9 research outputs found

    FT-IR Characterization of the Light-Induced Ni-L2 and Ni-L3 States of [NiFe] Hydrogenase from Desulfovibrio vulgaris Miyazaki F

    No full text
    Different light-induced Ni-L states of [NiFe] hydrogenase from its Ni-C state have previously been observed by EPR spectroscopy. Herein, we succeeded in detecting simultaneously two Ni-L states of [NiFe] hydrogenase from Desulfovibrio vulgaris Miyazaki F by FT-IR spectroscopy. A new light-induced ν<sub>CO</sub> band at 1890 cm<sup>–1</sup> and ν<sub>CN</sub> bands at 2034 and 2047 cm<sup>–1</sup> were detected in the FT-IR spectra of the H<sub>2</sub>-activated enzyme under N<sub>2</sub> atmosphere at basic conditions, in addition to the 1910 cm<sup>–1</sup> ν<sub>CO</sub> band and 2047 and 2061 cm<sup>–1</sup> ν<sub>CN</sub> bands of the Ni-L2 state. The new bands were attributed to the Ni-L3 state by comparison of the FT-IR and EPR spectra. The ν<sub>CO</sub> and ν<sub>CN</sub> frequencies of the Ni-L3 state are the lowest frequencies observed among the corresponding frequencies of standard-type [NiFe] hydrogenases in various redox states. These results indicate that a residue, presumably Ni-coordinating Cys546, is protonated and deprotonated in the Ni-L2 and Ni-L3 states, respectively. Relatively small Δ<i>H</i> (6.4 ± 0.8 kJ mol<sup>–1</sup>) and Δ<i>S</i> (25.5 ± 10.3 J mol<sup>–1</sup> K<sup>–1</sup>) values were obtained for the conversion from the Ni-L2 to Ni-L3 state, which was in agreement with the previous proposals that deprotonation of Cys546 is important for the catalytic reaction of the enzyme

    Domain Swapping of the Heme and N‑Terminal α‑Helix in <i>Hydrogenobacter thermophilus</i> Cytochrome <i>c</i><sub>552</sub> Dimer

    No full text
    Oxidized horse cytochrome <i>c</i> (cyt <i>c</i>) has been shown to oligomerize by domain swapping its C-terminal helix successively. We show that the structural and thermodynamic properties of dimeric <i>Hydrogenobacter thermophilus</i> (HT) cytochrome <i>c</i><sub>552</sub> (cyt <i>c</i><sub>552</sub>) and dimeric horse cyt <i>c</i> are different, although both proteins belong to the cyt <i>c</i> superfamily. Optical absorption and circular dichroism spectra of oxidized dimeric HT cyt <i>c</i><sub>552</sub> were identical to the corresponding spectra of its monomer. Dimeric HT cyt <i>c</i><sub>552</sub> exhibited a domain-swapped structure, where the N-terminal α-helix together with the heme was exchanged between protomers. Since a relatively strong H-bond network was formed at the loop around the heme-coordinating Met, the C-terminal α-helix did not dissociate from the rest of the protein in dimeric HT cyt <i>c</i><sub>552</sub>. The packing of the amino acid residues important for thermostability in monomeric HT cyt <i>c</i><sub>552</sub> were maintained in its dimer, and thus, dimeric HT cyt <i>c</i><sub>552</sub> exhibited high thermostability. Although the midpoint redox potential shifted from 240 ± 2 to 213 ± 2 mV by dimerization, it was maintained relatively high. Ethanol has been shown to decrease both the activation enthalpy and activation entropy for the dissociation of the dimer to monomers from 140 ± 9 to 110 ± 5 kcal/mol and 310 ± 30 to 270 ± 20 cal/(mol·K), respectively. Enthalpy change for the dissociation of the dimer to monomers was positive (14 ± 2 kcal/mol per protomer unit). These results give new insights into factors governing the swapping region and thermodynamic properties of domain swapping

    CD spectra and small angle X-ray scattering curves of WT and M61A PA cyt <i>c</i><sub>551</sub>.

