Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 20 de 30
Filter
1.
J Chem Phys ; 147(11): 114301, 2017 Sep 21.
Article in English | MEDLINE | ID: mdl-28938810

ABSTRACT

To facilitate direct spectroscopic observation of hydrogen chloride anions (HCl-), electron bombardment of CH3Cl diluted in excess Ar during matrix deposition was used to generate this anion. Subsequent characterization were performed by IR spectroscopy and quantum chemical calculations. Moreover the band intensity of HCl- decays slowly when the matrix sample is maintained in the dark for a prolonged time. High-level ab inito calculation suggested that HCl- is only weakly bound. Atom-in-molecule charge analysis indicated that both atoms of HCl- are negatively charged and the Cl atom is hypervalent.

2.
Inorg Chem ; 52(19): 10992-1003, 2013 Oct 07.
Article in English | MEDLINE | ID: mdl-24066790

ABSTRACT

A series of luminescent dinuclear platinum(II) complexes incorporating diphenylpyrazine-based bridging ligands (L(n)H2) has been prepared. Both 2,5-diphenylpyrazine (L(2)H2) and 2,3-diphenylpyrazine (L(3)H2) are able to undergo cyclometalation of the two phenyl rings, with each metal ion binding to the two nitrogen atoms of the central heterocycle, giving, after treatment with the anion of dipivaloyl methane (dpm), complexes of formula {Pt(dpm)}2L(n). These compounds are isomers of the analogous complex of 4,6-diphenylpyrimidine (L(1)H2). Related complexes of dibenzo(f,h)quinoxaline (L(4)H2), 2,3-diphenyl-quinoxaline (L(5)H2), and dibenzo[3,2-a:2',3'-c]phenazine (L(6)H2) have also been prepared, allowing the effects of strapping together the phenyl rings (L(4)H2 and L(6)H2) and/or extension of the conjugation from pyrazine to quinoxaline (L(5)H2 and L(6)H2) to be investigated. In all cases, the corresponding mononuclear complexes, Pt(dpm)L(n)H, have been isolated too. All 12 complexes are phosphorescent in solution at ambient temperature. Emission spectra of the dinuclear complexes are consistently red shifted compared to their mononuclear analogues, as are the lowest energy absorption bands. Electrochemical data and TD-DFT calculations suggest that this effect arises primarily from stabilization of the LUMO. Introduction of the second metal ion also has the effect of substantially increasing the molar absorptivity and, in most cases, the radiative rate constants. Meanwhile, extension of conjugation in the heterocycle of L(5)H2 and L(6)H2 and planarization of the aromatic system favored by interannular bond formation in L(4)H2 and L(6)H2 leads to further red shifts of the absorption and emission spectra to wavelengths that are unusually long for cyclometalated platinum(II) complexes. The results may offer a versatile design strategy for tuning and optimizing the optical properties of d-block metal complexes for contemporary applications.

3.
Org Biomol Chem ; 11(2): 309-17, 2013 Jan 14.
Article in English | MEDLINE | ID: mdl-23188177

ABSTRACT

Our recent work has provided new insights into the equilibria and species that exist in aqueous solution at different pHs for the boric acid - hydrogen peroxide system, and the role of these species in oxidation reactions. Most recently, (M. C. Durrant, D. M. Davies and M. E. Deary, Org. Biomol. Chem., 2011, 9, 7249-7254), we have produced strong theoretical and experimental evidence for the existence of a previously unreported monocyclic three membered peroxide species, dioxaborirane, that is the likely catalytic species in borate mediated electrophilic reactions of hydrogen peroxide in alkaline solution. In the present paper, we extend our study of the borate-peroxide system to look at a wide range of substrates that include substituted dimethyl anilines, methyl-p-tolyl sulfoxide, halides, hydrogen sulfide anion, thiosulfate, thiocyanate, and hydrazine. The unusual selectivity-reactivity pattern of borate catalysed reactions compared with hydrogen peroxide and inorganic or organic peracids previously observed for the organic sulfides (D. M. Davies, M. E. Deary, K. Quill and R. A. Smith, Chem.-Eur. J., 2005, 11, 3552-3558) is also seen with substituted dimethyl aniline nucleophiles. This provides evidence that the pattern is not due to any latent electrophilic tendency of the organic sulfides and further supports dioxaborirane being the likely reactive intermediate, thus broadening the applicability of this catalytic system. Moreover, density functional theory calculations on our proposed mechanism involving dioxaborirane are consistent with the experimental results for these substrates. Results obtained at high concentrations of both borate and hydrogen peroxide require the inclusion the diperoxodiborate dianion in the kinetic analysis. A scheme detailing our current understanding of the borate-peroxide system is presented.


