Academic literature on the topic 'Hydroxylation'

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the lists of relevant articles, books, theses, conference reports, and other scholarly sources on the topic 'Hydroxylation.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Journal articles on the topic "Hydroxylation"

1

Hilinski, Michael, Shea Johnson, and Logan Combee. "Organocatalytic Atom-Transfer C(sp3)–H Oxidation." Synlett 29, no. 18 (June 27, 2018): 2331–36. http://dx.doi.org/10.1055/s-0037-1610432.

Full text
Abstract:
Predictably site-selective catalytic methods for intermolecular C(sp3)–H hydroxylation and amination hold great promise for the synthesis and late-stage modification of complex molecules. Transition-metal catalysis has been the most common approach for early investigations of this type of reaction. In comparison, there are far fewer ­reports of organocatalytic methods for direct oxygen or nitrogen insertion into C–H bonds. Herein, we provide an overview of early efforts in this area, with particular emphasis on our own recent development of an iminium salt that catalyzes both oxygen and nitrogen insertion.1 Introduction2 Background: C–H Oxidation Capabilities of Heterocyclic Oxidants3 Oxaziridine-Mediated Catalytic Hydroxylation4 Dioxirane-Mediated Catalytic Hydroxylation5 Iminium Salt Catalysis of Hydroxylation and Amination6 Conclusion and Outlook
APA, Harvard, Vancouver, ISO, and other styles
2

Holland, Herbert L., Frances M. Brown, P. Chinna Chenchaiah, and J. Appa Rao. "Hydroxylation of prostanoids by fungi. Synthesis of (−)-15-deoxy-19-(R)-hydroxy-PGE1 and (−)-15-deoxy-18-(S)-hydroxy-PGE1." Canadian Journal of Chemistry 68, no. 2 (February 1, 1990): 282–93. http://dx.doi.org/10.1139/v90-039.

Full text
Abstract:
A series of racemic substituted cyclopentanones, with alkyl groups corresponding to the upper prostanoid side chain and (or) the lower prostanoid side chain without the C-15 alcohol, has been synthesized. Using a steroid template for the prostanoid molecule as a basis for selection, fungi capable of hydroxylating steroids have been used to biotransform the prostanoid substrates. The predominant products were hydroxylated at the prostanoid C-18 and C-19 positions. The hydroxylations were enantioselective, with excesses in the range 10–60%, and in most cases the predominant configuration corresponded to that of the natural prostanoids. The stereochemistry of the C-19 hydroxyl group was found to be R by degradation of products to methyl 6-acetoxyheptanoate and comparison of that material with a resolved sample, obtained via crystallization of the brucine salt of ethyl 6-phthaloxyheptanoate. Hydroxylation at C-18 gave the S configuration of alcohol. Hydroxylation at prostanoid C-15 was observed, but in all cases this was accompanied by other reactions. Hydroxylation of Rhizopusarrhizus has been used in a preparation of (−)-15-deoxy-19-(R)-hydroxy-PGE1 and (−)-15-deoxy-18-(S)-hydroxy-PGE1. Keywords: biotransformation, hydroxylation, prostaglandins, prostanoids.
APA, Harvard, Vancouver, ISO, and other styles
3

Masferrer-Rius, Eduard, Raoul M. Hopman, Jishai van der Kleij, Martin Lutz, and Robertus J. M. Klein Gebbink. "On the Ability of Nickel Complexes Derived from Tripodal Aminopyridine Ligands to Catalyze Arene Hydroxylations." CHIMIA International Journal for Chemistry 74, no. 6 (June 24, 2020): 489–94. http://dx.doi.org/10.2533/chimia.2020.489.

Full text
Abstract:
The development of catalysts for the selective hydroxylation of aromatic C–H bonds is an essential challenge in current chemical research. The accomplishment of this goal requires the discovery of powerful metal-based oxidizing species capable of hydroxylating inert aromatic bonds in a selective manner, avoiding the generation of non-selective oxygen-centered radicals. Herein we show an investigation on the ability of nickel(ii) complexes supported by tripodal tetradentate aminopyridine ligands to catalyze the direct hydroxylation of benzene to phenol with H2O2 as oxidant. We have found that modifications on the ligand structure of the nickel complex do not translate into different reactivity, which differs from previous findings for nickel-based arene hydroxylations. Besides, several nickel(ii) salts have been found to be effective in the oxidation of aromatic C–H bonds. The use of fluorinated alcohols as solvent has been found to result in an increase in phenol yield; however, showing no more than two turn-overs per nickel. These findings raise questions on the nature of the oxidizing species responsible for the arene hydroxylation reaction.
APA, Harvard, Vancouver, ISO, and other styles
4

Doostzadeh, J., and R. Morfin. "Effects of cytochrome P450 inhibitors and of steroid hormones on the formation of 7-hydroxylated metabolites of pregnenolone in mouse brain microsomes." Journal of Endocrinology 155, no. 2 (November 1, 1997): 343–50. http://dx.doi.org/10.1677/joe.0.1550343.

Full text
Abstract:
Hydroxylations of pregnenolone (PREG) at the 7 alpha- and 7 beta-positions have been reported in numerous murine tissues and organs and responsible cytochrome P450 (CYP) species await identification. Using thin layer chromatography and gas chromatography-mass spectrometry, we report identification of 7 alpha-hydroxy-PREG and 7 beta-hydroxy-PREG metabolites produced in mouse brain microsome digests and kinetic studies of their production with apparent KM values of 0.5 +/- 0.1 microM and 5.1 +/- 0.6 microM for 7 alpha- and 7 beta-hydroxylation respectively. Investigation of CYP inhibitors and of steroid hormone effects on both 7 alpha- and 7 beta-hydroxylations of PREG showed that: (i) different CYP were involved in 7 alpha- and 7 beta-hydroxylation of PREG because solely 7 alpha-hydroxylation was extensively inhibited by metyrapone, alpha-naphthoflavone, ketoconazole and 3 beta-hydroxysteroids, (ii) CYP 1A2, 2D6, 2B1 and 2B11 were not responsible for 7 alpha- and 7 beta-hydroxylation of PREG because respective specific inhibitors furafylline, quinidine and chloramphenicol triggered no inhibition, (iii) CYP 1A1 was responsible for only part of the 7 beta-hydroxylation of PREG because use of alpha-naphthoflavone, which inhibits specifically CYP 1A1, did not suppress entirely 7 beta-hydroxylation, while ketoconazole, metyrapone and antipyrine, which do not inhibit CYP 1A1, decreased part of the 7 beta-hydroxylation, (iv) 7 alpha-hydroxylation of PREG may be shared with other 3 beta-hydroxysteroids such as isoandrosterone and 5-androstene-3 beta,17 beta-diol which were strong inhibitors, but not with dehydroepiandrosterone which was a non-competitive inhibitor as weak as 3-oxosteroids, and (v) 7 beta-hydroxylation of PREG was not markedly changed by other steroids. Taken together, these findings will be of use for identification of the CYP species responsible for 7 alpha- and 7 beta-hydroxylation of PREG and for studies of their activities in brain.
APA, Harvard, Vancouver, ISO, and other styles
5

