Network


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

Hotspot


Dive into the research topics where Sam Hart is active.

Publication


Featured researches published by Sam Hart.


ChemBioChem | 2013

Structure and Activity of Nadph-Dependent Reductase Q1Eqe0 from Streptomyces Kanamyceticus, which Catalyses the R-Selective Reduction of an Imine Substrate.

María Rodríguez‐Mata; Annika Frank; Elizabeth Wells; Friedemann Leipold; Nicholas J. Turner; Sam Hart; Johan P. Turkenburg; Gideon Grogan

NADPH‐dependent oxidoreductase Q1EQE0 from Streptomyces kanamyceticus catalyzes the asymmetric reduction of the prochiral monocyclic imine 2‐methyl‐1‐pyrroline to the chiral amine (R)‐2‐methylpyrrolidine with >99 % ee, and is thus of interest as a potential biocatalyst for the production of optically active amines. The structures of Q1EQE0 in native form, and in complex with the nicotinamide cofactor NADPH have been solved and refined to a resolution of 2.7 Å. Q1EQE0 functions as a dimer in which the monomer consists of an N‐terminal Rossman‐fold motif attached to a helical C‐terminal domain through a helix of 28 amino acids. The dimer is formed through reciprocal domain sharing in which the C‐terminal domains are swapped, with a substrate‐binding cleft formed between the N‐terminal subunit of monomer A and the C‐terminal subunit of monomer B. The structure is related to those of known β‐hydroxyacid dehydrogenases, except that the essential lysine, which serves as an acid/base in the (de)protonation of the nascent alcohol in those enzymes, is replaced by an aspartate residue, Asp187 in Q1EQE0. Mutation of Asp187 to either asparagine or alanine resulted in an inactive enzyme.


Inorganic Chemistry | 2013

Iridium(III) hydrido N-heterocyclic carbene-phosphine complexes as catalysts in magnetization transfer reactions.

Marianna Fekete; Oliver W. Bayfield; Simon B. Duckett; Sam Hart; Ryan E. Mewis; Natalie E. Pridmore; Peter J. Rayner; Adrian C. Whitwood

The hyperpolarization (HP) method signal amplification by reversible exchange (SABRE) uses para-hydrogen to sensitize substrate detection by NMR. The catalyst systems [Ir(H)2(IMes)(MeCN)2(R)]BF4 and [Ir(H)2(IMes)(py)2(R)]BF4 [py = pyridine; R = PCy3 or PPh3; IMes = 1,3-bis(2,4,6-trimethylphenyl)imidazol-2-ylidene], which contain both an electron-donating N-heterocyclic carbene and a phosphine, are used here to catalyze SABRE. They react with acetonitrile and pyridine to produce [Ir(H)2(NCMe)(py)(IMes)(PPh3)]BF4 and [Ir(H)2(NCMe)(py)(IMes)(PCy3)]BF4, complexes that undergo ligand exchange on a time scale commensurate with observation of the SABRE effect, which is illustrated here by the observation of both pyridine and acetonitrile HP. In this study, the required symmetry breaking that underpins SABRE is provided for by the use of chemical inequivalence rather than the previously reported magnetic inequivalence. As a consequence, we show that the ligand sphere of the polarization transfer catalyst itself becomes hyperpolarized and hence that the high-sensitivity detection of a number of reaction intermediates is possible. These species include [Ir(H)2(NCMe)(py)(IMes)(PPh3)]BF4, [Ir(H)2(MeOH)(py)(IMes)(PPh3)]BF4, and [Ir(H)2(NCMe)(py)2(PPh3)]BF4. Studies are also described that employ the deuterium-labeled substrates CD3CN and C5D5N, and the labeled ligands P(C6D5)3 and IMes-d22, to demonstrate that dramatically improved levels of HP can be achieved as a consequence of reducing proton dilution and hence polarization wastage. By a combination of these studies with experiments in which the magnetic field experienced by the sample at the point of polarization transfer is varied, confirmation of the resonance assignments is achieved. Furthermore, when [Ir(H)2(pyridine-h5)(pyridine-d5)(IMes)(PPh3)]BF4 is examined, its hydride ligand signals are shown to become visible through para-hydrogen-induced polarization rather than SABRE.