Subject(s)
Borates/chemistry , Boron Compounds/chemistry , Hydrogen Peroxide/chemistry , Models, Chemical , Catalysis , Hydrazines/chemistry , Kinetics , Thermodynamics
4.
Inorg Chem ; 50(13): 6304-13, 2011 Jul 04.
Article in English | MEDLINE | ID: mdl-21627075

ABSTRACT

The proligand 4,6-di-(4-tert-butylphenyl)pyrimidine LH(2) can undergo cycloplatination with K(2)PtCl(4) at one of the two aryl rings to give, after treatment with sodium acetylacetonate, a mononuclear complex Pt(N^C-LH)(acac) (denoted Pt). If an excess of K(2)PtCl(4) is used, a dinuclear complex of the form [Pt(acac)](2){µ-(N^C-L-N^C)} (Pt(2)) is obtained instead, where the pyrimidine ring acts as a bridging unit. Alternatively, the mononuclear complex can undergo cyclometalation with a different metal ion. Thus, reaction of Pt with IrCl(3)·3H(2)O (2:1 ratio) leads, after treatment with sodium acetylacetonate, to an unprecedented mixed-metal complex of the form Ir{µ-(N^C-L-N^C)Pt(acac)}(2)(acac) (Pt(2)Ir). The mononuclear iridium complex Ir(N^C-LH)(2)(acac) (Ir) has also been prepared for comparison. The UV-visible absorption and photoluminesence properties of the four complexes and of the proligand have been investigated. The complexes are all highly luminescent, with quantum yields of around 0.5 in solution at room temperature. The introduction of the additional metal centers is found to lead to a substantial red-shift in absorption and emission, with λ(max) in the order Pt < Pt(2) < Ir < Pt(2)Ir. The trend is interpreted with the aid of electrochemical data and density functional theory calculations, which suggest that the red-shift is due primarily to a progressive stabilization of the lowest unoccupied molecular orbital (LUMO). The radiative decay constant is also increased. This versatile design strategy may offer a new approach for tuning and optimizing the luminescence properties of d-block metal complexes for contemporary applications.


Subject(s)
Iridium/chemistry , Luminescence , Organometallic Compounds/chemistry , Platinum/chemistry , Pyrimidines/chemistry , Cyclization , Molecular Structure , Organometallic Compounds/chemical synthesis
5.
Org Biomol Chem ; 9(20): 7249-54, 2011 Oct 21.
Article in English | MEDLINE | ID: mdl-21881663

