首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
During the cource of the investigation of ribotidation of purine and pyrimidine bases by Brevibacterium ammoniagenes ATCC 6872, it was found that a large amount of uridine 5′-monophosphate (UMP) was accumulated in the culture broth when the organism was incubated in a medium containing uracil or orotic acid. The yields of UMP were 83% (4.8 mg/ml) from uracil and 100% (4.3 mg/ml) from orotic acid when each substrate was added at the concentration of 2 mg/ml.

Addition of 6-azauracil or 5-hydroxyuracil to the culture of the organism during cultivation led to the accumulation of both orotidine 5′-monophosphate (OMP) and UMP. The accumulation of OMP seemed to be due to the inhibition of OMP decarboxylase (E. C. 4.1.1.23) by the ribotide formed from each base. The OMP accumulation was enhanced by the addition of orotic acid in addition to 6-azauracil. When 6-azauracil was added to the medium before inoculation, UMP was predominantly accumulated, and when it was added after one day incubation, OMP was predominantly accumulated. A largest accumulation (3.6 mg/ml) of OMP was obtained when 6-azauracil was added on the 1st day and orotic acid was added on the 3rd day.

UMP and OMP accumulated in the medium were isolated from the cultured broth and identified by usual methods.  相似文献   

2.
Flies mutant for one or both of the last two enzymes of de novo pyrimidine biosynthesis express a number of phenotypes that are also expressed by mutants of the first four pathway enzymes (r and Dhod-null mutants). However, r-1 flies also express two phenotypes, mottled eyes and poor viability, that are not usually expressed by r and Dhod-null flies. Chemical determinations show that orotic acid, a substrate for the fifth pathway enzyme, accumulates in r-1 individuals but not in r and wild-type individuals. Moreover, flies simultaneously mutant for r and r-1 do not express the mottled-eye phenotype, showing that r is epistatic to r-1 for this r-1-specific phenotype. When genotypically wild-type flies are cultured on a medium containing 6-azauracil, the base of a potent inhibitor of the last enzyme of de novo pyrimidine biosynthesis, phenocopies are obtained that include the mottled-eye as well as the wing phenotypes of r-1 flies. These results support hypotheses that the phenotypes common to r, Dhod-null, and r-1 flies are consequences of uridylic acid deficiency, whereas the r-1-specific phenotypes result from orotic acid accumulation in flies lacking either or both of the last two enzymes of de novo pyrimidine biosynthesis.This research was supported by NSF Research Grant PCM 78-14164, an NSF predoctoral fellowship award to T. Conner, and an NIH research career development award to J. Rawls.  相似文献   

3.
Uridine monophosphate (UMP) synthase is a bifunctional enzyme catalyzing the last two steps of de novo pyrimidine biosynthesis, orotate phosphoribosyltransferase (OPRT) and orotidine-5'-monophosphate decarboxylase (ODC). Loss of either enzymatic activity results in hereditary orotic aciduria, a rare autosomal recessive disorder characterized by retarded growth, anemia, and excessive urinary excretion of orotic acid. We have isolated the UMP synthase chromosomal gene from a lambdaEMBL-3 human genomic library and report a single-copy gene spanning approximately 15 kb. The UMP synthase genomic structure encodes six exons ranging in size from 115 bp to 672 bp, and all splicing junctions adhere to the canonical GT/AG rule. Cognate promoter elements implicated in glucocorticoid- and cAMP-mediated regulation as well as in liver-, myeloid-, and lymphocyte-specific expression are located within the 5' flanking sequence. Molecular investigation of UMP synthase deficiency in a Japanese orotic aciduria patient revealed mutations R96G (A-to-G transition; nt 286) and G429R (G-to-C transversion; nt 1285) in one allele and V109G (T-to-G transversion; nt 326) in the other allele. Expression of human UMP synthase cDNAs containing these mutations in pyrimidine auxotrophic Escherichia coli and in recombinant baculovirus-infected Sf21 cells demonstrates impaired activity presumably associated with the urinary orotic acid substrate accumulations observed in vivo. We further establish the identity of two polymorphisms, G213A (v = .26) and 440Gpoly (v = .27) located in exons 3 and 6, respectively, which did not significantly compromise either OPRT or ODC function.  相似文献   