ChemBioChem | 2015

Structure, Activity and Stereoselectivity of NADPH-Dependent Oxidoreductases Catalysing the S-Selective Reduction of the Imine Substrate 2-Methylpyrroline.

Henry Man; Elizabeth Wells; Shahed Hussain; Friedemann Leipold; Sam Hart; Johan P. Turkenburg; Nicholas J. Turner; Gideon Grogan

Oxidoreductases from Streptomyces sp. GF3546 [3546‐IRED], Bacillus cereus BAG3X2 (BcIRED) and Nocardiopsis halophila (NhIRED) each reduce prochiral 2‐methylpyrroline (2MPN) to (S)‐2‐methylpyrrolidine with >95 % ee and also a number of other imine substrates with good selectivity. Structures of BcIRED and NhIRED have helped to identify conserved active site residues within this subgroup of imine reductases that have S selectivity towards 2MPN, including a tyrosine residue that has a possible role in catalysis and superimposes with an aspartate in related enzymes that display R selectivity towards the same substrate. Mutation of this tyrosine residue—Tyr169—in 3546‐IRED to Phe resulted in a mutant of negligible activity. The data together provide structural evidence for the location and significance of the Tyr residue in this group of imine reductases, and permit a comparison of the active sites of enzymes that reduce 2MPN with either R or S selectivity.


ChemBioChem | 2012

A Flavoprotein Monooxygenase that Catalyses a Baeyer-Villiger Reaction and Thioether Oxidation Using Nadh as the Nicotinamide Cofactor.

Chantel N. Jensen; Jared Cartwright; Jonathan S. Ward; Sam Hart; Johan P. Turkenburg; Sohail T. Ali; Michael J. Allen; Gideon Grogan

A gene from the marine bacterium Stenotrophomonas maltophilia encodes a 38.6 kDa FAD‐containing flavoprotein (Uniprot B2FLR2) named S. maltophilia flavin‐containing monooxygenase (SMFMO), which catalyses the oxidation of thioethers and also the regioselective Baeyer–Villiger oxidation of the model substrate bicyclo[3.2.0]hept‐2‐en‐6‐one. The enzyme was unusual in its ability to employ either NADH or NADPH as nicotinamide cofactor. The KM and kcat values for NADH were 23.7±9.1 μM and 0.029 s−1 and 27.3±5.3 μM and 0.022 s−1 for NADPH. However, kcat/KM value for the ketone substrate in the presence of 100 μM cofactor was 17 times greater for NADH than for NADPH. SMFMO catalysed the quantitative conversion of 5 mM ketone in the presence of substoichiometric concentrations of NADH with the formate dehydrogenase cofactor recycling system, to give the 2‐oxa and 3‐oxa lactone products of Baeyer–Villiger reaction in a ratio of 5:1, albeit with poor enantioselectivity. The conversion with NADPH was 15 %. SMFMO also catalysed the NADH‐dependent transformation of prochiral aromatic thioethers, giving in the best case, 80 % ee for the transformation of p‐chlorophenyl methyl sulfide to its R enantiomer. The structure of SMFMO reveals that the relaxation in cofactor specificity appears to be accomplished by the substitution of an arginine residue, responsible for recognition of the 2′‐phosphate on the NADPH ribose in related NADPH‐dependent FMOs, with a glutamine residue in SMFMO. SMFMO is thus representative of a separate class of single‐component, flavoprotein monooxygenases that catalyse NADH‐dependent oxidations from which possible sequences and strategies for developing NADH‐dependent biocatalysts for asymmetric oxygenation reactions might be identified.