ABSTRACT

This paper reports on a kinetic and theoretical study into the borate mediated reaction of dimethyl sulfide with hydrogen peroxide in both acid and alkaline conditions. At high pH, whilst the kinetic data is consistent with the catalytic species being monoperoxoborate, formed from the rapid equilibrium between hydrogen peroxide and boric acid, DFT calculations show that this species is in fact less reactive than hydrogen peroxide, requiring us to seek an alternative catalytic mechanism. DFT provides an important insight for this, showing that although boric acid and peroxoboric acid are primarily Lewis acids, they can exhibit a small degree of Brønsted acidity, allowing, respectively, the B(O)(OH)(2)(-) and HOOB(OH)(O)(-) anions to exist in small concentrations. Whilst the peroxoborate anion, HOOB(OH)(O)(-), is predicted to have only marginal catalytic activity, its tautomer, dioxaborirane, (HO)(2)BO(2)(-), a three membered cyclic peroxide, has a very low activation barrier of 2.8 kcal/mol. Hence, even though dioxaborirane is likely to be present in very low concentrations, it is still sufficiently reactive for overall rate enhancements to be observed for this system. This is the first literature report of this species. The observed low selectivity observed for borate catalysed reactions of hydrogen peroxide with a range of substituted phenyl methyl sulfides in our previous study (D. M. Davies, M. E. Deary, K. Quill and R. A. Smith, Chem.-Eur. J. 2005, 11, 3552-3558) is further evidence in favour of a highly reactive catalytic species. At low pH, kinetic data shows that borate catalyses the reaction between hydrogen peroxide and dimethyl sulfide; this is supported by DFT calculations that predict peroxoboric acid to be an effective catalytic intermediate, with an energy barrier of 7.4 kcal mol(-1) compared to 10.1 kcal mol(-1) for the uncatalysed system. Nevertheless, the overall contribution of this pathway is small because of the unfavourable equilibrium between hydrogen peroxide and boric acid to form peroxoboric acid.

6.
Nature ; 438(7070): 1013-6, 2005 Dec 15.
Article in English | MEDLINE | ID: mdl-16355224

ABSTRACT

Root hairs are cellular protuberances extending from the root surface into the soil; there they provide access to immobile inorganic ions such as phosphate, which are essential for growth. Their cylindrical shape results from a polarized mechanism of cell expansion called tip growth in which elongation is restricted to a small area at the surface of the hair-forming cell (trichoblast) tip. Here we identify proteins that spatially control the sites at which cell growth occurs by isolating Arabidopsis mutants (scn1) that develop ectopic sites of growth on trichoblasts. We cloned SCN1 and showed that SCN1 is a RhoGTPase GDP dissociation inhibitor (RhoGDI) that spatially restricts the sites of growth to a single point on the trichoblast. We showed previously that localized production of reactive oxygen species by RHD2/AtrbohC NADPH oxidase is required for hair growth; here we show that SCN1/AtrhoGDI1 is a component of the mechanism that focuses RHD2/AtrbohC-catalysed production of reactive oxygen species to hair tips during wild-type development. We propose that the spatial organization of growth in plant cells requires the local RhoGDI-regulated activation of the RHD2/AtrbohC NADPH oxidase.


Subject(s)
Arabidopsis Proteins/metabolism , Arabidopsis/metabolism , Cation Transport Proteins/metabolism , Guanine Nucleotide Dissociation Inhibitors/metabolism , Plant Roots/cytology , Arabidopsis/enzymology , Arabidopsis/genetics , Arabidopsis Proteins/genetics , Cation Transport Proteins/genetics , Cloning, Molecular , Guanine Nucleotide Dissociation Inhibitors/genetics , Humans , Morphogenesis , Mutation/genetics , NADPH Oxidases/genetics , NADPH Oxidases/metabolism , Phenotype , Plant Roots/metabolism , Reactive Oxygen Species/metabolism , Suppression, Genetic/genetics , rho GTP-Binding Proteins/metabolism , rho Guanine Nucleotide Dissociation Inhibitor alpha , rho-Specific Guanine Nucleotide Dissociation Inhibitors
7.
Dalton Trans ; 47(43): 15364-15381, 2018 Nov 21.
Article in English | MEDLINE | ID: mdl-30298161

