Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 15 de 15
Filter
Add more filters










Publication year range
1.
Biochemistry ; 40(46): 13816-25, 2001 Nov 20.
Article in English | MEDLINE | ID: mdl-11705370

ABSTRACT

The nitrogenase MoFe protein contains the active site metallocluster called FeMo-cofactor [7Fe-9S-Mo-homocitrate] that exhibits an S = 3/2 EPR signal in the resting state. No interaction with FeMo-cofactor is detected when either substrates or inhibitors are incubated with MoFe protein in the resting state. Rather, the detection of such interactions requires the incubation of the MoFe protein together with its obligate electron donor, called the Fe protein, and MgATP under turnover conditions. This indicates that a more reduced state of the MoFe protein is required to accommodate substrate or inhibitor interaction. In the present work, substitution of an arginine residue (alpha-96(Arg)) located next to the active site FeMo-cofactor in the MoFe protein by leucine, glutamine, alanine, or histidine is found to result in MoFe proteins that can interact with acetylene or cyanide in the as-isolated, resting state without the need for the Fe protein, or MgATP. The dithionite-reduced, resting states of the alpha-96(Leu)-, alpha-96(Gln)-, alpha-96(Ala)-, or alpha-96(His)-substituted MoFe proteins show an S = 3/2 EPR signal (g = 4.26, 3.67, 2.00) similar to that assigned to FeMo-cofactor in the wild-type MoFe protein. However, in contrast to the wild-type MoFe protein, the alpha-96-substituted MoFe proteins all exhibit changes in their EPR spectra upon incubation with acetylene or cyanide. The alpha-96(Leu)-substituted MoFe protein was representative of the other alpha-96-substituted MoFe proteins examined. The incubation of acetylene with the alpha-96(Leu) MoFe protein decreased the intensity of the normal FeMo-cofactor signal with the appearance of a new EPR signal having inflections at g = 4.50 and 3.50. Incubation of cyanide with the alpha-96(Leu) MoFe protein also decreased the FeMo-cofactor EPR signal with concomitant appearance of a new EPR signal having an inflection at g = 4.06. The acetylene- and cyanide-dependent EPR signals observed for the alpha-96(Leu)-substituted MoFe protein were found to follow Curie law 1/T dependence, consistent with a ground-state transition as observed for FeMo-cofactor. The microwave power dependence of the EPR signal intensity is shifted to higher power for the acetylene- and cyanide-dependent signals, consistent with a change in the relaxation properties of the spin system of FeMo-cofactor. Finally, the alpha-96(Leu)-substituted MoFe protein incubated with (13)C-labeled cyanide displays a (13)C ENDOR signal with an isotropic hyperfine coupling of 0.42 MHz in Q-band Mims pulsed ENDOR spectra. This indicates the existence of some spin density on the cyanide, and thus suggests that the new component of the cyanide-dependent EPR signals arise from the direct bonding of cyanide to the FeMo-cofactor. These data indicate that both acetylene and cyanide are able to interact with FeMo-cofactor contained within the alpha-96-substituted MoFe proteins in the resting state. These results support a model where effective interaction of substrates or inhibitors with FeMo-cofactor occurs as a consequence of both increased reactivity and accessibility of FeMo-cofactor under turnover conditions. We suggest that, for the wild-type MoFe protein, the alpha-96(Arg) side chain acts as a gatekeeper, moving during turnover in order to permit accessibility of acetylene or cyanide to a specific [4Fe-4S] face of FeMo-cofactor.


