Network


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

Hotspot


Dive into the research topics where Roger N. F. Thorneley is active.

Publication


Featured researches published by Roger N. F. Thorneley.


Journal of the American Chemical Society | 2001

Mechanism of reaction of hydrogen peroxide with horseradish peroxidase: identification of intermediates in the catalytic cycle.

José Neptuno Rodríguez-López; David J. Lowe; Josefa Hernández-Ruiz; Alexander N. P. Hiner; Francisco García-Cánovas; Roger N. F. Thorneley

The mechanism of the reaction of horseradish peroxidase isoenzyme C (HRPC) with hydrogen peroxide to form the reactive enzyme intermediate compound I has been studied using electronic absorbance, rapid-scan stopped-flow, and electron paramagnetic resonance (EPR) spectroscopies at both acid and basic pH. The roles of the active site residues His42 and Arg38 in controlling heterolytic cleavage of the H(2)O(2) oxygen-oxygen bond have been probed with site-directed mutant enzymes His42 --> Leu (H42L), Arg38 --> Leu (R38L), and Arg38 --> Gly (R38G). The biphasic reaction kinetics of H42L with H(2)O(2) suggested the presence of an intermediate species and, at acid pH, a reversible second step, probably due to a neutral enzyme-H(2)O(2) complex and the ferric-peroxoanion-containing compound 0. EPR also indicated the formation of a protein radical situated more than approximately 10 A from the heme iron. The stoichiometry of the reaction of the H42L/H(2)O(2) reaction product and 2,2-azinobis(3-ethylbenzothiazolinesulfonic acid) (ABTS) was concentration dependent and fell from a value of 2 to 1 above 0.7 mM ABTS. These data can be explained if H(2)O(2) undergoes homolytic cleavage in H42L. The apparent rate of compound I formation by H42L, while low, was pH independent in contrast to wild-type HRPC where the rate falls at acid pH, indicating the involvement of an ionizable group with pK(a) approximately 4. In R38L and R38G, the apparent pK(a) was shifted to approximately 8 but there is no evidence that homolytic cleavage of H(2)O(2) occurs. These data suggest that His42 acts initially as a proton acceptor (base catalyst) and then as a donor (acid catalyst) at neutral pH and predict the observed slower rate and lower efficiency of heterolytic cleavage observed at acid pH. Arg38 is influential in lowering the pK(a) of His42 and additionally in aligning H(2)O(2) in the active site, but it does not play a direct role in proton transfer.


Cancer Research | 2005

The Antifolate Activity of Tea Catechins

Enma Navarro-Perán; Juan Cabezas-Herrera; Francisco García-Cánovas; Marcus C. Durrant; Roger N. F. Thorneley; José Neptuno Rodríguez-López

A naturally occurring gallated polyphenol isolated from green tea leaves, (-)-epigallocatechin gallate (EGCG), has been shown to be an inhibitor of dihydrofolate reductase (DHFR) activity in vitro at concentrations found in the serum and tissues of green tea drinkers (0.1-1.0 micromol/L). These data provide the first evidence that the prophylactic effect of green tea drinking on certain forms of cancer, suggested by epidemiologic studies, is due to the inhibition of DHFR by EGCG and could also explain why tea extracts have been traditionally used in alternative medicine as anticarcinogenic/antibiotic agents or in the treatment of conditions such as psoriasis. EGCG exhibited kinetics characteristic of a slow, tight-binding inhibitor of 7,8-dihydrofolate reduction with bovine liver DHFR (K(I) = 0.109 micromol/L), but of a classic, reversible, competitive inhibitor with chicken liver DHFR (K(I) = 10.3 micromol/L). Structural modeling showed that EGCG can bind to human DHFR at the same site and in a similar orientation to that observed for some structurally characterized DHFR inhibitor complexes. The responses of lymphoma cells to EGCG and known antifolates were similar, that is, a dose-dependent inhibition of cell growth (IC50 = 20 micromol/L for EGCG), G0-G1 phase arrest of the cell cycle, and induction of apoptosis. Folate depletion increased the sensitivity of these cell lines to antifolates and EGCG. These effects were attenuated by growing the cells in a medium containing hypoxanthine-thymidine, consistent with DHFR being the site of action for EGCG.


