Network


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

Hotspot


Dive into the research topics where Lata Prasad is active.

Publication


Featured researches published by Lata Prasad.


Acta Crystallographica Section D-biological Crystallography | 2004

The structure of a universally employed enzyme: V8 protease from Staphylococcus aureus

Lata Prasad; Yvonne Leduc; Koto Hayakawa; Louis T. J. Delbaere

V8 protease, an extracellular protease of Staphylococcus aureus, is related to the pancreatic serine proteases. The enzyme cleaves peptide bonds exclusively on the carbonyl side of aspartate and glutamate residues. Unlike the pancreatic serine proteases, V8 protease possesses no disulfide bridges. This is a major evolutionary difference, as all pancreatic proteases have at least two disulfide bridges. The structure of V8 protease shows structural similarity with several other serine proteases, specifically the epidermolytic toxins A and B from S. aureus and trypsin, in which the conformation of the active site is almost identical. V8 protease is also unique in that the positively charged N-terminus is involved in determining the substrate-specificity of the enzyme.


Journal of Biological Chemistry | 2007

Activation of the SspA serine protease zymogen of Staphylococcus aureus proceeds through unique variations of a trypsinogen-like mechanism and is dependent on both autocatalytic and metalloprotease-specific processing.

Nicholas N. Nickerson; Lata Prasad; Latha Jacob; Louis T. J. Delbaere; Martin J. McGavin

The serine and cysteine proteases SspA and SspB of Staphylococcus aureus are secreted as inactive zymogens, zSspA and zSspB. Mature SspA is a trypsin-like glutamyl endopeptidase and is required to activate zSspB. Although a metalloprotease Aureolysin (Aur) is in turn thought to contribute to activation of zSspA, a specific role has not been demonstrated. We found that pre-zSspA is processed by signal peptidase at ANA29 ↓, releasing a Leu30 isoform that is first processed exclusively through autocatalytic intramolecular cleavage within a glutamine-rich propeptide segment, 40QQTQSSKQQTPKIQ53. The preferred site is Gln43 with secondary processing at Gln47 and Gln53. This initial processing is necessary for optimal and subsequent Aur-dependent processing at Leu58 and then Val69 to release mature SspA. Although processing by Aur is rate-limiting in zSspA activation, the first active molecules of Val69SspA promote rapid intermolecular processing of remaining zSspA at Glu65, producing an N-terminal 66HANVILP isoform that is inactive until removal of the HAN tripeptide by Aur. Modeling indicated that His66 of this penultimate isoform blocks the active site by hydrogen bonding to Ser237 and occlusion of substrate. Binding of glutamate within the active site of zSspA is energetically unfavorable, but glutamine fits into the primary specificity pocket and is predicted to hydrogen bond to Thr232 proximal to Ser237, permitting autocatalytic cleavage of the glutamine-rich propeptide segment. These and other observations suggest that zSspA is activated through a trypsinogen-like mechanism where supplementary features of the propeptide must be sequentially processed in the correct order to allow efficient activation.


Proteins | 2006

Substrate-free structure of a monomeric NADP isocitrate dehydrogenase: An open conformation phylogenetic relationship of isocitrate dehydrogenase