ABSTRACT

During the red blood cell phase of their life cycle, malaria parasites digest their host's haemoglobin, with concomitant release of potentially toxic iron(iii) protoporphyrin IX (FePPIX). The parasites' strategy for detoxification of FePPIX involves its crystallization to haemozoin, such that the build-up of free haem in solution is avoided. Antimalarial drugs of both historical importance and current clinical use are known to be capable of disrupting the growth of crystals of ß-haematin, which is the synthetic equivalent of haemozoin. Hence, the disruption of haemozoin crystal growth is implicated as a possible mode of action of such drugs. However, the details of ß-haematin crystal poisoning at the molecular level have yet to be fully elucidated. In this study, we have used a combination of density functional theory (DFT) and molecular modelling to examine the possible modes of action of ten different antimalarial drugs, including quinine-type aliphatic alcohols, amodiaquine-type phenols, and chloroquine-type aliphatic diamines. The DFT calculations indicate that each of the drugs can form at least one molecular complex with FePPIX. These complexes have 1 : 1 or 2 : 1 FePPIX : drug stoichiometries and all of them incorporate Fe-O bonds, formed either by direct coordination of a zwitterionic form of the drug, or by deprotonation of water. Most of the drugs can form more than one such complex. We have used the DFT model structures to explore the possible formation of a monolayer of each drug-haem complex on four of the ß-haematin crystal faces. In all cases, the drug complexes can form a monolayer on the fast-growing {001} and {011} faces, but not on the slower growing {010} and {100} faces. Additional modelling of the chloroquine and quinidine complexes shows that individual molecules of these species can also obstruct the growth of new layers on other crystal faces. The implications of these observations for antimalarial drug development are discussed.


Subject(s)
Antimalarials/chemistry , Antimalarials/pharmacology , Hemin/chemistry , Models, Molecular , Crystallization , Iron/chemistry , Molecular Conformation , Protoporphyrins/chemistry
8.
Biochimie ; 89(4): 500-7, 2007 Apr.
Article in English | MEDLINE | ID: mdl-17276574

ABSTRACT

Microcin B17 (MccB17) is a peptide-based bacterial toxin that targets DNA gyrase, the bacterial enzyme that introduces supercoils into DNA. The site and mode of action of MccB17 on gyrase are unclear. We review what is currently known about MccB17-gyrase interactions and summarise approaches to understanding its mode of action that involve modification of the toxin. We describe experiments in which treatment of the toxin at high pH leads to the deamidation of two asparagine residues to aspartates. The modified toxin was found to be inactive in vivo and in vitro, suggesting that the Asn residues are essential for activity. Following on from these studies we have used molecular modelling to suggest a 3D structure for microcin B17. We discuss the implications of this model for MccB17 action and investigate the possibility that it binds metal ions.


Subject(s)
Bacteriocins/pharmacology , DNA Gyrase/metabolism , Amino Acid Sequence , Bacterial Proteins/metabolism , Bacteriocins/chemistry , Bacteriocins/isolation & purification , DNA Gyrase/chemistry , Escherichia coli , Models, Molecular , Molecular Sequence Data , Protein Conformation , Spectrometry, Mass, Matrix-Assisted Laser Desorption-Ionization
9.
Biochem J ; 397(2): 261-70, 2006 Jul 15.
Article in English | MEDLINE | ID: mdl-16566750

ABSTRACT

Although it is generally accepted that the active site of nitrogenase is located on the FeMo-cofactor, the exact site(s) of N2 binding and reduction remain the subject of continuing debate, with both molybdenum and iron atoms being suggested as key players. The current consensus favours binding of acetylene and some other non-biologically relevant substrates to the central iron atoms of the FeMo-cofactor [Dos Santos, Igarashi, Lee, Hoffman, Seefeldt and Dean (2005) Acc. Chem. Res. 38, 208-214]. The reduction of N2 is, however, a more demanding process than reduction of these alternative substrates because it has a much higher activation energy and does not bind until three electrons have been accumulated on the enzyme. The possible conversion of bidentate into monodentate homocitrate on this three electron-reduced species has been proposed to free up a binding site for N2 on the molybdenum atom. One of the features of this hypothesis is that alpha-Lys426 facilitates chelate ring opening and subsequent orientation of the monodentate homocitrate by forming a specific hydrogen bond to the homocitrate -CH2CH2CO2- carboxylate group. In support of this concept, we show that mutation of alpha-Lys426 can selectively perturb N2 reduction without affecting acetylene reduction. We interpret our experimental observations in the light of a detailed molecular mechanics modelling study of the wild-type and altered MoFe-nitrogenases.


