Lactate inhibits citrulline and arginine synthesis from proline in pig enterocytes. (1/132)

Hypocitrullinemia and hypoargininemia but hyperprolinemia are associated with elevated plasma concentration of lactate in infants. Because the small intestine may be a major organ for initiating proline catabolism via proline oxidase in the body and is the major source of circulating citrulline and arginine in neonates, we hypothesized that lactate is an inhibitor of intestinal synthesis of citrulline and arginine from proline. To test this hypothesis, jejunum was obtained from 14-day-old suckling pigs for preparation of enterocyte mitochondria and metabolic studies. Mitochondria were used for measuring proline oxidase activity in the presence of 0-10 mM L-lactate. For metabolic studies, enterocytes were incubated at 37 degrees C for 30 min in Krebs bicarbonate buffer (pH 7.4) containing 5 mM D-glucose, 2 mM L-glutamine, 2 mM L-[U-14C]proline, and 0, 1, 5, or 10 mM L-lactate. Kinetics analysis revealed noncompetitive inhibition of intestinal proline oxidase by lactate (decreased maximal velocity and unaltered Michaelis constant). Lactate had no effect on either activities of other enzymes for arginine synthesis from proline or proline uptake by enterocytes but decreased the synthesis of ornithine, citrulline, and arginine from proline in a concentration-dependent manner. These results demonstrate that lactate decreased intestinal synthesis of citrulline and arginine from proline via an inhibition of proline oxidase and provide a biochemical basis for explaining hyperprolinemia, hypocitrullinemia, and hypoargininemia in infants with hyperlactacidemia.  (+info)

Regulation of flavin dehydrogenase compartmentalization: requirements for PutA-membrane association in Salmonella typhimurium. (2/132)

PutA is a multifunctional, peripheral membrane protein which functions both as an autogenous transcriptional repressor and the enzyme which catalyzes the two-step conversion of proline to glutamate in Salmonella typhimurium and Escherichia coli. To understand how PutA associates with the membrane, we determined the role of FAD redox and membrane components in PutA-membrane association. Reduction of the tightly bound FAD is required for both derepression of the put operon and membrane association of PutA. FADH(2) alters the conformation of PutA, resulting in an increased hydrophobicity. Previous studies used enzymatic activity as an assay for membrane association and concluded that electron transfer from the reduced FAD in PutA to the membrane is required for the PutA-membrane interaction. However, direct physical assays of PutA association with membrane vesicles from quinone deficient mutants demonstrated that although electron transfer is essential for proline dehydrogenase activity, it is not required for PutA-membrane association per se. Furthermore, PutA efficiently associated with liposomes, indicating that PutA-membrane association does not require interactions with other membrane proteins. PutA enzymatic activity can be efficiently reconstituted with liposomes containing ubiquinone and cytochrome bo, confirming that proline dehydrogenase can pass electrons directly to the quinone pool. These results indicate that PutA-membrane association is due strictly to a protein-lipid interaction initiated by reduction of FAD.  (+info)

Proline catabolism by Pseudomonas putida: cloning, characterization, and expression of the put genes in the presence of root exudates. (3/132)

Pseudomonas putida KT2442 is a root-colonizing strain which can use proline, one of the major components in root exudates, as its sole carbon and nitrogen source. A P. putida mutant unable to grow with proline as the sole carbon and nitrogen source was isolated after random mini-Tn5-Km mutagenesis. The mini-Tn5 insertion was located at the putA gene, which is adjacent to and divergent from the putP gene. The putA gene codes for a protein of 1,315 amino acid residues which is homologous to the PutA protein of Escherichia coli, Salmonella enterica serovar Typhimurium, Rhodobacter capsulatus, and several Rhizobium strains. The central part of P. putida PutA showed homology to the proline dehydrogenase of Saccharomyces cerevisiae and Drosophila melanogaster, whereas the C-terminal end was homologous to the pyrroline-5-carboxylate dehydrogenase of S. cerevisiae and a number of aldehyde dehydrogenases. This suggests that in P. putida, both enzymatic steps for proline conversion to glutamic acid are catalyzed by a single polypeptide. The putP gene was homologous to the putP genes of several prokaryotic microorganisms, and its gene product is an integral inner-membrane protein involved in the uptake of proline. The expression of both genes was induced by proline added in the culture medium and was regulated by PutA. In a P. putida putA-deficient background, expression of both putA and putP genes was maximal and proline independent. Corn root exudates collected during 7 days also strongly induced the P. putida put genes, as determined by using fusions of the put promoters to 'lacZ. The induction ratio for the putA promoter (about 20-fold) was 6-fold higher than the induction ratio for the putP promoter.  (+info)

Metabolite repression and inducer exclusion in the proline utilization gene cluster of Aspergillus nidulans. (4/132)

The clustered prnB, prnC, and prnD genes are repressed by the simultaneous presence of glucose and ammonium. A derepressed mutation inactivating a CreA-binding site acts in cis only on the permease gene (prnB) while derepression of prnD and prnC is largely the result of reversal of inducer exclusion.  (+info)

Sinorhizobium meliloti putA gene regulation: a new model within the family Rhizobiaceae. (5/132)

