Hostname: page-component-8448b6f56d-c47g7 Total loading time: 0 Render date: 2024-04-20T02:25:27.963Z Has data issue: false hasContentIssue false

Type I and type II fatty acid biosynthesis in Eimeria tenella: enoyl reductase activity and structure

Published online by Cambridge University Press:  13 August 2007

J. Z. LU
Affiliation:
Department of Molecular Microbiology & Immunology, Johns Hopkins Bloomberg School of Public Health, Baltimore, MD 21205, USA
S. P. MUENCH
Affiliation:
Department of Molecular Biology and Biotechnology, University of Sheffield, Sheffield S10 2TN, UK
M. ALLARY
Affiliation:
Department of Molecular Microbiology & Immunology, Johns Hopkins Bloomberg School of Public Health, Baltimore, MD 21205, USA
S. CAMPBELL
Affiliation:
Strathclyde Institute of Biomedical Sciences, University of Strathclyde, Glasgow G4 0NR, UK
C. W. ROBERTS
Affiliation:
Strathclyde Institute of Biomedical Sciences, University of Strathclyde, Glasgow G4 0NR, UK
E. MUI
Affiliation:
Department of Ophthalmology and Visual Sciences, University of Chicago, Chicago, IL 60637, USA
R. L. McLEOD
Affiliation:
Department of Ophthalmology and Visual Sciences, University of Chicago, Chicago, IL 60637, USA Department of Pediatrics (Infectious Diseases), and Pathology and Committees on Genetics, Molecular Medicine and Immunology and the College, University of Chicago, Chicago, IL 60637, USA
D. W. RICE
Affiliation:
Department of Molecular Biology and Biotechnology, University of Sheffield, Sheffield S10 2TN, UK
S. T. PRIGGE*
Affiliation:
Department of Molecular Microbiology & Immunology, Johns Hopkins Bloomberg School of Public Health, Baltimore, MD 21205, USA
*
*Corresponding author: Department of Molecular Microbiology & Immunology, Johns Hopkins Bloomberg School of Public Health, Baltimore, MD 21205, USA. Tel: +1 443 287 4822. Fax: +1 410 955 0105. E-mail address: sprigge@jhsph.edu

Summary

Apicomplexan parasites of the genus Eimeria are the major causative agent of avian coccidiosis, leading to high economic losses in the poultry industry. Recent results show that Eimeria tenella harbours an apicoplast organelle, and that a key biosynthetic enzyme, enoyl reductase, is located in this organelle. In related parasites, enoyl reductase is one component of a type II fatty acid synthase (FAS) and has proven to be an attractive target for antimicrobial compounds. We cloned and expressed the mature form of E. tenella enoyl reductase (EtENR) for biochemical and structural studies. Recombinant EtENR exhibits NADH-dependent enoyl reductase activity and is inhibited by triclosan with an IC50 value of 60 nm. The crystal structure of EtENR reveals overall similarity with other ENR enzymes; however, the active site of EtENR is unoccupied, a state rarely observed in other ENR structures. Furthermore, the position of the central beta-sheet appears to block NADH binding and would require significant movement to allow NADH binding, a feature not previously seen in the ENR family. We analysed the E. tenella genomic database for orthologues of well-characterized bacterial and apicomplexan FAS enzymes and identified 6 additional genes, suggesting that E. tenella contains a type II FAS capable of synthesizing saturated, but not unsaturated, fatty acids. Interestingly, we also identified sequences that appear to encode multifunctional type I FAS enzymes, a feature also observed in Toxoplasma gondii, highlighting the similarity between these apicomplexan parasites.