4.
Arginine deficiency is associated with a mild orotic aciduria. Liver slices from rats fed a purified l-amino acid diet with (control) and without arginine supplementation were used for studies of [14C]bicarbonate incorporation into orotic acid. The nanomoles of orotic acid synthesized in isolated liver slices from both control and arginine-deficient animals increased linearly with time. Orotic acid biosynthesis was significantly greater in liver slices than slices of heart, muscle, kidney, and minced spleen. The order of orotate biosynthesis from [14C]bicarbonate was liver > spleen = kidney > muscle > heart. Arginine deficiency resulted in a significant stimulation of liver orotic acid biosynthesis. This stimulation in pyrimidine biosynthesis can account for a major portion of the orotic aciduria. Orotic acid synthesis from spleens isolated from arginine-deficient rats was also enhanced compared with controls. Although the rate of orotic acid biosynthesis is small relative to liver production, the spleen may contribute slightly to increased orotic aciduria in the arginine-deficient rat. Arginine supplementation in vitro to livers from rats fed either the control of arginine-deficient diet resulted in a significant reduction in synthesis of orotic acid. Dietary arginine may play a key role in regulating mitochondrial carbamoyl phosphate utilization into both pyrimidine and urea biosynthesis.  相似文献   

5.
The mechanism whereby Escherichia coli K12 accumulates orotic acid in culture fluid was studied. Pyrimidine compounds were incorporated effectively into cells of E. coli K12, stimulated the growth, and depressed the accumulation; while purine compounds were not so much consumed by the microorganism for its growth, and affected the accumulation to a lesser extent. On the other hand, E. coli B unable to accumulate orotic acid utilized less effectively pyrimidine compounds for its growth than strain K12.

It is supposed, therefore, that in the de novo pathway for pyrimidine synthesis in E. coli K12 the step from orotic acid to 5′-UMP is genetically depressed so that orotic acid is accumulated when pyrimidine compounds, that would cause a feedback inhibition of orotic acid synthesis upon incorporation, are not supplemented.  相似文献   

6.
The bifunctional enzyme uridine monophosphate (UMP) synthase catalyzes the last two steps in de novo pyrimidine biosynthesis. A genetic deficiency in the activity of this enzyme causes the inherited human disease orotic aciduria. We used a human cDNA probe to localize the gene for UMP synthase to human chromosome region 3q13 by the technique of in situ hybridization.  相似文献   

7.
Jones GE 《Plant physiology》1984,75(1):161-165
6-Azauracil-resistant variants of Haplopappus gracilis (Nutt.) Gray and Datura innoxia Mill. lack activity of uracil phosphoribosyltransferase, a pyrimidine salvage enzyme that catalyzes the conversion of uracil and 6-azauracil to uridine-5′-monophosphate and 6-azauridine-5′-monophosphate, respectively. Resistant cells are competent to take up uracil from their growth medium but do not convert it into a form that can be used for macromolecular synthesis. In extracts from resistant cells, orotate monophosphate decarboxylase, a target enzyme of 6-azauridine monophosphate, is fully sensitive to the phosphorylated analog. These results strongly suggest that uracil phosphoribosyltransferase is the major pathway of pyrimidine salvage in cells of these species and that loss of this enzyme activity confers on the variants resistance to 6-azauracil.  相似文献   

8.
To study the regulation of lysine and threonine metabolism in plants, we have transformed Arabidopsis thaliana with chimeric genes encoding the two bacterial enzymes dihydrodipicolinate synthase (DHPS) and aspartate kinase (AK). These bacterial enzymes are much less sensitive to feedback inhibition by lysine and threonine than their plant counterparts. Transgenic plants expressing the bacterial DHPS overproduced lysine, but lysine levels were quite variable within and between transgenic genotypes and there was no direct correlation between the levels of free lysine and the activity of DHPS. The most lysine-overproducing plants also exhibited abnormal phenotypes. However, these phenotypes were detected only at early stages of plant growth, while at later stages, new buds emerged that looked completely normal and set seeds. Wild-type plants exhibited relatively high levels of free threonine, suggesting that in Arabidopsis AK regulation may be more relaxed than in other plants. This was also supported by the fact that expression of the bacterial AK did not cause any dramatic elevation in this amino acid. Yet, the relaxed regulation of threonine synthesis in Arabidopsis was not simply due to a reduced sensitivity of the endogenous AK to feedback inhibition by lysine and threonine because growth of wild-type plants, but not of transgenic plants expressing the bacterial AK, was arrested in media containing these two amino acids. The present results, combined with previous studies from our laboratory, suggest that the regulation of lysine and threonine metabolism is highly variable among plant species and is subject to complex biochemical, physiological and environmental controls. The suitability of these transgenic Arabidopsis plants for molecular and genetic dissection of lysine and threonine metabolism is also discussed.  相似文献   

