Network


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

Hotspot


Dive into the research topics where Lynda J. Donald is active.

Publication


Featured researches published by Lynda J. Donald.


Biochemical Pharmacology | 1998

Mechanisms of Inhibition of the Thioredoxin Growth Factor System by Antitumor 2-Imidazolyl Disulfides

D. Lynn Kirkpatrick; Miles Kuperus; Marla Dowdeswell; Noelle Potier; Lynda J. Donald; Mark Kunkel; Margareta Berggren; Miguel Angulo; Garth Powis

The interactions of a series of 2-imidazolyl disulfide antitumor compounds with the thioredoxin reductase(TR)/thioredoxin (hTrx) redox system have been studied. Disulfides III-2 (n-butyl 2-mercaptoimidazolyl disulfide) and VI-2 (ethyl 2-mercaptoimidazolyl disulfide) were substrates for reduction by TR with Km values of 43 and 48 microM. Disulfides IV-2 (1-methylpropyl 2-mercaptoimidazolyl disulfide) and DLK-36 (benzyl 2-mercaptoimidazolyl disulfide) were competitive inhibitors of the reduction of hTrx by TR with Ki values of 31 microM. None of the disulfides were substrates for reduction by human glutathione reductase. The disulfides caused reversible thioalkylation of hTrx at the redox catalytic site as shown by the fact that there was no thioalkylation of a mutant hTrx where both the catalytic site Cys32 and Cys35 residues were replaced by Ser. In addition, the disulfides caused a slower irreversible inactivation of hTrx as a substrate for reduction by TR, with half-lives for III-2 of 30 min, for IV-2 of 4 hr, and for IX-2 (t-butyl 2-mercaptoimidazolyl disulfide) of 24 hr. This irreversible inactivation of hTrx occurred at concentrations of the disulfides an order of magnitude below those that inhibited TR, and involved the Cys73 of hTrx, which is outside the conserved redox catalytic site, as shown by the resistance to inactivation of a mutant hTrx where Cys73 was replaced by Ser. Electrophoretic and mass spectral analyses of the products of the reaction between the disulfides and hTrx show that modification of 1-3 Cys residues of the protein occurred in a concentration-dependent fashion. The disulfides inhibited the hTrx-dependent proliferation of MCF-7 breast cancer cells with IC50 values for III-2 and IV-2 of 0.2 and 1.2 microM, respectively. The results show that although the catalytic sites of TR and hTrx are reversibly inhibited by the 2-imidazolyl disulfides, it is the irreversible thioalkylation of Cys73 of hTrx by the disulfides that most probably accounts for the inhibition of thioredoxin-dependent cell growth by the disulfides.


Journal of Biological Chemistry | 2004

Catalase-peroxidases (KatG) Exhibit NADH Oxidase Activity

Rahul Singh; Ben Wiseman; Taweewat Deemagarn; Lynda J. Donald; Harry W. Duckworth; Xavier Carpena; Ignacio Fita; Peter C. Loewen

Catalase-peroxidases (KatG) produced by Burkholderia pseudomallei, Escherichia coli, and Mycobacterium tuberculosis catalyze the oxidation of NADH to form NAD+ and either H2O2 or superoxide radical depending on pH. The NADH oxidase reaction requires molecular oxygen, does not require hydrogen peroxide, is not inhibited by superoxide dismutase or catalase, and has a pH optimum of 8.75, clearly differentiating it from the peroxidase and catalase reactions with pH optima of 5.5 and 6.5, respectively, and from the NADH peroxidase-oxidase reaction of horseradish peroxidase. B. pseudomallei KatG has a relatively high affinity for NADH (Km = 12 μm), but the oxidase reaction is slow (kcat = 0.54 min-1) compared with the peroxidase and catalase reactions. The catalase-peroxidases also catalyze the hydrazinolysis of isonicotinic acid hydrazide (INH) in an oxygen- and H2O2-independent reaction, and KatG-dependent radical generation from a mixture of NADH and INH is two to three times faster than the combined rates of separate reactions with NADH and INH alone. The major products from the coupled reaction, identified by high pressure liquid chromatography fractionation and mass spectrometry, are NAD+ and isonicotinoyl-NAD, the activated form of isoniazid that inhibits mycolic acid synthesis in M. tuberculosis. Isonicotinoyl-NAD synthesis from a mixture of NAD+ and INH is KatG-dependent and is activated by manganese ion. M. tuberculosis KatG catalyzes isonicotinoyl-NAD formation from NAD+ and INH more efficiently than B. pseudomallei KatG.