Adams, J. S., and M. A. Gacad. "Characterization of 1 alpha-hydroxylation of vitamin D3 sterols by cultured alveolar macrophages from patients with sarcoidosis." Journal of Experimental Medicine 161, no. 4 (April 1, 1985): 755–65. http://dx.doi.org/10.1084/jem.161.4.755.

Full text
Abstract:
We investigated the 1 alpha-hydroxylation of vitamin D3 sterols by cultured pulmonary alveolar macrophages (PAM) from patients with sarcoidosis with or without clinically abnormal calcium homeostasis. Like the naturally occurring renal 1 alpha-hydroxylase, the PAM 1 alpha-hydroxylation reaction exhibited a high affinity for 25-hydroxyvitamin D3 (25-OH-D3) and a preference for substrates containing a 25-hydroxyl group in the side chain of the sterol. Unlike the renal enzyme, the PAM 1 alpha-hydroxylating mechanism was not accompanied by 24-hydroxylating activity, even after preincubation with 75 nM 1,25-dihydroxyvitamin D3 [1,25-(OH)2-D3] or exposure to high concentrations of substrate (500 nM 25-OH-D3). The PAM 25-OH-D3-1 alpha-hydroxylation reaction was stimulated by gamma interferon and inhibited by exposure to the glucocorticoid dexamethasone. The characteristics of the PAM hydroxylation process in vitro appear to reflect the efficiency of the extrarenal production of 1,25-(OH)2-D3 and the therapeutic efficacy of glucocorticoids in patients with sarcoidosis and disordered calcium metabolism.
APA, Harvard, Vancouver, ISO, and other styles
6

Dahlbäck, H., and K. Wikvall. "25-Hydroxylation of vitamin D3 by a cytochrome P-450 from rabbit liver mitochondria." Biochemical Journal 252, no. 1 (May 15, 1988): 207–13. http://dx.doi.org/10.1042/bj2520207.

Full text
Abstract:
A cytochrome P-450 catalysing 25-hydroxylation of vitamin D3 was purified from liver mitochondria of untreated rabbits. The enzyme fraction contained 9 nmol of cytochrome P-450/mg of protein and showed only one protein band with an apparent Mr of 52,000 upon SDS/polyacrylamide-gel electrophoresis. The preparation showed a single protein spot with an apparent isoelectric point of 7.8 and an Mr of approx. 52,000 upon two-dimensional isoelectric-focusing-polyacrylamide-gel electrophoresis. The purified cytochrome P-450 catalysed 25-hydroxylation of vitamin D3 up to 5000 times more efficiently than did the mitochondria. The cytochrome P-450 required both ferredoxin and ferredoxin reductase for catalytic activity. Microsomal NADPH-cytochrome P-450 reductase could not replace ferredoxin and ferredoxin reductase. The cytochrome P-450 catalysed, in addition to 25-hydroxylation of vitamin D3, the 25-hydroxylation of 1 alpha-hydroxyvitamin D3 and the 26-hydroxylation of 5 beta-cholestane-3 alpha, 7 alpha, 12 alpha-triol. The enzyme did not catalyse side-chain cleavage of cholesterol, 11 beta-hydroxylation of deoxycorticosterone, 1 alpha-hydroxylation of 25-hydroxyvitamin D3, hydroxylations of lauric acid and testosterone or demethylation of benzphetamine. The results raise the possibility that the 25-hydroxylation of vitamin D3 and the 26-hydroxylation of C27 steroids are catalysed by the same species of cytochrome P-450 in liver mitochondria. The possible role of the liver mitochondrial cytochrome P-450 in the metabolism of vitamin D3 is discussed.
APA, Harvard, Vancouver, ISO, and other styles
7

Ramirez, Leyla C., and Bernard F. Maume. "Regulation of 11β/18-steroid hydroxylation in newborn rat adrenal cells in primary culture." Acta Endocrinologica 111, no. 1 (January 1986): 106–15. http://dx.doi.org/10.1530/acta.0.1110106.

Full text
Abstract:
Abstract. In newborn rat adrenal cells in primary culture, the level of activity of the 11β/18-steroid hydroxylase system involved in the last step of the corticosteroid biosynthesis is increased by ACTH. A parallel study of 11β- and 18-hydroxylation showed the same apparent Km values (64 μm) for both hydroxylations. The Vmax values differed: 11.5 μg/106 cells/h for corticosterone and 6.9 μg/106 cells/h for 18-hydroxyDOC. A dose response study of the ACTH effect, measured by the bioconversion of deoxycorticosterone to corticosterone and 18-hydroxyDOC, showed maximum hydroxylation with a dose of 2.2 mU of ACTH/ml. Addition of ACTH after several weeks in culture produced a smaller increase in 1 1β/18-hydroxylation. Removal of ACTH after several weeks of treatment produced an immediate decrease in corticosteroid production; readdition of ACTH produced an increase to the previous level in the case of the 22 mU/ml dose, but not in the case of the 2.2 mU/ml dose. The use of actinomycin D demonstrated that ACTH affects mainly the biosynthesis of protein which must be renewed approximately every 24 h. Finally, the effect of pretreatment or co-treatment with various concentrations of the end products of the reaction showed no inhibition or destruction of the I 1β/18-hydroxylating enzyme system. Therefore, the regulation of the 11β/18-steroid hydroxylase system in these cell cultures seems to be accomplished through the induction by ACTH of the transcription involved in the biosynthesis of cytochrome P45011β and the amount of available precursor furnished by endogenous steroidogenesis.
APA, Harvard, Vancouver, ISO, and other styles
8

