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.
The degree of fluoresence polarization, P, of unoriented and magnetically oriented spinach chloroplasts as a function of excitation (400–680 nm) and emission wavelengths (675–750 nm) is reported. For unoriented chloroplasts P can be divided into two contributions, PIN and PAN. The latter arises from the optical anisotropy of the membranes which is due to the orientation with respect to the membrane plane of pigment molecules in vivo. The intrinsic polarization PIN, which reflects the energy transfer between different pigment molecules and their degree of mutual orientation, can be measured unambiguously only if (1) oriented membranes are used and the fluorescence is viewed along a direction normal to the membrane planes, and (2) the excitation is confined to the Qy (≈ 660−680 nm) absorption band of chlorophyll in vivo. With 670–680 nm excitation, values of P using unoriented chloroplasts can be as high as +14%, mostly reflecting the orientational anisotropy of the pigments. Using oriented chloroplasts, PIN is shown to be +5±1%. The excitation wavelength dependence studies of PIN indicate that the carotenoid and chlorophyll Qy transition moments tend to be partially oriented with respect to each other on a local level (within a given photosynthetic unit or its immediate neighbors).  相似文献   

2.
Using ELISAs for B-50/GAP43 and neurofilament (NF), we tested ACTH(1–24), -MSH, ACTH(4–10), and an ACTH(4–9) analogue (ORG2766) for their ability to induce sprouting and neuritogenesis from spinal and sensory neurons. Dissociated fetal rat spinal cord neurons or neonatal rat dorsal root ganglion (DRG) cells were cultured with peptide and assayed after 24, 48, or 96 h. In spinal neurons, -MSH and ACTH(1–24) induced the expression of B-50 dose dependently. After 24 h -MSH had a stimulatory effect (from 10 nM onwards), with a maximum at 100 μM (36% increase). After 96 h the maximal effect of 100 μM -MSH on B-50/GAP43 was lower (19%). ACTH(1–24) (100 μM) stimulated B-50/GAP43 by 19%. Neurofilament levels (96 h) were elevated maximally by 64% at 100 μM -MSH. In DRG neurons a bell-shaped dose-response curve was found for -MSH, the maximal effect being observed after 48 h at 100 nM: 54% for B-50/GAP43 and 22% for NF. In both culture systems neither ACTH(4–10) nor ORG2766 was effective. We conclude that -MSH stimulates the expression of B-50/GAP43 (sprouting) and the formation of NF (neurite elongation) and may therefore be considered a neurotrophic factor.  相似文献   

3.
Stearidonic acid (18:4ω3), which is reported to be of rare occurrence in the plant kingdom and which is of considerable dietary and pharmaceutical interest has been found in three closely related Primula species. It occurs, together with γ-linolenic acid (3–4% of the seed oil total fatty acids), in significant percentages in Primula florindae (11%), P. sikkimensis (14%) and P. alpicola (14%). 18:4(ω3 may also be of chemotaxonomic interest in the genus Primula, as high levels may be typical for section Sikkimensis. The only commercial plant source of stearidonic acid known so far is the seed oil of Ribes nigrum.  相似文献   

4.
The influence of using an anaerobically pre-treated baker’s yeast on the reduction of (R)-1-hydroxy-1-phenyl-2-propanone (2) and (S)-2-hydroxy-1-phenyl-1-propanone (4) was investigated in comparison with non-pre-treated baker’s yeast reduction (control experiments). We observed that there is no significant difference between the anaerobically pre-treated yeast and the control experiment on the reduction rates of 2. On the other hand, the rate of reduction of 4 mediated by the anaerobically pre-treated yeast is much slower than the aerobic experiment. To improve the regioselectivity of reduction of 1-phenyl-1,2-propanedione (1), a baker’s yeast suspension was pre-treated with nitrogen (60 min) followed by oxygen (20 min), to give 2 in 28–31% of yields (96% e.e.) and 3 in 42–62% (>99% e.e.) after 75–90 min of reaction.  相似文献   

