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.
When seven crop species were grown under identical environmental conditions, decreased sink:source ratio led to a decreased photosynthetic rate within 1 to 3 days in Cucumis sativus L., Gossypium hirsutum L., and Raphanus sativus L., but not in Capsicum annuum L., Solanum melongena L., Phaseolus vulgaris L., or Ricinus communis L. The decrease was not associated with stomatal closure. In cotton and cucumber, sink removal led to an increase in starch and sugar content, in glucose 6-phosphate and fructose 6-phosphate pools, and in the proportion of 14C detected in sugar phosphates and UDPglucose following 14CO2 supply. When mannose was supplied to leaf discs to sequester cytoplasmic inorganic phosphate, promotion of starch synthesis, and inhibition of CO2 fixation, were observed in control discs, but not in discs from treated plants. Phosphate buffer reduced starch synthesis in the latter, but not the former discs. The findings suggest that sink removal led to a decreased ratio inorganic phosphate:phosphorylated compounds. In beans 14C in sugar phosphates increased following sink removal, but without sucrose accumulation, suggesting tighter feedback control of sugar level. Starch accumulated to higher levels than in the other plants, but CO2 fixation rate was constant for several days.  相似文献   

2.
《Insect Biochemistry》1989,19(3):323-326
High resolution 31P nuclear magnetic resonance spectroscopy (NMR) was successfully applied to 5th instar larvae of Manduca sexta. Conditions for in vivo analysis under non-saturating conditions are described. The 31P NMR spectrum of intact larvae was composed of six peaks. Their resonance frequencies are reported relative to orthophosphoric acid. Analysis of tissue extracts demonstrated the in vivo peaks to be composed of the β phosphorus resonance of nucleotide triphosphates (NTP) at −19.36 ppm; α phosphorus of NTP and nucleotide diphosphates (NDP) at −10.51 ppm; β and γ phosphorus of NDP and NTP, respectively, at −5.42 ppm; phosphoarginine (PA) at −3.45 ppm; inorganic phosphate (Pi) at +2.76 ppm and sugar phosphates at +3.34 ppm. The major sugar phosphate present in fat body extracts was trehalose-6-phosphate and this was the major phosphorus component of the spectrum of hemolymph. The spin-lattice relaxation times for each in vivo peak were determined.Titration of aqueous fat body and hemolymph extracts was carried out and the relationship between the chemical shift of Pi and pH determined. On this basis the pH of the hemolymph was estimated at approx. 6.7.The metabolic inhibitors, iodoacetate and dinitrophenol, had significant effects on the 31P NMR spectrum of intact larvae. Administration of iodoacetate caused a rapid increase in the levels of sugar phosphates together with decreases in NTP and PA. Dinitrophenol also caused declines in the relative levels of NTP and PA but sugar phosphates decreased as well. The experiments demonstrated the potential of in vivo NMR analysis for metabolic studies on high energy phosphate metabolites in M. sexta.  相似文献   

3.
It has been well established that sugars can be used to stabilize liposomes during drying by a mechanism that involves the formation of a glassy state by the sugars as well as by a direct interaction between the sugar and the phospholipid head groups. We have investigated the protective effect of phosphate on solute retention and storage stability of egg phosphatidylcholine (egg PC) liposomes that were dried (air-dried and freeze-dried) in the presence of sugars and phosphate. The protective effect of phosphate was tested using both glucose (low Tg) and sucrose (high Tg) by measuring leakage of carboxyfluorescein (CF), which was incorporated inside the vesicles. Liposomes that were dried with glucose or phosphate alone showed complete leakage after rehydration. However, approximately 30% CF-retention was obtained using mixtures of phosphate and glucose. Approximately 75% CF-retention was observed with liposomes that were dried with sucrose. The solute retention further increased to 85% using mixtures of phosphate and sucrose. The pH of the phosphate buffer prior to drying was found to have a strong effect on the solute retention. Fourier transform infrared spectroscopy studies showed that phosphate and sugars form a strong hydrogen bonding network, which dramatically increased the Tg. The HPO42− form of phosphate was found to interact stronger with sugars than the H2PO4 form. The increased solute retention of liposomes dried in the sugar phosphate mixtures did not coincide with improved storage stability. At temperatures below 60 °C the rate of solute-leakage was found to be strikingly higher in the presence of phosphate, indicating that phosphate impairs storage stability of dried liposomes.  相似文献   