Veronese, M. E., C. J. Doecke, P. I. Mackenzie, M. E. McManus, J. O. Miners, D. L. Rees, R. Gasser, U. A. Meyer, and D. J. Birkett. "Site-directed mutation studies of human liver cytochrome P-450 isoenzymes in the CYP2C subfamily." Biochemical Journal 289, no. 2 (January 15, 1993): 533–38. http://dx.doi.org/10.1042/bj2890533.

Full text
Abstract:
Evidence from human studies in vivo and in vitro strongly suggests that the methylhydroxylation of tolbutamide and the 4-hydroxylation of phenytoin, the major pathways in the elimination of these two drugs, are catalysed by the same cytochrome P-450 isoenzyme(s). In the present study we used site-directed mutagenesis and cDNA expression in COS cells to characterize in detail the kinetics of tolbutamide and phenytoin hydroxylations by seven CYP2C proteins (2C8, 2C9 and variants, and 2C10) in order to define the effects of small changes in amino acid sequences and the likely proteins responsible in the metabolism of these two drugs in man. Tolbutamide was hydroxylated to varying extents by all expressed cytochrome P-450 isoenzymes, although activity was much lower for the expressed 2C8 protein. While the apparent Km values for the 2C9/10 isoenzymes (71.6-131.7 microM) were comparable with the range of apparent Km values previously observed in human liver microsomes, the apparent Km for 2C8 (650.5 microM) was appreciably higher. The 2C8 enzyme also showed quite different sulphaphenazole inhibition characteristics. The 4-hydroxylation of phenytoin was also more efficiently catalysed by the 2C9/10 enzymes. These enzymes showed similarities in kinetics of phenytoin hydroxylation and sulphaphenazole inhibition compared with human liver phenytoin hydroxylase. Also of interest was the observation that, among the 2C9 variants, small differences in amino acid composition could appreciably affect both tolbutamide and phenytoin hydroxylations. The amino acid substitution Cys-144->Arg increased both the rates of tolbutamide and phenytoin hydroxylations, while the Leu-359->Ile change had a greater effect on phenytoin hydroxylation. We conclude that: (1) although 2C8 and 2C9/10 proteins metabolize tolbutamide. only 2C9/10 proteins play a major role in human liver; (2) 2C9/10 proteins also appear to be chiefly responsible for phenytoin hydroxylation; and (3) subtle differences in the amino acid composition of these 2C9/10 proteins can affect the functional specificities towards both tolbutamide and phenytoin.
APA, Harvard, Vancouver, ISO, and other styles
9

Axén, E., T. Bergman, and K. Wikvall. "Purification and characterization of a vitamin D3 25-hydroxylase from pig liver microsomes." Biochemical Journal 287, no. 3 (November 1, 1992): 725–31. http://dx.doi.org/10.1042/bj2870725.

Full text
Abstract:
A cytochrome P-450 which catalyses 25-hydroxylation of vitamin D3 has been purified to apparent homogeneity from pig liver microsomes. The specific content of cytochrome P-450 was 12 nmol.mg of protein-1, and the preparation showed a single band with an apparent M(r) of 50,500 upon SDS/PAGE. A monoclonal antibody raised against the vitamin D3 25-hydroxylase reacted strongly with the purified 25-hydroxylating cytochrome P-450 from pig kidney microsomes [Bergman & Postlind (1990) Biochem. J. 270, 345-350]. The liver enzyme showed structural and functional properties very similar to those of the kidney enzyme. The two enzymes differed with respect to only one of the first 16 N-terminal amino acids. The vitamin D3 25-hydroxylase in pig liver microsomes exhibited a turnover and an apparent Km for 25-hydroxylation of vitamin D3 which were of the same order of magnitude as those of a well-characterized male-specific 25-hydroxylating cytochrome P-450 in rat liver microsomes. The two enzymes differed structurally. The pig liver enzyme was, in contrast to the rat liver enzyme, not sex-specific, and did not catalyse 16 alpha-hydroxylation of testosterone. These properties of the 25-hydroxylase in rat liver microsomes have led to questions on the role of microsomal 25-hydroxylation of vitamin D3. It is concluded that studies on microsomal 25-hydroxylation with the rat may be misleading. The results of the present study show that the pig appears to be a representative species for evaluation of vitamin D3 hydroxylases in other mammals, including man.
APA, Harvard, Vancouver, ISO, and other styles
10

Webb, James D., Andrea Murányi, Christopher W. Pugh, Peter J. Ratcliffe, and Mathew L. Coleman. "MYPT1, the targeting subunit of smooth-muscle myosin phosphatase, is a substrate for the asparaginyl hydroxylase factor inhibiting hypoxia-inducible factor (FIH)." Biochemical Journal 420, no. 2 (May 13, 2009): 327–36. http://dx.doi.org/10.1042/bj20081905.