9.
10.
11.
Pyrimidine metabolism was investigated at various stages ofsomatic embryo development of white spruce (Picea glauca). The contribution of thede novo and the salvage pathways of pyrimidine biosynthesis to nucleotide and nucleic acid formation and the catabolism of pyrimidine was estimated by the exogenously supplied [6-14C]orotic acid, an intermediate of thede novo pathway, and with [2-14C]uridine and [2-14C]uracil, substrates of the salvage pathways. Thede novo pathway was very active throughout embryo development. More than 80 percnt; of [6-14C]orotic acid taken up by the tissue was utilized for nucleotide and nucleic acid synthesis in all stages of this process. The salvage pathways of uridine and uracil were also operative. Relatively high nucleic acid biosynthesis from uridine was observed, whereas the contribution of uracil salvage to the pyrimidine nucleotide and nucleic acid synthesis was extremely limited. A large proportion of uracil was degraded as 14CO2, probably via β-ureidopropionate. Among the enzymes of pyrimidine metabolism, orotate phosphoribosyltransferase was high during the initial phases of embryo development, after which it gradually declined. Uridine kinase, responsible for the salvage of uridine, showed an opposite pattern, since its activity increased as embryos developed. Low activities of uracil phosphoribosyltransferase and non-specific nucleoside phosphotransferase were also detected throughout the developmental period. These results suggest that the flux of thede novo and salvage pathways of pyrimidine nucleotide biosynthesisin vivo is roughly controlled by the amount of these enzymes. However, changing patterns of enzyme activity during embryo development that were measuredin vitro did not exactly correlate with the flux estimated by the radioactive precursors. Therefore, other fine control mechanisms, such as the fluctuation of levels of substrates and/or effectors may also participate to the real control of pyrimidine metabolism during white spruce somatic embryo development.  相似文献   

12.
UMP pyrophosphorylase (EC 2.4.2.9, UMP:pyrophosphate phosphoribosyltransferase) was purified approximately 85-fold from exponentially growing cells of Tetrahymena pyriformis GL-7. It was found to have a molecular weight of 36,000, and was active over a broad pH range, with an optimum at 7.5. The enzyme exhibited a temperature optimum at 40 °C, above which irreversible inactivation began to occur. The apparent Km values for uracil and phosphoribosyl pyrophosphate (PRPP) were 0.4 and 6.9 m, respectively. The pyrophosphorylase exhibited a pyrimidine base specificity for uracil, although 5-fluorouracil was utilized by the enzyme. Neither cytosine, orotic acid, nor 6-azauracil competed with uracil for the enzyme or inhibited the production of UMP from uracil and PRPP. Although most triphosphates had little effect on pyrophosphorylase activity, UTP and dUTP, each at a concentration of 1 mm, depressed UMP formation by 86 and 59%, respectively. Thus, UMP pyrophosphorylase may be sensitive to feedback inhibition by the product of the pathway it initiates. UMP pyrophosphorylase specific activity in extracts of Tetrahymena grown in a medium containing uracil as the sole pyrimidine source was threefold higher than that in extracts of cells grown on uridine or UMP.  相似文献   

13.
Type 1 phosphoprotein Ser/Thr phosphatases (PP1) are highly conserved enzymes found in all eukaryotes. These enzymes have multiple functions in fungal and animal cells but little is known of their function and regulation in plants. Previous studies in our laboratory indicated that maize and Arabidopsis contain a family of PP1 genes and/or pseudogenes. In this study, we report the isolation of five distinct Arabidopsis cDNA clones (TOPP1, TOPP2, TOPP3, TOPP4 and TOPP5) which encode the catalytic subunit (PP1c) of type 1 protein phosphatases. Genomic Southern blot analyses indicate that these clones are the products of five distinct genes and that an additional 2–3 PP1c genes and/or pseudogenes may be present in the Arabidopsis genome. The derived amino acid sequences of the TOPP clones are very similar to published sequences of PP1c from animals, fungi and plants. Four of the TOPP amino acid sequences show unique structural features not observed in other PP1c sequences from fungi or animals. All of the TOPP genes are expressed in Arabidopsis roots, rosettes and flowers, although TOPP1, TOPP2 and TOPP3 appear to be expressed at higher levels in these tissues than TOPP4 and TOPP5.  相似文献   