4.
The structures of the cell wall teichoic acids (TA) from some species of the genus Nocardiopsis were established by chemical and NMR spectroscopic methods. The cell walls of Nocardiopsis synnemataformans VKM Ac-2518T and Nocardiopsis halotolerans VKM Ac-2519T both contain two TA with unique structures—poly(polyol phosphate-glycosylpolyol phosphate)—belonging to the type IV TA. In both organisms, the minor TA have identical structures: poly(glycerol phosphate-N-acetyl-β-galactosaminylglycerol phosphate) with the phosphodiester bond between C-3 of glycerol and C-4 of the amino sugar. This structure is found for the first time. The major TA of N. halotolerans has a hitherto unknown structure: poly(glycerol phosphate-N-acetyl-β-galactosaminylglycerol phosphate), the N-acetyl-β-galactosamine being acetalated with pyruvic acid at positions 4 and 6. The major TA of N. synnemataformans is a poly(glycerol phosphate-N-acetyl-β-galactosaminylglycerol phosphate) with the phosphodiester bond between C-3 of glycerol and C-3 of the amino sugar. The cell walls of Nocardiopsis composta VKM Ac-2520 and N. composta VKM Ac-2521T contain only one TA, namely 1,3-poly(glycerol phosphate) partially substituted with N-acetyl-α-glucosamine. The cell wall of Nocardiopsis metallicus VKM Ac-2522T contains two TA. The major TA is 1,5-poly(ribitol phosphate), each ribitol unit carrying a pyruvate ketal group at positions 2 and 4. The structure of the minor TA is the same as that of N. composta. The results presented correlate well with the phylogenetic grouping of strains and confirm the species and strain specific features of cell wall TA in members of the genus Nocardiopsis.  相似文献   

5.
The biological and chemical basis of vanadium action in fungi is relatively poorly understood. In the present study, we investigate the influence of vanadate (V5+) on phosphate metabolism of Phycomyces blakesleeanus. Addition of V5+ caused increase of sugar phosphates signal intensities in 31P NMR spectra in vivo. HPLC analysis of mycelial phosphate extracts demonstrated increased concentrations of glucose 6 phosphate, fructose 6 phosphate, fructose 1, 6 phosphate and glucose 1 phosphate after V5+ treatment. Influence of V5+ on the levels of fructose 2, 6 phosphate, glucosamine 6 phosphate and glucose 1, 6 phosphate (HPLC), and polyphosphates, UDPG and ATP (31P NMR) was also established. Increase of sugar phosphates content was not observed after addition of vanadyl (V4+), indicating that only vanadate influences its metabolism. Obtained results from in vivo experiments indicate catalytic/inhibitory vanadate action on enzymes involved in reactions of glycolysis and glycogenesis i.e., phosphoglucomutase, phosphofructokinase and glycogen phosphorylase in filamentous fungi.  相似文献   