Subject(s)
Acetylene/metabolism , Amino Acid Substitution , Cyanides/metabolism , Molybdoferredoxin/chemistry , Nitrogenase/chemistry , Arginine/metabolism , Azotobacter vinelandii/enzymology , Binding Sites , Carbon Monoxide/chemistry , Catalytic Domain , Electron Spin Resonance Spectroscopy , Enzyme Inhibitors/chemistry , Glutamine/metabolism , Histidine/metabolism , Leucine/metabolism , Molybdoferredoxin/antagonists & inhibitors , Molybdoferredoxin/metabolism , Nitrogenase/antagonists & inhibitors , Nitrogenase/metabolism , Substrate Specificity , Thermodynamics
2.
Plant Mol Biol ; 43(1): 113-20, 2000 May.
Article in English | MEDLINE | ID: mdl-10949378

ABSTRACT

The phytobilin chromophores of phycobiliproteins and phytochromes are biosynthesized from heme in a pathway that begins with the opening of the tetrapyrrole macrocycle of protoheme to form biliverdin IXalpha, in a reaction catalyzed by heme oxygenase. An Arabidopsis thaliana hy1 mutant was previously shown to be deficient in phytochrome responses, and these responses were regained when the plants were administered biliverdin IXalpha. A heme oxygenase-encoding gene, ho1, was recently cloned from the cyanobacterium Synechocystis sp. PCC 6803. When ho1 was expressed in Escherichia coli, the cells produced active ferredoxin-dependent soluble heme oxygenase. The open reading frame of ho1 was fused in frame with a chloroplast transit peptide-encoding sequence from the oli gene of Antirrhinum majus. This construct was placed in a binary plasmid vectorcontaining a kanamycin resistance marker and a cauliflower mosaic virus 35S promoter to control expression of the chimeric oli-ho1 gene and used to transform A. thaliana hy1 plants. Two independent transformed lines were obtained that had the phenotype of the parental Landsberg erecta line and expressed the chimeric gene, as indicated by detection of its mRNA by reverse transcriptase-polymerase chain reaction. The results indicate that Synechocystis sp. PCC 6803 heme oxygenase encoded by ho1 can substitute for the defective HY1 gene product and that the only required enzyme activity of the HY1 gene product is heme oxygenase.


Subject(s)
Bacterial Proteins/biosynthesis , Cyanobacteria/metabolism , Plant Proteins/biosynthesis , Amino Acid Sequence , Arabidopsis/genetics , Arabidopsis/metabolism , Cyanobacteria/genetics , DNA, Recombinant/genetics , Gene Expression , Genetic Complementation Test , Heme Oxygenase (Decyclizing)/genetics , Light-Harvesting Protein Complexes , Molecular Sequence Data , Mutation , Phenotype , Phytochrome/metabolism , Plasmids , RNA, Messenger/genetics , RNA, Messenger/metabolism , Sequence Alignment , Sequence Homology, Amino Acid , Transformation, Genetic
3.
J Mol Biol ; 292(4): 871-91, 1999 Oct 01.
Article in English | MEDLINE | ID: mdl-10525412

ABSTRACT

The X-ray crystal structure of Klebsiella pneumoniae nitrogenase component 1 (Kp1) has been determined and refined to a resolution of 1.6 A, the highest resolution reported for any nitrogenase structure. Models derived from three 1.6 A resolution X-ray data sets are described; two represent distinct oxidation states, whilst the third appears to be a mixture of both oxidized and reduced states (or perhaps an intermediate state). The structures of the protein and the iron-molybdenum cofactor (FeMoco) appear to be largely unaffected by the redox status, although the movement of Ser beta90 and a surface helix in the beta subunit may be of functional significance. By contrast, the 8Fe-7S P-cluster undergoes discrete conformational changes involving the movement of two iron atoms. Comparisons with known component 1 structures reveal subtle differences in the FeMoco environment, which could account for the lower midpoint potential of this cluster in Kp1. Furthermore, a non-proline- cis peptide bond has been identified in the alpha subunit that may have a functional role. It is within 10 A of the FeMoco and may have been overlooked in other component 1 models. Finally, metal-metal and metal-sulphur distances within the metal clusters agree well with values derived from EXAFS studies, although they are generally longer than the values reported for the closely related protein from Azotobacter vinelandii. A number of bonds between the clusters and their ligands are distinctly longer than the EXAFS values, in particular, those involving the molybdenum atom of the FeMoco.