5.
S.C. Huber  G.E. Edwards   《BBA》1976,449(3):420-433
1. Cyclic photophosphorylation driven by white light, as followed by 14CO2 fixation by mesophyll chloroplast preparations of the C4 plant Digitaria sanguinalis, was specifically inhibited by disalicylidenepropanediamine (DSPD), antimycin A, 2,5-dibromo-3-methyl-6-isopropyl-p-benzoquinone (DBMIb), 1-ethyl-3(3-dimethylaminopropyl)-carbodiimide (EDAC), and KCN suggesting that ferredoxin, cytochrome b563, plastoquinone, cytochrome f, and plastocyanin are obligatory intermediates of cyclic electron flow. It was found that 0.2 μM DCMU and 40 μM o-phenanthroline blocked noncyclic electron flow, stimulated cyclic photophosphorylation, and caused a partial reversal (40–100%) of the inhibition by DBMIB and antimycin A, but not DSPD.

2. Cyclic photophosphorylation could also be activated using only far-red illumination. Under this condition, however, cyclic photophosphorylation was much less sensitive to the inhibitors DBMIB, EDAC and antimycin A, but remained completely sensitive to DSPD and KCN. Inhibition in far-red light was not increased by preincubating the chloroplasts with the various inhibitors for several minutes in white light.

3. The striking correspondence between the effects of photosystem II inhibitors, DCMU and o-phenanthroline, on cyclic photophosphorylation under white light and cyclic photophosphorylation under far-red light (in the absence of photosystem II inhibitors) suggests that electrons flowing from photosystem II may regulate the pathway of cyclic electron flow.  相似文献   


6.
George Papageorgiou  Govindjee 《BBA》1971,234(3):428-432
The pH of the suspension medium was found to have a remarkable influence on the “slow” (min) time course of Chlorophyll a fluorescence yield in the green alga Chlorella pyrenoidosa and in the blue-green alga Anacystis nidulans. In Chlorella, the decay of fluorescence yield, in the 1- to 5-min region, is strongly retarded at alkaline pH; this decay rate shows an optimum at pH 6–7. In Anacystis, the rise of fluorescence yield, in the same time range, is decreased optimally at pH 6–7; poisoning with 3(3,4-dichlorophenyl)-1,1-dimethylurea reverses the direction of this pH effect. These observations suggest a correlation of the H+ status (or the processes associated with it such as photophosphorylation and resulting conformational changes) of the chloroplast to the yield of chlorophyll a fluorescence in vivo.  相似文献   

7.
Thysanolaena maxima (Roxb) (tiger or broom grass), is a tall reed-like perennial grass, grown on shady slopes in forests in India and the Nicobar Islands. Culms are solid, smooth and rounded and up to a height of 4 m. The leaves and the tips are used as fodder; the bushy, fox-tail-like panicles (30–90 cm) are used for making brooms. When the panicles are cut, the stem portion (3–4 m) is left out in the field and is burnt. Fibres, of average 1.25 mm in length at 45% yield (unbleached), could be obtained from this grass. The laboratory handmade paper sheets exhibited good properties, with a burst factor of 30, a breaking length of 3555 m and a tear factor of 106. Hence, it can be suggested that T. maxima may become a potential source of raw material for pulp- and paper-making either alone or in combination with the conventional pulp- and paper-making raw materials. This material could help to meet the future demand for pulp- and paper-making raw material, if properly exploited.  相似文献   

8.
9.
Preillumination of intact cells of the eukaryotic, halotolerant, cell-wall-less green alga Dunaliella salina induces a dark ATPase activity the magnitude of which is about 3–5-fold higher than the ATPase activity observed in dark-adapted cells. The light-induced activity arises from the activation and stabilization in vivo of chloroplast coupling factor 1 (CF1). This activity, 150–300 μmol ATP hydrolyzed/mg Chl per h, rapidly decays (with a half-time of about 6 min at room temperature) in intact cells but only slowly decays (with a half-time of about 45 min at room temperature) if the cells are lysed by osmotic shock immediately after illumination. The activated form of the ATPase in lysed cells is inhibited if the membranes are treated with ferri- but not ferrocyanide, suggesting that the stabilization of the activated form of CF1 is due to the reduction of the enzyme in vivo in the light.  相似文献   