Full text
Abstract:
The asparaginyl hydroxylase FIH [factor inhibiting HIF (hypoxia-inducible factor)] was first identified as a protein that inhibits transcriptional activation by HIF, through hydroxylation of an asparagine residue in the CAD (C-terminal activation domain). More recently, several ARD [AR (ankyrin repeat) domain]-containing proteins were identified as FIH substrates using FIH interaction assays. Although the function(s) of these ARD hydroxylations is unclear, expression of the ARD protein Notch1 was shown to compete efficiently with HIF CAD for asparagine hydroxylation and thus to enhance HIF activity. The ARD is a common protein domain with over 300 examples in the human proteome. However, the extent of hydroxylation among ARD proteins, and the ability of other members to compete with HIF–CAD for FIH, is not known. In the present study we assay for asparagine hydroxylation in a bioinformatically predicted FIH substrate, the targeting subunit of myosin phosphatase, MYPT1. Our results confirm hydroxylation both in cultured cells and in endogenous protein purified from animal tissue. We show that the extent of hydroxylation at three sites is dependent on FIH expression level and that hydroxylation is incomplete under basal conditions even in the animal tissue. We also show that expression of MYPT1 enhances HIF–CAD activity in a manner consistent with competition for FIH and that this property extends to other ARD proteins. These results extend the range of FIH substrates and suggest that cross-competition between ARDs and HIF–CAD, and between ARDs themselves, may be extensive and have important effects on hypoxia signalling.
APA, Harvard, Vancouver, ISO, and other styles

Dissertations / Theses on the topic "Hydroxylation"

1

Tang, Campbell. "Microbial hydroxylation of the morphinans." Thesis, University of Cambridge, 2005. http://ethos.bl.uk/OrderDetails.do?uin=uk.bl.ethos.614764.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

Djaneye-Boundjou, Gbandi. "Hydroxylation catalytique de composés aromatiques." Poitiers, 1989. http://www.theses.fr/1989POIT2324.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Rautenbach, Daniel. "The electrochemical hydroxylation of aromatic substrates." Thesis, Port Elizabeth Technikon, 2002. http://hdl.handle.net/10948/94.

Full text
Abstract:
The electrochemical hydroxylation of aromatic substrates was investigated in some detail, with the view to develop a method, which could produce dihydroxybenzenes in acceptable yields. Of particular interest was the selectivity and yield of the 1,4-dihydroxybenzenes. Two distinctly different methods were investigated in order to achieve this goal, acyloxylation and direct electrochemical hydroxylation. Acyloxylation is the process where radical cations generated at the anode undergoes nucleophilic attack by acetate anions. The resulting aromatic acetates so produced can then be hydrolysed to the phenolic compounds. Two nucleophile systems were considered in the investigation, acetates (acetoxylation) and trifluoro-acetates (trifluoro-acetoxylation). These investigations were conducted under a variety of conditions using phenol and phenyl acetate as starting materials. From the results it was, however, concluded that the acetoxylation of these aromatic compounds occurs in unacceptable product and current yields. Trifluoro-acetoxylation on the other hand showed promise, but due to the nature and cost of the reagents it was deemed to be an impractical process. Direct electrochemical hydroxylation: in which the radical cations produced at the anode undergoes nucleophilic attack by water producing the corresponding dihydroxybenzenes. These dihydroxybenzenes are then further oxidised to the benzoquinones, which then undergo reduction at the cathode in order to produce the corresponding dihydroxybenzene. In this process phenol, 2-tert-butylphenol and 2,6-di-tert-butylphenol were investigated as substrates. The results indicated that the yield towards the 1,4-dihdroxybenzenes increased as the degree of substitution on the ring increased.
APA, Harvard, Vancouver, ISO, and other styles
4

Gqogqa, Pumeza. "Hydroxylation of aromatic compounds over zeolites." Thesis, Stellenbosch : University of Stellenbosch, 2009. http://hdl.handle.net/10019.1/2564.

Full text
Abstract:
Thesis (MScEng (Process Engineering))--University of Stellenbosch, 2009.
Aromatic precursor compounds are derivatives that play an important role in biosystems and are useful in the production of fine chemicals. This work focuses on the catalytic synthesis of 2-methyl-1, 4-naphthoquinone and cresols (para- and ortho) using aqueous hydrogen peroxide as an oxidant in liquidphase oxidation of 2-methylnaphthalene and toluene over titanium-substituted zeolite TS-1 or Ti-MCM-41. Catalysts synthesised in this work were calcined at 550°C, extensively characterised using techniques such as X-ray Fluorescence for determining the catalyst chemical composition; BET for surface area, pore size and micropore volume; Powder X-ray diffraction for determining their crystallinity and phase purity and SEM was used to investigate the catalyst morphologies. The BET surface areas for Ti-MCM-41 showed a surface area of 1025 m2/g, and a 0.575 cm3/g micropore volume. However, zeolite TS-1 showed a BET surface area of 439 m2/g and a 0.174 cm3/g micropore volume. The initial experiments on 2-methylnaphthalene hydroxylation were performed using the normal batch method. After a series of batch runs, without any success as no products were generated as confirmed by GC, a second experimental tool was proposed. This technique made use of the reflux system at reaction conditions similar to that of the batch system. After performing several experimental runs and optimising the system to various reactor operating conditions and without any products formed, the thought of continuing using the reflux was put on hold. Due to this, a third procedure was brought into perspective. This process made use of PTFE lined Parr autoclave. The reactor operating conditions were changed in order to suit the specifications and requirements of the autoclave. This process yielded promising results and the formation of 2-MNQ was realised. There was a drawback when using an autoclave as only one data point was obtained, at the end of each run. Therefore, it was not possible to investigate reaction kinetics in terms of time. Addition of aqueous hydrogen peroxide (30 wt-%) solution in the feed was done in one lot at the beginning of each reaction in all oxidation reactions, to a reactor containing 2-methylnaphthalene and the catalyst in an appropriate solvent of choice (methanol, acetonitrile, 2-propanol, 1-propanol, 1-pentanol, and butanol), with sample withdrawal done over a period of 6 hours (excluding catalytic experiments done with a Parr autoclave as sampling was impossible). As expected, 2-methylnaphthalene oxidation reactions with medium pore zeolite TS-1 yielded no formation of 2-methyl-1, 4-naphthoquinone using various types of solvents, with a batch reactor, reflux system, or a Parr PTFE autoclave. This was attributed to the fact that 2-methylnaphthalene is a large compound and hinders diffusion into zeolite channels. With the use of an autoclave, Ti-MCM-41 catalysed reactions showed that the choice of a solvent and reaction temperature strongly affect 2- methylnaphthalene conversion and product selectivity. This was proven after comparing a series of different solvents (such as methanol, isopropanol, npropanol, isobutanol, n-pentanol and acetonitrile) at different temperatures. Only reactions using acetonitrile as a solvent showed 2-MNQ. Formation of 2- MNQ, indicating that acetonitrile is an appropriate choice of solvent for this system. The highest 2-methylnaphthalene conversion (92%) was achieved at 120 ˚C, with a relative product selectivity of 51.4 %. Temperature showed a major effect on 2-MN conversion as at lower reaction temperature 100˚C, the relative product selectivity (72%) seems to enhance; however, the drawback is the fact that lower 2-methylnaphthalene conversions (18%) are attained. Another important point to note is the fact that using an autoclave (with acetonitrile as a solvent), 2-methyl-1-naphthol was generated as a co-product. In conclusion, it has been shown that the hydroxylation of different aromatic compounds over zeolites conducted in this study generated interesting findings. In 2-MN hydroxylation over Ti-MCM-41 as a catalyst, only acetonitrile is an appropriate choice of solvent using an autoclave. In addition, zeolite TS-1 is not a suitable catalyst for 2-MN hydroxylation reactions. It is ideal to optimise an autoclave in order to investigate reaction kinetics and optimum selectivity. Toluene hydroxylation reactions yielded para and ortho-cresol as expected with either water or acetonitrile as a solvent. No meta-cresol was formed. The kinetic model fitted generated a good fit with water as a solvent or excess toluene, with acetonitrile as a solvent generating a reasonable fit.
APA, Harvard, Vancouver, ISO, and other styles
5

