acid (10%), whereas Ga. intermedius and Acetobacter pasteurianus resisted up to 6% of acetic acid. In media with different concentrations of acetic acid, the maximal acetic acid production rate of Ga. europaeus slowly increased, but specific growth rates decreased concomitant with increased concentration of acetic acid in medium. The lag phase of A. pasteurianus was twice and four times longer in comparison to the lag phases of Ga. europaeus and Ga. intermedius, respectively. PQQ-dependent ADH activity was twice as high in Ga. europaeus and Ga. intermedius as in A. pasteurinus. The purified enzymes showed almost the same specific activity to each other, but in the presence of acetic acid, the enzyme activity decreased faster in A. pasteurianus and Ga. intermedius than in Ga. europaeus. These results suggest that high ADH activity in the Ga. europaeus cells and high acetic acid stability of the purified enzyme represent two of the unique features that enable this species to grow and stay metabolically active at extremely high concentrations of acetic acid.


Journal of Bacteriology | 2005

Acetobacter aceti Possesses a Proton Motive Force-Dependent Efflux System for Acetic Acid

Kazunobu Matsushita; Taketo Inoue; Osao Adachi; Hirohide Toyama

Acetic acid bacteria are obligate aerobes able to oxidize ethanol, sugar alcohols, and sugars into their corresponding acids. Among them, Acetobacter and Gluconacetobacter species have very high ethanol oxidation capacity, leading to accumulation of vast amounts of acetic acid outside the cell. Since these bacteria are able to grow in media with high concentrations of acetic acid, they must possess a specific mechanism such as an efflux pump by which they can resist the toxic effects of acetic acid. In this study, the efflux pump of Acetobacter aceti IFO 3283 was examined using intact cells and membrane vesicles. The accumulation of acetic acid/acetate in intact cells was increased by the addition of a proton uncoupler and/or cyanide, suggesting the presence of an energy-dependent efflux system. To confirm this, right-side-out and inside-out membrane vesicles were prepared from A. aceti IFO 3283, and the accumulation of acetic acid/acetate in the vesicles was examined. Upon the addition of a respiratory substrate, the accumulation of acetic acid/acetate in the right-side-out vesicles was largely decreased, while its accumulation was very much increased in the inside-out vesicles. These respiration-dependent phenomena observed in both types of membrane vesicles were all sensitive to a proton uncoupler. Acetic acid/acetate uptake in the inside-out membrane vesicles was dependent not on ATP but on the proton motive force. Furthermore, uptake was shown to be rather specific for acetic acid and to be pH dependent, because higher uptake was observed at lower pH. Thus, A. aceti IFO 3283 possesses a proton motive force-dependent efflux pump for acetic acid.


Bioscience, Biotechnology, and Biochemistry | 2001

Membrane-bound Quinoprotein D-Arabitol Dehydrogenase of Gluconobacter suboxydans IFO 3257: A Versatile Enzyme for the Oxidative Fermentation of Various Ketoses

Osao Adachi; Yoshikazu Fujii; Mohamed F. Ghaly; Hirohide Toyama; Emiko Shinagawa; Kazunobu Matsushita

Solubilization of membrane-bound quinoprotein D-arabitol dehydrogenase (ARDH) was done successfully with the membrane fraction of Gluconobacter suboxydans IFO 3257. In enzyme solubilization and subsequent enzyme purification steps, special care was taken to purifiy ARDH as active as it was in the native membrane, after many disappointing trials. Selection of the best detergent, keeping ARDH as the holoenzyme by the addition of PQQ and Ca2+, and of a buffer system involving acetate buffer supplemented with Ca2+, were essential to treat the highly hydrophobic and thus labile enzyme. Purification of the enzyme was done by two steps of column chromatography on DEAE-Toyopearl and CM-Toyopearl in the presence of detergent and Ca2+. ARDH was homogenous and showed a single sedimentation peak in analytical ultracentrifugation. ARDH was dissociated into two different subunits upon SDS-PAGE with molecular masses of 82kDa (subunit I) and 14kDa (subunit II), forming a heterodimeric structure. ARDH was proven to be a quinoprotein by detecting a liberated PQQ from SDS-treated ARDH in HPLC chromatography. More preliminarily, an EDTA-treated membrane fraction lost the enzyme activity and ARDH activity was restored to the original level by the addition of PQQ and Ca2+. The most predominant unique character of ARDH, the substrate specificity, was highly versatile and many kinds of substrates were oxidized irreversibly by ARDH, not only pentitols but also other polyhydroxy alcohols including D-sorbitol, D-mannitol, glycerol, meso-erythritol, and 2,3-butanediol. ARDH may have its primary function in the oxidative fermentation of ketose production by acetic acid bacteria. ARDH contained no heme component, unlike the type II or type III quinoprotein alcohol dehydrogenase (ADH) and did not react with primary alcohols.


