Hostname: page-component-76fb5796d-22dnz Total loading time: 0 Render date: 2024-04-28T00:13:04.025Z Has data issue: false hasContentIssue false

Purification and characterization of guanylate kinase, a nucleoside monophosphate kinase of Brugia malayi

Published online by Cambridge University Press:  20 May 2014

SMITA GUPTA
Affiliation:
Division of Biochemistry, CSIR-Central Drug Research Institute, Lucknow - 226031, Uttar Pradesh, India
SUNITA YADAV
Affiliation:
Division of Biochemistry, CSIR-Central Drug Research Institute, Lucknow - 226031, Uttar Pradesh, India
NIDHI SINGH
Affiliation:
Division of Molecular and Structural Biology, CSIR-Central Drug Research Institute, Lucknow - 226031, Uttar Pradesh, India
ANITA VERMA
Affiliation:
Division of Biochemistry, CSIR-Central Drug Research Institute, Lucknow - 226031, Uttar Pradesh, India
IMRAN SIDDIQI
Affiliation:
Division of Biochemistry, CSIR-Central Drug Research Institute, Lucknow - 226031, Uttar Pradesh, India
JITENDRA K. SAXENA*
Affiliation:
Division of Biochemistry, CSIR-Central Drug Research Institute, Lucknow - 226031, Uttar Pradesh, India
*
*Corresponding author: Division of Biochemistry Central Drug Research Institute BS10/1, Sector 10, Jankipuram Extension, Sitapur Road, Lucknow - 226031, Uttar Pradesh, India. E-mail: jkscdri@yahoo.com

Summary

Guanylate kinase, a nucleoside monophosphate kinase of Brugia malayi which is involved in reversible transfer of phosphate groups from ATP to GMP, was cloned, expressed and characterized. The native molecular mass of BmGK was found to be 45 kDa as determined by size exclusion chromatography and glutaraldehyde cross-linking which revealed that the protein is homodimer in nature. This is a unique characteristic among known eukaryotic GKs. GMP and ATP served as the most effective phosphate acceptor and donor, respectively. Recombinant BmGK utilized both GMP and dGMP, as substrates showing Km values of 30 and 38 μm, respectively. Free Mg+2 (un-complexed to ATP) and GTP play a regulatory role in catalysis of BmGK. The enzyme showed higher catalytic efficiency as compared with human GK and showed ternary complex (BmGK-GMP-ATP) formation with sequential substrate binding. The secondary structure of BmGK consisted of 45% α-helices, 18% β-sheets as revealed by CD analysis. Homology modelling and docking with GMP revealed conserved substrate binding residues with slight differences. Differences in kinetic properties and oligomerization of BmGK compared with human GK can provide the way for design of parasite-specific inhibitors.