Deng, Yifan. "Flavin-dependent hydroxylation of aromatic compounds." Thesis, Sorbonne université, 2019. http://www.theses.fr/2019SORUS601.

Full text
Abstract:
Le marché mondial des enzymes industrielles, estimé à 17 milliards USD en 2025, continue à croître. La forte demande attire de plus en plus l'attention en raison des avantages de ces enzymes à efficacité élevée, faible coût et respect de l'environnement. Les biocatalyseurs sont développés en raison de la demande dans les domaines pharmaceutique, de la recherche, de la biotechnologie et du diagnostic. L'objectif de ce projet est de réaliser un système in vitro capable de catalyser une réaction d'hydroxylation aromatique. Ce projet est principalement divisé en deux parties. Après une introduction, les détails de l’établissement du système seront décrits et suivis d’une étude sur les propriétés biochimiques. Afin d’obtenir un rendement plus élevé en produits à haute valeur ajoutée, l’activité sur des substrats non naturels doit être renforcée. Une approche cristallographie-structure-mutagenèse est appliquée, basée sur la structure modèle de l'enzyme, certains résidus sont mutés en acide aminé spécifique et l'activité de ces mutants est testée sur des non-substrats. L'hydroxylation est une réaction redox, des cofacteurs sont nécessaires. Pour un système in vitro, il faut réduire les coûts de production en évitant d’utiliser des cofacteurs de forme réduite qui est à prix élevé. Un système de régénération sera introduit. Le système de régénération enzymatique actuellement utilisé dans la production industrielle reste une méthode efficace. De nouvelles méthodes telles que la régénération électrochimique, la régénération catalysée par des complexes organométalliques montrent une activité comparable au système enzymatique. Afin de faciliter le recyclage du catalyseur, un processus d’hétérogénéisation du catalyseur est appliqué. Un exemple d'immobilisation d'un complexe organométallique à base de Rh sur une Bipyridine-Périodique Mésoporeuse Organosilica (Bpy-PMO) est étudié. Ce catalyseur hétérogène montre une activité de régénération et une recyclabilité relativement intéressante
The growing global market of industrial enzymes is estimated to attain 17 billion USD in 2025. The strong demand and many advantages such as high efficiency, low cost, environmentally friendly procedures draw more attention to its use. Besides the regular utilization, specialty enzymes including biocatalysts are quickly developing due to the demand in pharmaceutics, research & biotechnology, and diagnostics. The objective of this project is to realize an in vitro system which can catalyze an aromatic hydroxylation reaction. This project is divided into two parts. After an introduction, the details of system establishment will be described, followed by a study on biochemical properties. In order to achieve a higher yield on high added value products, activity on non-natural substrates needs to be enhanced. A crystallography-structure-mutagenesis approach is applied, based on the model structure of the enzyme, chosen residues are mutated into specific amino acids and activity of these mutants is tested on non-substrates. Considering hydroxylation is a redox reaction, cofactors are needed. For an in vitro system, in order to lower production cost by avoiding using high price reduced cofactors, a regeneration system will be introduced. The enzymatic regeneration system currently used in industrial production is still considered to be an efficient method. Nevertheless, new methods such as electrochemical regeneration, organometallic complexes catalyzed regeneration show a comparable activity to the aforementioned enzymatic regeneration system. With the goal of facilitating catalyst recycling, a catalyst heterogenization process is applied. Immobilization of a Rh based organometallic complex on Bipyridine-Periodic Mesoporous Organosilica (Bpy-PMO) is investigated as an example. This heterogeneous catalyst shows relatively interesting cofactors regeneration activity and recyclability
APA, Harvard, Vancouver, ISO, and other styles
6

Bhowmick, Rupa. "Transition metal-ion mediated hydroxylation reactions." Thesis, University of North Bengal, 1993. http://hdl.handle.net/123456789/867.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

Fedkenheuer, Michael Gerald. "Structural and Mutational Analyses of Aspergillus fumigatus SidA: A Flavin-Dependent N-hydroxylating Enzyme." Thesis, Virginia Tech, 2012. http://hdl.handle.net/10919/76837.