6.
THEORIES of the molecular structure of nucleic acids have so far been based on evidence from the crystal structures of monomeric units such as nucleosides and mononucleotides, the interpretation of diffraction patterns of oriented nucleic acid fibres and molecular model building1–6. Such approaches can help to suggest structures of periodic molecules such as helices, but they are insufficient for predicting and understanding nonrepetitive structures such as the loops in transfer RNA (tRNA), presumably associated with many of the functions of tRNA. To understand the geometry of nucleic acids and possible constraints on their conformation, it is therefore essential to know the detailed conformation of the sugar residues and the conformational relationship between the sugar residue, the base and the phosphate group7–9. The simplest molecule which contains this information is a 3´5´-dinucleoside phosphate. We now report the structure of uridine-3´,5´-adenosine phosphate (UpA). This is the first naturally occurring dinucleoside phosphate whose crystal structure has been determined by X-ray diffraction. The only other dinucleoside phosphate with known crystal structure is adenosine-2´,5´-uridine phosphate10, but it does not have the naturally occurring 3´5´ sugar phosphate linkage.  相似文献   

7.
《Inorganica chimica acta》1987,135(3):207-210
An FT-IR spectroscopic study concerning changes in the conformation of sugar in the dinucleotides; GpC and CpG, on platination and intercalation is presented. The results are compared with the FT-IR spectral data of 5′-CMP, 5′-GMP, 3′-GMP and their metal adducts. The spectra of free GpC, free CpG, proflavine-GpC, proflavine-CpG, and cis-[Pt(NH3)2(GpC)2]2+ exhibit the diagnostic band at 800 cm−1 which was assigned to a sugar phosphate vibrational mode and diagnostic of C3′-endo sugar pucker. In the case of 9-aminoacridine-GpC and cis-[Pt(NH3)2(CpG]+ the diagnostic bands of the C2′-endo and C3′-endo conformations are observed at 810–820 cm−1 and near 800 cm−1 respectively. The results are in good agreement with X-ray data. The infrared diagnostic bands are important for distinguishing the sugar pucker conformational changes.  相似文献   

8.
《Insect Biochemistry》1988,18(6):531-538
Studies were made on 13C and 31P NMR in larvae of two species of silkworm, Bombyx mori and Philosamia cynthia ricini, in vivo as well as in vitro to determine the pathways of glucose utilization, especially those to amino acids as components of silk fibroin. Results showed that the 13C of [1-13C]glucose administered orally into 5th instar larvae of both species was incorporated into glucose-1-phosphate, glucose-6-phosphate and trehalose. Serine, glutamate, glutamine, citrate, malate, trehalose and sorbitol-6-phosphate were detected in the hemolymphs of these larvae as metabolites of [1-13C]glucose. Two days after [1-13C]glucose administration, labeled alanine, glycine, serine, urea, glycogen, trehalose and glycerol were clearly detected in Bombyx larvae. Starvation caused rapid consumption of administered [1-13C]glucose with very little accumulation of 13C in glycogen or trehalose. In the in vivo31P NMR spectra of Bombyx larvae, ATP, arginine phosphate, sorbitol-6-phosphate, uridine diphosphoglucose, phosphoenolpyruvate and inorganic phosphate were detected with some sugar phosphates, such as glucose-1-phosphate and glucose-6-phosphate. During starvation, the intensity of the signal of inorganic phosphate increased and those of sugar phosphate other than sorbitol-6-phosphate decreased, but these changes were reversed by oral administration of glucose.  相似文献   

9.
The extent of phosphate uptake measured by the relative changes in cytoplasmic Pi, vacuolar Pi, ATP, glucose-6-phosphate, and UDPG was determined using in vivo31P nuclear magnetic resonance spectroscopy. Maize (Zea mays) root tips were perfused with a solution containing 0.5 or 1.0 millimolar phosphate at pH ~6.5 under different conditions. In the aerated state, phosphate uptake resulted in a significant increase (>80%) in vacuolar Pi, but cytoplasmic Pi only transiently increased by 10%. Under N2, the cytoplasmic Pi increased ~150% which could be attributed to a large extent to the breakdown of ATP, sugar phosphates and UDPG. Vacuolar Pi increased but only to the extent of ~10% of that seen under aerobic conditions. 2-deoxyglucose pretreatment was utilized to decrease the level of cytoplasmic Pi. When pretreated with the 2-deoxyglucose, the excised maize roots absorbed phosphate from the perfusate with a significant increase in the cytoplasmic Pi. The increase could only be traced to external phosphate since the concentrations of other phosphorus containing species remained constant during the uptake period. With 2-deoxyglucose pretreatment, phosphate uptake under anaerobic conditions was substantially inhibited with only the vacuolar phosphate showing a slight increase. When roots were treated with carbonyl cyanide m-chlorophenyl hydrazone, no detectable Pi uptake was found. These results were used to propose a H+-ATPase related transport mechanism for phosphate uptake and compartmentation in corn root cells.  相似文献   