Bioscience, Biotechnology, and Biochemistry | 2001

Membrane-bound Sugar Alcohol Dehydrogenase in Acetic Acid Bacteria catalyzes L-Ribulose Formation and NAD-Dependent Ribitol Dehydrogenase is Independent of the Oxidative Fermentation

Osao Adachi; Yoshikazu Fujii; Yoshitaka Ano; Duangtip Moonmangmee; Hirohide Toyama; Emiko Shinagawa; Gunjana Theeragool; Napha Lotong; Kazunobu Matsushita

To identify the enzyme responsible for pentitol oxidation by acetic acid bacteria, two different ribitol oxidizing enzymes, one in the cytosolic fraction of NAD(P)-dependent and the other in the membrane fraction of NAD(P)-independent enzymes, were examined with respect to oxidative fermentation. The cytoplasmic NAD-dependent ribitol dehydrogenase (EC 1.1.1.56) was crystallized from Gluconobacter suboxydans IFO 12528 and found to be an enzyme having 100kDa of molecular mass and 5 s as the sedimentation constant, composed of four identical subunits of 25 kDa. The enzyme catalyzed a shuttle reversible oxidoreduction between ribitol and D-ribulose in the presence of NAD and NADH, respectively. Xylitol and L-arabitol were well oxidized by the enzyme with reaction rates comparable to ribitol oxidation. D-Ribulose, L-ribulose, and L-xylulose were well reduced by the enzyme in the presence of NADH as cosubstrates. The optimum pH of pentitol oxidation was found at alkaline pH such as 9.5-10.5 and ketopentose reduction was found at pH 6.0. NAD-Dependent ribitol dehydrogenase seemed to be specific to oxidoreduction between pentitols and ketopentoses and D-sorbitol and D-mannitol were not oxidized by this enzyme. However, no D-ribulose accumulation was observed outside the cells during the growth of the organism on ribitol. L-Ribulose was accumulated in the culture medium instead, as the direct oxidation product catalyzed by a membrane-bound NAD(P)-independent ribitol dehydrogenase. Thus, the physiological role of NAD-dependent ribitol dehydrogenase was accounted to catalyze ribitol oxidation to D-ribulose in cytoplasm, taking D-ribulose to the pentose phosphate pathway after being phosphorylated. L-Ribulose outside the cells would be incorporated into the cytoplasm in several ways when need for carbon and energy sources made it necessary to use L-ribulose for their survival. From a series of simple experiments, membrane-bound sugar alcohol dehydrogenase was concluded to be the enzyme responsible for L-ribulose production in oxidative fermentation by acetic acid bacteria.


Bioscience, Biotechnology, and Biochemistry | 2006

High Shikimate Production from Quinate with Two Enzymatic Systems of Acetic Acid Bacteria

Osao Adachi; Yoshitaka Ano; Hirohide Toyama; Kazunobu Matsushita

3-Dehydroshikimate was formed with a yield of 57–77% from quinate via 3-dehydroquinate by two successive enzyme reactions, quinoprotein quinate dehydrogenase (QDH) and 3-dehydroquinate dehydratase, in the cytoplasmic membranes of acetic acid bacteria. 3-Dehydroshikimate was then reduced to shikimate (SKA) with NADP-dependent SKA dehydrogenase (SKDH) from the same organism. When SKDH was coupled with NADP-dependent D-glucose dehydrogenase (GDH) in the presence of excess D-glucose as an NADPH re-generating system, SKDH continued to produce SKA until 3-dehydroshikimate added initially in the reaction mixture was completely converted to SKA. Based on the data presented, a strategy for high SKA production was proposed.


Structure | 2002

Structure at 1.9 Å Resolution of a Quinohemoprotein Alcohol Dehydrogenase from Pseudomonas putida HK5

Zhi-wei Chen; Kazunobu Matsushita; Tetsuo Yamashita; Takaaki Fujii; Hirohide Toyama; Osao Adachi; Henry D. Bellamy; F. Scott Mathews

The type II quinohemoprotein alcohol dehydrogenase of Pseudomonas putida is a periplasmic enzyme that oxidizes substrate alcohols to the aldehyde and transfers electrons first to pyrroloquinoline quinone (PQQ) and then to an internal heme group. The 1.9 A resolution crystal structure reveals that the enzyme contains a large N-terminal eight-stranded beta propeller domain (approximately 60 kDa) similar to methanol dehydrogenase and a small C-terminal c-type cytochrome domain (approximately 10 kDa) similar to the cytochrome subunit of p-cresol methylhydoxylase. The PQQ is bound near the axis of the propeller domain about 14 A from the heme. A molecule of acetone, the product of the oxidation of isopropanol present during crystallization, appears to be bound in the active site cavity.

Collaboration


Dive into the Hirohide Toyama's collaboration.

Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Duangtip Moonmangmee

King Mongkut's University of Technology Thonburi

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Researchain Logo
Decentralizing Knowledge