10.
The coral-inhabiting snails, Coralliophila violacea (Lamarck), are widely distributed in the Indo–Pacific region. It has been inferred from field data that sex changes have occurred and might be influenced by neighbors. In this study, we designed a field experiment with 2 treatments, i.e. ‘single males' and ‘male–female pairings' to test the hypothesis of sex changes as well as the role of neighbors on sex changes. Snails were collected by SCUBA diving from the surface of massive coral Porites lobata in shallow waters (<6 m) in southern Taiwan. A total of 73 males and 32 females were marked and returned to the surface of a colony of the massive coral, P. lobata. After 4–5 months, a total of 25 marked males and eight marked females were recaptured. In the treatment of ‘single males', eight of 17 males had changed their sex. None of eight recaptured males in the treatment of ‘male–female pairings' had changed their sex. The occurrence of sex changes is dependent on the presence or absence of a female neighbor (P<0.05, Fisher's exact test). In this study, we have provided direct evidence of socially controlled sex changes in Coralliophila violacea.  相似文献   

11.
1. The wavelength dependence of the fluorescence polarization (FP) ratio and dichroism has been studied with magneto-oriented (10–13 kG) whole cells of Chlorella pyrenoidosa, Scenedesmus obliquus, Euglena gracilis and spinach chloroplasts suspended in their aqueous growth media (or Tris-buffered sucrose solution in the case of the chloroplasts) under physiological conditions. The FP ratio is defined as the fluorescence intensity polarized parallel divided by the intensity polarized perpendicular to the membrane planes.

2. The FP ratio is typically in the range of 1.2–1.9 in Chlorella, 1.20–1.25 in Scenedesmus and 1.4–1.5 in spinach chloroplasts at fluorescence wavelengths above 690 nm. Below 690 nm the FP ratio decreases steadily with decreasing wavelength and may be as low as approx. 1.05 at 660 nm. These results are interpreted in terms of the orientation of the Qy transition moment vectors of the different spectroscopic forms of chlorophyll. For the chlorophyll a 680 form these vectors are inclined at angles of 30° or less (in Chlorella) with respect to the membrane planes, while the shorter wavelength chlorophyll a 670 forms appear to be not nearly as well oriented.

3. The Euglena fluorescence peak is red shifted to 714 nm (in the other algae and chloroplasts it is situated at 685 nm) and the FP ratio is approx. 1.20 in the 720–730 nm region and decreases with decreasing wavelength below 720 nm and is only 1.05 at 690 nm. This wavelength dependence is in good qualitative agreement with the fluorescence microscope studies of single chloroplasts of Euglena by Olson, R. A., Butler, W. H. and Jennings, W. H. ((1961) Biochim. Biophys. Acta, 54, 615–617).

4. By means of a model calculation it is shown that the high FP ratios observed with Chlorella are entirely consistent with the low values of the degree of polarization (0.01–0.06) determined by previous workers with unoriented cell suspensions.

5. The influence of reabsorption and the resulting distortion in the wavelength dependence of the FP ratio are described. The possibility that the fluorescence is polarized by scattering artifacts, rather than being a result of the intrinsic orientation of chlorophyll, is considered.

6. Linear dichroism studies with Chlorella and spinach chloroplasts confirm the orientation of the Qy transition moment vectors deduced from the FP ratio. Furthermore, it appears that the porphyrin rings are tilted out of the membrane plane and that the carotenoid molecules tend to lie with their long axes in the lamellar plane.

7. In Euglena, dichroism studies indicate that chlorophyll a 680 is unoriented, while chlorophyll a 695 appears to be oriented similar to chlorophyll a 680 in Chlorella or spinach chloroplasts, a result which is also in accord with the measured FP ratio of Euglena.