    No full text
    <p>(A) CD spectra of oxidized monomeric WT (red) and M61A (green) PA cyt <i>c</i><sub>551</sub>. Measurement conditions: Sample concentration, 10 μM (heme unit); buffer, 50 mM potassium phosphate buffer; pH, 7.0; temperature, room temperature. (B) Small angle X-ray scattering curves of oxidized monomeric WT (red) and M61A (green) PA cyt <i>c</i><sub>551</sub> shown by Kratky plots. The intensities are normalized at their maximum intensities. Measurement conditions: sample concentration, 500 μM (heme unit); buffer, 50 mM potassium phosphate buffer; pH, 7.0; temperature, 20°C.</p

    Fe–His16 and Fe–Met61 distances in monomeric and dimeric WT PA cyt <i>c</i><sub>551</sub>.

    No full text
    <p><sup>a</sup> PDB ID: 351C.</p><p><sup>b</sup> There are two independent WT PA cyt <i>c</i><sub>551</sub> molecules in the asymmetric unit of dimeric WT PA cyt <i>c</i><sub>551</sub> crystal.</p><p>Fe–His16 and Fe–Met61 distances in monomeric and dimeric WT PA cyt <i>c</i><sub>551</sub>.</p

    Active site structures of monomeric and dimeric WT PA cyt <i>c</i><sub>551</sub>.

    No full text
    <p>(A) Structure of monomeric WT PA cyt <i>c</i><sub>551</sub> (PDB ID: 351C). (B) Structure of dimeric WT PA cyt <i>c</i><sub>551</sub> (PDB ID: 3X39). The heme and side-chains of amino acid residues near the heme (Phe7, Cys12, Ala14, Cys15, His16, Val23, Pro25, Val30, Leu44, Arg47, Ile48, Ser52, Trp56, Pro60, Met61, Pro62, Pro63, Asn64, Leu74, and Val78) are shown as stick models. The sulfur atoms of the heme axial Met ligand and heme-linked Cys are shown in yellow, and the nitrogen atoms of the heme axial His ligand are shown in blue. The cyan strand in the dimeric structure is a region from another molecule. The hemes and Thr20–Met22 residues (hinge loop) are depicted in dark and pale colors, respectively.</p

    Refined Regio- and Stereoselective Hydroxylation of l‑Pipecolic Acid by Protein Engineering of l‑Proline <i>cis</i>-4-Hydroxylase Based on the X‑ray Crystal Structure

    No full text
    Enzymatic regio- and stereoselective hydroxylation are valuable for the production of hydroxylated chiral ingredients. Proline hydroxylases are representative members of the nonheme Fe<sup>2+</sup>/α-ketoglutarate-dependent dioxygenase family. These enzymes catalyze the conversion of l-proline into hydroxy-l-prolines (Hyps). l-Proline <i>cis</i>-4-hydroxylases (<i>cis</i>-P4Hs) from <i>Sinorhizobium meliloti</i> and <i>Mesorhizobium loti</i> catalyze the hydroxylation of l-proline, generating <i>cis</i>-4-hydroxy-l-proline, as well as the hydroxylation of l-pipecolic acid (l-Pip), generating two regioisomers, <i>cis</i>-5-Hypip and <i>cis</i>-3-Hypip. To selectively produce <i>cis</i>-5-Hypip without simultaneous production of two isomers, protein engineering of <i>cis</i>-P4Hs is required. We therefore carried out protein engineering of <i>cis</i>-P4H to facilitate the conversion of the majority of l-Pip into the <i>cis</i>-5-Hypip isomer. We first solved the X-ray crystal structure of <i>cis</i>-P4H in complex with each of l-Pro and l-Pip. Then, we conducted three rounds of directed evolution and successfully created a <i>cis</i>-P4H triple mutant, V97F/V95W/E114G, demonstrating the desired regioselectivity toward <i>cis</i>-5-Hypip

    Formation of Oligomeric Cytochrome <i>c</i> during Folding by Intermolecular Hydrophobic Interaction between N- and C‑Terminal α‑Helices