14.
Zhong  Xiao  Feng  Pei  Ma  Qiqi  Zhang  Yi  Yang  Yazhen  Zhang  Jianmin 《Plant Molecular Biology Reporter》2021,39(1):251-261

Despite the involvement of many members of the chitinase family in the plant immune system, the exact functions of most chitinases remain poorly understood, especially in plant defense responses to phytophagous insects. Here, the gene GhChi6, which encodes a chitinase protein in Gossypium hirsutum, was shown to be induced by cotton aphid feeding and mechanical wounding. Overexpression of GhChi6 in Arabidopsis plants improved their defense response to aphids. The activities of chitinase and PPO in GhChi6 transgenic Arabidopsis plants were higher than those in wild-type plants. Callose deposition in leaves from GhChi6 transgenic Arabidopsis plants was clearly increased compared with wild-type plants. The levels of AtEDS1, AtPAD4, and AtEDS5 in the SA signaling pathway were higher in GhChi6 transgenic Arabidopsis Line4 than those in wild-type plants, while the expression levels of AtLOX2 in the JA signaling pathway and AtEIN2 in the ethylene signaling pathway were lower in GhChi6 transgenic Arabidopsis Line4 than those in wild-type plants. These results collectively showed that the cotton chitinase gene GhChi6 modulated the plant defense response to aphid attack, which may help guide strategies for improving cotton aphid prevention.

  相似文献   

15.
Mannitol, a sugar alcohol that may serve as a compatible solute to cope with salt stress, is synthesized via the action of a mannose‐6‐phosphate reductase (M6PR) in celery (Apium graveolens L). In contrast to previous approaches that have used a bacterial gene to engineer mannitol biosynthesis in plants and other organisms, Arabidopsis thaliana, a non‐mannitol producer, was transformed with the celery leaf M6PR gene under control of the CaMV 35S promotor. In all independent Arabidopsis M6PR transformants, mannitol accumulated throughout the plants in amounts ranging from 0·5 to 6 µmol g?1 fresh weight. A novel compound, not found in either celery or Arabidopsis, 1‐O‐β‐d ‐glucopyranosyl‐d ‐mannitol, also accumulated in vegetative tissues of mature plants in amounts up to 4 µmol g?1 fresh weight, but not in flowers and seeds. In the absence of NaCl, all transformants were phenotypically the same as the wild type; however, in the presence of NaCl, mature transgenic plants showed a high level of salt tolerance, i.e. growing, completing normal development, flowering, and producing seeds in soil irrigated with 300 mm NaCl in the nutrient solution. These results demonstrate a major role in developing salt‐tolerant plants by means of introducing mannitol biosynthesis using M6PR.  相似文献   

16.
ABSTRACT

Orotate (OA) is well-known as a precursor in biosynthesis of pyrimidines; in mammals it is released from the mitochondrial dihydroorotate dehydrogenase (DHODH) for conversion to UMP by the cytoplasmic UMP synthase enzyme. OA is also a normal part of the diet, being found in milk and dairy products, and it is converted to uridine for use in the pyrimidine salvage pathway predominantly in liver, kidney and erythrocytes. Early research into nutrition identified orotate as “vitamin B13,” and its use as a complex with organic cations or metal ions was promulgated in body-building, and in assisting therapies of metabolic syndromes. It has recently been established that the amelioration of gout by dairy products arises from the competition of orotate and urate at the hURAT1 transporter. The orotic aciduria that arises in children with defective UMP synthase can be rescued by oral uridine therapy, since UMP is the end-product and also a feedback inhibitor of the de novo pathway. In contrast, Miller (dysmorphology) syndrome is connected with defects in DHODH, and hence in the supply of OA, and cannot be helped by uridine. Other models of dysmorphisms are connected with enzymes early in the pyrimidine de novo pathway. We conclude that the OA molecule is itself required for the regulation of genes that are important in the development of cells, tissues and organisms.  相似文献   