Type
Research Article
Copyright
Copyright © Cambridge University Press 2014 

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Agarwal, K., Miech, R. and Parks, R. (1978). Guanylate kinases from human erythrocytes, hog brain, and rat liver. Methods in Enzymology 51, 483490.Google Scholar
Andrade, M., Chacon, P., Merelo, J. and Moran, F. (1993). Evaluation of secondary structure of proteins from UV circular dichroism spectra using an unsupervised learning neural network. Protein Engineering 6, 383390.Google Scholar
Auvynet, C., Topalis, D., Caillat, C., Munier Lehmann, H., Seclaman, E., Balzarini, J., Agrofoglio, L. A., Kaminski, P. A., Meyer, P. and Deville Bonne, D. (2009). Phosphorylation of dGMP analogs by vaccinia virus TMP kinase and human GMP kinase. Biochemical and Biophysical Research Communications 388, 611.Google Scholar
Berg, J. M., Tymoczko, J. L. and Stryer, L. (2002). Biochemistry, 5th Edn. W.H. Freeman, New York, NY, USA.Google Scholar
Bockarie, M. J. and Deb, R. M. (2010). Elimination of lymphatic filariasis: do we have the drugs to complete the job? Current Opinion in Infectious Diseases 23, 617620.Google Scholar
Bradford, M. M. (1976). A rapid and sensitive method for the quantification of microgram quantities of protein utilizing the principle of protein dye binding. Analytical Biochemistry 72, 248254.Google Scholar
Brady, W. A., Kokoris, M. S., Fitzgibbon, M. and Black, M. E. (1996). Cloning, characterization, and modeling of mouse and human guanylate kinases. Journal of Biological Chemistry 271, 1673416740.Google Scholar
Buccino, R. J. and Roth, J. S. (1969). Partial purification and properties of ATP: GMP phosphotransferase from rat liver. Archives of Biochemistry and Biophysics 132, 4961.CrossRefGoogle ScholarPubMed
Chung, W. C. and Kermode, J. C. (2005). Suramin disrupts receptor-G protein coupling by blocking association of G protein α and βγ subunits. Journal of Pharmacology and Experimental Therapeutics 313, 191198.Google Scholar
Crompton, D. W. T. (2010). First WHO Report on Neglected Tropical Diseases: Working to Overcome the Global Impact of Neglected Tropical Diseases. World Health Organization, Geneva, Switzerland.Google Scholar
Eftimie, A., Toma, F., Costache, A. Z. and Bucurenci, N. (2007). Characterization of guanylate kinase from gram positive and gram negative microorganisms; preliminary results. Roumanian Archives of Microbiology and Immunology 66, 22.Google Scholar
Fidock, D. A., Rosenthal, P. J., Croft, S. L., Brun, R. and Nwaka, S. (2004). Antimalarial drug discovery: efficacy models for compound screening. Nature Reviews Drug Discovery 3, 509520.Google Scholar
Gentry, D., Bengra, C., Ikehara, K. and Cashel, M. (1993). Guanylate kinase of Escherichia coli K-12. Journal of Biological Chemistry 268, 1431614321.CrossRefGoogle ScholarPubMed
Gokhale, R. S., Soumya, S., Balaram, H. and Balaram, P. (1999). Unfolding of Plasmodium falciparum triosephosphate isomerase in urea and guanidinium chloride: evidence for a novel disulfide exchange reaction in a covalently cross-linked mutant. Biochemistry 38, 423431.Google Scholar
Hible, G., Renault, L., Schaeffer, F., Christova, P., Zoe Radulescu, A., Evrin, C., Gilles, A. M. and Cherfils, J. (2005). Calorimetric and crystallographic analysis of the oligomeric structure of Escherichia coli GMP kinase. Journal of Molecular Biology 352, 10441059.Google Scholar
Hible, G., Christova, P., Renault, L., Seclaman, E., Thompson, A., Girard, E., Munier Lehmann, H. and Cherfils, J. (2006 a). Unique GMP binding site in Mycobacterium tuberculosis guanosine monophosphate kinase. Proteins: Structure, Function, and Bioinformatics 62, 489500.Google Scholar
Hible, G., Daalova, P., Gilles, A. M. and Cherfils, J. (2006 b). Crystal structures of GMP kinase in complex with ganciclovir monophosphate and Ap5G. Biochimie 88, 11571164.Google Scholar
Jain, A. N. (2003). Surflex: fully automatic flexible molecular docking using a molecular similarity-based search engine. Journal of Medicinal Chemistry 46, 499511.Google Scholar
Kandeel, M. and Kitade, Y. (2011). Binding dynamics and energetic insight into the molecular forces driving nucleotide binding by guanylate kinase. Journal of Molecular Recognition 24, 322332.Google Scholar
Kandeel, M., Nakanishi, M., Ando, T., Shazly, K., Yosef, T., Ueno, Y. and Kitade, Y. (2008). Molecular cloning, expression, characterization and mutation of Plasmodium falciparum guanylate kinase. Molecular and Biochemical Parasitology 159, 130133.Google Scholar