Journal of Biological Chemistry | 2010

Isonicotinic Acid Hydrazide Conversion to Isonicotinyl-NAD by Catalase-peroxidases

Ben Wiseman; Xavier Carpena; Miguel Feliz; Lynda J. Donald; Miquel Pons; Ignacio Fita; Peter C. Loewen

Activation of the pro-drug isoniazid (INH) as an anti-tubercular drug in Mycobacterium tuberculosis involves its conversion to isonicotinyl-NAD, a reaction that requires the catalase-peroxidase KatG. This report shows that the reaction proceeds in the absence of KatG at a slow rate in a mixture of INH, NAD+, Mn2+, and O2, and that the inclusion of KatG increases the rate by >7 times. Superoxide, generated by either Mn2+- or KatG-catalyzed reduction of O2, is an essential intermediate in the reaction. Elimination of the peroxidatic process by mutation slows the rate of reaction by 60% revealing that the peroxidatic process enhances, but is not essential for isonicotinyl-NAD formation. The isonicotinyl-NAD•+ radical is identified as a reaction intermediate, and its reduction by superoxide is proposed. Binding sites for INH and its co-substrate, NAD+, are identified for the first time in crystal complexes of Burkholderia pseudomallei catalase-peroxidase with INH and NAD+ grown by co-crystallization. The best defined INH binding sites were identified, one in each subunit, on the opposite side of the protein from the entrance to the heme cavity in a funnel-shaped channel. The NAD+ binding site is ∼20 Å from the entrance to the heme cavity and involves interactions primarily with the AMP portion of the molecule in agreement with the NMR saturation transfer difference results.


Protein Science | 2008

Mass spectrometric study of the Escherichia coli repressor proteins, IcIR and GcIR, and their complexes with DNA

Lynda J. Donald; David J. Hosfield; Susan L. Cuvelier; Werner Ens; Kenneth G. Standing; Harry W. Duckworth

In Escherichia coli, the IclR protein regulates both the aceBAK operon and its own synthesis. Database homology searches have identified many IclR‐like proteins, now known as the IclR family, which can be identified by a conserved C‐terminal region. We have cloned and purified one of these proteins, which we have named GclR (glyoxylate carboligase repressor). Although purification is straightforward, both the IclR and GclR proteins are difficult to manipulate, requiring high salt (up to 0.6 M KCl) for solubility. With the advent of nanospray ionization, we could transfer the proteins into much higher concentrations of volatile buffer than had been practical with ordinary electrospray. In 0.5 M ammonium bicarbonate buffer, both proteins were stable as tetramers, with a small amount of dimer. In a separate experiment, we found that IclR protein selected from a random pool a sequence which matched exactly that of the presumed binding region of the GclR protein, although IclR does not regulate the gcl gene. We designed a 29 bp synthetic DNA to which IclR and GclR bind, and with which we were able to form noncovalent DNA‐protein complexes for further mass spectrometry analysis. These complexes were far more stable than the proteins alone, and we have evidence of a stoichiometry which has not been described previously with (protein monomer : dsDNA) = (4 : 1).


Journal of Molecular Biology | 2010

Crystal Structure of the AmpR Effector Binding Domain Provides Insight into the Molecular Regulation of Inducible AmpC β-Lactamase

Misty D. Balcewich; Thomas M. Reeve; Evan Orlikow; Lynda J. Donald; David J. Vocadlo; Brian L. Mark