Proline dehydrogenase (PutA) is a bifunctional enzyme that catalyzes the oxidation of proline to glutamate. In Sinorhizobium meliloti, as in other microorganisms, the putA gene is transcriptionally activated in response to proline. In Rhodobacter capsulatus, Agrobacterium, and most probably in Bradyrhizobium, this activation is dependent on an Lrp-like protein encoded by the putR gene, located immediately upstream of putA. Interestingly, sequence and genetic analysis of the region upstream of the S. meliloti putA gene did not reveal such a putR locus or any other encoded transcriptional activator of putA. Furthermore, results obtained with an S. meliloti putA null mutation indicate the absence of any proline-responsive transcriptional activator and that PutA serves as an autogenous repressor. Therefore, the model of S. meliloti putA regulation completely diverges from that of its Rhizobiaceae relatives and resembles more that of enteric bacteria. However, some differences have been found with the latter model: (i) S. meliloti putA gene is not catabolite repressed, and (ii) the gene encoding for the major proline permease (putP) does not form part of an operon with the putA gene.  (+info)

Differential gene expression in p53-mediated apoptosis-resistant vs. apoptosis-sensitive tumor cell lines. (6/132)

Induction of wild-type p53 in the ECV-304 bladder carcinoma cell line by infection with a p53 recombinant adenovirus (Ad5CMV-p53) resulted in extensive apoptosis and eventual death of nearly all of the cells. As a strategy to determine the molecular events important to p53-mediated apoptosis in these transformed cells, ECV-304 cells were selected for resistance to p53 by repeated infections with Ad5CMV-p53. We compared the expression of 5,730 genes in p53-resistant (DECV) and p53-sensitive ECV-304 cells by reverse transcription-PCR, Northern blotting, and DNA microarray analysis. The expression of 480 genes differed by 2-fold or more between the two p53-infected cell lines. A number of potential targets for p53 were identified that play roles in cell cycle regulation, DNA repair, redox control, cell adhesion, apoptosis, and differentiation. Proline oxidase, a mitochondrial enzyme involved in the proline/pyrroline-5-carboxylate redox cycle, was up-regulated by p53 in ECV but not in DECV cells. Pyrroline-5-carboxylate (P5C), a proline-derived metabolite generated by proline oxidase, inhibited the proliferation and survival of ECV-304 and DECV cells and induced apoptosis in both cell lines. A recombinant proline oxidase protein tagged with a green fluorescent protein at the amino terminus localized to mitochondria and induced apoptosis in p53-null H1299 non-small cell lung carcinoma cells. The results directly implicate proline oxidase and the proline/P5C pathway in p53-induced growth suppression and apoptosis.  (+info)

Proline oxidase, encoded by p53-induced gene-6, catalyzes the generation of proline-dependent reactive oxygen species. (7/132)

The p53-dependent initiation of apoptosis is accompanied by the induction of proline oxidase (POX), a mitochondrial enzyme catalyzing the conversion of proline to pyrroline-5-carboxylate with the concomitant transfer of electrons to cytochrome c. However, the contribution of increased POX activity to apoptosis, if any, remains unknown. Using Adriamycin to initiate p53-dependent apoptosis, we showed that the expression of POX is up-regulated in a time- and dose-dependent manner in a human colon cancer cell line (LoVo). In cells expressing POX, the addition of proline increases reactive oxygen species (ROS) generation in a concentration-dependent manner; glutamate, a downstream product of proline oxidation, had no effect. Induction of POX was dependent on the p53 status of the cell. In the conditionally immortalized murine colonic epithelial cell line YAMC, where the p53 phenotype can be modulated by temperature, proline oxidase expression and ROS production could only be induced when the cells were phenotypically p53-positive. To confirm that the observed ROS production was not secondary to some other effect of p53, we also conditionally expressed POX in a p53-negative colon cancer line. Again, we found a proline-dependent ROS increase with POX expression. We hypothesize that proline oxidation supports the generation of ROS by donating reducing potential to an electron transport chain altered either by p53-dependent mechanisms or by overexpression of POX.  (+info)

Purification, characterization, and application of a novel dye-linked L-proline dehydrogenase from a hyperthermophilic archaeon, Thermococcus profundus. (8/132)

The distribution of dye-linked L-amino acid dehydrogenases was investigated in several hyperthermophiles, and the activity of dye-linked L-proline dehydrogenase (dye-L-proDH, L-proline:acceptor oxidoreductase) was found in the crude extract of some Thermococcales strains. The enzyme was purified to homogeneity from a hyperthermophilic archaeon, Thermococcus profundus DSM 9503, which exhibited the highest specific activity in the crude extract. The molecular mass of the enzyme was about 160 kDa, and the enzyme consisted of heterotetrameric subunits (alpha(2) beta(2)) with two different molecular masses of about 50 and 40 kDa. The N-terminal amino acid sequences of the alpha-subunit (50-kDa subunit) and the beta-subunit (40-kDa subunit) were MRLTEHPILDFSERRGRKVTIHF and XRSEAKTVIIGGGIIGLSIAYNLAK, respectively. Dye-L-proDH was extraordinarily stable among the dye-linked dehydrogenases under various conditions: the enzyme retained its full activity upon incubation at 70 degrees C for 10 min, and ca. 40% of the activity still remained after heating at 80 degrees C for 120 min. The enzyme did not lose the activity upon incubation over a wide range of pHs from 4.0 to 10.0 at 50 degrees C for 10 min. The enzyme exclusively catalyzed L-proline dehydrogenation using 2,6-dichloroindophenol (Cl2Ind) as an electron acceptor. The Michaelis constants for L-proline and Cl2Ind were determined to be 2.05 and 0.073 mM, respectively. The reaction product was identified as Delta(1)-pyrroline-5-carboxylate by thin-layer chromatography. The prosthetic group of the enzyme was identified as flavin adenine dinucleotide by high-pressure liquid chromatography. In addition, the simple and specific determination of L-proline at concentrations from 0.10 to 2.5 mM using the stable dye-L-proDH was achieved.  (+info)