Subject(s)
Azotobacter vinelandii/metabolism , Molybdoferredoxin/chemistry , Nitrogen Fixation , Nitrogen/chemistry , Tricarboxylic Acids/chemistry , Catalysis , Hydrogen Bonding , Lysine/chemistry , Models, Chemical , Models, Molecular , Molybdenum/chemistry
10.
Cancer Res ; 65(6): 2059-64, 2005 Mar 15.
Article in English | MEDLINE | ID: mdl-15781612

ABSTRACT

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.


Subject(s)
Catechin/analogs & derivatives , Catechin/chemistry , Catechin/pharmacology , Folic Acid Antagonists/pharmacology , Tea/chemistry , Tetrahydrofolate Dehydrogenase/chemistry , Tetrahydrofolate Dehydrogenase/metabolism , Animals , Catechin/metabolism , Cattle , Chickens , Enzyme Inhibitors/chemistry , Enzyme Inhibitors/metabolism , Enzyme Inhibitors/pharmacology , Folic Acid Antagonists/chemistry , Folic Acid Antagonists/metabolism , Humans , Kinetics , Leukemia L1210/drug therapy , Leukemia L1210/enzymology , Liver/enzymology , Mice , Models, Molecular , Quinazolines/chemistry , Quinazolines/metabolism
11.
Chem Sci ; 7(5): 3448-3449, 2016 May 01.
Article in English | MEDLINE | ID: mdl-29999044

ABSTRACT

The Lewis and quantum mechanical theories of chemical bonding are compared and contrasted, with a view to clarifying the relationship between Harcourt's 'increased valence' quantum approach and the recently proposed quantitative definition of hypervalency.

12.
Steroids ; 113: 95-102, 2016 09.
Article in English | MEDLINE | ID: mdl-27421190

ABSTRACT

Bioconversion of the aromatase inhibitor formestane (4-hydroxyandrost-4-ene-3,17-dione) (1) by the fungus Rhizopus oryzae ATCC 11145 resulted in a new minor metabolite 3,5α-dihydroxyandrost-2-ene-4,17-dione (2) and the known 4ß,5α-dihydroxyandrostane-4,17-dione (3) as the major product. The structural elucidation and bioactivities of these metabolites are reported herein. Molecular modeling studies of the interactions between these metabolites and the aromatase protein indicated that acidic (D309), basic (R115), polar (T310), aromatic (F134, F221, and W224), and non-polar (I133, I305, A306, V369, V370, L372, V373, M374, and L477) amino acid residues contribute important interactions with the steroidal substrates. These combined experimental and theoretical studies provide fresh insights for the further development of more potent aromatase inhibitors.


Subject(s)
Aromatase Inhibitors/chemistry , Aromatase Inhibitors/metabolism , Biotransformation , Models, Molecular , Molecular Docking Simulation , Rhizopus/metabolism
13.
Dalton Trans ; 45(16): 6949-62, 2016 Apr 28.
Article in English | MEDLINE | ID: mdl-26983757

ABSTRACT

A new family of eight dinuclear iridium(iii) complexes has been prepared, featuring 4,6-diarylpyrimidines L(y) as bis-N^C-coordinating bridging ligands. The metal ions are also coordinated by a terminal N^C^N-cyclometallating ligand L(X) based on 1,3-di(2-pyridyl)benzene, and by a monodentate chloride or cyanide. The general formula of the compounds is {IrL(X)Z}2L(y) (Z = Cl or CN). The family comprises examples with three different L(X) ligands and five different diarylpyrimidines L(y), of which four are diphenylpyrimidines and one is a dithienylpyrimidine. The requisite proligands have been synthesised via standard cross-coupling methodology. The synthesis of the complexes involves a two-step procedure, in which L(X)H is reacted with IrCl3·3H2O to form dinuclear complexes of the form [IrL(X)Cl(µ-Cl)]2, followed by treatment with the diarylpyrimidine L(y)H2. Crucially, each complex is formed as a single compound only: the strong trans influence of the metallated rings dictates the relative disposition of the ligands, whilst the use of symmetrically substituted tridentate ligands eliminates the possibility of Λ and Δ enantiomers that are obtained when bis-bidentate units are linked through bridging ligands. The crystal structure of one member of the family has been obtained using a synchrotron X-ray source. All of the complexes are very brightly luminescent, with emission maxima in solution varying over the range 517-572 nm, according to the identity of the ligands. The highest-energy emitter is the cyanide derivative whilst the lowest is the complex with the dithienylpyrimidine. The trends in both the absorption and emission energies as a function of ligand substituent have been rationalised accurately with the aid of TD-DFT calculations. The lowest-excited singlet and triplet levels correlate with the trend in the HOMO-LUMO gap. All the complexes have quantum yields that are close to unity and phosphorescence lifetimes - of the order of 500 ns - that are unusually short for complexes of such brightness. These impressive properties stem from an unusually high rate of radiative decay, possibly due to spin-orbit coupling pathways being facilitated by the second metal ion, and to low non-radiative decay rates that may be related to the rigidity of the dinuclear scaffold.