Type
Research Article
Copyright
Copyright © Cambridge University Press 2007

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

REFERENCES

Allary, M., Lu, J. Z., Zhu, L. and Prigge, S. T. (2007). Scavenging of the cofactor lipoate is essential for the survival of the malaria parasite Plasmodium falciparum. Molecular Microbiology 63, 13311344.CrossRefGoogle ScholarPubMed
Altschul, S. F., Gish, W., Miller, W., Myers, E. W. and Lipman, D. J. (1990). Basic local alignment search tool. Journal of Molecular Biology 215, 403410.Google Scholar
Barta, J. R., Martin, D. S., Carreno, R. A., Siddall, M. E., Profous-Juchelkat, H., Hozza, M., Powles, M. A. and Sundermann, C. (2001). Molecular phylogeny of the other tissue coccidia: Lankesterella and Caryospora. Journal of Parasitology 87, 121127.Google Scholar
Bergler, H., Fuchsbichler, S., Hogenauer, G. and Turnowsky, F. (1996). The enoyl-[acyl-carrier-protein] reductase (FabI) of Escherichia coli, which catalyzes a key regulatory step in fatty acid biosynthesis, accepts NADH and NADPH as cofactors and is inhibited by palmitoyl-CoA. European Journal of Biochemistry 242, 689694.Google Scholar
Bergler, H., Wallner, P., Ebeling, A., Leitinger, B., Fuchsbichler, S., Aschauer, H., Kollenz, G., Hogenauer, G. and Turnowsky, F. (1994). Protein EnvM is the NADH-dependent enoyl-ACP reductase (FabI) of Escherichia coli. Journal of Biological Chemistry 269, 54935496.Google Scholar
Brunger, A. T., Adams, P. D., Clore, G. M., DeLano, W. L., Gros, P., Grosse-Kunstleve, R. W., Jiang, J. S., Kuszewski, J., Nilges, M., Pannu, N. S., Read, R. J., Rice, L. M., Simonson, T. and Warren, G. L. (1998). Crystallography & NMR System: A new software suite for macromolecular structure determination. Acta Crystallographica D Biological Crystallography 54, 905921.Google Scholar
Campbell, J. W. and Cronan, J. E. Jr. (2001). Bacterial fatty acid biosynthesis: targets for antibacterial drug discovery. Annual Review Microbiology 55, 305332.CrossRefGoogle ScholarPubMed
CCP4 (1994). The CCP4 suite: programs for protein crystallography. Acta Crystallographica D Biological Crystallography 50, 760763.Google Scholar
Chang, S. I. and Hammes, G. G. (1989). Homology analysis of the protein sequences of fatty acid synthases from chicken liver, rat mammary gland, and yeast. Proceedings of the National Academy of Sciences, USA 86, 83738376.Google Scholar
Coppens, I. (2006). Contribution of host lipids to Toxoplasma pathogenesis. Cell Microbiology 8, 19.Google Scholar
Crawford, M. J., Thomsen-Zieger, N., Ray, M., Schachtner, J., Roos, D. S. and Seeber, F. (2006). Toxoplasma gondii scavenges host-derived lipoic acid despite its de novo synthesis in the apicoplast. EMBO Journal 25, 32143222.Google Scholar
Fast, N. M., Kissinger, J. C., Roos, D. S. and Keeling, P. J. (2001). Nuclear-encoded, plastid-targeted genes suggest a single common origin for apicomplexan and dinoflagellate plastids. Molecular Biology and Evolution 18, 418426.CrossRefGoogle ScholarPubMed
Ferguson, D. J., Birch-Andersen, A., Hutchison, W. M. and Siim, J. C. (1976). Ultrastructural studies on the endogenous development of Eimeria brunetti. Acta Pathologica et Microbiologica Scandinavica. Section B, Microbiology 84B, 401413.Google Scholar