Chemistry: A European Journal | 2014

Halogen- and Hydrogen-Bonded Salts and Co-crystals Formed from 4-Halo-2,3,5,6-tetrafluorophenol and Cyclic Secondary and Tertiary Amines: Orthogonal and Non-orthogonal Halogen and Hydrogen Bonding, and Synthetic Analogues of Halogen-Bonded Biological Systems

Akihiro Takemura; Linda J. McAllister; Sam Hart; Natalie E. Pridmore; Peter B. Karadakov; Adrian C. Whitwood; Duncan W. Bruce

Co-crystallisation of, in particular, 4-iodotetrafluorophenol with a series of secondary and tertiary cyclic amines results in deprotonation of the phenol and formation of the corresponding ammonium phenate. Careful examination of the X-ray single-crystal structures shows that the phenate anion develops a C=O double bond and that the C–C bond lengths in the ring suggest a Meissenheimer-like delocalisation. This delocalisation is supported by the geometry of the phenate anion optimised at the MP2(Full) level of theory within the aug-cc-pVDZ basis (aug-cc-pVDZ-PP on I) and by natural bond orbital (NBO) analyses. With sp2 hybridisation at the phenate oxygen atom, there is strong preference for the formation of two non-covalent interactions with the oxygen sp2 lone pairs and, in the case of secondary amines, this occurs through hydrogen bonding to the ammonium hydrogen atoms. However, where tertiary amines are concerned, there are insufficient hydrogen atoms available and so an electrophilic iodine atom from a neighbouring 4-iodotetrafluorophenate group forms an I⋅⋅⋅O halogen bond to give the second interaction. However, in some co-crystals with secondary amines, it is also found that in addition to the two hydrogen bonds forming with the phenate oxygen sp2 lone pairs, there is an additional intermolecular I⋅⋅⋅O halogen bond in which the electrophilic iodine atom interacts with the C=O π-system. All attempts to reproduce this behaviour with 4-bromotetrafluorophenol were unsuccessful. These structural motifs are significant as they reproduce extremely well, in low-molar-mass synthetic systems, motifs found by Ho and co-workers when examining halogen-bonding interactions in biological systems. The analogy is cemented through the structures of co-crystals of 1,4-diiodotetrafluorobenzene with acetamide and with N-methylbenzamide, which, as designed models, demonstrate the orthogonality of hydrogen and halogen bonding proposed in Ho’s biological study.


Journal of Biological Chemistry | 2009

The 1.5-A structure of XplA-heme, an unusual cytochrome P450 heme domain that catalyzes reductive biotransformation of royal demolition explosive.

Federico Sabbadin; Rosamond G. Jackson; Kamran Haider; Girish Tampi; Johan P. Turkenburg; Sam Hart; Neil C. Bruce; Gideon Grogan

XplA is a cytochrome P450 of unique structural organization, consisting of a heme- domain that is C-terminally fused to its native flavodoxin redox partner. XplA, along with flavodoxin reductase XplB, has been shown to catalyze the breakdown of the nitramine explosive and pollutant hexahydro-1,3,5-trinitro-1,3,5-triazine (royal demolition explosive) by reductive denitration. The structure of the heme domain of XplA (XplA-heme) has been solved in two crystal forms: as a dimer in space group P21 to a resolution of 1.9 Å and as a monomer in space group P21212 to a resolution of 1.5 Å, with the ligand imidazole bound at the heme iron. Although it shares the overall fold of cytochromes P450 of known structure, XplA-heme is unusual in that the kinked I-helix that traverses the distal face of the heme is broken by Met-394 and Ala-395 in place of the well conserved Asp/Glu plus Thr/Ser, important in oxidative P450s for the scission of the dioxygen bond prior to substrate oxygenation. The heme environment of XplA-heme is hydrophobic, featuring a cluster of three methionines above the heme, including Met-394. Imidazole was observed bound to the heme iron and is in close proximity to the side chain of Gln-438, which is situated over the distal face of the heme. Imidazole is also hydrogen-bonded to a water molecule that sits in place of the threonine side-chain hydroxyl exemplified by Thr-252 in Cyt-P450cam. Both Gln-438 → Ala and Ala-395 → Thr mutants of XplA-heme displayed markedly reduced activity compared with the wild type for royal demolition explosive degradation when combined with surrogate electron donors.