Full text
Abstract:
SidA from Aspergillus fumigatus is an N-hydroxylating monooxygenase that catalyzes the committed step in siderophore biosynthesis. This gene is essential for virulence making it an excellent drug target. In order to design an inhibitor against SidA a greater understanding of the mechanism and structure is needed. We have determined the crystal structure of SidA in complex with NADP+, Ornithine, and FAD at 1.9 ? resolution. The crystal structure has provided insight into substrate and coenzyme selectivity as well as residues essential for catalysis. In particular, we have chosen to study the interactions of Arg 279, shown to interact with the 2'phosphate of the adenine moiety of NADP+ as well as the adenine ring itself. The mutation of this residue to alanine makes the enzyme have little to no selectivity between coenzymes NADPH and NADH which supports the importance of the ionic interaction between Arg279 and the 2'phosphate. Additionally, the mutant enzyme is significantly more uncoupled than WT enzyme with NADPH. We see that the interactions of the guanadinyl group of Arg279 and the adenine ring are also important because KM and Kd values for the mutant enzyme are shifted well above those of wild type with coenzyme NADH. The data is further supported by studies on the reductive and oxidative half reactions. We have also explored the allosteric effect of L-arginine. We provide evidence that an enzyme/coenzyme/L-arginine complex is formed which improves coupling, oxygen reactivity, and reduction in SidA; however more work is needed to fully understand the role of L-arginine as an allosteric effector.
Master of Science in Life Sciences
APA, Harvard, Vancouver, ISO, and other styles
8

Frémy, Nicolas. "Etude des premiers stades d'oxydation de NiA1(001) : croissance d'un film ultra-mince d'alumine et réactivité à la vapeur d'eau de l'oxyde formé." Paris 6, 2004. http://www.theses.fr/2004PA066455.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Karabacak, Elife Ozlem. "Aspergillus Niger Mediated A-hydroxylation Of Cyclic Ketones." Master's thesis, METU, 2006. http://etd.lib.metu.edu.tr/upload/3/12608088/index.pdf.

Full text
Abstract:
Chiral a -hydroxy ketones are important structural units in many natural products, biologically active compounds and the hydroxyl group has frequently been used as a reagent directing group, such as for the selective elaboration of aldol products. In this work, enzymatic synthesis of both enantiomers of the a -hydroxy ketones (2-hydroxy indanone, 2-hydroxy tetralone) using Aspergillus niger by selective &
#945
-oxidation of ketones (1-indanone, 1-tetralone) was studied. The &
#945
-oxidation of ketones was carried out by using whole cells of Aspergillus niger in different growth media. A. niger whole cell catalyzed reactions afforded (S)-configurated 2- hydroxy-1-tetralone with %87 e.e. in DMSO at pH 5.0. In addition to this,while (S)-configurated 2-hydroxy-1-indanone with %33 e.e. in pH 8.0 (in DMSO) was synthesized, (R)-configurated-2-hyroxy-1-indanone with %32 e.e. in pH 7.0 ( in DMSO) was synthesized.
APA, Harvard, Vancouver, ISO, and other styles
10

Kim, Namhoon. "Epoxidation and di-hydroxylation of camelina sativa oil." Thesis, Kansas State University, 2014. http://hdl.handle.net/2097/18252.

Full text
Abstract:
Master of Science
Department of Grain Science and Industry
Xiuzhi Susan Sun
Plant oil-based raw materials have become more attractive alternatives in polymer industry as fossil resources depletion and environmental concerns continue to arise. Camelina (camelina sativa L.) seed contains about 45% of oil and about 90% of the oil is unsaturated fatty acids such as linoleic acid, α-linolenic acid, and erucic acids. It also provides the advantages of low cost and low fertilizer demand. Functionalized oils such as epoxidized camelina oil (ECO) and di-hydroxyl camelina oil (DCO) can be used for resins, adhesives, coatings, etc. The objectives of this work were to synthesize and characterize ECO and DCO from camelina oil. The epoxidation reaction of camelina oil was completed with formic acid and hydrogen peroxide. Catalyst ratio, reaction time, and temperature effects on the epoxidation reaction were studied. The optimum epoxy content of 7.52 wt% with a conversion rate of 76.34% was obtained from camelina oil using excess hydrogen peroxide and a molar ratio of formic acid of less than 1 for 5 hours in 50 °C. Camelina oil yields higher epoxy content (7.52 wt%) than soybean oil (6.53 wt%); however, soybean oil had a higher conversion rate of 80.16% compared to camelina oil because of uniform fatty acids distribution. In this study, we found that epoxidation efficiency is significantly affected by fatty acids composition, structure, and distribution. DCO was synthesized from ECO with different reaction parameters. The ring opening of ECO was performed with water, perchloric acid, and THF as proton donor, catalyst, and solvent respectively. Hydroxyl value of DCO was measured, and the maximal hydroxyl value was 369.24 mg KOH/g. physical properties of DCO were characterized by acid value and moisture content; thermal properties of DCO were obtained using different scanning calorimeter (DSC), thermalgravimetric analysis (TGA). Amount of solvent and acid catalyst addition affected the hydroxyl value and residual acid in DCO. Heat capacity, phase transition temperatures, and thermal stability of DCO were obtained and showed higher values than ECO’s. The DCO showed higher peel adhesion when it was formulated with epoxidized soybean oils through UV curing because camelina oil allows higher epoxy content, which results in higher hydroxyl values.
APA, Harvard, Vancouver, ISO, and other styles

Books on the topic "Hydroxylation"

1

Yoshizawa, Kazunari, ed. Direct Hydroxylation of Methane. Singapore: Springer Singapore, 2020. http://dx.doi.org/10.1007/978-981-15-6986-9.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

Cooke, Robert Grahame *. Hydroxylation of tricyclic antidepressants in vivo clinical importance. 1988.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
3

Elshafei, Harbi Abdallah Ibrahim. Removal of light hydrocarbons in the preparation of ultra-pure ¹⁴CO tracer for measuring hydroxyl radicals in the atmosphere. 1987.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
4

Yoshizawa, Kazunari. Direct Hydroxylation of Methane: Interplay Between Theory and Experiment. Springer Singapore Pte. Limited, 2021.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
5

Yoshizawa, Kazunari. Direct Hydroxylation of Methane: Interplay Between Theory and Experiment. Springer Singapore Pte. Limited, 2020.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
6

Schofield, Christopher, Martin J. Bollinger, Carsten Krebs, Robert Hausinger, and C. David Garner. 2-Oxoglutarate-Dependent Oxygenases. Royal Society of Chemistry, The, 2015.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
7

Bollinger, J. Martin, Christopher Schofield, Robert Hausinger, and C. David Garner. 2-Oxoglutarate-Dependent Oxygenases. Royal Society of Chemistry, The, 2015.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
8

Andersson, Inger, Robert P. Hausinger, Robert Hausinger, Graham Moran, and Tina Muller. 2-Oxoglutarate-Dependent Oxygenases. Royal Society of Chemistry, The, 2015.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
9

The caffeine 8-hydroxylation in humans: A predominantly intestinal phenomenn. Ottawa: National Library of Canada, 2001.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
10

Shaik, Sason, Samuel P. de Visser, Devesh Kumar, Andrew W. Munro, and Saptaswa Sen. Iron-Containing Enzymes: Versatile Catalysts of Hydroxylation Reactions in Nature. Royal Society of Chemistry, The, 2011.