10.
The interaction between the plant hormone, 3-indoleacetic acid (IAA), and some phospholipids in CDCL3 has been studied by 1H, 13C and 31P nuclear magnetic resonance (NMR) spectroscopy. Upon interaction with IAA, significant changes occurred in resonance positions of the phospholipid head group nuclei. Alteration of the fatty acid composition influenced the effects of IAA on these nuclei. These effects were observed in the ethanolamine and phosphate groups of the phosphatidylethanolamines, and in the choline, phosphate and glycerol groups of the phosphatidylcholines. Changes in resonance positions of the phospholipid head group nuclei were used for the determination of dissociation constants (Kd). In all cases, Kd values were approx. 10?2 molal for 1 : 1 interaction. The NMR results suggest an interaction orientation in which the aromatic ring system of IAA interacts with the quaternary nitrogen function of the head group, and the phosphate group becomes hydrogen-bonded to the NH or carboxyl proton of 1AA.  相似文献   

11.
Two kinds of conformational variability have been reported for left-handed Z-DNA: the Z to Z′ transition, which involves a change in guanine sugar pucker from C-3′-endo to C-1′-exo, and the ZIto ZII transition, which corresponds to a simple three-atom phosphate-group rotation. Neither of these motions substantially affects base stacking or helical twist, and this is because the degree of independent motion of phosphate groups, sugar molecules and base-pairs is greater in the left-handed Z helix than in right-handed B-DNA. Detailed considerations of Z helix geometry suggest that ZI, ZIIand Z′ are not separate species, but only samplings of the full range of conformation open to Z-DNA.  相似文献   

12.
Abstract

Molecular mechanics studies are performed on single stranded as well as base paired forms of dinucleoside methylphosphonates comprising different base sequences for both the Sand R-isomers of methylphosphonate (MP). S-MP produces noticeable distortions in the geometry, locally at the phosphate center, and this enables the stereochemical feasibility of compact g? g? phosphodiester. Besides, it tends to perturb the conformations around the P- 03′ and glycosyl bonds, causing minor variations in stacking interactions. In single stranded dinucleosides, the gain in adjacent base stacking interaction energies seems to be sufficient to overcome the barrier to P-03′ bond rotation arising due to S-MP…sugar interaction, and this results in transition to a compact phosphodiester (BI-type) from an initial extended phosphodiester (BII-type) conformation. Such a thing seems rather difficult in base pair constrained duplexes. Dinucleosides with R-MP behave analogous to normal phosphate duplexes as the methyl group is away from the sugar. It is found that dinucleoside methylphosphonates are energetically less favoured than the corresponding dinucleoside phosphates mainly due to the depletion of contributions from electrostatic attractive interactions involving the base and sugar with the methylphosphonate consequent to the nonionic nature of the latter. Neither S-MP nor R-MP seem to significantly alter the stereochemistry of duplex structure.  相似文献   

13.
A convenient method for the synthesis of N2-dimethylaminomethylene-2′-O-methylguanosine (1), which is a useful intermediate for oligonucleotide construction, was developed. We chose the di-tert-butylsilyl group and the triisopropylbenzenesulfonyl group as sugar and base protecting groups, respectively. These protecting groups were stable during the 2′-O-methylation step with MeI and NaH. Our six-step synthesis of 1 is easy to perform using commercially available reagents, and requires only three chromatographic purifications. Compound 1 was obtained in 56% yield from guanosine.  相似文献   