14.
Mol Plant Microbe Interact ; 18(1): 24-32, 2005 Jan.
Article in English | MEDLINE | ID: mdl-15672815

ABSTRACT

Root nodule extensins (RNEs) are highly glycosylated plant glycoproteins localized in the extracellular matrix of legume tissues and in the lumen of Rhizobium-induced infection threads. In pea and other legumes, a family of genes encode glycoproteins of different overall length but with the same basic composition. The predicted polypeptide sequence reveals repeating and alternating motifs characteristic of extensins and arabinogalactan proteins. In order to monitor the behavior of individual RNE gene products in the plant extracellular matrix, the coding sequence of PsRNE1 from Pisum sativum was expressed in insect cells and in tobacco leaves. RNE products extracted from tobacco tissues were of high molecular weight (in excess of 80 kDa), indicating extensive glycosylation similar to that in pea tissues. Epitope-tagged derivatives of PsRNE1 could be localized in cell walls. However, the introduction of epitope tags at the C-terminus of RNE altered the behavior of RNE in the extracellular matrix, apparently preventing intermolecular crosslinking of RNE molecules and their covalent association with other cell wall components. These observations are discussed in the light of a computational model for the RNE glycoprotein that is consistent with an extended rod-like structure. It is proposed that RNE can undergo three classes of tyrosine-based crosslinking. Intramolecular crosslinking of vicinal Tyr residues is rod stiffening, end-to-end linkage is rod lengthening, and side-to-side intermolecular crosslinking is rod bundling. The control of these interconversions could have important implications for the biomechanics of infection thread growth.


Subject(s)
Glycoproteins/physiology , Pisum sativum/physiology , Plant Proteins/physiology , Amino Acid Sequence , Animals , Cell Line , Cell Wall/chemistry , Epitopes , Fabaceae/chemistry , Glycoproteins/chemistry , Models, Molecular , Molecular Sequence Data , Organisms, Genetically Modified , Plant Proteins/chemistry , Protein Conformation , Nicotiana/genetics
15.
J Mol Biol ; 334(5): 933-47, 2003 Dec 12.
Article in English | MEDLINE | ID: mdl-14643658

ABSTRACT

The key protein in the initiation of Helicobacter pylori chromosome replication, DnaA, has been characterized. The amount of the DnaA protein was estimated to be approximately 3000 molecules per single cell; a large part of the protein was found in the inner membrane. The H.pylori DnaA protein has been analysed using in vitro (gel retardation assay and surface plasmon resonance (SPR)) as well as in silico (comparative computer modeling) studies. DnaA binds a single DnaA box as a monomer, while binding to the fragment containing several DnaA box motifs, the oriC region, leads to the formation of high molecular mass nucleoprotein complexes. In comparison with the Escherichia coli DnaA, the H.pylori DnaA protein exhibits lower DNA-binding specificity; however, it prefers oriC over non-box DNA fragments. As determined by gel retardation techniques, the H.pylori DnaA binds with a moderate level of affinity to its origin of replication (4nM). Comparative computer modelling showed that there are nine residues within the binding domain which are possible determinants of the reduced H.pylori DnaA specificity. Of these, the most interesting is probably the triad PTL; all three residues show significant divergence from the consensus, and Thr398 is the most divergent residue of all.