    No full text
    We have previously shown that horse cytochrome <i>c</i> (cyt <i>c</i>) forms oligomers by domain swapping its C-terminal α-helix when interacting with ethanol. Although folding of cyt <i>c</i> has been studied extensively, formation of domain-swapped oligomers of cyt <i>c</i> during folding has never been reported. We found that domain-swapped oligomeric cyt <i>c</i> is produced during refolding from its guanidinium ion-induced unfolded state at high protein concentrations and low temperatures. The obtained dimer exhibited a domain-swapped structure exchanging the C-terminal α-helical region between molecules. The extent of dimer formation decreased significantly for the folding of C-terminal cyt <i>c</i> mutants with reduced hydrophobicity achieved by replacement of hydrophobic residues with Gly in the C-terminal region, whereas a large amount of heterodimers was generated for the folding of a mixture of N- and C-terminal mutants. These results show that cyt <i>c</i> oligomers are formed through intermolecular hydrophobic interaction between the N- and C-terminal α-helices during folding. A slow phase (4–5 s) was observed in addition to a 400–500 ms phase during folding of a high concentration of cyt <i>c</i> in the presence of 1.17 M guanidine hydrochloride. The fast phase is attributed to the intramolecular ligand exchange process, and we attribute the slow phase to the ligand exchange process in oligomers. These results show that it is important to consider formation of domain-swapped oligomeric proteins when folding at high protein concentrations

    Light-Driven Hydrogen Production by Hydrogenases and a Ru-Complex inside a Nanoporous Glass Plate under Aerobic External Conditions

    No full text
    Hydrogenases are powerful catalysts for light-driven H<sub>2</sub> production using a combination of photosensitizers. However, except oxygen-tolerant hydrogenases, they are immediately deactivated under aerobic conditions. We report a light-driven H<sub>2</sub> evolution system that works stably even under aerobic conditions. A [NiFe]-hydrogenase from <i>Desulfovibrio vulgaris</i> Miyazaki F was immobilized inside nanoporous glass plates (PGPs) with a pore diameter of 50 nm together with a ruthenium complex and methyl viologen as a photosensitizer and an electron mediator, respectively. After immersion of PGP into the medium containing the catalytic components, an anaerobic environment automatically established inside the nanopores even under aerobic external conditions upon irradiation with solar-simulated light; this system constantly evolved H<sub>2</sub> with an efficiency of 3.7 μmol H<sub>2</sub> m<sup>–2</sup> s<sup>–1</sup>. The PGP system proposed in this work represents a promising first step toward the development of an O<sub>2</sub>-tolerant solar energy conversion system

    Formation of Domain-Swapped Oligomer of Cytochrome <i>c</i> from Its Molten Globule State Oligomer

    No full text
    Many proteins, including cytochrome <i>c</i> (cyt <i>c</i>), have been shown to form domain-swapped oligomers, but the factors governing the oligomerization process remain unrevealed. We obtained oligomers of cyt <i>c</i> by refolding cyt <i>c</i> from its acid molten globule state to neutral pH state under high protein and ion concentrations. The amount of oligomeric cyt <i>c</i> obtained depended on the nature of the anion (chaotropic or kosmotropic) in the solution: ClO<sub>4</sub><sup>–</sup> (oligomers, 11% ± 2% (heme unit)), SCN<sup>–</sup> (10% ± 2%), I<sup>–</sup> (6% ± 2%), NO<sub>3</sub><sup>–</sup> (3% ± 1%), Br<sup>–</sup> (2% ± 1%), Cl<sup>–</sup> (2% ± 1%), and SO<sub>4</sub><sup>2–</sup> (3% ± 1%) for refolding of 2 mM cyt <i>c</i> (anion concentration 125 mM). Dimeric cyt <i>c</i> obtained by refolding from the molten globule state exhibited a domain-swapped structure, in which the C-terminal α-helices were exchanged between protomers. According to small-angle X-ray scattering measurements, approximately 25% of the cyt <i>c</i> molecules were dimerized in the molten globule state containing 125 mM ClO<sub>4</sub><sup>–</sup>. These results indicate that a certain amount of molten globule state oligomers of cyt <i>c</i> convert to domain-swapped oligomers during refolding and that the intermolecular interactions necessary for domain swapping are present in the molten globule state
    corecore