Hyperproduction of AmpC beta-lactamase (AmpC) is a formidable mechanism of resistance to penicillins and cephalosporins in Gram-negative bacteria such as Pseudomonas aeruginosa and Enterobacteriaceae. AmpC expression is regulated by the LysR-type transcriptional regulator AmpR. ampR and ampC genes form a divergent operon with overlapping promoters to which AmpR binds and regulates the transcription of both genes. AmpR induces ampC by binding to one member of the family of 1,6-anhydro-N-acetylmuramyl peptides, which are cytosolic catabolites of peptidoglycan that accumulate during beta-lactam challenge. To gain structural insights into AmpR regulation, we determined the crystal structure of the effector binding domain (EBD) of AmpR from Citrobacter freundii up to 1.83 A resolution. The AmpR EBD is dimeric and each monomer comprises two subdomains that adopt alpha/beta Rossmann-like folds. Located between the monomer subdomains is a pocket that was found to bind the crystallization buffer molecule 2-(N-morpholino)ethanesulfonic acid. The pocket, together with a groove along the surface of subdomain I, forms a putative effector binding site into which a molecule of 1,6-anhydro-N-acetylmuramyl pentapeptide could be modeled. Amino acid substitutions at the base of the interdomain pocket either were found to render AmpR incapable of inducing ampC (Thr103Val, Ser221Ala and Tyr264Phe) or resulted in constitutive ampC expression (Gly102Glu). While the substitutions that prevented ampC induction did not alter the overall AmpR EBD structure, circular dichroism spectroscopy revealed that the nonconservative Gly102Glu mutation affected EBD secondary structure, confirming previous work suggesting that Gly102Glu induces a conformational change to result in constitutive AmpC production.


Rapid Communications in Mass Spectrometry | 1997

Use of a non‐porous polyurethane membrane as a sample support for matrix‐assisted laser desorption/ionization time‐of‐flight mass spectrometry of peptides and proteins

Mark E. McComb; Richard D. Oleschuk; Darren M. Manley; Lynda J. Donald; Art Chow; Joe D. O'Neil; Werner Ens; Kenneth G. Standing; Hélène Perreault

Matrix-assisted laser desorption/ionization time-of-flight mass spectrometry (MALDI-TOFMS) of proteins and peptides was performed on samples deposited onto non-porous ether-type polyurethane (PU) membranes. Spectra obtained using PU membranes showed that mass resolution and accuracy were equivalent to values observed using a metal target, and superior to those obtained using poly(vinylidene difluoride) (PVDF) membranes. A small apparent increase in the mass of proteins and also loss of resolution were observed at very high laser irradiance due to charging, but were not observed under normal conditions. Analysis of NaCl-doped standards demonstrated that PU membranes yielded better results than a metallic target for salt-containing solutions. Relatively strong hydrophobic interactions between the proteins and peptides and the PU membrane allowed the incorporation of a washing step. This step allowed for the removal of salts and buffer components and thus provided an increase in resolution and mass accuracy. Digestion of citrate synthase (a protein of molecular weight 47,886) with trypsin was performed directly on the surface of the membrane for variable periods of time, and characteristic peptide fragments were observed by MALDI-TOFMS. Delayed extraction was used to increase the resolution and to permit more accurate mass assignments for those fragments. The use of PU membranes for MALDI-TOFMS analysis of proteins with higher molecular weights is also demonstrated.


Proteins | 2003

Structure of the Clade 1 catalase, CatF of Pseudomonas syringae, at 1.8 Å resolution

Xavi Carpena; Manuel Soriano; Martin G. Klotz; Harry W. Duckworth; Lynda J. Donald; William Melik-Adamyan; Ignacio Fita; Peter C. Loewen

Catalase CatF of Pseudomonas syringae has been identified phylogenetically as a clade 1 catalase, closely related to plant catalases, a group from which no structure has been determined. The structure of CatF has been refined at 1.8 Å resolution by using X‐ray synchrotron data collected from a crystal flash‐cooled with liquid nitrogen. The crystallographic agreement factors R and Rfree are, respectively, 18.3% and 24.0%. The asymmetric unit of the crystal contains a whole molecule that shows accurate 222‐point group symmetry. The crystallized enzyme is a homotetramer of subunits with 484 residues, some 26 residues shorter than predicted from the DNA sequence. Mass spectrometry analysis confirmed the absence of 26 N‐terminal residues, possibly removed by a periplasmic transport system. The core structure of the CatF subunit was closely related to seven other catalases with root‐mean‐square deviations (RMSDs) of 368 core Cα atoms of 0.99–1.30 Å. The heme component of CatF is heme b in the same orientation that is found in Escherichia coli hydroperoxidase II, an orientation that is flipped 180° with respect the orientation of the heme in bovine liver catalase. NADPH is not found in the structure of CatF because key residues required for nucleotide binding are missing; 2129 water molecules were refined into the model. Water occupancy in the main or perpendicular channel of CatF varied among the four subunits from two to five in the region between the heme and the conserved Asp150. A comparison of the water occupancy in this region with the same region in other catalases reveals significant differences among the catalases. Proteins 2003;50:423–436.