8. The possibility that the magnetic field gives rise to the reorientation of individual chlorophyll molecules is shown to be highly unlikely.  相似文献   


12.
1986. In vitro excystrnent of the metacercaria of Plagiorchis species 1 (Trematoda, Plagiorchiidae). International Journal for Parasitology 16: 641–645. An optimal hatching success of Plagiorchis species 1 metacercariae (100% excystment, active metacercariae, mean hatching speed 2–10 min, lowest variance of the mean speed) was observed after pretreatment in an HCl-pepsin solution at pH 2.0 and 42°C for 60–70 min, and incubation in a hatching medium at 42 °C and pH 7.3–8.0 with a bile salt (Nacholate), NaHCO3, and a reductant (cysteine with 100% N2). The minimum conditions for nearly 100% excystment with lower hatching speeds and higher variances were the presence of NaHCO3, an oxygen concentration reduced to about 3% in the gas phase, pH> 7.3 and a temperature near 30°C if Na-cholate was absent, or in the presence of the bile salt, a phosphate buffer at pH> 5.0 and room temperature only. Obviously some hatching factors acted interchangeably with compensation for missing stimuli by others. The effect of the bile salt was comparable with that of other surfactants. The metacercariae excysted in nonenzymatic media, which implies an active hatching mechanism.  相似文献   

13.
The chemical substance 2-(2-nitro-4-trifluoromethylbenzoyl)-1,3-cyclohexanedione (NTBC) is in clinical use for the treatment of hereditary tyrosinemia type 1. In the present study, the plasma concentration of NTBC was determined by a coupled column liquid chromatographic method. A 20-μl volume of plasma was diluted with phosphate buffer, pH 2, and injected into a small precolumn (BioTrapAcid C18) with a mobile phase containing sulfuric acid. The precolumn was based on the restricted access principle, i.e., retention of NTBC within the lipophilic pores, while polar and large endogenous compounds were eluted with the void volume. NTBC was transferred to the analytical column using a mobile phase with a high content of acetonitrile. The compound was monitored by UV detection at 278 nm. The standard curve was linear between 0.3 and 69 μM, and the between-day precision (RSD) was 3% (n=6 days) at 13.8 μM and 14% (n=6 days) at 0.3 μM NTBC in plasma. The quantitation limit was approximately 0.3 μM using 20 μl of plasma.  相似文献   

14.
In order to solve discrepancies between earlier assignments we have reinvestigated the stereoisomerism of the spheroidene molecule bound to reaction centers (RC) of Rhodobacter sphaeroides. A stable cis isomer could be extracted and purified from the reaction centres by working at very low ambient light. Resonance Raman spectroscopy showed that this cis isomer assumed the same configuration as that of the RC-bound molecule. Proton-NMR spectroscopy of the extracted isomer permitted to assign it the 15–15′ mono cis configuration. Comparisons between resonance Raman spectra of the native form and of the 15 cis extract showed that, in the reaction center, 15 cis spheroidene is in addition twisted into a non-planar conformation. Comparisons of extraction-induced changes in relative intensities of Raman bands of the 760–1060 cm−1 regions, which largely correspond to out-of-plane modes, further indicated that the out-of-plane twist of RC-bound spheroidene should predominantly affect C8–C12 and/or C8′–C12′ regions of the molecule rather than the central region. Comparisons between difference electronic absorption spectra of RC-bound spheroidene and of RC-bound methoxyneurosporene showed that the out-of-plane twisting of both these native forms results in a drastic weakening of their 1C ← 1A electronic transitions, compared with those of the planar, 15 cis forms. Finally, it is proposed, on the basis of their resonance Raman spectra, that spirilloxanthin bound to RCs of Rhodospirillum rubrum as well as dihydroneurosporene or dihydrolycopene bound to RCs of Rhodopseudomonas viridis shares 15 cis configurations and out-of-plane twisting with carotenoids bound to RCs of various strains of Rb. sphaeroides.  相似文献   