Ferguson, D. J., Campbell, S. A., Henriquez, F. L., Phan, L., Mui, E., Richards, T. A., Muench, S. P., Allary, M., Lu, J. Z., Prigge, S. T., Tomley, F., Shirley, M. W., Rice, D. W., McLeod, R. and Roberts, C. W. (2007). Enzymes of type II fatty acid synthesis and apicoplast differentiation and division in Eimeria tenella. International Journal for Parasitology 37, 3351.Google Scholar
Ferguson, D. J., Henriquez, F. L., Kirisits, M. J., Muench, S. P., Prigge, S. T., Rice, D. W., Roberts, C. W. and McLeod, R. L. (2005). Maternal inheritance and stage-specific variation of the apicoplast in Toxoplasma gondii during development in the intermediate and definitive host. Eukaryotic Cell 4, 814826.Google Scholar
Fleige, T., Fischer, K., Ferguson, D. J., Gross, U. and Bohne, W. (2007). Carbohydrate metabolism in the Toxoplasma gondii apicoplast: localization of three glycolytic isoenzymes, the single pyruvate dehydrogenase complex and a plastid phosphate translocator. Eukaryotic Cell [Epub].Google Scholar
Foth, B. J., Stimmler, L. M., Handman, E., Crabb, B. S., Hodder, A. N. and McFadden, G. I. (2005). The malaria parasite Plasmodium falciparum has only one pyruvate dehydrogenase complex, which is located in the apicoplast. Molecular Microbiology 55, 3953.Google Scholar
He, C. Y., Striepen, B., Pletcher, C. H., Murray, J. M. and Roos, D. S. (2001). Targeting and processing of nuclear-encoded apicoplast proteins in plastid segregation mutants of Toxoplasma gondii. Journal of Biological Chemistry 276, 2843628442.CrossRefGoogle ScholarPubMed
Heath, R. J., Rubin, J. R., Holland, D. R., Zhang, E., Snow, M. E. and Rock, C. O. (1999). Mechanism of triclosan inhibition of bacterial fatty acid synthesis. Journal of Biological Chemistry 274, 1111011114.CrossRefGoogle ScholarPubMed
Hee Lee, S., Stephens, J. L. and Englund, P. T. (2007). A fatty-acid synthesis mechanism specialized for parasitism. Nature Reviews. Microbiology 5, 287297.CrossRefGoogle Scholar
Holm, L. and Sander, C. (1995). Dali: a network tool for protein structure comparison. Trends in Biochemical Sciences 20, 478480.Google Scholar
Huang, W., Jia, J., Edwards, P., Dehesh, K., Schneider, G. and Lindqvist, Y. (1998). Crystal structure of beta-ketoacyl-acyl carrier protein synthase II from E. coli reveals the molecular architecture of condensing enzymes. EMBO Journal 17, 11831191.CrossRefGoogle Scholar
Jelenska, J., Crawford, M. J., Harb, O. S., Zuther, E., Haselkorn, R., Roos, D. S. and Gornicki, P. (2001). Subcellular localization of acetyl-CoA carboxylase in the apicomplexan parasite Toxoplasma gondii. Proceedings of the National Academy of Sciences, USA 98, 27232728.Google Scholar
Kapoor, M., Reddy, C. C., Krishnasastry, M. V., Surolia, N. and Surolia, A. (2004). Slow-tight-binding inhibition of enoyl-acyl carrier protein reductase from Plasmodium falciparum by triclosan. The Biochemical Journal 381, 719724.CrossRefGoogle ScholarPubMed
Kapust, R. B. and Waugh, D. S. (2000). Controlled intracellular processing of fusion proteins by TEV protease. Protein Expression and Purification 19, 312318.CrossRefGoogle ScholarPubMed
Levy, C. W., Baldock, C., Wallace, A. J., Sedelnikova, S., Viner, R. C., Clough, J. M., Stuitje, A. R., Slabas, A. R., Rice, D. W. and Rafferty, J. B. (2001). A study of the structure-activity relationship for diazaborine inhibition of Escherichia coli enoyl-ACP reductase. Journal of Molecular Biology 309, 171180.Google Scholar