Journal of Biological Chemistry | 2015

The β-lactamase gene regulator AmpR is a tetramer that recognizes and binds the D-Ala-D-Ala motif of its repressor UDP-N-acetylmuramic acid (MurNAc)-pentapeptide.

Grishma Vadlamani; Misty D. Thomas; Trushar R. Patel; Lynda J. Donald; Thomas M. Reeve; Jörg Stetefeld; Kenneth G. Standing; David J. Vocadlo; Brian L. Mark

Background: Peptidoglycan metabolites regulate AmpR-mediated control of ampC β-lactamase expression. Results: AmpR binds DNA as a tetramer and interacts with up to four copies of its repressor UDP-MurNAc-pentapeptide via its d-Ala-d-Ala motif. Conclusion: d-Ala-d-Ala recognition supports 1,6-anhydroMurNAc-pentapeptide as an AmpR activator, probably through competitive binding with UDP-MurNAc-pentapeptide. Significance: Understanding how peptidoglycan metabolites modulate AmpR provides insight into how β-lactam classes differentially induce ampC expression. Inducible expression of chromosomal AmpC β-lactamase is a major cause of β-lactam antibiotic resistance in the Gram-negative bacteria Pseudomonas aeruginosa and Enterobacteriaceae. AmpC expression is induced by the LysR-type transcriptional regulator (LTTR) AmpR, which activates ampC expression in response to changes in peptidoglycan (PG) metabolite levels that occur during exposure to β-lactams. Under normal conditions, AmpR represses ampC transcription by binding the PG precursor UDP-N-acetylmuramic acid (MurNAc)-pentapeptide. When exposed to β-lactams, however, PG catabolites (1,6-anhydroMurNAc-peptides) accumulate in the cytosol, which have been proposed to competitively displace UDP-MurNAc-pentapeptide from AmpR and convert it into an activator of ampC transcription. Here we describe the molecular interactions between AmpR (from Citrobacter freundii), its DNA operator, and repressor UDP-MurNAc-pentapeptide. Non-denaturing mass spectrometry revealed AmpR to be a homotetramer that is stabilized by DNA containing the T-N11-A LTTR binding motif and revealed that it can bind four repressor molecules in an apparently stepwise manner. A crystal structure of the AmpR effector-binding domain bound to UDP-MurNAc-pentapeptide revealed that the terminal d-Ala-d-Ala motif of the repressor forms the primary contacts with the protein. This observation suggests that 1,6-anhydroMurNAc-pentapeptide may convert AmpR into an activator of ampC transcription more effectively than 1,6-anhydroMurNAc-tripeptide (which lacks the d-Ala-d-Ala motif). Finally, small angle x-ray scattering demonstrates that the AmpR·DNA complex adopts a flat conformation similar to the LTTR protein AphB and undergoes only a slight conformational change when binding UDP-MurNAc-pentapeptide. Modeling the AmpR·DNA tetramer bound to UDP-MurNAc-pentapeptide predicts that the UDP-MurNAc moiety of the repressor participates in modulating AmpR function.