Plant Physiology | 1997

Purification and Characterization of a Novel Class III Peroxidase Isoenzyme from Tea Leaves

M. Kvaratskhelia; C. Winkel; Roger N. F. Thorneley

A novel, basic (isoelectric point > 10), heme peroxidase isoenzyme (TP; relative molecular weight = 34,660 [plus or minus] 10, mean [plus or minus] SE) that can account for a significant part of the ascorbate peroxidase activity in tea (Camellia sinensis) leaves has been purified to homogeneity. The ultraviolet/visible absorption spectrum is typical of heme-containing plant peroxidases, with a Soret peak at 406 nm ([epsilon] = 115 mM-1 cm-1) and an A406/A280 value of 3.4. The enzyme has a high specific activity for ascorbate oxidation (151 [mu]mol min-1 mg-1), with a pH optimum in the range of 4.5 to 5.0. Substrate-specificity studies have revealed significant differences between TP and other class III peroxidases, as well as similarities with class I ascorbate peroxidases. TP, like ascorbate peroxidase, exhibits a preference for ascorbate over guaiacol, whereas other class III isoenzymes are characterized by 2-orders-of-magnitude higher activity for guaiacol than for ascorbate. TP also forms an unstable porphyrin [pi] cation radical-type compound I, which is converted to compound II within approximately 2 min in the absence of added reductant. Amino acid sequence data show TP to be the first example, to our knowledge, of a class III peroxidase with a high specificity for ascorbate as an electron donor.


Biochemistry | 2000

Stopped-flow and steady-state study of the diphenolase activity of mushroom tyrosinase.

José Neptuno Rodríguez-López; Lorena G. Fenoll; Pedro Antonio García-Ruiz; R. Varón; José Tudela; Roger N. F. Thorneley; Francisco García-Cánovas

The reaction of mushroom (Agaricus bisporus) tyrosinase with dioxygen in the presence of several o-diphenolic substrates has been studied by steady-state and transient-phase kinetics in order to elucidate the rate-limiting step and to provide new insights into the mechanism of oxidation of these substrates. A kinetic analysis has allowed for the first time the determination of individual rate constants for several of the partial reactions that comprise the catalytic cycle. Mushroom tyrosinase rapidly reacts with dioxygen with a second-order rate constant k(+8) = 2.3 x 10(7) M(-)(1) s(-)(1), which is similar to that reported for hemocyanins [(1.3 x 10(6))-(5.7 x 10(7)) M(-)(1) s(-)(1)]. Deoxytyrosinase binds dioxygen reversibly at the binuclear Cu(I) site with a dissociation constant K(D)(O)()2 = 46.6 microM, which is similar to the value (K(D)(O)()2 = 90 microM) reported for the binding of dioxygen to Octopus vulgaris deoxyhemocyanin [Salvato et al. (1998) Biochemistry 37, 14065-14077]. Transient and steady-state kinetics showed that o-diphenols such as 4-tert-butylcatechol react significantly faster with mettyrosinase (k(+2) = 9.02 x 10(6) M(-)(1) s(-)(1)) than with oxytyrosinase (k(+6) = 5.4 x 10(5) M(-)(1) s(-)(1)). This difference is interpreted in terms of differential steric and polar effects that modulate the access of o-diphenols to the active site for these two forms of the enzyme. The values of k(cat) for several o-diphenols are also consistent with steric and polar factors controlling the mobility, orientation, and thence the reactivity of substrates at the active site of tyrosinase.


Journal of Inorganic Biochemistry | 2002

Mechanisms of compound I formation in heme peroxidases

Alexander N. P. Hiner; Emma Lloyd Raven; Roger N. F. Thorneley; Francisco García-Cánovas; José Neptuno Rodríguez-López