Subject(s)
Klebsiella pneumoniae/enzymology , Molybdoferredoxin/chemistry , Nitrogenase/chemistry , Amino Acid Sequence , Binding Sites , Crystallization , Crystallography, X-Ray , Electrons , Iron/chemistry , Iron/metabolism , Ligands , Models, Molecular , Molecular Sequence Data , Molybdenum/chemistry , Molybdenum/metabolism , Molybdoferredoxin/metabolism , Nitrogenase/metabolism , Oxidation-Reduction , Proline/chemistry , Proline/metabolism , Protein Conformation , Structure-Activity Relationship , Sulfur/chemistry , Sulfur/metabolism
4.
Klin Monbl Augenheilkd ; 208(4): 203-11, 1996 Apr.
Article in German | MEDLINE | ID: mdl-8778488

ABSTRACT

BACKGROUND: Several methods are available for treatment of orbital floor fractures: antral balloon catheters, orbital implants, combinations of both methods, or no treatment at all. This study compares the outcome of the different surgical approaches. PATIENTS AND METHODS: In a retrospective study 153 patients with unilateral orbital floor fractures were analysed with particular emphasis on the extend of the fracture regarding to Motility, enophthalmos, globe position and loss of sensitivity as analytical parameters. RESULTS: After treatment with antral balloon catheters development of motility in the upper field of gaze was significantly better (p < 0.03) than using combinations of antral balloon catheters and implants. In addition least enophthalmos and bulbus deviation resulted from antral balloon catheters when treating small and medium size fractures. A significant difference was found in enophthalmos (p < 0.02) and vertical globe position (p < 0.01) between implantation and no treatment at all. While exophthalmos and lifting of the bulbus are more often associated with implantation, enophthalmos and vertical depression of the bulbus may result when no treatment is given. Implantation results in least loss of sensitivity of N. V/2 (49%). In contrast the combination of implantation and antral balloon catheter shows the worst outcome (65%). CONCLUSIONS: In small and medium fractures best results are achieved when using the antral balloon catheter. However high filling-pressures should be avoided and preferably anatomically shaped catheters used. Implantations are only indicated when treating major fractures. In the presence of possible prion transmission alloplastic or autologous materials should be used. The combination of antral balloon catheters and implants are only indicated in exceptional circumstances, furthermore even in small orbital floor fractures surgical treatment is always indicated.


Subject(s)
Catheterization/instrumentation , Orbital Fractures/surgery , Prostheses and Implants , Adult , Combined Modality Therapy , Enophthalmos/etiology , Eye Movements/physiology , Female , Follow-Up Studies , Fracture Healing/physiology , Humans , Male , Postoperative Complications/etiology , Retrospective Studies
5.
Arch Biochem Biophys ; 312(1): 203-9, 1994 Jul.
Article in English | MEDLINE | ID: mdl-7913310