Levy, C. W., Roujeinikova, A., Sedelnikova, S., Baker, P. J., Stuitje, A. R., Slabas, A. R., Rice, D. W. and Rafferty, J. B. (1999). Molecular basis of triclosan activity. Nature, London 398, 383384.CrossRefGoogle ScholarPubMed
Los, D. A. and Murata, N. (1998). Structure and expression of fatty acid desaturases. Biochimica et Biophysica Acta 1394, 315.Google Scholar
Lu, J. Z., Lee, P. J., Waters, N. C. and Prigge, S. T. (2005). Fatty Acid synthesis as a target for antimalarial drug discovery. Combinatorial Chemistry and High Throughput Screen 8, 1526.Google Scholar
Magnuson, K., Jackowski, S., Rock, C. O. and Cronan, J. E. Jr. (1993). Regulation of fatty acid biosynthesis in Escherichia coli. Microbiological Reviews 57, 522542.Google Scholar
Mazumdar, J., Wilson, E. H., Masek, K., Hunter, C. A. and Striepen, B. (2006). Apicoplast fatty acid synthesis is essential for organelle biogenesis and parasite survival in Toxoplasma gondii. Proceedings of the National Academy of Sciences, USA 103, 1319213197.Google Scholar
McLeod, R., Muench, S. P., Rafferty, J. B., Kyle, D. E., Mui, E. J., Kirisits, M. J., Mack, D. G., Roberts, C. W., Samuel, B. U., Lyons, R. E., Dorris, M., Milhous, W. K. and Rice, D. W. (2001). Triclosan inhibits the growth of Plasmodium falciparum and Toxoplasma gondii by inhibition of apicomplexan Fab I. International Journal for Parasitology 31, 109113.CrossRefGoogle ScholarPubMed
McRee, D. E. (1999). XtalView/Xfit–A versatile program for manipulating atomic coordinates and electron density. Journal of Structural Biology 125, 156165.CrossRefGoogle ScholarPubMed
Muench, S. P., Prigge, S. T., McLeod, R., Rafferty, J. B., Kirisits, M. J., Roberts, C. W., Mui, E. J. and Rice, D. W. (2007). Studies of Toxoplasma gondii and Plasmodium falciparum enoyl acyl carrier protein reductase and implications for the development of antiparasitic agents. Acta Crystallographica D Biological Crystallography 63, 328338.Google Scholar
Muench, S. P., Prigge, S. T., Zhu, L., Kirisits, M. J., Roberts, C. W., Wernimont, S., McLeod, R. and Rice, D. W. (2006). Expression, purification and preliminary crystallographic analysis of the Toxoplasma gondii enoyl reductase. Acta Crystallographica. Section F, Structural Biology and Crystallization Communications, 62, 604606.Google Scholar
Muench, S. P., Rafferty, J. B., McLeod, R., Rice, D. W. and Prigge, S. T. (2003). Expression, purification and crystallization of the Plasmodium falciparum enoyl reductase. Acta Crystallographica D Biological Crystallography 59, 12461248.Google Scholar
Otwinowski, Z. and Minor, W. (1997). Processing of X-ray diffraction data collected in oscillation mode. Methods in Enzymology. Macromolecular Crystallography Part A, 307326.Google Scholar
Perozzo, R., Kuo, M., Sidhu, A. S., Valiyaveettil, J. T., Bittman, R., Jacobs, W. R. Jr., Fidock, D. A. and Sacchettini, J. C. (2002). Structural elucidation of the specificity of the antibacterial agent triclosan for malarial enoyl acyl carrier protein reductase. Journal of Biological Chemistry 277, 1310613114.Google Scholar
Pettersen, E. F., Goddard, T. D., Huang, C. C., Couch, G. S., Greenblatt, D. M., Meng, E. C. and Ferrin, T. E. (2004). UCSF Chimera – a visualization system for exploratory research and analysis. Journal of Computational Chemistry 25, 16051612.Google Scholar
Price, A. C., Choi, K. H., Heath, R. J., Li, Z., White, S. W. and Rock, C. O. (2001). Inhibition of beta-ketoacyl-acyl carrier protein synthases by thiolactomycin and cerulenin. Structure and mechanism. Journal of Biological Chemistry 276, 65516559.Google Scholar