Journal of the American Chemical Society | 2010

A Covalent Succinylcysteine-like Intermediate in the Enzyme-Catalyzed Transformation of Maleate to Fumarate by Maleate Isomerase

Florian Fisch; Carlos Martinez Fleites; Marie Delenne; Nina Baudendistel; Bernhard Hauer; Johan P. Turkenburg; Sam Hart; Neil C. Bruce; Gideon Grogan

Maleate isomerase (MI), a member of the Asp/Glu racemase superfamily, catalyzes cis-trans isomerization of the C2-C3 double bond in maleate to yield fumarate. Mutational studies, in conjunction with the structure of the C194A mutant of Nocardia farcinica MI cocrystallized with maleate, have revealed an unprecedented mode of catalysis for the superfamily in which the isomerization reaction is initiated by nucleophilic attack of cysteine at the double bond, yielding a covalent succinylcysteine-like intermediate.


Catalysis Science & Technology | 2012

Mutational Analysis of Phenolic Acid Decarboxylase from Bacillus Subtilis (Bspad), which Converts Bio-Derived Phenolic Acids to Styrene Derivatives

Annika Frank; William Eborall; Ralph Hyde; Sam Hart; Johan P. Turkenburg; Gideon Grogan

Phenolic acid decarboxylase from Bacillus subtilis (BsPAD) catalyses the decarboxylation of phenolic acids such as coumaric acid to give vinyl phenols, which are of interest as possible polymer precursors and flavour/fragrance compounds. The structure of the Tyr19Ala mutant of BsPAD has been solved in complex with coumaric acid. In the active site, the substrate carboxylate is bound by Tyr11 and Tyr13, and the phenolic hydroxyl by the NE atom of Arg41. A comparison of the mutant complex with the wild-type apoenzyme reveals that the β1–β2 loop, running from Tyr11 to Ala19, closes over the active site in the presence of substrate, shielding it from bulk solvent. The complex structure, in conjunction with an activity study of point mutants of BsPAD, provides support for a mechanism for PADs, proposed by Mancheno and co-workers for the homologue from Lactobacillus plantarum [Proteins, 2010, 78, 1662–1676]. In this mechanism, a quinone methide intermediate results from deprotonation of the phenolic hydroxyl of the substrate by Glu64, assisted by Arg41. Decarboxylation of the substrate is effected through binding of the carboxylate by Tyr11 and Tyr13, the latter being brought into contact with the substrate as a result of the movement of the β1–β2 loop on substrate binding.


Catalysis Science & Technology | 2016

Homogeneous and silica-supported zinc complexes for the synthesis of propylene carbonate from propane-1,2-diol and carbon dioxide

James W. Comerford; Sam Hart; Michael North; Adrian C. Whitwood

Three organozinc complexes have been synthesised and found to catalyse the carbonylation of propylene glycol with carbon dioxide to form propylene carbonate. A similar tethered organozinc complex was supported onto high loading aminopropyl functionalised hexagonal mesoporous silica and was also found to be catalytically active.


Acta Crystallographica Section F-structural Biology and Crystallization Communications | 2012

Structures of a γ-aminobutyrate (GABA) transaminase from the s-triazine-degrading organism Arthrobacter aurescens TC1 in complex with PLP and with its external aldimine PLP-GABA adduct.

H. Bruce; A. Nguyen Tuan; J. Mangas Sánchez; Charlotte Leese; J. Hopwood; Ralph Hyde; Sam Hart; Johan P. Turkenburg; Gideon Grogan

Two complex structures of the γ-aminobutyrate (GABA) transaminase A1R958 from Arthrobacter aurescens TC1 are presented. The first, determined to a resolution of 2.80 Å, features the internal aldimine formed by reaction between the ℇ-amino group of Lys295 and the cofactor pyridoxal phosphate (PLP); the second, determined to a resolution of 2.75 Å, features the external aldimine adduct formed between PLP and GABA in the first half-reaction. This is the first structure of a microbial GABA transaminase in complex with its natural external aldimine and reveals the molecular determinants of GABA binding in this enzyme.

Collaboration


Dive into the Sam Hart'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
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Researchain Logo
Decentralizing Knowledge