ABSTRACT

In the biosynthetic conversion of glutamate to the tetrapyrrole precursor, delta-aminolevulinic acid (ALA), glutamate is activated at C-1 by glutamyl-tRNA synthetase-catalyzed ligation to tRNAGlu. Glutamyl-tRNA reductase next catalyzes reduction of the activated glutamate to glutamate-1-semialdehyde (GSA), which is then converted to ALA by GSA aminotransferase. Glutamyl-tRNA synthetase is known to require a divalent metal (usually Mg2+) for activity, but it has not been established whether Mg2+ or another metal ion is also required for glutamyl-tRNA reductase or GSA aminotransferase, because these enzymes have previously been assayed in combined incubations containing all factors required for conversion of glutamate to ALA. We now report the metal requirements individually for each of the three enzyme reactions. Glutamyl-tRNA reductase activity in extracts from both Chlorella vulgaris and Synechocystis sp. PCC 6803 was stimulated by Mg2+ and inhibited by EDTA. EDTA-pretreated Chlorella glutamyl-tRNA reductase-containing fraction had very little activity in the absence of added Mg2+, but recovered full activity in incubations containing added Mg2+. The divalent metal requirement could be met by Mg2+, Mn2+, or Ca2+. Maximum activity was reached at approximately 15 mM concentration of each of these metals, and higher concentrations were inhibitory. Zn2+ was inhibitory at micromolar concentrations. Chlorella glutamyl-tRNA synthetase showed a metal requirement that could be met by Mg2+ or Mn2+, but not Ca2+. Maximum activity was reached at approximately 15 mM Mg2+ or Mn2+. Although the presence of 10 mM Ca2+ did not affect the Mg2+ concentration optimum, Ca2+ increased the effectiveness of low concentrations of Mg2+. In contrast to glutamyl-tRNA synthetase and glutamyl-tRNA reductase, Chlorella GSA aminotransferase did not show a metal requirement or inhibition by EDTA. However, EDTA decreased nonenzymatic transformation of GSA to ALA.


Subject(s)
Aldehyde Oxidoreductases/drug effects , Cations, Divalent/pharmacology , Eukaryota/enzymology , Glutamate-tRNA Ligase/drug effects , Intramolecular Transferases , Isomerases/drug effects , Aminolevulinic Acid/metabolism , Cell-Free System , Chlorella/enzymology , Cyanobacteria/enzymology , Glutamates/metabolism , Glutamic Acid , Pyrroles/metabolism , Tetrapyrroles
6.
Plant Physiol ; 101(3): 1029-38, 1993 Mar.
Article in English | MEDLINE | ID: mdl-7906043

ABSTRACT

delta-Aminolevulinic acid (ALA), the universal biosynthetic precursor of tetrapyrrole pigments, is synthesized from glutamate in plants, algae, and many bacteria via a three-step process that begins with activation by ligation of glutamate to tRNA(Glu), followed by reduction to glutamate-1-semialdehyde (GSA) and conversion of GSA to ALA. The GSA aminotransferase step requires no substrate other than GSA. A previous study examined whether the aminotransferase reaction proceeds via intramolecular or intermolecular N transfer and concluded that the reaction catalyzed by Chlamydomonas extracts occurs via intermolecular N transfer (Y.-H.L. Mau and W.-Y. Wang [1988] Plant Physiol 86: 793-797). However, in that study the possibility was not excluded that the result was a consequence of N exchange among product ALA molecules during the incubation, rather than intermolecular N transfer during the conversion of GSA to ALA. Therefore, this question was reexamined in another species and with additional controls. A gel-filtered extract of Chlorella vulgaris cells was incubated with ATP, Mg2+, NADPH, tRNA, and a mixture of L-glutamate molecules, one-half of which were labeled with 15N and the other half with 13C at C-1. The ALA product was purified, derivatized, and analyzed by gas chromatography-mass spectrometry. A significant fraction of the ALA molecules was heavy by two mass units, indicating incorporation of both 15N and 13C. These results show that the N and C atoms of each ALA molecule were derived from different glutamate molecules. Control experiments indicated that the results could not be attributed to exchange of N atoms between glutamate or ALA molecules during the incubation. These results confirm the earlier conclusion that GSA is converted to ALA via intermolecular N transfer and extend the results to another species. The labeling results, combined with the results of kinetic and inhibitor studies, support a model for the GSA aminotransferase reaction in which a single molecule of GSA is converted to ALA via an enzyme-bound 4,5-diaminovaleric acid intermediate.