15.
The effect of eye carotenoid content, light conditions and retinoid supply on the biosynthesis of opsin, as well as the ability to isomerize of exogenous all-trans-retinal to 11-cis-retinal were investigated in the photoreceptors of blowfly Calliphora. SDS-PAGE of digitonin extracts from isolated rhabdom fractions of carotenoid-fortified and carotenoid-deficient animals revealed, on heavily loaded gels, that in both cases opsin forms faint minor bands similar to the patterns of “R-flies” described as “vitamin A-deficient flies” (Paulsen and Schwemer, Biochim. biophys Acta 557, 358–390, 1979) or “carotenoid-deficient flies” (Paulsen and Schwemer, Eur. J. Biochem. 137, 609–614, 1983. Similar opsin patterns were obtained in flies subjected to continuous 72 h blue or yellow-green light or to a 12 h: 12 h white light:dark cycle, or total darkness, irrespective of the carotenoid content in their eyes. 11-cis-retinal and, to a lesser extent, all-trans-retinal, when included in the diet or painted on the cornea, were found to stimulate opsin biosynthesis in the dark. 11-cis-retinal in the dark or all-trans-retinal after illumination of the flies with blue light were the most effective, as compared with the effect of all-trans-retinal in the dark. Exogenous all-trans-retinal in the dark can be partially converted into a mixture of 11-cis-retinal (26%) and 13-cis-retinal (74%) by fly retina homogenate in vitro. It was concluded that Calliphora opsin biosynthesis is not strongly dependent on the carotenoid supply or on light: dark conditions and is triggered by the 11-cis aldehyde form of the chromophore, which can be produced in the fly retina either by an isomerase system in the dark or as a result of photoisomerization.  相似文献   

16.
The environmental record from Lake Baikal, Russia, from 310 to 50 ky BP (MIS 9a to MIS 3) was interpreted using rock magnetic, UV–Vis spectral, mineralogical, and diatom analyses. The age model was based on a correlation of the diatom and chemical weathering records and the summer insolation curve at 55°N and checked against an age model based on the proxy of relative palaeointensity of the Earth's magnetic field. Peaks in chemical weathering within the watershed, inferred from maximum concentration of magnetic and coloured minerals and mica, the lowest mean Fe oxidation state in silicates and highs in expandable clay minerals correlated with the Northern Hemisphere summer insolation minima at 55°N. Reconstructed changes in weathering intensity are better correlated to insolation patterns than to global ice volume records. We propose a scheme of yet missing palaeoenvironmental interpretation of the diatom assemblage, including also some extinct species. Aulacoseira baicalensis and Aulacoseira skvortzowii were abundant in the early stages of lake flora recovery immediately after deglaciation and during MIS 7e and MIS 5e; periods of more pronounced continental climate and peak chemical weathering. Stephanodiscus formosus var. minor, Cyclotella minuta and Cyclotella ornata dominated in intervals of decreased seasonality and decreased humidity at the end of most interglacial/interstadial diatom zones. Stephanodiscus grandis, Stephanodiscus carconeiformis and Stephanodiscus formosus were ubiquitous between MIS 8 and MIS 5, an interval marked by high seasonality, i.e., large differences between winter and summer insolation, and low humidity revealed by a low hydrolysis of expandable clay minerals in the watershed. Diatom concentrations peaked in the climatic optima of MIS 7e and MIS 5e and in the short periods marked by shifts to warmer conditions in the upper sections of MIS 5: MIS 5c (103–99 ky BP), MIS 5b (90–88 ky BP), and MIS 5a (84–79 ky BP) in which increased humidity resulted in enhanced hydrolysis of clay minerals. No such short similar climatic optimums were found from MIS 9a to MIS 6. Sharp climate deteriorations recorded in the diatom and clay mineral records at 107, 94, and 87 ky BP, however, occurred within 1–2 ky of cold extremes in North Atlantic sea surface temperature emphasizing the strong teleconnections between the two localities.  相似文献   

