首页 | 本学科首页   官方微博 | 高级检索  
     


Evidence that Ubiquinone Is a Required Intermediate for Rhodoquinone Biosynthesis in Rhodospirillum rubrum
Authors:Brian C. Brajcich  Andrew L. Iarocci  Lindsey A. G. Johnstone  Rory K. Morgan  Zachary T. Lonjers  Matthew J. Hotchko  Jordan D. Muhs  Amanda Kieffer  Bree J. Reynolds  Sarah M. Mandel  Beth N. Marbois  Catherine F. Clarke  Jennifer N. Shepherd
Affiliation:Department of Chemistry and Biochemistry, Gonzaga University, Spokane, Washington 99258,1. Department of Chemistry and Biochemistry and the Molecular Biology Institute, UCLA, Los Angeles, California 900952.
Abstract:Rhodoquinone (RQ) is an important cofactor used in the anaerobic energy metabolism of Rhodospirillum rubrum. RQ is structurally similar to ubiquinone (coenzyme Q or Q), a polyprenylated benzoquinone used in the aerobic respiratory chain. RQ is also found in several eukaryotic species that utilize a fumarate reductase pathway for anaerobic respiration, an important example being the parasitic helminths. RQ is not found in humans or other mammals, and therefore inhibition of its biosynthesis may provide a parasite-specific drug target. In this report, we describe several in vivo feeding experiments with R. rubrum used for the identification of RQ biosynthetic intermediates. Cultures of R. rubrum were grown in the presence of synthetic analogs of ubiquinone and the known Q biosynthetic precursors demethylubiquinone, demethoxyubiquinone, and demethyldemethoxyubiquinone, and assays were monitored for the formation of RQ3. Data from time course experiments and S-adenosyl-l-methionine-dependent O-methyltransferase inhibition studies are discussed. Based on the results presented, we have demonstrated that Q is a required intermediate for the biosynthesis of RQ in R. rubrum.Rhodospirillum rubrum is a well-characterized and metabolically diverse member of the family of purple nonsulfur bacteria (29, 61). R. rubrum is typically found in aquatic environments and can adapt to a variety of growth conditions by using photosynthesis, respiration, or fermentation pathways (28, 70). In the light, R. rubrum exhibits photoheterotrophic growth using organic substrates or photoautotrophic growth using CO2 and H2 (15, 70). In the dark, R. rubrum can utilize either aerobic respiration (70, 73) or anaerobic respiration with a fumarate reduction pathway or with nonfermentable substrates in the presence of oxidants such as dimethyl sulfoxide (DMSO) or trimethylamine oxide (15, 58, 73). R. rubrum can also grow anaerobically in the dark by fermentation of sugars in the presence of bicarbonate (58). The focus of this work was the biosynthesis of quinones used by R. rubrum for aerobic and anaerobic respiration.Rhodoquinone (RQ; compound 1 in Fig. Fig.1)1) is an aminoquinone structurally similar to ubiquinone (coenzyme Q or Q [compound 2]) (44); however, the two differ considerably in redox potential (that of RQ is −63 mV, and that of Q is +100 mV) (2). Both RQ and Q have a fully substituted benzoquinone ring and a polyisoprenoid side chain that varies in length (depending on the species; see Fig. Fig.11 for examples). The only difference between the structures is that RQ has an amino substituent (NH2) instead of a methoxy substituent (OCH3) on the quinone ring. While Q is a ubiquitous lipid component involved in aerobic respiratory electron transport (9, 36, 60), RQ functions in anaerobic respiration in R. rubrum (19) and in several other phototrophic purple bacteria (21, 22, 41) and is also present in a few aerobic chemotrophic bacteria, including Brachymonas denitrificans and Zoogloea ramigera (23). In these varied species of bacteria, RQ has been proposed to function in fumarate reduction to maintain NAD+/NADH redox balance, either during photosynthetic anaerobic metabolism (12, 15-18, 64) or in chemotrophic metabolism when the availability of oxygen as a terminal oxidant is limiting (23). Another recent finding is that RQH2 is capable of inducing Q-cycle bypass reactions in the cytochrome bc1 complex in Saccharomyces cerevisiae, resulting in superoxide formation (7). If RQ/RQH2 coexists in the cytoplasmic membrane with Q/QH2 in R. rubrum, it might serve as both a substrate for and an inhibitor of the bc1 complex (47).Open in a separate windowFIG. 