Journal of Bacteriology | 2008

RhaU of Rhizobium leguminosarum Is a Rhamnose Mutarotase

Jason S. Richardson; Xavi Carpena; Jack Switala; Rosa Pérez-Luque; Lynda J. Donald; Peter C. Loewen; Ivan J. Oresnik

Of the nine genes comprising the L-rhamnose operon of Rhizobium leguminosarum, rhaU has not been assigned a function. The construction of a Delta rhaU strain revealed a growth phenotype that was slower than that of the wild-type strain, although the ultimate cell yields were equivalent. The transport of L-rhamnose into the cell and the rate of its phosphorylation were unaffected by the mutation. RhaU exhibits weak sequence similarity to the formerly hypothetical protein YiiL of Escherichia coli that has recently been characterized as an L-rhamnose mutarotase. To characterize RhaU further, a His-tagged variant of the protein was prepared and subjected to mass spectrometry analysis, confirming the subunit size and demonstrating its dimeric structure. After crystallization, the structure was refined to a 1.6-A resolution to reveal a dimer in the asymmetric unit with a very similar structure to that of YiiL. Soaking a RhaU crystal with L-rhamnose resulted in the appearance of beta-L-rhamnose in the active site.


Biochemistry | 2014

Identifying the elusive sites of tyrosyl radicals in cytochrome c peroxidase: implications for oxidation of substrates bound at a site remote from the heme.

Kyle D. Miner; Thomas D. Pfister; Parisa Hosseinzadeh; Nadime Karaduman; Lynda J. Donald; Peter C. Loewen; Yi Lu; Anabella Ivancich

The location of the Trp radical and the catalytic function of the [Fe(IV)=O Trp191•+] intermediate in cytochrome c peroxidase (CcP) are well-established; however, the unambiguous identification of the site(s) for the formation of tyrosyl radical(s) and their possible biological roles remain elusive. We have now performed a systematic investigation of the location and reactivity of the Tyr radical(s) using multifrequency Electron Paramagnetic Resonance (EPR) spectroscopy combined with multiple-site Trp/Tyr mutations in CcP. Two tyrosines, Tyr71 and Tyr236, were identified as those contributing primarily to the EPR spectrum of the tyrosyl radical, recorded at 9 and 285 GHz. The EPR characterization also showed that the heme distal-side Trp51 is involved in the intramolecular electron transfer between Tyr71 and the heme and that formation of Tyr71• and Tyr236• is independent of the [Fe(IV)=O Trp191•+] intermediate. Tyr71 is located in an optimal position to mediate the oxidation of substrates binding at a site, more than 20 Å from the heme, which has been reported recently in the crystal structures of CcP with bound guaicol and phenol [Murphy, E. J., et al. (2012) FEBS J. 279, 1632–1639]. The possibility of discriminating the radical intermediates by their EPR spectra allowed us to identify Tyr71• as the reactive species with the guaiacol substrate. Our assignment of the surface-exposed Tyr236 as the other radical site agrees well with previous studies based on MNP labeling and protein cross-linking [Tsaprailis, G., and English, A. M. (2003) JBIC, J. Biol. Inorg. Chem. 8, 248–255] and on its covalent modification upon reaction of W191G CcP with 2-aminotriazole [Musah, R. A., and Goodin, D. B. (1997) Biochemistry 36, 11665–11674]. Accordingly, while Tyr71 acts as a true reactive intermediate for the oxidation of certain small substrates that bind at a site remote from the heme, the surface-exposed Tyr236 would be more likely related to oxidative stress signaling, as previously proposed. Our findings reinforce the view that CcP is the monofunctional peroxidase that most closely resembles its ancestor enzymes, the catalase-peroxidases, in terms of the higher complexity of the peroxidase reaction [Colin, J., et al. (2009) J. Am. Chem. Soc. 131, 8557–8563]. The strategy used to identify the elusive Tyr radical sites in CcP may be applied to other heme enzymes containing a large number of Tyr and Trp residues and for which Tyr (or Trp) radicals have been proposed to be involved in their peroxidase or peroxidase-like reaction.

Collaboration


Dive into the Lynda J. Donald's collaboration.

Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Werner Ens

University of Manitoba

View shared research outputs
Top Co-Authors

Avatar

Ignacio Fita

Spanish National Research Council

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Xavi Carpena

Spanish National Research Council

View shared research outputs
Top Co-Authors

Avatar

Ayeda Ayed

University of Manitoba

View shared research outputs
Top Co-Authors

Avatar

Ben Wiseman

University of Manitoba

View shared research outputs
Researchain Logo
Decentralizing Knowledge