17.
Backgrounds and aims: skin lesions in cutaneous porphyrias appear to be determined by the structural properties of the porphyrins accumulated. To better understand the relationship between the structure and physicochemical properties of porphyrins and their specific effect on protein configuration, the action of a whole range of 8 to 2 carboxylic porphyrins has been studied. Materials and methods: δ-aminolevulinic acid dehydratase (ALA-D) and porphobilinogen deaminase (PBG-D) partially purified from bovine liver, were exposed to 10 μM uroporphyrin (Uro), phyriaporphyrin (Phyria), hexaporphyrin (Hexa), pentaporphyrin (Penta), coproporphyrin (Copro) or protoporphyrin (Proto), either in the dark or under UV light. All experiments were performed in the enzyme solutions after removing the porphyrins. Results: under both illuminating conditions, all porphyrins inactivated the enzymes (20–70% under control values), indicating photodynamic action mediated by oxidative reactions and conformational changes due to direct binding of porphyrins to the protein. Total thiol content in ALA-D was not significantly changed by most porphyrins under UV light, while all porphyrins increase total sulfhydryl groups in PBG-D (23–52% over the control values) indicating changes in the redox status of SH residues. Free amino groups were reduced by all porphyrins in ALA-D (23–56% under controls), instead they were enhanced in PBG-D (23–51% over controls), suggesting protein fragmentation. The formation of molecular aggregates would be the consequence of cross-links between oxidation products, while fragmentation can be attributed to either rupture of disulphur bridges and/or enhancement of free amino groups on the protein enzyme. Conclusions: the effect of the porphyrins on enzyme activity, total SH groups and free amino groups content, was different for ALA-D and PBG-D, even under the same illuminating conditions. On the basis of these results, no correlation between enzyme alterations and the physico-chemical properties of porphyrins could be established.  相似文献   

18.
From Emerson enhancement measurements of O2 evolution in Chlorella pyrenoidosa, it was possible to establish a relationship between the concentration of photosystem II open reaction centers (E) and the distribution of photons between photosystems I and II [(1 − )/] during steady state. The superposition of lights of two different wavelengths (1 and 2) gives concentrations of E and intermediate between those obtained with light 1 and 2 separately. This relationship extends a previous one based on quantum yield measurements. It has been expressed here by a curve corresponding to a fixed value of the intersystem apparent equilibrium constant (K). Up to 700 nm, K remains equal to 6. Above this wavelength, although the margin of error is rather great, K apparently increases to 12 or more.

The possibility of “spill-over” of light absorbed by System II to System I was studied. There is no probability that this spill-over, if any, exceeds 25% in Chlorella.

The apparent equilibrium constant is decreased by 3(3,4-dichlorophenyl)-1,1-dimethylurea. This is not in favor of the hypothesis of fully independent electron-transfer chains in photosynthesis; it is therefore likely that some communication between those chains exists.  相似文献   


19.
20.
Caveolin-1 (Cav1), a structural protein of caveolae, plays cell- and context-dependent roles in signal transduction pathway regulation. We have generated a knockout mouse homozygous for a null mutation of the Cav1 gene. Cav1 knockout mice exhibited impaired urinary bladder contractions in vivo during cystometry. Contractions of male bladder strips were evoked with electric and pharmacologic stimulation (5–40 Hz, 1–10 μM carbachol, 10 mM ,β-methylene ATP, 100 mM KCl). Acetylcholine (ACh) and norepinephrine (NE) release from bladder strips were measured with a radiochemical method by incubating the strips with 14C-choline and 3H-NE prior to electric stimulation, whereas ATP release was measured using the luciferin-luciferase assay with a luminometer. A 60–75% decline in contractility was observed when Cav1 knockout muscle strips were stimulated with electric current or carbachol, compared to wildtype muscle strips. No difference in contractility was noted when contractions were evoked either by the purinergic agonist ,β-methylene ATP, or by extracellular potassium. To investigate the relative contribution of non-cholinergic activity to bladder contractility, the amplitude of the electric stimulation-evoked contractions was compared in the presence of the muscarinic antagonist atropine (1 μM). While the non-muscarinic (purinergic) response was unaltered, muscarinic cholinergic response was principally disrupted in Cav1 knockout mice. The loss of Cav1 gene expression was also associated with a 70% reduction in ACh release. NE and ATP release was not altered. It is concluded that the loss of caveolin-1 is associated with disruption of M3 muscarinic cholinergic activity in the bladder. Both pre-junctional (acetylcholine neurotransmitter release from neuromuscular junctions) and post-junctional (M3 receptor-mediated signal transduction in bladder smooth muscles) mechanisms are disrupted, resulting in impaired bladder contraction.  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