Subject(s)
Aminolevulinic Acid/metabolism , Chlorella/metabolism , Glutamates/metabolism , Nitrogen/metabolism , Amination , Chlorella/enzymology , Gas Chromatography-Mass Spectrometry , Glutamic Acid , Kinetics , Pyridoxal Phosphate/metabolism
7.
Plant Physiol ; 101(2): 657-665, 1993 Feb.
Article in English | MEDLINE | ID: mdl-12231722

ABSTRACT

In plants, algae, and many bacteria, the heme and chlorophyll precursor, [delta]-aminolevulinic acid (ALA), is synthesized from glutamate in a reaction involving a glutamyl-tRNA intermediate and requiring ATP and NADPH as cofactors. In particulate-free extracts of algae and chloroplasts, ALA synthesis is inhibited by heme. Inclusion of 1.0 mM glutathione (GSH) in an enzyme and tRNA extract, derived from the green alga Chlorella vulgaris, lowered the concentration of heme required for 50% inhibition approximately 10-fold. The effect of GSH could not be duplicated with other reduced sulfhydryl compounds, including mercaptoethanol, dithiothreitol, and cysteine, or with imidazole or bovine serum albumin, which bind to heme and dissociate heme dimers. Absorption spectroscopy indicated that heme was fully reduced in incubation medium containing dithiothreitol, and addition of GSH did not alter the heme reduction state. Oxidized GSH was as effective in enhancing heme inhibition as the reduced form. Co-protoporphyrin IX inhibited ALA synthesis nearly as effectively as heme, and 1.0 mM GSH lowered the concentration required for 50% inhibition approximately 10-fold. Because GSH did not influence the reduction state of heme in the incubation medium, and because GSH could not be replaced by other reduced sulfhydryl compounds or ascorbate, the effect of GSH cannot be explained by action as a sulfhydryl protectant or heme reductant. Preincubation of enzyme extract with GSH, followed by rapid gel filtration, could not substitute for inclusion of GSH with heme during the reaction. The results suggest that GSH must specifically interact with the enzyme extract in the presence of the inhibitor to enhance the inhibition.

8.
Plant Physiol ; 99(2): 482-7, 1992 Jun.
Article in English | MEDLINE | ID: mdl-16668911

ABSTRACT

Euglena gracilis cells synthesize the key tetrapyrrole precursor, delta-aminolevulinic acid (ALA), by two routes: plastid ALA is formed from glutamate via the transfer RNA-dependent five-carbon route, and ALA that serves as the precursor to mitochondrial hemes is formed by ALA synthase-catalyzed condensation of succinyl-coenzyme A and glycine. The biosynthetic source of succinyl-coenzyme A in Euglena is of interest because this species has been reported not to contain alpha-ketoglutarate dehydrogenase and not to use succinyl-coenzyme A as a tricarboxylic acid cycle intermediate. Instead, alpha-ketoglutarate is decarboxylated to form succinic semialdehyde, which is subsequently oxidized to form succinate. Desalted extract of Euglena cells catalyzed ALA formation in a reaction that required coenzyme A and GTP but did not require exogenous succinyl-coenzyme A synthetase. GTP could be replaced with ATP. Cell extract also catalyzed glycine-and alpha-ketoglutarate-dependent ALA formation in a reaction that required coenzyme A and GTP, was stimulated by NADP(+), and was inhibited by NAD(+). Succinyl-coenzyme A synthetase activity was detected in extracts of dark- and light-grown wild-type and nongreening mutant cells. In vitro succinyl-coenzyme A synthetase activity was at least 10-fold greater than ALA synthase activity. These results indicate that succinyl-coenzyme A synthetase is present in Euglena cells. Even though the enzyme may play no role in the transformation of alpha-ketoglutarate to succinate in the atypical tricarboxylic acid cycle, it catalyzes succinyl-coenzyme A formation from succinate for use in the biosynthesis of ALA and possibly other products.