Subject(s)
Bacterial Proteins/metabolism , DNA, Bacterial/metabolism , DNA-Binding Proteins/metabolism , Helicobacter pylori/metabolism , Bacterial Proteins/chemistry , Base Sequence , DNA Primers , DNA-Binding Proteins/chemistry , Models, Molecular , Protein Binding , Surface Plasmon Resonance
16.
Chem Sci ; 6(11): 6614-6623, 2015 Nov 01.
Article in English | MEDLINE | ID: mdl-30090275

ABSTRACT

From the inception of Lewis' theory of chemical bonding, hypervalency has remained a point of difficulty that has not been fully resolved by the currently accepted qualitative definition of this term. Therefore, in this work, a quantitative measure of hypervalency has been developed. The only required input is the atomic charge map, which can be obtained from either quantum calculations or from experiment. Using this definition, it is found that well-known species such as O3, CH2N2 and ClO4-, are indeed hypervalent, whilst others such as XeF4, PCl5 and SO42-, are not. Quantitative analysis of known species of general formulae XF nm-, XCl nm-, and XO nm- shows that there are no fundamental differences in chemical bonding for hypervalent and non-hypervalent species. Nevertheless, hypervalency is associated with chemical instability, as well as a high degree of covalent rather than ionic bonding. The implications for accepted Lewis structure conventions are discussed.

17.
Dalton Trans ; 43(25): 9754-65, 2014 Jul 07.
Article in English | MEDLINE | ID: mdl-24846575

ABSTRACT

The search for novel anti-malarial drugs that can disrupt biomineralization of ferriprotoporphyrin IX to haemozoin requires an understanding of the fundamental chemistry of the porphyrin's iron(iii) centre at the water-lipid interface. Towards this end, the binding affinities for a diverse set of 31 small ligands with iron(iii) porphine have been calculated using density functional theory, in the gas phase and also with implicit solvent corrections for both water and n-octanol. In addition, the binding of hydroxide, chloride, acetate, methylamine and water to ferriprotoporphyrin IX has been studied, and very similar trends are observed for the smaller and larger models. Anionic ligands generally give stronger binding than neutral ones; the strongest binding is observed for RO(-) and OH(-) ligands, whilst acetate binds relatively weakly among the anions studied. Electron-rich nitrogen donors tend to bind more strongly than electron-deficient ones, and the weakest binding is found for neutral O and S donors such as oxazole and thiophene. In all cases, ligand binding is stronger in n-octanol than in water, and the differences in binding energies for the two solvents are greater for ionic ligands than for neutrals. Finally, dimerization of ferriprotoporphyrin IX by means of iron(iii)-carboxylate bond formation has been modelled. The results are discussed in terms of haemozoin crystal growth and its disruption by known anti-malarial drugs.


Subject(s)
Computational Biology , Coordination Complexes/chemistry , Iron/chemistry , Organometallic Compounds/chemistry , Porphyrins/chemistry , Protoporphyrins/chemistry , 1-Octanol/chemistry , Antimalarials/chemistry , Electrons , Hemeproteins/chemistry , Hydroxides/chemistry , Ligands , Models, Chemical , Models, Molecular , Quantum Theory , Solvents/chemistry , Water/chemistry
18.
J Mol Graph Model ; 39: 108-17, 2013 Feb.
Article in English | MEDLINE | ID: mdl-23261880