14.
Supplementation with phytase is an effective way to increase the availability of phosphorus in seed-based animal feed. The biochemical characteristics of an ideal phytase for this application are still largely unknown. To extend the biochemical characterization of wild-type phytases, the catalytic properties of a series of fungal phytases, as well as Escherichia coli phytase, were determined. The specific activities of the fungal phytases at 37°C ranged from 23 to 196 U · (mg of protein)−1, and the pH optima ranged from 2.5 to 7.0. When excess phytase was used, all of the phytases were able to release five phosphate groups of phytic acid (myo-inositol hexakisphosphate), which left myo-inositol 2-monophosphate as the end product. A combination consisting of a phytase and Aspergillus niger pH 2.5 acid phosphatase was able to liberate all six phosphate groups. When substrate specificity was examined, the A. niger, Aspergillus terreus, and E. coli phytases were rather specific for phytic acid. On the other hand, the Aspergillus fumigatus, Emericella nidulans, and Myceliophthora thermophila phytases exhibited considerable activity with a broad range of phosphate compounds, including phenyl phosphate, p-nitrophenyl phosphate, sugar phosphates, α- and β-glycerophosphates, phosphoenolpyruvate, 3-phosphoglycerate, ADP, and ATP. Both phosphate liberation kinetics and a time course experiment in which high-performance liquid chromatography separation of the degradation intermediates was used showed that all of the myo-inositol phosphates from the hexakisphosphate to the bisphosphate were efficiently cleaved by A. fumigatus phytase. In contrast, phosphate liberation by A. niger or A. terreus phytase decreased with incubation time, and the myo-inositol tris- and bisphosphates accumulated, suggesting that these compounds are worse substrates than phytic acid is. To test whether broad substrate specificity may be advantageous for feed application, phosphate liberation kinetics were studied in vitro by using feed suspensions supplemented with 250 or 500 U of either A. fumigatus phytase or A. niger phytase (Natuphos) per kg of feed. Initially, phosphate liberation was linear and identical for the two phytases, but considerably more phosphate was liberated by the A. fumigatus phytase than by the A. niger phytase at later stages of incubation.  相似文献   

15.
Abstract

The mode of base-base stacking, the handedness and the sugar(dGpA)phosphate backbone conformation of deoxyguanosyl 3′-5′ deoxyadenosine and its diastereomeric ethyl phosphotriester analogs were studied by 1H NMR, UV and CD spectroscopy. The results indicate the three dimers are left-handed, while the sugar phosphate backbone is comprised predominantly of C2′-endo, gg (C4′-C5′) and g′g′ (C5′-O) conformers. The two bases are extensively stacked and interact about 90° along the dyad axes. The extent of base overlap in dGpA is slightly greater than in either ethyl phosphotriester analog. The absolute configurations of the two ethyl phosphotriester diastereoisomers of dGpA can be assigned by one-dimensional and two-dimensional 1H NMR nuclear Overhauser enhancement experiments.  相似文献   

16.
1. Changes in the amounts of mucus secreted by the epidermis of the slug in response to an irritant chemical substance (NaCl) and the effect of mechanical stimulation on high-energy phosphorylation metabolism were studied.2. 31P signals obtained from the slug with in vivo31P nuclear magnetic resonance (NMR) spectroscopy included phosphonate, sugar phosphate (SP), inorganic phosphate (Pi), arginine phosphate (Arg-P), and ATP (γ, α and β).3. When the slug epidermis came into contact with NaCl or was exposed to mechanical stimulation, the concentration of Arg-P decreased and that of Pi increased along with mucus secretion.4. The change of 31P signals in response to chemical stimulation was much more variable than that in response to mechanical stimulation.5. The β-ATP concentration was dependent on the amount of mucus secreted.  相似文献   