9.
Plant Physiol ; 97(3): 1094-102, 1991 Nov.
Article in English | MEDLINE | ID: mdl-16668494

ABSTRACT

Wild-type Euglena gracillis cells synthesize the key chlorophyll precursor, delta-aminolevulinic acid (ALA), from glutamate in their plastids. The synthesis requires transfer RNA(Glu) (tRNA(Glu)) and the three enzymes, glutamyl-tRNA synthetase, glutamyl-tRNA reductase, and glutamate-1-semialdehyde aminotransferase. Non-greening mutant Euglena strain W(14)ZNaIL does not synthesize ALA from glutamate and is devoid of the required tRNA(Glu). Other cellular tRNA(Glu)s present in the mutant cells were capable of being charged with glutamate, but the resulting glutamyl-tRNAs did not support ALA synthesis. Surprisingly, the mutant cells contain all three of the enzymes, and their cell extracts can convert glutamate to ALA when supplemented with tRNA(Glu) obtained from wild-type cells. Activity levels of the three enzymes were measured in extracts of cells grown under a number of light conditions. All three activities were diminished in extracts of cells grown in complete darkness, and full induction of activity required 72 hours of growth in the light. A light intensity of 4 microeinsteins per square meter per second was sufficient for full induction. Blue light was as effective as white light, but red light was ineffective, in inducing extractable enzyme activity above that of cells grown in complete darkness, indicating that the light control operates via the nonchloroplast blue light receptor in the mutant cells. Of the three enzyme activities, the one that is most acutely affected by light is glutamate-1-semialdehyde aminotransferase, as has been previously shown for wild-type Euglena cells. These results indicate that the enzymes required for ALA synthesis from glutamate are present in an active form in the nongreening mutant cells, even though they cannot participate in ALA formation in these cells because of the absence of the required tRNA(Glu), and that the activity of all three enzymes is regulated by light. Because the absence of plastid tRNA(Glu) precludes the synthesis of proteins within the plastids, the three enzymes must be synthesized in the cytoplasm and their genes encoded in the nucleus in Euglena.

10.
Plant Physiol ; 94(3): 1365-75, 1990 Nov.
Article in English | MEDLINE | ID: mdl-16667842

ABSTRACT

Chlorophyll synthesis in Euglena, as in higher plants, occurs only in the light. The key chlorophyll precursor, delta-aminolevulinic acid (ALA), is formed in Euglena, as in plants, from glutamate in a reaction sequence catalyzed by three enzymes and requiring tRNA(Glu). ALA formation from glutamate occurs in extracts of light-grown Euglena cells, but activity is very low in dark-grown cell extracts. Cells grown in either red (650-700 nanometers) or blue (400-480 nanometers) light yielded in vitro activity, but neither red nor blue light alone induced activity as high as that induced by white light or red and blue light together, at equal total fluence rates. Levels of the individual enzymes and the required tRNA were measured in cell extracts of light- and dark-grown cells. tRNA capable of being charged with glutamate was approximately equally abundant in extracts of light- and dark-grown cells. tRNA capable of supporting ALA synthesis was approximately three times more abundant in extracts of light-grown cells than in dark-grown cell extracts. Total glutamyl-tRNA synthetase activity was nearly twice as high in extracts of light-grown cells as in dark-grown cell extracts. However, extracts of both light- and dark-grown cells were able to charge tRNA(Glu) isolated from light-grown cells to form glutamyl-tRNA that could function as substrate for ALA synthesis. Glutamyl-tRNA reductase, which catalyzes pyridine nucleotide-dependent reduction of glutamyl-tRNA to glutamate-1-semialdehyde (GSA), was approximately fourfold greater in extracts of light-grown cells than in dark-grown cell extracts. GSA aminotransferase activity was detectable only in extracts of light-grown cells. These results indicate that both the tRNA and enzymes required for ALA synthesis from glutamate are regulated by light in Euglena. The results further suggest that ALA formation from glutamate in dark-grown Euglena cells may be limited by the absence of GSA aminotransferase activity.