Fumie Imabayashi; Sanjukta Aich; Lata Prasad; Louis T. J. Delbaere

Both monomeric and dimeric NADP+‐dependent isocitrate dehydrogenase (IDH) belong to the metal‐dependent β‐decarboxylating dehydrogenase family and catalyze the oxidative decarboxylation from 2R,3S‐isocitrate to yield 2‐oxoglutarate, CO2, and NADPH. It is important to solve the structures of IDHs from various species to correlate with its function and evolutionary significance. So far, only two crystal structures of substrate/cofactor‐bound (isocitrate/NADP) NADP+‐dependent monomeric IDH from Azotobacter vinelandii (AvIDH) have been solved. Herein, we report for the first time the substrate/cofactor‐free structure of a monomeric NADP+‐dependent IDH from Corynebacterium glutamicum (CgIDH) in the presence of Mg2+. The 1.75 Å structure of CgIDH‐Mg2+ showed a distinct open conformation in contrast to the closed conformation of AvIDH‐isocitrate/NADP+ complexes. Fluorescence studies on CgIDH in the presence of isocitrate/or NADP+ suggest the presence of low energy barrier conformers. In CgIDH, the amino acid residues corresponding to the Escherichia coli IDH phosphorylation‐loop are α‐helical compared with the more flexible random‐coil region in the E. coli protein where IDH activation is controlled by phosphorylation. This more structured region supports the idea that activation of CgIDH is not controlled by phosphorylation. Monomeric NADP+‐specific IDHs have been identified from about 50 different bacterial species, such as proteobacteria, actinobacteria, and planctomycetes, whereas, dimeric NADP+‐dependent IDHs are diversified in both prokaryotes and eukaryotes. We have constructed a phylogenetic tree based on amino acid sequences of all bacterial monomeric NADP+‐dependent IDHs and also another one with specifically chosen species which either contains both monomeric and dimeric NADP+‐dependent IDHs or have monomeric NADP+‐dependent, as well as NAD+‐dependent IDHs. This is done to examine evolutionary relationships. Proteins 2006.


Journal of Bacteriology | 2003

Mechanisms of Activation of Phosphoenolpyruvate Carboxykinase from Escherichia coli by Ca2+ and of Desensitization by Trypsin

Athena M. Sudom; Robert Walters; Landon Pastushok; Douglas Goldie; Lata Prasad; Louis T. J. Delbaere; Hughes Goldie

The 1.8-A resolution structure of the ATP-Mg(2+)-Ca(2+)-pyruvate quinary complex of Escherichia coli phosphoenolpyruvate carboxykinase (PCK) is isomorphous to the published complex ATP-Mg(2+)-Mn(2+)-pyruvate-PCK, except for the Ca(2+) and Mn(2+) binding sites. Ca(2+) was formerly implicated as a possible allosteric regulator of PCK, binding at the active site and at a surface activating site (Glu508 and Glu511). This report found that Ca(2+) bound only at the active site, indicating that there is likely no surface allosteric site. (45)Ca(2+) bound to PCK with a K(d) of 85 micro M and n of 0.92. Glu508Gln Glu511Gln mutant PCK had normal activation by Ca(2+). Separate roles of Mg(2+), which binds the nucleotide, and Ca(2+), which bridges the nucleotide and the anionic substrate, are implied, and the catalytic mechanism of PCK is better explained by studies of the Ca(2+)-bound structure. Partial trypsin digestion abolishes Ca(2+) activation (desensitizes PCK). N-terminal sequencing identified sensitive sites, i.e., Arg2 and Arg396. Arg2Ser, Arg396Ser, and Arg2Ser Arg396Ser (double mutant) PCKs altered the kinetics of desensitization. C-terminal residues 397 to 540 were removed by trypsin when wild-type PCK was completely desensitized. Phe409 and Phe413 interact with residues in the Ca(2+) binding site, probably stabilizing the C terminus. Phe409Ala, DeltaPhe409, Phe413Ala, Delta397-521 (deletion of residues 397 to 521), Arg396(TAA) (stop codon), and Asp269Glu (Ca(2+) site) mutations failed to desensitize PCK and, with the exception of Phe409Ala, appeared to have defects in the synthesis or assembly of PCK, suggesting that the structure of the C-terminal domain is important in these processes.