The formation of compound I is the first step in the reaction mechanism of plant heme peroxidases. This intermediate stores two oxidizing equivalents from hydrogen peroxide as an oxyferryl iron center and a radical, either on the porphyrin ring or on a tryptophan residue. Site-directed mutagenesis has proved to be a most useful tool for the identification of the intermediates involved and the resulting nature of the compound I formed. Although there is no doubt that an acid-base mechanism operates in heme peroxidase during the formation of compound I, the roles of several distal pocket residues are currently the subject of intensive research. It is now generally accepted that the conserved distal histidine in the active site of heme peroxidases is the acid-base catalyst that promotes the heterolytic cleavage of hydrogen peroxide. Other residues, such as the distal arginine and asparagine, participate in a range of roles assisting catalysis by the distal histidine. Recent advances in the elucidation of the mechanism at the molecular level are discussed. Another aspect related to the nature of compound I is the location of the radical center. Novel radical species have been detected in the reactions of ascorbate peroxidase, lignin peroxidase and several mutants of horseradish peroxidase. Detailed kinetic and spectroscopic studies of these radical species have provided important insights about the factors that control porphyrin-protein radical exchange. The wide range of data being obtained on compound I will lead to an understanding of its vital function in peroxidase catalysis and the physiological roles played by these enzymes.


Journal of Biological Inorganic Chemistry | 1996

Recombinant horseradish peroxidase isoenzyme C: the effect of distal haem cavity mutations (His42→Leu and Arg38→Leu) on compound I formation and substrate binding

Jose Neptuno Rodriguez-Lopez; Andrew T. Smith; Roger N. F. Thorneley

Abstractu2002Horseradish peroxidase isoenzyme C (HRPC) mutants were constructed in order to understand the role of two key distal haem cavity residues, histidine 42 and arginine 38, in the formation of compound I and in substrate binding. The role of these residues as general acid-base catalysts, originally proposed for cytochrome c peroxidase by Poulos and Kraut in 1980 was assessed for HRPC. Replacement of histidine 42 by leucine [(H42L)HRPC*] decreased the apparent bimolecular rate constant for the reaction with hydrogen peroxide by five orders of magnitude (k1u2009=u20091.4×102 M–1s–1) compared with both native-glycosylated and recombinant forms of HRPC (k1u2009=u20091.7×107 M–1s–1). The first-order rate constant for the heterolytic cleavage of the oxygen-oxygen bond to form compound I was estimated to be four orders of magnitude slower for this variant. Replacement of arginine 38 by leucine [(R38L)HRPC*] decreased the observed pseudo-first-order rate constant for the reaction with hydrogen peroxide by three orders of magnitude (k1u2009=u20091.1×104 M–1s–1), while the observed rate constant of oxygen bond scission was decreased sixfold (k2u2009=u2009142 s–1). These rate constants are consistent with arginine 38 having two roles in catalysing compound I formation: firstly, promotion of proton transfer to the imidazole group of histidine 42 to facilitate peroxide anion binding to the haem, and secondly, stabilisation of the transition state for the heterolytic cleavage of the oxygen-oxygen bond. These roles for arginine 38 explain, in part, why dioxygen-binding globins, which do not have an arginine in the distal cavity, are poor peroxidases. Binding studies of benzhydroxamic acid to (H42L)HRPC* and (R38L)HRPC* indicate that both histidine 42 and arginine 38 are involved in the modulation of substrate affinity.


Journal of Biological Chemistry | 1997

The Inactivation and Catalytic Pathways of Horseradish Peroxidase with m-Chloroperoxybenzoic Acid A SPECTROPHOTOMETRIC AND TRANSIENT KINETIC STUDY

Jose Neptuno Rodriguez-Lopez; Josefa Hernández-Ruiz; Francisco García-Cánovas; Roger N. F. Thorneley; Manuel Acosta; Marino B. Arnao

The kinetics of the catalytic cycle and irreversible inactivation of horseradish peroxidase C (HRP-C) reacting with m-chloroperoxybenzoic acid (mCPBA) have been studied by conventional and stopped-flow spectrophotometry. mCPBA oxidized HRP-C to compound I with a second order-rate constant k1 = 3.6 × 107 M−1 s−1 at pH 7.0, 25°C. Excess mCPBA subsequently acted as a one-electron reducing substrate, converting compound I to compound II and compound II to resting, ferric enzyme. In both of these reactions, spectrally distinct, transient forms of the enzyme were observed (λmax = 411 nm, ε = 45 mM−1 cm−1 for compound I with mCPBA, and λmax = 408 nm, ε = 77 mM−1 cm−1 for compound II with mCPBA). The compound I-mCPBA intermediate (shown by near infrared spectroscopy to be identical to P965) decayed either to compound II in a catalytic cycle (k3 = 6.4 × 10−3 s−1) or, in a competing inactivation reaction, to verdohemoprotein (ki = 3.3 × 10−3 s−1). Thus, a partition ratio of r = 2 is obtained for the inactivation of ferric HRP-C by mCPBA. The intermediate formed from compound II with mCPBA is not part of the inactivation pathway and only decays via the catalytic cycle to give resting, ferric enzyme (k5 = 1.0 × 10−3 s−1). The data are compared with those from earlier steady-state kinetic studies and demonstrate the importance of single turnover experiments. The results are discussed in terms of the physiologically relevant reactions of plant peroxidases with hydrogen peroxide.