11.
J Biol Chem ; 262(26): 12541-9, 1987 Sep 15.
Article in English | MEDLINE | ID: mdl-2442164

ABSTRACT

Glutamate was converted to the chlorophyll and heme precursor delta-aminolevulinic acid in soluble extracts of Euglena gracilis. delta-Aminolevulinic acid-forming activity depended on the presence of native enzyme, glutamate, ATP, Mg2+, NADPH or NADH, and RNA. The requirement for reduced pyridine nucleotide was observed only if, prior to incubation, the enzyme extract was filtered through activated carbon to remove firmly bound reductant. Dithiothreitol was also required for activity after carbon treatment. delta-Aminolevulinic acid formation was stimulated by RNA from various plant tissues and algal cells, including greening barley leaves and members of the algal groups Chlorophyta (Chlorella vulgaris, Chlamydomonas reinhardtii), Rhodophyta (Cyanidium caldarium), Cyanophyta (Anacystis nidulans, Synechocystis sp. PCC 6803), and Prochlorophyta (Prochlorothrix hollandica), but not by RNA derived from Escherichia coli, yeast, wheat germ, bovine liver, and Methanobacterium thermoautotrophicum. E. coli glutamate-specific tRNA was inhibitory. Several of the RNAs that did not stimulate delta-aminolevulinic acid formation nevertheless became acylated when incubated with glutamate in the presence of Euglena enzyme extract. RNA extracted from nongreen dark-grown wild-type Euglena cells was about half as stimulatory as that from chlorophyllous light-grown cells, and RNA from aplastidic mutant cells stimulated only slightly. delta-Aminolevulinic acid-forming enzyme activity was present in extracts of light-grown wild-type cells, but undetectable in extracts of aplastidic mutant and dark-grown wild-type cells. Gabaculine inhibited delta-aminolevulinic acid formation at submicromolar concentration. Heme inhibited 50% at 25 microM, but protoporphyrin IX, Mg-protoporphyrin IX, and protochlorophyllide inhibited only slightly at this concentration.


Subject(s)
Aminolevulinic Acid/metabolism , Euglena gracilis/metabolism , Glutamates/metabolism , Levulinic Acids/metabolism , Adenosine Triphosphate/metabolism , Escherichia coli/metabolism , Euglena gracilis/drug effects , Euglena gracilis/radiation effects , Glutamic Acid , Hydrogen-Ion Concentration , Light , Magnesium/metabolism , NADP/metabolism , RNA/pharmacology , RNA, Bacterial/metabolism , RNA, Transfer, Amino Acyl/metabolism , Temperature
12.
Plant Physiol ; 84(2): 244-50, 1987 Jun.
Article in English | MEDLINE | ID: mdl-16665424

ABSTRACT

Extracts from plant chloroplasts and algae catalyze the conversion of glutamate to delta-aminolevulinic acid (ALA) in the first committed step of the tetrapyrrole biosynthetic pathway leading to chlorophylls, hemes, and bilins. The conversion requires ATP, Mg(2+), and NADPH as cofactors. Soluble extracts from Chlorella vulgaris have now been resolved into four macromolecular fractions, all of which are required to reconstitute activity. One fraction contains a low molecular weight RNA which can be separated from the protein components in an active high-speed supernatant by treatment with 1 molar NaCl followed by precipitation of the proteins with (NH(4))(2)SO(4) at 70% saturation. The proteins recovered from the (NH(4))(2)SO(4) precipitate are reactivated by addition of a fraction containing tRNAs isolated from Chlorella by phenol-chloroform extraction and DEAE cellulose chromatography. Three required protein fractions were resolved from the RNA-depleted (NH(4))(2)SO(4) precipitate by serial affinity chromatography on Reactive Blue 2-Sepharose and 2',5'-ADP-agarose. Glycerol was found to stabilize the enzyme activity during the separation process. The majority of the glutamate:tRNA ligase activity was associated with the fraction which was retained by Blue-Sepharose and not retained by ADP-agarose, in agreement with the reported properties of the affinity ligands. The active material in the fraction not retained by Blue-Sepharose eluted as a single component on gel filtration chromatography, with an apparent molecular weight of 67,000. The active component in the RNA fraction also eluted as a single component on gel filtration chromatography.