Molecular Immunology | 1998

Conversion of an antibody into an enzyme which cleaves the protein HPr

E. Liu; Lata Prasad; Louis T. J. Delbaere; E.b. Waygood; Jeremy S. Lee

Jel 42 is an IgG which binds to the small bacterial protein, HPr and the structure of the complex is known at high resolution. The IgG was expressed as a single chain variable fragment (scFv) and the binding to HPr was assessed by fluorescence polarization of fluorescein-labelled HPr. The binding constant for the IgG was about 20-fold higher than the scFv. Inspection of the structure of the complex suggested that it might be possible to convert the scFv into a bond-specific protease by the introduction of three catalytic residues: a glutamate to increase the nucleophilicity of a nearby water molecule, a lysine to increase the polarizability of the carbonyl group and a histidine to provide a proton to convert the amine into a better leaving group. By trial and error it was found that a fourth residue had to be converted into glycine in order to maintain the integrity of complimentarity-determining region three of the heavy chain (CDRH3) at the binding interface. The resulting quadruple mutant still bound to HPr and unlike other mutants, showed weak protease activity as judged from the fluorescence polarization assay. The activity was maximum at pH 6 consistent with a requirement for a protonated histidine residue. With the aid of HPr fluorescein-labelled at two different positions, it was demonstrated that the size of the products was consistent with cleavage occurring in the vicinity of the target peptide bond. The activity was specific for HPr since an excess of bovine serum albumin did not interfere with the reaction.


The International Journal of Biochemistry & Cell Biology | 2008

Structure of a GTP-dependent Bacterial PEP-carboxykinase from Corynebacterium glutamicum

Sanjukta Aich; Lata Prasad; Louis T. J. Delbaere

GTP-dependent phosphoenolpyruvate carboxykinase (PCK) is the key enzyme that controls the blood glucose level during fasting in higher animals. Here we report the first substrate-free structure of a GTP-dependent phosphoenolpyruvate (PEP) carboxykinase from a bacterium, Corynebacterium glutamicum (CgPCK). The protein crystallizes in space group P2(1) with four molecules per asymmetric unit. The 2.3A resolution structure was solved by molecular replacement using the human cytosolic PCK (hcPCK) structure (PDB ID: 1KHF) as the starting model. The four molecules in the asymmetric unit pack as two dimers, and is an artifact of crystal packing. However, the P-loop and the guanine binding loop of the substrate-free CgPCK structure have different conformations from the other published GTP-specific PCK structures, which all have bound substrates and/or metal ions. It appears that a change in the P-loop and guanine binding loop conformation is necessary for substrate binding in GTP-specific PCKs, as opposed to overall domain movement in ATP-specific PCKs.


Acta Crystallographica Section F-structural Biology and Crystallization Communications | 2010

Crystallization and preliminary X-ray diffraction analysis of the Thermoanaerobacter ethanolicus secondary alcohol dehydrogenase I86A mutant

Carla Protsko; Claire Vieille; Maris Laivenieks; Lata Prasad; David A. R. Sanders; Louis T. J. Delbaere

The Thermoanaerobacter ethanolicus secondary alcohol dehydrogenase I86A mutant is stereospecific for (R)-alcohols instead of (S)-alcohols. Pyramidal crystals grown in the presence of (R)-phenylethanol via the hanging-drop vapour-diffusion method diffracted to 3.2 A resolution at the Canadian Light Source. The crystal belonged to the orthorhombic space group P2(1)2(1)2(1), with unit-cell parameters a = 80.23, b = 124.90, c = 164.80 A. The structure was solved by molecular replacement using the structure of T. brockii SADH (PDB entry 1ykf).


Acta Crystallographica Section F-structural Biology and Crystallization Communications | 2009

Crystallization and preliminary X-ray studies of the N-domain of the Wilson disease associated protein.

Lili Liu; Christopher O'Grady; Sean A. Dalrymple; Lata Prasad; Oleg Y. Dmitriev; Louis T. J. Delbaere

Wilson disease associated protein (ATP7B) is essential for copper transport in human cells. Mutations that affect ATP7B function result in Wilsons disease, a chronic copper toxicosis. Disease-causing mutations within the N-domain of ATP7B (WND) are known to disrupt ATP binding, but a high-resolution X-ray structure of the ATP-binding site has not been reported. The N-domain was modified to delete the disordered loop comprising residues His1115-Asp1138 (WNDDelta(1115-1138)). Unlike the wild-type N-domain, WNDDelta(1115-1138) formed good-quality crystals. Synchrotron diffraction data have been collected from WNDDelta(1115-1138) at the Canadian Light Source. A native WNDDelta(1115-1138) crystal diffracted to 1.7 A resolution and belonged to space group P4(2)2(1)2, with unit-cell parameters a = 39.2, b = 39.2, c = 168.9 A. MAD data were collected to 2.7 A resolution from a SeMet-derivative crystal with unit-cell parameters a = 38.4, b = 38.4, c = 166.7 A. The WNDDelta(1115-1138) structure is likely to be solved by phasing from multiwavelength anomalous diffraction (MAD) experiments.