Find full text
APA, Harvard, Vancouver, ISO, and other styles

Book chapters on the topic "Hydroxylation"

1

Yoshizawa, Kazunari, and Mayuko Miyanishi. "Orbital Concept for Methane Activation." In Direct Hydroxylation of Methane, 1–22. Singapore: Springer Singapore, 2020. http://dx.doi.org/10.1007/978-981-15-6986-9_1.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

Shiota, Yoshihito, and Kazunari Yoshizawa. "Theoretical Study of the Direct Conversion of Methane by First-Row Transition-Metal Oxide Cations in the Gas Phase." In Direct Hydroxylation of Methane, 23–44. Singapore: Springer Singapore, 2020. http://dx.doi.org/10.1007/978-981-15-6986-9_2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Yumura, Takashi, Takehiro Ohta, and Kazunari Yoshizawa. "Enzymatic Methane Hydroxylation: sMMO and pMMO." In Direct Hydroxylation of Methane, 45–73. Singapore: Springer Singapore, 2020. http://dx.doi.org/10.1007/978-981-15-6986-9_3.

Full text
APA, Harvard, Vancouver, ISO, and other styles
4

Mahyuddin, Muhammad Haris, Hermawan Kresno Dipojono, and Kazunari Yoshizawa. "Mechanistic Understanding of Methane Hydroxylation by Cu-Exchanged Zeolites." In Direct Hydroxylation of Methane, 75–86. Singapore: Springer Singapore, 2020. http://dx.doi.org/10.1007/978-981-15-6986-9_4.

Full text
APA, Harvard, Vancouver, ISO, and other styles
5

Mahyuddin, Muhammad Haris. "Oxidative Activation of Metal-Exchanged Zeolite Catalysts for Methane Hydroxylation." In Direct Hydroxylation of Methane, 87–100. Singapore: Springer Singapore, 2020. http://dx.doi.org/10.1007/978-981-15-6986-9_5.

Full text
APA, Harvard, Vancouver, ISO, and other styles
6

Tsuji, Yuta, Masashi Saito, and Kazunari Yoshizawa. "Dynamics and Energetics of Methane on the Surfaces of Transition Metal Oxides." In Direct Hydroxylation of Methane, 101–33. Singapore: Springer Singapore, 2020. http://dx.doi.org/10.1007/978-981-15-6986-9_6.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

Toyao, Takashi, Ichigaku Takigawa, and Ken-ichi Shimizu. "Machine Learning Predictions of Adsorption Energies of CH4-Related Species." In Direct Hydroxylation of Methane, 135–49. Singapore: Springer Singapore, 2020. http://dx.doi.org/10.1007/978-981-15-6986-9_7.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

Hori, Yuta, and Tsukasa Abe. "Theoretical Approach to Homogeneous Catalyst of Methane Hydroxylation: Collaboration with Computation and Experiment." In Direct Hydroxylation of Methane, 151–65. Singapore: Springer Singapore, 2020. http://dx.doi.org/10.1007/978-981-15-6986-9_8.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Holland, Herbert L. "Hydroxylation and Dihydroxylation." In Biotechnology, 475–533. Weinheim, Germany: Wiley-VCH Verlag GmbH, 2008. http://dx.doi.org/10.1002/9783527620906.ch10.

Full text
APA, Harvard, Vancouver, ISO, and other styles
10

Li, Jie Jack. "Sharpless asymmetric amino hydroxylation." In Name Reactions, 364–65. Berlin, Heidelberg: Springer Berlin Heidelberg, 2003. http://dx.doi.org/10.1007/978-3-662-05336-2_271.

Full text
APA, Harvard, Vancouver, ISO, and other styles

Conference papers on the topic "Hydroxylation"

1

Nair, Renjini M., B. Bindhu, and Susmi Anna Thomas. "Hydroxylation of boron nitride nanosheets." In DAE SOLID STATE PHYSICS SYMPOSIUM 2019. AIP Publishing, 2020. http://dx.doi.org/10.1063/5.0017307.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

ARNHOLD, J., and J. GEYER. "IRON-MEDIATED HYDROXYLATION OF PHTHALIC HYDRAZIDE." In Bioluminescence and Chemiluminescence - Progress and Current Applications - 12th International Symposium on Bioluminescence (BL) and Chemiluminescence (CL). WORLD SCIENTIFIC, 2002. http://dx.doi.org/10.1142/9789812776624_0029.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Zhu, Ming, Ruiqing Peng, Xin Liang, Zhengdao Lan, Meng Tang, Pingping Hou, Jian H. Song, et al. "Yap1 Hydroxylation Suppress Prostate Cancer Metastasis." In Leading Edge of Cancer Research Symposium. The University of Texas at MD Anderson Cancer Center, 2022. http://dx.doi.org/10.52519/00102.

Full text
APA, Harvard, Vancouver, ISO, and other styles
4

Peterle, Marcos M., Marcelo V. Marques, and Marcus M. Sá. "α-Hydroxylation of malonates under mild reaction conditions." In 15th Brazilian Meeting on Organic Synthesis. São Paulo: Editora Edgard Blücher, 2013. http://dx.doi.org/10.5151/chempro-15bmos-bmos2013_2013820152632.