13.
Plant Physiol ; 82(4): 1096-101, 1986 Dec.
Article in English | MEDLINE | ID: mdl-16665140

ABSTRACT

Formation of the chlorophyll and heme precursor delta-aminolevulinic acid (ALA) from glutamate in soluble extracts of Chlorella vulgaris, Euglena gracilis, and Cyanidium caldarium was stimulated by addition of low molecular weight RNA derived from greening algae or plant tissue. Enzyme extracts were prepared for the ALA formation assay by high-speed centrifugation, partial RNA depletion, and gel filtration through Sephadex G-25. RNA was extracted from greening barley epicotyls, greening cucumber cotyledon chloroplasts, and growing cells of Chlorella, Euglena, Chlamydomonas reinhardtii, and Anacystis nidulans, freed of protein, and fractionated on DEAE-cellulose to yield an active component corresponding to the tRNA-containing fraction. RNA from homologous and heterologous species stimulated ALA formation when added to enzyme extracts, and the degree of stimulation was proportional to the amount of RNA added. Algal enzyme extracts were stimulated by algal RNAs interchangeably, with the exception of RNA prepared from aplastidic Euglena, which did not stimulate ALA production. RNA from greening cucumber cotyledon chloroplasts and greening barley epicotyls stimulated ALA formation in algal enzyme incubations. In contrast, tRNA from Escherichia coli, both nonspecific and glutamate-specific, as well as wheat germ, bovine liver, and yeast tRNA, failed to reconstitute ALA formation. Moreover, E. coli tRNA inhibited ALA formation by algal extracts, both in the presence and absence of added algal RNA. Chlorella extracts were capable of catalyzing aminoacyl bond formation between glutamate and both the activity reconstituting and nonreconstituting RNAs, indicating that the inability of some RNAs to stimulate ALA formation was not due to their inability to serve as glutamyl acceptors. The first step in the ALA-forming reaction sequence has been proposed to be activation of glutamate via aminoacyl bond formation with a specific tRNA, analogous to the first step in peptide bond formation. Our results suggest that the RNA that is required for ALA formation may be functionally distinct from the glutamyl-tRNA species involved in protein synthesis.

14.
Cancer Chemother Rep ; 59(5): 929-33, 1975.
Article in English | MEDLINE | ID: mdl-1203897

ABSTRACT

A comparison has been made of the effects of three nitrosoureas (BCNU, CCNU, and methyl-CCNU) on the cells of Lewis lung carcinoma and B16 melanoma in vivo using a new in vitro method for assaying colony-forming cells. In addition, the effect on early hemopoietic precursors has been studied using the in vivo spleen colony assay and the in vitro agar colony assay. The results show that the three nitrosoureas are selective against Lewis lung carcinoma and B16 melanoma relative to normal hemopoietic precursors. Furthermore, the effect of these nitrosoureas is unchanged when the rate of proliferation of the hemopoietic precursor cells is increased.


Subject(s)
Bone Marrow Cells , Bone Marrow/drug effects , Hematopoiesis/drug effects , Lung Neoplasms/pathology , Melanoma/pathology , Nitrosourea Compounds/pharmacology , Animals , Carmustine/pharmacology , Cell Survival/drug effects , Cell Survival/radiation effects , Clone Cells/drug effects , Gamma Rays , Lomustine/pharmacology , Mice , Neoplasms, Experimental/pathology , Radiation Effects , Semustine/pharmacology , Spleen/cytology
15.
Ment Retard ; 6(5): 53, 1968 Oct.
Article in English | MEDLINE | ID: mdl-5725066
SELECTION OF CITATIONS
SEARCH DETAIL