Qiu, X., Janson, C. A., Court, R. I., Smyth, M. G., Payne, D. J. and Abdel-Meguid, S. S. (1999 a). Molecular basis for triclosan activity involves a flipping loop in the active site. Protein Science 8, 25292532.CrossRefGoogle ScholarPubMed
Qiu, X., Janson, C. A., Konstantinidis, A. K., Nwagwu, S., Silverman, C., Smith, W. W., Khandekar, S., Lonsdale, J. and Abdel-Meguid, S. S. (1999 b) Crystal structure of beta-ketoacyl-acyl carrier protein synthase III. A key condensing enzyme in bacterial fatty acid biosynthesis. Journal of Biological Chemistry 274, 3646536471.CrossRefGoogle ScholarPubMed
Rafferty, J. B., Simon, J. W., Baldock, C., Artymiuk, P. J., Baker, P. J., Stuitje, A. R., Slabas, A. R. and Rice, D. W. (1995). Common themes in redox chemistry emerge from the X-ray structure of oilseed rape (Brassica napus) enoyl acyl carrier protein reductase. Structure 3, 927938.Google Scholar
Ralph, S. A., Van Dooren, G. G., Waller, R. F., Crawford, M. J., Fraunholz, M. J., Foth, B. J., Tonkin, C. J., Roos, D. S. and McFadden, G. I. (2004). Tropical infectious diseases: Metabolic maps and functions of the Plasmodium falciparum apicoplast. Nature Review. Microbiology 2, 203216.Google Scholar
Roberts, C. W., McLeod, R., Rice, D. W., Ginger, M., Chance, M. L. and Goad, L. J. (2003). Fatty acid and sterol metabolism: potential antimicrobial targets in apicomplexan and trypanosomatid parasitic protozoa. Molecular and Biochemical Parasitology 126, 129142.Google Scholar
Roujeinikova, A., Sedelnikova, S., de Boer, G. J., Stuitje, A. R., Slabas, A. R., Rafferty, J. B. and Rice, D. W. (1999). Inhibitor binding studies on enoyl reductase reveal conformational changes related to substrate recognition. Journal of Biological Chemistry 274, 3081130817.Google Scholar
Samuel, B. U., Hearn, B., Mack, D., Wender, P., Rothbard, J., Kirisits, M. J., Mui, E., Wernimont, S., Roberts, C. W., Muench, S. P., Rice, D. W., Prigge, S. T., Law, A. B. and McLeod, R. (2003). Delivery of antimicrobials into parasites. Proceedings of the National Academy of Sciences, USA 100, 1428114286.Google Scholar
Seefeld, M. A., Miller, W. H., Newlander, K. A., Burgess, W. J., DeWolf, W. E. Jr., Elkins, P. A., Head, M. S., Jakas, D. R., Janson, C. A., Keller, P. M., Manley, P. J., Moore, T. D., Payne, D. J., Pearson, S., Polizzi, B. J., Qiu, X., Rittenhouse, S. F., Uzinskas, I. N., Wallis, N. G. and Huffman, W. F. (2003). Indole naphthyridinones as inhibitors of bacterial enoyl-ACP reductases FabI and FabK. Journal of Medicinal Chemistry 46, 16271635.Google Scholar
Shirley, M. W., Smith, A. L. and Tomley, F. M. (2005). The biology of avian Eimeria with an emphasis on their control by vaccination. Advances in Parasitology 60, 285330.CrossRefGoogle ScholarPubMed
Sivaraman, S., Zwahlen, J., Bell, A. F., Hedstrom, L. and Tonge, P. J. (2003). Structure-activity studies of the inhibition of FabI, the enoyl reductase from Escherichia coli, by triclosan: kinetic analysis of mutant FabIs. Biochemistry 42, 44064413.CrossRefGoogle ScholarPubMed
Smith, S. (1994). The animal fatty acid synthase: one gene, one polypeptide, seven enzymes. FASEB Journal 8, 12481259.CrossRefGoogle ScholarPubMed