1.Proposed pathways for RQ biosynthesis. The number of isoprene units (n) varies by species (in S. cerevisiae, n = 6; in E. coli, n = 8; in C. elegans, n = 9; in helminth parasites, n = 9 or 10; in R. rubrum, n = 10; in humans, n = 10). RQ is not found in S. cerevisiae, E. coli, or humans. Known Coq (from S. cerevisiae) and Ubi (from E. coli) gene products required for the biosynthesis of ubiquinone (Q, compound 2) are labeled. A polyisoprenyl diphosphate (compound 5) is assembled from dimethylallyl disphosphate (compound 3) and isopentyl diphosphate (compound 4). Coupling of compound 5 with p-hydroxybenzoic acid (compound 6) yields 3-polyprenyl-4-hydroxybenzoic acid (compound 7). The next three steps differ between S. cerevisiae and E. coli. However, they merge at the common intermediate (compound 8), which is oxidized to demethyldemethoxyubiquinone (DDMQn, compound 9). RQ (compound 1) has been proposed to arise from compound 9, demethoxyubiquinone (DMQn; compound 10), demethylubiquinone (DMeQn; compound 11), or compound 2 (by pathway A, B, C, or D). Results presented in this work support pathway D as the favored route for RQ biosynthesis in R. rubrum.RQ is also found in the mitochondrial membrane of eukaryotic species capable of fumarate reduction, such as the flagellate Euglena gracilis (25, 53), the free-living nematode Caenorhabditis elegans (62), and the parasitic helminths (65, 66, 68, 72). Similar to R. rubrum, these species can adapt their metabolism to both aerobic and anaerobic conditions throughout their life cycle. For example, most adult parasitic species (e.g., Ascaris suum, Fasciola hepatica, and Haemonchus contortus) rely heavily on fumarate reduction for their energy generation while inside a host organism, where the oxygen tension is very low (30, 65, 72). Under these conditions, the biosynthesis of RQ is upregulated; however, during free-living stages of their life cycle, the helminth parasites use primarily aerobic respiration, which requires Q (30, 65, 72). The anaerobic energy metabolism of the helminthes has been reviewed (63, 67). Humans and other mammalian hosts use Q for aerobic energy metabolism but do not produce or require RQ; therefore, selective inhibition of RQ biosynthesis may lead to highly specific antihelminthic drugs that do not have a toxic effect on the host (35, 48).R. rubrum is an excellent facultative model system for the study of RQ biosynthesis. The complete genome of R. rubrum has recently been sequenced by the Department of Energy Joint Genome Institute, finished by the Los Alamos Finishing Group, and further validated by optical mapping (57). The 16S rRNA sequence of R. rubrum is highly homologous to cognate eukaryotic mitochondrial sequences (46). Due to the similarities in structure, the biosynthetic pathways of RQ and Q have been proposed to diverge from a common precursor (67). Proposed pathways for RQ biosynthesis (A to D), in conjunction with the known steps in Q biosynthesis, are outlined in Fig. Fig.11 (31, 34, 60). Parson and Rudney previously showed that when R. rubrum was grown anaerobically in the light in the presence of [U-14C]p-hydroxybenzoate, 14C was incorporated into both Q10 and RQ10 (50). In their growth experiments, the specific activity of Q10 was measured at its maximal value 15 h after inoculation and then began to decrease. However, the specific activity of RQ10 continued to increase for 40 h before declining. These results suggested that Q10 was a biosynthetic precursor of RQ10, although this was not directly demonstrated using radiolabeled Q10; hence, the possibility remained that the labeled RQ10 was derived from another radiolabeled lipid species. We have done this feeding experiment with a synthetic analog of Q where n = 3 (Q3) and monitored for the production of RQ3. The synthesis and use of farnesylated quinone and aromatic intermediates for characterization of the Q biosynthetic pathway in S. cerevisiae and Escherichia coli has been well documented (4, 5, 38, 52, 59). The other proposed precursors of RQ shown in Fig. Fig.11 were also fed to R. rubrum, and the lipid extracts from these assays were analyzed for the presence of RQ3, i.e., demethyldemethoxyubiquinone-3 (DDMQ3; compound 9), demethoxyubiquinone-3 (DMQ3; compound 10), and demethylubiquinone-3 (DMeQ3; compound 11).In S. cerevisiae and E. coli, the last O-methylation step in Q biosynthesis is catalyzed by the S-adenosyl-l-methionine (SAM)-dependent methyltransferases Coq3 and UbiG, respectively (26, 52); this final methylation step converts DMeQ to Q. Using the NCBI Basic Local Alignment Search Tool, an O-methyltransferase (GeneID no. 3834724 Rru_A0742) that had 41% and 59% sequence identity with Coq3 and UbiG, respectively, was identified in R. rubrum. S-Adenosyl-l-homocysteine (SAH) is a well-known inhibitor of SAM-dependent methyltransferases (13, 24). Because SAH is the transmethylation by-product of SAM-dependent methyltransferases, it is not readily taken up by cells and must be generated in vivo (24). SAH can be produced in vivo from S-adenosine and l-homocysteine thiolactone by endogenous SAH hydrolase (SAHH) (37, 71). A search of the R. rubrum genome also confirmed the presence of a gene encoding SAHH (GeneID no. 3836896 Rru_A3444). It was proposed that if DMeQ is the immediate precursor of RQ, then SAH inhibition of the methyltransferase required for Q biosynthesis should have little effect on RQ production. Conversely, if Q is required for RQ synthesis, then inhibition of Q biosynthesis should have a significant effect on RQ production. Assays were designed to quantify the levels of RQ3 produced from DMeQ3 and Q3 in R. rubrum cultures at various concentrations of SAH.
Keywords:
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号