17.
Recent studies from this laboratory have suggested that rat-liver Golgi apparatus derived membranes contain different proteins which can translocate in vitro CMP-N-acetylneuraminic acid, GDP-fucose and adenosine 3′-phosphate 5′-phosphosulfate from an external compartment into a lumenal one. The aim of this study was to define the role of the nucleotide, sugar and sulfate moieties of sugar nucleotides and adenosine 3′-phosphate 5′-phosphosulfate in translocation of these latter compounds across Golgi vesicle membranes. Indirect evidence was obtained suggesting that the nucleotide (but not sugar or sulfate) is a necessary recognition feature for binding to the Golgi membrane (measured as inhibition of translocation) but is not sufficient for overall translocation; this latter event also depends on the type of sugar. Important recognition features for inhibition of translocation of the above sugar nucleotides and adenosine 3′-phosphate 5′-phosphosulfate were found to be the type of nucleotide base (purine or pyrimidine) and the position of the phosphate group in the ribose. Thus, UMP and CMP were found to be competitive inhibitors of translocation of CMP-N-acetylneuraminic acid, while AMP did not inhibit. Structural features of the nucleotides which were less important in inhibition of translocation (and thus presumably in binding) of the above sugar nucleotides and adenosine 3′-phosphate 5′-phosphosulfate were the number of phosphate groups in the nucleotide (CDP and CMP inhibited to a similar extent), the presence of ribose or deoxyribose in the nucleotide, a replacement of hydrogen in positions 5 of pyrimidines or 8 in purines by halogens or an azido group. The sugar or sulfate did not inhibit translocation of the above sugar nucleotides and adenosine 3′-phosphate 5′-phosphosulfate into Golgi vesicles and therefore appear not to be involved in their binding to the Golgi membrane.  相似文献   

18.
To better understand the mechanism of sugar unloading and sugar concentration in hexose- and sucrose-accumulating tomato fruits (Lycopersicon chmielewskii and L. esculentum, respectively) and to determine the causes of the late accumulation of sucrose present in sucrose-accumulating tomato fruits, the assimilation of [3H](fructosyl)-sucrose was studied. Key enzymes involved in carbohydrate metabolism were also assayed. The results demonstrated that the low level of sucrose present in young fruits accumulates directly without undergoing hydrolysis, suggesting a symplastic pathway for sucrose unloading. By contrast, the large quantity of the sucrose present in ripe sucrose-accumulating fruits originates from hydrolysis and resynthesis, suggesting an apoplastic pathway for sucrose unloading. The increase in sucrose level observed in sucrose-accumulating fruits is associated with a gradual decline in invertase activity and an increase in sucrose phosphate synthase activity. This latter enzyme seems to play a key biochemical role in the accumulation of sucrose and the establishment of a high sugar content in tomato fruits.  相似文献   

19.
Effects of inoculum size and total sugar content on both l-phenylalanine productivity and titre have been investigated using a tyrosine auxotrophic regulatory mutant of Escherichia coli. Fermentations were carried out in a 500 litre pilot fermenter with intermittent feeding of d-glucose plus phosphate. It was found that the productivity was not greatly affected by inoculum size. However, the l-phenylalanine titre was significantly affected by total sugar content. Relatively high productivities of up to 0.35–0.40 g l-phenylalanine l?1 h?1 have been achieved at l-phenylalanine titres of 14–15 g l?1.  相似文献   

20.
At the GABAA receptor, low concentrations of etomidate potentiate the inhibitory effect of GABA on specific binding of the closed channel ligand [3H]ethynylpropylbicycloorthobenzoate ([3H]EBOB). Here, we present SARs for etomidate and structurally related compounds inducing this effect. In the absence of GABA, similar SARs, but 14–20 times weaker potencies were observed. We discuss these SARs in comparison to the much higher potencies of these compounds as inhibitors of 11β-hydroxylase.  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