Biochemistry | 2008

Structural Investigation of a Phosphorylation-Catalyzed, Isoaspartate-Free, Protein Succinimide: Crystallographic Structure of Post-Succinimide His15Asp Histidine-Containing Protein

Scott Napper; Lata Prasad; Louis T. J. Delbaere

Aspartates and asparagines can spontaneously cyclize with neighboring main-chain amides to form succinimides. These succinimides hydrolyze to a mixture of isoaspartate and aspartate products. Phosphorylation of aspartates is a common mechanism of protein regulation and increases the propensity for succinimide formation. Although typically regarded as a form of protein damage, we hypothesize succinimides could represent an effective mechanism of phosphoaspartate autophosphatase activity, provided hydrolysis is limited to aspartate products. We previously reported the serendipitous creation of a protein, His15Asp histidine-containing protein (HPr), which undergoes phosphorylation-catalyzed formation of a succinimide whose hydrolysis is seemingly exclusive for aspartate formation. Here, through the high-resolution structure of postsuccinimide His15Asp HPr, we confirm the absence of isoaspartate residues and propose mechanisms for phosphorylation-catalyzed succinimide formation and its directed hydrolysis to aspartate. His15Asp HPr represents the first characterized protein example of an isoaspartate-free succinimide and lends credence to the hypothesis that intramolecular cyclization could represent a physiological mechanism of autophosphatase activity. Furthermore, this indicates that current strategies for succinimide evaluation, based on isoaspartate detection, underestimate the frequencies of these reactions. This is considerably significant for evaluation of protein stability and integrity.


Advances in Space Research | 1992

Protein crystal growth aboard the U.S. space shuttle flights STS-31 and STS-32

Lawrence J. DeLucas; Craig D. Smith; Daniel C. Carter; Pam Twigg; Xiao-min He; Robert S. Snyder; Patricia C. Weber; J.V. Schloss; Howard M. Einspahr; L.L. Clancy; Alexander McPherson; Stanley Koszelak; Margaret Vandonselaar; Lata Prasad; J. Wilson Quail; Louis T. J. Delbaere; Charles E. Bugg

The first microgravity protein crystal growth experiments were performed on Spacelab I by Littke and John. These experiments indicated that the space grown crystals, which were obtained using a liquid-liquid diffusion system, were larger than crystals obtained by the same experimental system on earth. Subsequent experiments were performed by other investigators on a series of space shuttle missions from 1985 through 1990. The results from two of these shuttle flights (STS-26 and STS-29) have been described previously. The results from these missions indicated that the microgravity grown crystals for a number of different proteins were larger, displayed more uniform morphologies, and yielded diffraction data to significantly higher resolutions than the best crystals of these proteins grown on earth. This paper presents the results obtained from shuttle flight STS-32 (flown in January, 1990) and preliminary results from the most recent shuttle flight, STS-31 (flown in April, 1990).

Collaboration


Dive into the Lata Prasad's collaboration.

Top Co-Authors

Avatar
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Athena M. Sudom

University of Saskatchewan

View shared research outputs
Top Co-Authors

Avatar

Hughes Goldie

University of Saskatchewan

View shared research outputs
Top Co-Authors

Avatar

J. Wilson Quail

University of Saskatchewan

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Jeremy S. Lee

University of Saskatchewan

View shared research outputs
Top Co-Authors

Avatar

Jim Xiang

University of Saskatchewan

View shared research outputs
Top Co-Authors

Avatar

Yvonne Leduc

University of Saskatchewan

View shared research outputs
Top Co-Authors

Avatar

Charles E. Bugg

University of Alabama at Birmingham

View shared research outputs
Researchain Logo
Decentralizing Knowledge