Konrad, M. (1992). Cloning and expression of the essential gene for guanylate kinase from yeast. Journal of Biological Chemistry 267, 2565225655.Google Scholar
Kumar, V., Spangenberg, O. and Konrad, M. (2001). Cloning of the guanylate kinase homologues AGK1 and AGK2 from Arabidopsis thaliana and characterization of AGK1. European Journal of Biochemistry 267, 606615.Google Scholar
Laemmli, U. K. (1970). Cleavage of structural proteins during the assembly of the head of bacteriophage T4. Nature 227, 680685.Google Scholar
Li, Y., Zhang, Y. and Yan, H. (1996). Kinetic and thermodynamic characterizations of yeast guanylate kinase. Journal of Biological Chemistry 271, 2803828044.Google Scholar
Maenpuen, S., Sopitthummakhun, K., Yuthavong, Y., Chaiyen, P. and Leartsakulpanich, U. (2009). Characterization of Plasmodium falciparum serine hydroxymethyltransferase – A potential antimalarial target. Molecular and Biochemical Parasitology 168, 6373.Google Scholar
Mikkelsen, N. E., Johansson, K., Karlsson, A., Knecht, W., Andersen, G., Piskur, J., Munch Petersen, B. and Eklund, H. (2003). Structural basis for feedback inhibition of the deoxyribonucleoside salvage pathway: studies of the Drosophila deoxyribonucleoside kinase. Biochemistry 42, 57065712.Google Scholar
Miller, W. and Miller, R. (1980). Phosphorylation of acyclovir (acycloguanosine) monophosphate by GMP kinase. Journal of Biological Chemistry 255, 72047207.Google Scholar
Oeschger, M. P. and Bessman, M. J. (1966). Purification and properties of guanylate kinase from Escherichia coli . Journal of Biological Chemistry 241, 54525460.Google Scholar
Okada, N. and Koizumi, S. (1995). A neuroprotective compound, aurin tricarboxylic acid, stimulates the tyrosine phosphorylation cascade in PC12 cells. Journal of Biological Chemistry 270, 1646416469.CrossRefGoogle ScholarPubMed
Sali, A. and Blundell, T. L. (1993). Comparative protein modelling by satisfaction of spatial restraints. .Journal of Molecular Biology 234, 779815.Google Scholar
Sambrook, J., Fritsch, E. F. and Maniatis, T. (1989). Molecular Cloning: A Laboratory Manual, 2nd Edn. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY, USA.Google Scholar
Sekulic, N., Shuvalova, L., Spangenberg, O., Konrad, M. and Lavie, A. (2002). Structural characterization of the closed conformation of mouse guanylate kinase. Journal of Biological Chemistry 277, 3023630243.Google Scholar
Shimono, H. and Sugino, Y. (1971). Metabolism of deoxyribonucleotides. European Journal of Biochemistry 19, 256263.Google Scholar
Singh, A. R., Joshi, S., Arya, R., Kayastha, A. M., Srivastava, K. K., Tripathi, L. M. and Saxena, J. K. (2008). Molecular cloning and characterization of Brugia malayi hexokinase. Parasitology International 57, 354361.Google Scholar
Stein, C., LaRocca, R., Thomas, R., McAtee, N. and Myers, C. E. (1989). Suramin: an anticancer drug with a unique mechanism of action. Journal of Clinical Oncology 7, 499508.Google Scholar
Tan, Y. W., Hanson, J. A. and Yang, H. (2009). Direct Mg+2 binding activates adenylate kinase from Escherichia coli . Journal of Biological Chemistry 284, 33063313.Google Scholar
Vonrhein, C., Schlauderer, G. J. and Schulz, G. E. (1995). Movie of the structural changes during a catalytic cycle of nucleoside monophosphate kinases. Structure 3, 483490.Google Scholar
Willmon, C., Krabbenhoft, E. and Black, M. (2006). A guanylate kinase/HSV-1 thymidine kinase fusion protein enhances prodrug-mediated cell killing. Gene Therapy 13, 13091312.Google Scholar
World Health Organization (2010). Progress Report 2000–2009 and Strategic Plan 2010–2020 of the Global Programme to Eliminate Lymphatic Filariasis: Halfway Towards Eliminating Lymphatic Filariasis. World Health Organization, Geneva, Switzerland.Google Scholar
Yan, H. and Tsai, M. D. (1999). Nucleoside monophosphate kinases: structure, mechanism, and substrate specificity. Advanced Enzymology Related Areas Molecular Biology 73, 103134.Google Scholar
Zhang, Y. L., Zhou, J. M. and Tsou, C. L. (1993). Inactivation precedes conformation change during thermal denaturation of adenylate kinase. Biochimica et Biophysica Acta (BBA) – Protein Structure and Molecular Enzymology 1164, 6167.Google Scholar
Supplementary material: Image

Gupta Supplementary Material

Figure S1

Download Gupta Supplementary Material(Image)
Image 1.4 MB
Supplementary material: Image

Gupta Supplementary Material

Figure S2

Download Gupta Supplementary Material(Image)
Image 2.1 MB
Supplementary material: Image

Gupta Supplementary Material

Figure S3

Download Gupta Supplementary Material(Image)
Image 3.7 MB