Journal of Biological Chemistry | 1997

Salicylic Acid Is a Reducing Substrate and Not an Effective Inhibitor of Ascorbate Peroxidase

Mamuka Kvaratskhelia; Simon J. George; Roger N. F. Thorneley

This communication describes the interactions of salicylic acid (SA) with plant ascorbate peroxidase (APX). Contrary to a recent report (Durner, J., and Klessig, D. F. (1995) Proc. Natl. Acad. Sci. U.u2009S.u2009A. 92, 11312–11316) we show conclusively that ascorbate oxidation by APX is not inhibited by SA (10 mm), but that SA is a slow reducing substrate of this enzyme. The suggestion that SA-dependent inhibition of APXin planta may result in the elevation of H2O2 levels, which in turn acts as a second messenger in systemic acquired resistance signaling, is therefore not tenable. We conclude that APX remains a key antioxidant during systemic acquired resistance following pathogenic infection of plants. The transient products of SA oxidation by APX appear to be SA free radicals that undergo subsequent chemistry. APX-dependent oxidation of SA could be essential for diminishing the detrimental effects of this phenolic acid on plant cells.


Journal of Biological Inorganic Chemistry | 1996

Nitrogenase: substrate binding and activation

Roger N. F. Thorneley; David J. Lowe

Abstractu2002This commentary discusses the structure and function of the MoFe-protein of nitrogenase. The aim is to help the reader appreciate the ways in which the biomimetic and theoretical chemistry described in the four following commentaries contribute to understanding how protons, dinitrogen, substrate analogues and inhibitors are bound and subsequently reduced at the metal clusters present in nitrogenase.


Biochemical Journal | 1999

Reversible alkaline inactivation of lignin peroxidase involves the release of both the distal and proximal site calcium ions and bishistidine co-ordination of the haem.

Simon J. George; Mamuka Kvaratskhelia; M. J. Dilworth; Roger N. F. Thorneley

Phanerochaete chrysosporium lignin peroxidase isoenzyme H2 (LiP H2) exhibits a transition to a stable, inactive form at pH 9.0 with concomitant spectroscopic changes. The Söret peak intensity decreases some 55% with a red shift from 408 to 412 nm; the bands at 502 nm and 638 nm disappear and the peak at 536 nm increases. The EPR spectrum changes from a signal typical of high spin ferric haem to an exclusively low spin spectrum with g=2.92, 2.27, 1.50. These data indicate that the active pentaco-ordinated haem is converted into a hexaco-ordinated species at alkaline pH. Room temperature near-IR MCD data coupled with the EPR spectrum allow us to assign the haem co-ordination of alkali-inactivated enzyme as bishistidine. Re-acidification of the alkali-inactivated enzyme to pH 6 induces further spectroscopic changes and generates an irreversibly inactivated species. By contrast, a pH shift from 9.0 to 6.0 with simultaneous addition of 50 mM CaCl(2) results in the recovery of the initial activity together with the spectroscopic characteristics of the native ferric enzyme. Incubating with 50 mM CaCl(2) at a pH between 6.0 and 9.0 can also re-activate the enzyme. Divalent metals other than Ca(2+) do not result in restoration of activity. Experiments with (45)Ca indicate that two tightly bound calcium ions per enzyme monomer are lost during inactivation and reincorporated during subsequent re-activation, consistent with the presence of two structural Ca(2+) ions in LiP H2. It is concluded that both the structural Ca(2+) ions play key roles in the reversible alkaline inactivation of LiP H2.

Collaboration


Dive into the Roger N. F. Thorneley's collaboration.

Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Simon J. George

Lawrence Berkeley National Laboratory

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Chris Abell

University of Cambridge

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Researchain Logo
Decentralizing Knowledge