17.
18.
The transporter(s) that mediate uptake of nicotinate and its N-methyl derivative trigonelline are not known in plants, and certain mammalian nicotinate transporters also remain unidentified. Potential candidates for these missing transporters include proteins from the ubiquitous NiaP family. In bacteria, niaP genes often belong to NAD-related regulons, and genetic evidence supports a role for Bacillus subtilis and Acinetobacter baumannii NiaP proteins in uptake of nicotinate or nicotinamide. Other bacterial niaP genes are, however, not in NAD-related regulons but cluster on the chromosome with choline-related (e.g., Ralstonia solanacearum and Burkholderia xenovorans) or thiamin-related (e.g., Thermus thermophilus) genes, implying that they might encode transporters for these compounds. Radiometric uptake assays using Lactococcus lactis cells expressing NiaP proteins showed that B. subtilis, R. solanacearum, and B. xenovorans NiaP transport nicotinate via an energy-dependent mechanism. Likewise, NiaP proteins from maize (GRMZM2G381453, GRMZM2G066801, and GRMZM2G081774), Arabidopsis (At3g13050), and mouse (SVOP) transported nicotinate; the Arabidopsis protein also transported trigonelline. In contrast, T. thermophilus NiaP transported only thiamin. None of the proteins tested transported choline or the thiazole and pyrimidine products of thiamin breakdown. The maize and Arabidopsis NiaP proteins are the first nicotinate transporters reported in plants, the Arabidopsis protein is the first trigonelline transporter, and mouse SVOP appears to represent a novel type of mammalian nicotinate transporter. More generally, these results indicate that specificity for nicotinate is conserved widely, but not absolutely, among pro- and eukaryotic NiaP family proteins.  相似文献   

19.
Tomato spotted wilt virus (TSWV) is one of the most devastating plant viruses and often causes severe crop losses worldwide. Generally, mature plants become more resistant to pathogens, known as adult plant resistance. In this study, we demonstrated a new phenomenon involving developmentally regulated susceptibility of Arabidopsis thaliana to TSWV. We found that Arabidopsis plants become more susceptible to TSWV as plants mature. Most young 3-week-old Arabidopsis were not infected by TSWV. Infection of TSWV in 4-, 5-, and 6-week-old Arabidopsis increased from 9%, 21%, and 25%, respectively, to 100% in 7- to 8-week-old Arabidopsis plants. Different isolates of TSWV and different tospoviruses show a low rate of infection in young Arabidopsis but a high rate in mature plants. When Arabidopsis dcl2/3/4 or rdr1/2/6 mutant plants were inoculated with TSWV, similar results as observed for the wild-type Arabidopsis plants were obtained. A cell-to-cell movement assay showed that the intercellular movement efficiency of TSWV NSm:GFP fusion was significantly higher in 8-week-old Arabidopsis leaves compared with 4-week-old Arabidopsis leaves. Moreover, the expression levels of pectin methylesterase and β-1,3-glucanase, which play critical roles in macromolecule cell-to-cell trafficking, were significantly up-regulated in 8-week-old Arabidopsis leaves compared with 4-week-old Arabidopsis leaves during TSWV infection. To date, this mature plant susceptibility to pathogen infections has rarely been investigated. Thus, the findings presented here should advance our knowledge on the developmentally regulated mature host susceptibility to plant virus infection.  相似文献   

20.
A uracil phosphoribosyltransferase (UMP-pyrophosphorylase) was found in several angiosperms and was partially purified from epicotyls of pea (Pisum sativum L. cv. Alaska) seedlings. Its pH optimum was about 8.5; its required approximately 0.3 mm MgCl2 for maximum activity but was inhibited by MnCl2; its molecular weight determined by chromatography on Sephadex G-150 columns was approximately 100,000; its Km values for uracil and 5-phosphorylribose 1-pyrophosphate were 0.7 μm and 11 μm; and it was partially resolved from a similar phosphoribosyltransferase converting orotic acid to orotodine 5′-phosphate. Enzyme fractions containing both uracil phosphoribosyl transferase and orotate phosphoribosyltransferase converted 6-azauracil and 5-fluorouracil to products with chromatographic properties of 6-azauradine 5′-phosphate and 5-fluorouridine 5′-phosphate. Uracil phosphoribosyltransferase probably functions in salvage of uracil for synthesis of pyrimidine nucleotides.  相似文献   

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