Surolia, N. and Surolia, A. (2001). Triclosan offers protection against blood stages of malaria by inhibiting enoyl-ACP reductase of Plasmodium falciparum. Nature, Medicine 7, 167173.CrossRefGoogle ScholarPubMed
Waller, R. F., Keeling, P. J., Donald, R. G., Striepen, B., Handman, E., Lang-Unnasch, N., Cowman, A. F., Besra, G. S., Roos, D. S. and McFadden, G. I. (1998). Nuclear-encoded proteins target to the plastid in Toxoplasma gondii and Plasmodium falciparum. Proceedings of the National Academy of Sciences, USA 95, 1235212357.Google Scholar
Waller, R. F., Reed, M. B., Cowman, A. F. and McFadden, G. I. (2000). Protein trafficking to the plastid of Plasmodium falciparum is via the secretory pathway. EMBO Journal 19, 17941802.Google Scholar
Wang, J., Soisson, S. M., Young, K., Shoop, W., Kodali, S., Galgoci, A., Painter, R., Parthasarathy, G., Tang, Y. S., Cummings, R., Ha, S., Dorso, K., Motyl, M., Jayasuriya, H., Ondeyka, J., Herath, K., Zhang, C., Hernandez, L., Allocco, J., Basilio, A., Tormo, J. R., Genilloud, O., Vicente, F., Pelaez, F., Colwell, L., Lee, S. H., Michael, B., Felcetto, T., Gill, C., Silver, L. L., Hermes, J. D., Bartizal, K., Barrett, J., Schmatz, D., Becker, J. W., Cully, D. and Singh, S. B. (2006). Platensimycin is a selective FabF inhibitor with potent antibiotic properties. Nature, London 441, 358361.Google Scholar
Ward, W. H., Holdgate, G. A., Rowsell, S., McLean, E. G., Pauptit, R. A., Clayton, E., Nichols, W. W., Colls, J. G., Minshull, C. A., Jude, D. A., Mistry, A., Timms, D., Camble, R., Hales, N. J., Britton, C. J. and Taylor, I. W. (1999). Kinetic and structural characteristics of the inhibition of enoyl (acyl carrier protein) reductase by triclosan. Biochemistry 38, 1251412525.Google Scholar
Weppelman, R. M., Vandenheuvel, W. J. and Wang, C. C. (1976). Mass spectrometric analysis of the fatty acids and nonsaponifiable lipids of Eimeria tenella oocysts. Lipids 11, 209215.CrossRefGoogle ScholarPubMed
White, S. W., Zheng, J., Zhang, Y. M. and Rock, (2005). The structural biology of type II fatty acid biosynthesis. Annual Review of Biochemistry 74, 791831.Google Scholar
Williams, R. B. (1998). Epidemiological aspects of the use of live anticoccidial vaccines for chickens. International Journal of Parasitology 28, 10891098.Google Scholar
Williamson, D. H., Gardner, M. J., Preiser, P., Moore, D. J., Rangachari, K. and Wilson, R. J. (1994). The evolutionary origin of the 35 kb circular DNA of Plasmodium falciparum: new evidence supports a possible rhodophyte ancestry. Molecular & General Genetics 243, 249252.Google Scholar
Wilson, R. J. M. (2002). Progress with parasite plastids. Journal of Molecular Biology 319, 257274.Google Scholar
Zhu, G., Marchewka, M. J. and Keithly, J. S. (2000 a) Cryptosporidium parvum appears to lack a plastid genome. Microbiology 146, 315321.CrossRefGoogle Scholar
Zhu, G., Marchewka, M. J., Woods, K. M., Upton, S. J. and Keithly, J. S. (2000 b). Molecular analysis of a Type I fatty acid synthase in Cryptosporidium parvum. Molecular and Biochemical Parasitology 105, 253260.Google Scholar
Zuther, E., Johnson, J. J., Haselkorn, R., McLeod, R. and Gornicki, P. (1999). Growth of Toxoplasma gondii is inhibited by aryloxyphenoxypropionate herbicides targeting acetyl-CoA carboxylase. Proceedings of the National Academy of Sciences, USA 96, 1338713392.Google Scholar
Supplementary material: PDF

Lu Supplementary Material

Tables.pdf

Download Lu Supplementary Material(PDF)
PDF 87.9 KB