ABSTRACT

Human bone sialoprotein (BSP) is an essential component of the extracellular matrix of bone. It is thought to be the primary nucleator of hydroxyapatite crystallization, and is known to bind to hydroxyapatite, collagen, and cells. Mature BSP shows extensive post-translational modifications, including attachment of glycans, sulfation, and phosphorylation, and is highly flexible with no specific 2D or 3D structure in solution or the solid state. These features have severely limited the experimental characterization of the structure of this protein. We have therefore developed a 3D structural model for BSP, based on the available literature data, using molecular modelling techniques. The complete model consists of 301 amino acids, including six phosphorylated serines and two sulfated tyrosines, plus 92 N- and O-linked glycan residues. A notable feature of the model is a large acidic patch that provides a surface for binding Ca(2+) ions. Density functional theory quantum calculations with an implicit solvent model indicate that Ca(2+) ions are bound most strongly by the phosphorylated serines within BSP, along with reasonably strong binding to Asp and Glu, but weak binding to His and sulfated tyrosine. The process of early hydroxyapatite nucleation has been studied by molecular dynamics on an acidic surface loop of the protein; the results suggest that the cationic nature of the loop promotes nucleation by attracting Ca(2+) ions, while its flexibility allows for their rapid self-assembly with PO(4)(3-) ions, rather than providing a regular template for crystallization. The binding of a hydroxyapatite crystal at the protein's acidic patch has also been modelled. The relationships between hydroxyapatite, collagen and BSP are discussed.


Subject(s)
Integrin-Binding Sialoprotein/chemistry , Models, Molecular , Amino Acid Sequence , Bone and Bones/chemistry , Durapatite/chemistry , Humans , Integrin-Binding Sialoprotein/metabolism , Ions/chemistry , Ions/metabolism , Molecular Sequence Data , Protein Binding , Protein Conformation , Quantum Theory , Sequence Alignment , Structure-Activity Relationship
19.
J Mol Graph Model ; 30: 120-8, 2011 Sep.
Article in English | MEDLINE | ID: mdl-21798775

ABSTRACT

An empirical method for estimation of the boiling points of organic molecules based on density functional theory (DFT) calculations with polarized continuum model (PCM) solvent corrections has been developed. The boiling points are calculated as the sum of three contributions. The first term is calculated directly from the structural formula of the molecule, and is related to its effective surface area. The second is a measure of the electronic interactions between molecules, based on the DFT-PCM solvation energy, and the third is employed only for planar aromatic molecules. The method is applicable to a very diverse range of organic molecules, with normal boiling points in the range of -50 to 500 °C, and includes ten different elements (C, H, Br, Cl, F, N, O, P, S and Si). Plots of observed versus calculated boiling points gave R²=0.980 for a training set of 317 molecules, and R²=0.979 for a test set of 74 molecules. The role of intramolecular hydrogen bonding in lowering the boiling points of certain molecules is quantitatively discussed.


Subject(s)
Computer Simulation , Models, Chemical , Solvents/chemistry , Transition Temperature , Algorithms , Hydrogen Bonding , Organic Chemicals/chemistry
20.
Dalton Trans ; (46): 10223-30, 2009 Dec 14.
Article in English | MEDLINE | ID: mdl-19921057

ABSTRACT

The deprotonation energies of the water ligands in a set of 40 d-block metal complexes have been calculated using density functional theory with polarized continuum model solvent corrections. The complexes include 13 aqua ions [M(OH(2))(n)](2+/3+) and a variety of aqua complexes with organic co-ligands, whose experimental pK(a) values have been reported in the literature. For comparison, the deprotonation energies of a set of 60 organic and inorganic molecules with experimental pK(a) values ranging from -25 (HSbF(6)) to +52 (C(2)H(6)) have also been calculated. Three different classes of acids are identified as giving different slopes in plots of pK(a) versus deprotonation energies; namely non-hydroxy acids, hydroxy acids, and the metal complexes. The correlation coefficients for the straight lines obtained for these three classes are 0.96, 0.97 and 0.92 respectively. Better correlations are found for sub-sets of the complexes, such as the 31 first row complexes (correlation coefficient 0.95).For several of the complexes, comparison of the calculated and observed pK(a) values, together with changes in the geometry upon optimization, offer new insights into the possible solution structures. It is concluded that DFT calculations incorporating solvent corrections can be used to give reasonable estimates of pK(a) values for the aqua ligands in a range of complex types.


Subject(s)
Models, Molecular , Transition Elements/chemistry , Water/chemistry , Hydrogen-Ion Concentration , Ions , Ligands , Protons , Solvents
SELECTION OF CITATIONS
SEARCH DETAIL