Full text
APA, Harvard, Vancouver, ISO, and other styles
5

Kryuchkova, Ye V., G. L. Burygin, N. E. Gogoleva, Y. V. Gogolev, E. I. Shagimardanova, A. S. Balkin, A. Y. Muratova, and O. V. Turkovskaya. "A genomic analysis of the catabolism of aromatic compounds in Azospirillum brasilense SR80." In 2nd International Scientific Conference "Plants and Microbes: the Future of Biotechnology". PLAMIC2020 Organizing committee, 2020. http://dx.doi.org/10.28983/plamic2020.133.

Full text
Abstract:
We performed a genomic analysis of the presence and organization of oxygenases involved in the hydroxylation of various substrates, including the aromatic ring, and dioxygenases catalyzing a ring-cleavage of the formed hydroxylated intermediates in A. brasilense SR80.
APA, Harvard, Vancouver, ISO, and other styles
6

Weber, Juliane, Jacquelyn Bracco, Ke Yuan, Vitalii Starchenko, Peter Eng, Joanne Stubbs, Matthew Boebinger, Brittany Moseley, and Gabriela Camacho. "Effect of Impurities on Magnesium Oxide Hydroxylation and Carbonation." In Goldschmidt2023. France: European Association of Geochemistry, 2023. http://dx.doi.org/10.7185/gold2023.18053.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

Xiong, Gaofeng, Rachel Stewart, Jie Chen, Tianyan Gao, Timothy Scott, Luis Samayoa, Kathleen O’Connor, Andrew Lane, and Ren Xu. "Abstract 3035: Collagen hydroxylation promotes TNBC chemoresistance through stabilizing HIF1á." In Proceedings: AACR Annual Meeting 2019; March 29-April 3, 2019; Atlanta, GA. American Association for Cancer Research, 2019. http://dx.doi.org/10.1158/1538-7445.am2019-3035.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

Ye, Pengjin, Wenya Wang, and Lei Wang. "Study of heteropolyacids for the hydroxylation of Naphthalene to naphthol." In 5th International Conference on Information Engineering for Mechanics and Materials. Paris, France: Atlantis Press, 2015. http://dx.doi.org/10.2991/icimm-15.2015.302.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Xiong, Gaofeng, Rachel Stewart, Jie Chen, Tianyan Gao, Timothy Scott, Luis Samayoa, Kathleen O’Connor, Andrew Lane, and Ren Xu. "Abstract 3035: Collagen hydroxylation promotes TNBC chemoresistance through stabilizing HIF1á." In Proceedings: AACR Annual Meeting 2019; March 29-April 3, 2019; Atlanta, GA. American Association for Cancer Research, 2019. http://dx.doi.org/10.1158/1538-7445.sabcs18-3035.

Full text
APA, Harvard, Vancouver, ISO, and other styles
10

Ying-Tsang Lo, Tsan-Huang Shih, Han-Jia Lin, Tun-Wen Pai, and Margaret Dah-Tsyr Chang. "Cross-species identification of hydroxylation sites for ARD and FIH interaction." In 2011 IEEE International Conference on Systems Biology (ISB). IEEE, 2011. http://dx.doi.org/10.1109/isb.2011.6033176.

Full text
APA, Harvard, Vancouver, ISO, and other styles

Reports on the topic "Hydroxylation"

1

Parshikov, Igor. Microbial Transformation of Some Ethylpyridines by Fungi. Intellectual Archive, January 2022. http://dx.doi.org/10.32370/iaj.2635.

Full text
Abstract:
We were observed transformation 4-ethylpyridine and 2-methyl-5-ethylpyridine by fungus Beauveria bassiana ATCC 7159. Stereoselective oxidation of methylene group leading to the optically active (-)-(1-hydroxyethyl)pyridine was shown. Besides, the hydroxylation of methyl groups and the oxidation of the heterocyclic ring in the nitrogen atom to the respective primary alcohols and N-oxides were observed
APA, Harvard, Vancouver, ISO, and other styles
2

Gantt, Elisabeth, Avigad Vonshak, Sammy Boussiba, Zvi Cohen, and Amos Richmond. Carotenoid-Rich Algal Biomass for Aquaculture: Astaxanthin Production by Haematococcus Pluvialis. United States Department of Agriculture, August 1996. http://dx.doi.org/10.32747/1996.7613036.bard.

Full text
Abstract:
The synthesis of carotenoids has been studied toward enhancing the production of ketocarotenoids, since fish and crustaceans raised by aquaculture require astaxanthin and other ketocaroteinoids in their feed for desirable pigmentation. Notable progress has been made in attaining the goals of determining improved conditions for ketocarotenoid production in Haematococcus pluvialis and in elucidating the carotenoid biosynthetic pathway. For production of astaxanthin a number of strains of the green alga Haematococcus were evaluated, a strain CCAG was found to be optimal for photoautotrophic growth. Of four mutants, selected for enhanced carotenoid production, two hold considerable promise because caroteinoid accumulation occurs without encystment. The biosynthetic pathway of carotenoids was elucidated in photosynthetic organisms by characterizing novel genes encoding carotenoid enzymes and by examining the function of these enzymes in a bacterial complementation system. Two cyclases (b- and e-) were cloned that are at a critical branch point in the pathway. One branch leads to the formation of b-carotene and zeaxanthin and astaxanthin, and the other to the production of a-carotene and lutein. Cyclization of both endgroups of lycopene to yield b-carotene was shown to be catalyzed by a single gene product, b-lycopene cyclase in cyanobacteria and plants. The formation of a-carotene was found to require the e-cyclase gene product in addition to the b-cyclase. By cloning a b-hydroxylase gene we showed that a single gene product forms zeaxanthin by hydroxylatin of both b-carotene rings. It is expected that a second hydroxylase is required in the synthesis of astaxanthin, since canthaxanthin rather than zeaxanthin is the precursor. Evidence, from inhibitor studies, suggests that astaxanthin is formed from canthaxanthin and that b-carotene is a major precursor. Feasibility studies with the photobioreactors have shown that a two-stage system is the most practical, where Haematococcus cultures are first grown to high cell density and are then switched to high light for maximal astaxanthin production. The basic knowledge and molecular tools generated from this study will significantly enhance Haematococcus as a viable model for enhanced astaxanthin production.
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography