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 31 毫秒
1.
The activities of chlorophyllase, contents of pigments including chlorophyll a and b, chlorophyllide a and b, and phaeophorbide a during leaf senescence under low oxygen (0.5% O2) and control (air) were investigated in a non-yellowing mutant and wild-type leaves of snap beans (Phaseolus vulgaris L.). Chlorophyllase from leaf tissues had maximum activity when incubated at 40C in a mixture containing 50% acetone. In both mutant and wild type, chlorophyllase activity was the highest in freshly harvested non-senescent leaves and decreased sharply in the course of senescence, indicating that the loss of chlorophylls in senescing leaves is not directly related to the activity of chlorophyllase and that chlorophyllase activity is not altered in the mutant. The wild type had higher ratios of chlorophyll a to chlorophyll b than the mutant and chlorophyll a : b ratios increased during senescence in both types. In the senescent mutant leaves, accumulations of chlorophyllide a and chlorophyllide b were detected, but no phaeophorbide a was found. Chlorophyllide b had a greater accumulation than chlorophyllide a in the early stage of senescence. Low oxygen treatment not only delayed chlorophyll degradation but also enhanced the accumulations of chlorophyllide a and b and lowered the ratios of chlorophyll a to chlorophyll b.  相似文献   

2.
W. A. W. Moll  D. Stegwee 《Planta》1978,140(1):75-80
Chlorophyllase (chlorophyll-chlorophyllidohydrolase, EC 3.1.1.14) was isolated and purified from Phaseolus vulgaris L. chloroplasts and etioplasts dissolved in 1% Triton X-100 and 10% glycerol. A 100 and 40-fold purification, respectively, was achieved. Enzyme preparations from both sources had similar affinities for chlorophyll a when assayed in a Triton X-100 medium. When electrophoresed in sodium dodecyl sulphate polyacrylamide gels the major band in both preparations migrated as a peptide of 30,000 daltons. Chlorophyll containing liposomes were also used as a substrate for chlorophyllase. The rate of hydrolysis did not follow Michaelis-Menten kinetics. When chlorophyllide a or methyl chlorophyllide a was incorporated in the liposomes, then in the presence of phytol dissolved in methanol, methylchlorophyllide a and chlorophyll a were shown to be synthesized. Apparently the purified enzyme in the presence of lipids, is endowed with both synthetic and hydrolytic activity.Abbreviations DEAE diethylaminoethyl - MeOH methanol - SDS sodium dodecyl sulphate  相似文献   

3.
Chlorophyllide b and four chemically different chlorophyll b specieis, chlorophyllide b esterified with geranylgeraniol, dihydrogeranylgeraniol, tetrahydrogeranylgeraniol and phytol have been detected in addition to the same derivatives of chlorophyll a in the greening cotyledons of cucumber. These esters could be separated and determined by high-performance liquid chromatography. The results suggest that chlorophyll b phytol is formed from the esterification of chlorophyllide b and geranylgeraniol followed by three hydrogenations of the alcohol moiety, as in the case of chlorophyll a and protochlorophyll phytol formation  相似文献   

4.
Spectral changes and esterification (presumably with phytol) of newly formed chlorophyllide a in dark-grown leaves of wildtype bean (Phaseolus vulgaris) and barley (Hordeum vulgare) and a number of chloroplast mutants in barley, were studied by spectrofluorimetry on leaves and on solvent extracts. The shift of the fluorescence emission maximum from 692–694 to 678 nm (excitation shift: 682–684 to 672 nm) and esterification of chlorophyllide a have a similar time course, and both processes are temperature dependent in a similar manner. After completion of the spectral shift and esterification, the fluorescence efficiency of chlorophyll a increases with a subsequent reaccumulation of protochlorophyllide. In leaves of mutants where the shift of fluorescence from 692 to 678 nm is lacking, esterification and the subsequent processes are also blocked. In leaves of mutants with a rapid shift of the fluorescence from 692 to 678 nm, or with direct photoconversion to chlorophyllide a with the fluorescence at 678 nm, esterification is also rapid. The results are interpreted as a sequence of molecular events involving a conformational relaxation of the chlorophyllide holochrome and a translocation of chlorophyll a to reaction centers of the photosystems.  相似文献   

5.
Illumination of etiolated maize leaves with low-intensity light produces a chlorophyll/pheophytin-containing complex. The complex contains two native chlorophyll forms Chl 671/668 and Chl 675/668 as well as pheophytin Pheo 679/675 (with chlorophyll/pheophytin ratio of 2/1). The complex is formed in the course of two successive reactions: reaction of protochlorophyllide Pchlde 655/650 photoreduction resulted in chlorophyllide Chlde 684/676 formation, and the subsequent dark reaction of Chlde 684/676 involving Mg substitution by H2 in pigment chromophore and pigment esterification by phytol. Out data show that the reaction leading to chlorophyll/pheophytin-containing complex formation is not destructive. The reaction is in fact biosynthetic, and is competitive with the known reactions of biosynthesis of the bulk of chlorophyll molecules. The relationship between chlorophyll and pheophytin biosynthesis reactions is controlled by temperature, light intensity and exposure duration.The native complex containing pheophytin a and chlorophyll a is supposed to be a direct precursor of the PS II reaction centre in plant leaves.Abbreviations Chl chlorophyll - Chlde chlorophyllide - Pchl protochlorophyll - Pchlde protochloropyllide - Pheo pheophytin - PS II RC Photosystem II reaction centres. Abbreviations for native pigment forms: the first number after pigment symbol corresponds to the maximum position of low-temperature fluorescence band (nm); the second number corresponds to the maximum position of long wave absorption band  相似文献   

6.
The kinetics of formation of esterified chlorophyll in etiolated barley (Hordeum vulgare L.) leaves after illumination with a single flash was studied. It was found that after partial (14–24%) and after full photoreduction of protochlorophyllide, the same quantity of esterified products appear during the first 5 s after the flash. The rest of formed chlorophyllide was esterified in a slow process during at least 30 min at 15 °C. The product of fast esterification can be correlated with ‘short-wavelength’ chlorophyll, characterized by a fluorescence emission peak at 673–675 nm. This is the only chlorophyll form detectable within 20 s after partial (14%) photoconversion, and it appears at the same time as the shoulder of the chlorophyll(ide) fluorescence after full photoconversion. The main product after full photoconversion shows a fluorescence at 689 nm shifting in darkness within 15 s to 693 nm and then within 30 min to 682 nm (Shibata shift). The slow esterification proceeds with similar kinetics as the Shibata shift. We propose that the fast esterification of only part of total chlorophyllide after full photoconversion of protochlorophyllide in etiolated leaves reflects the restricted capacity of the esterifying system. The slow esterification of the residual chlorophyllide may be time-limited by its release from protochlorophyllide oxidoreductase, by disaggregation of prolamellar bodies and by diffusion of tetraprenyl diphosphates towards chlorophyll synthase. This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

7.
Chlorophyllide b could be extracted from the wild type of Scenedesmus obliquus and its pigment mutant C-2A'. Its identity was proved by absorption and fluorescence spectroscopy and by a positive hydroxylamine test. Chlorophyllide b could be transformed into pheophorbide b and methylpheophorbide b. The formation of chlorophyllide b from chlorophyll b by dephytylation with chlorophyllase could be ruled out. The stimulation of chlorophyllide b biosynthesis with o-phenanthroline, as described in the literature, could not be confirmed under physiological conditions.  相似文献   

8.
In this study, two extracting methods (sonication and dispersing) and three solvents (90% acetone, N,N′-dimethylformamide and methanol) were compared for their ability to extract chlorophyll a of freshwater phytoplankton. Measurements were performed with both spectrophotometry and high-performance liquid chromatography. Results showed that (i) cell disruption is essential and that (ii) the method of cell disruption and solvent applied differed significantly. Dispersing in acetone surpassed all other combinations. Sonication in N,N′-dimethylformamide was found less effective. N,N′-dimethylformamide and methanol seem to promote the formation of degradation products (chlorophyllide a, allomer, epimer and phaeophytin a) which lead to overestimates of chlorophyll a of about 10% by means of spectrophotometry.  相似文献   

9.
Gas chromatographic determinations revealed a certain amount of free phytol in dark-grown barley leaves. When a short light impulse or continuous light is given to the leaves, the phytol pool is partly emptied due to esterification of chlorophyllide a. The regeneration is slow during the first 2–3 hours. A pretreatment with light flashes followed by a dark period accelerates the regeneration, which stops however after approximately 30 min. Some evidence points to the existence of an acceptor for excess phytol entering at this stage. Connections between phytol changes during irradiation and the lag phase of chlorophyll formation are discussed.  相似文献   

10.
Chlorophyll(ide) spectroscopic properties and Photosystem II assembly, monitored by 77 K variable fluorescence, were studied in etiolated barley leaves as a function of the extent of protochlorophyllide photoreduction by a single millisecond light flash of different intensities. Variable fluorescence, measured 2 hours after the flash, was only detected when the extent of phototransformation was higher than a threshold value of 0.4. Its development paralleled the formation of a chlorophyll emission component at 685 nm, which itself derived from long-wavelength chlorophyllide with an emission maximum at 695 nm. At low flash intensities, short-wavelength chlorophyllide forms preferentially accumulated and no Photosystem II fluorescence was detected after 2 hours. Chlorophyllide esterification was independent of the extent of phototransformation. These results suggested that the formation of long-wavelength chlorophyllide was essential for further assembly of Photosystem II. This interpretation was strengthened by the observed inhibition of both long-wavelength chlorophyllide formation and of variable fluorescence development in leaves treated with -aminolevulinic acid or in untreated leaves subjected to repeated flashes of low intensity.  相似文献   

11.
Recently, some evidence for the occurence of a light-independent protochlorophyllide-reducing enzyme in greening barley plants has been presented. In the present work this problem was reinvestigated. -[14C] Aminolevulinic acid was fed to isolated barley shoots from plants which had been preilluminated for various lengths of time. Porphyrins which had been synthesized during the dark incubation were analyzed by high-performance liquid chromatography. There was no evidence for a light-independent synthesis of chlorophyll(ide). The 14C-labelled precursor was incorporated almost exclusively into protochlorophyllide. The reduction of labelled protochlorophyllide to chlorophyllide was strictly light-dependent. These results are not consistent with the existence of a light-independent protochlorophyllide-reductase in barley as proposed previously.Abbreviation HPLC high-performance liquid chromatography  相似文献   

12.
A bacteriochlorophyll a biosynthesis mutant of the purple photosynthetic bacterium Rhodobacter capsulatus was functionally complemented with a cosmid genomic library from Synechocystis sp. PCC 6803. The complemented R. capsulatus strain contains a defined mutation in the bchM gene that codes for Mg-protoporphyrin IX methyltransferase, the enzyme which converts Mg-protoporphyrin IX to Mg-protoporphyrin IX methylester using S-adenosyl-l-methionine as a cofactor. Since chlorophyll biosynthesis also requires the same methylation reaction, the Synechocystis genome should similarly code for a Mg-protoporphyrin IX methyltransferase. Sequence analysis of the complementing Synechocystis cosmid indicates that it contains an open reading frame exhibiting 29% sequence identity to BchM. In addition, expression of the Synechocystis gene in the R. capsulatus bchM mutant via the strong R. capsulatus puc promoter was shown to support nearly wild-type levels of bacteriochlorophyll a synthesis. To our knowledge, the Synechocystis sequence thus represents the first chlorophyll biosynthesis gene homolog of bchM. The complementing Synechocystis cosmid was also shown to code for a gene product that is a member of a highly conserved family of RNA binding proteins, the function of which in cyanobacteria remains undetermined.  相似文献   

13.
Hauke Holtorf  Klaus Apel 《Planta》1996,199(2):289-295
In etiolated barley (Hordeum vulgare L.) seedlings the light-induced accumulation of chlorophyll is controlled by two light-dependent NADPH-proto-chlorophyllide oxidoreductase (POR; EC 1.6.99.1) enzymes. While the concentration of one of these enzymes (POR A) and its mRNA rapidly decline during illumination, the second POR protein (POR B) and its mRNA remain at an approximately constant level during the transition from dark growth to the light. These results may suggest that only one of the enzymes, POR B, operates throughout the greening process and in light-adapted mature plants while the second enzyme, POR A, is active only in etiolated seedlings at the beginning of illumination. The fate of the two POR proteins and their mRNAs in fully green plants, however, has not been studied yet. In the present work we determined changes in the level of POR A and POR B proteins and mRNAs in green barley plants kept under a diurnal 12 h light/12 h dark cycle. In green barley plants, not only POR B is present but also trace amounts of POR A continue to reappear transiently at the end of a night period and seem to be involved in the synthesis and accumulation of chlorophyll at the beginning of each day.Abbreviations Chl chlorophyll - Chlide chlorophyllide - Lhcb light-harvesting chlorophyll a/b protein - Pchlide protochlorophyllide - POR NADPH-protochlorophyllide oxidoreductase Dedicated to Horst Senger on the occasion of his 65th birthday.We thank Dr. Dieter Rubli for photography and Renate Langjahr for typing. This work was supported by the Swiss National Science Foundation and the ETH-Zürich.  相似文献   

14.
Reported crystallographic data and calculated molecular models indicated that chlorophyll (Chl) a and bacteriochlorophyll (BChl) a tend to bind the fifth ligand on the side of the macrocycle where the C132-(R)-methoxycarbonyl moiety protrudes (denoting the ‘back’ side). The crystal structures of 34 photosynthetic proteins possessing (B)Chl cofactors revealed that most of Chl a and BChl a (and b) are coordinated by any peptidyl residue (e.g., histydyl-imidazolyl group), peptidyl backbone or water from the ‘back’ side. Almost all the cofactors that bind a water molecule as the fifth ligand in these proteins have a ‘back’ configuration. Theoretical model calculations for methyl chlorophyllide a (MeChlid a) and methyl bacteriochlorophyllide a (MeBChlid a) bound to an imidazole molecule indicated that the ‘back’ side is energetically favored for the ligand binding. These results are consistent with the fact that ethyl chlorophyllide a (EtChlid a) dihydrate crystal consists of the ‘back’ complex. The modeling also showed that both removal and stereochemical inverse of the C132-methoxycarbonyl group affect the relative stability between the ‘back’ and ‘face’ complexes. The effect of the C132-moiety on the choice of the macrocycle side for the ligand binding is discussed in relation to the function of P700. This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

15.
During leaf senescence, chlorophyll is removed from thylakoid membranes and converted in a multistep pathway to colorless breakdown products that are stored in vacuoles. Dephytylation, an early step of this pathway, increases water solubility of the breakdown products. It is widely accepted that chlorophyll is converted into pheophorbide via chlorophyllide. However, chlorophyllase, which converts chlorophyll to chlorophyllide, was found not to be essential for dephytylation in Arabidopsis thaliana. Here, we identify pheophytinase (PPH), a chloroplast-located and senescence-induced hydrolase widely distributed in algae and land plants. In vitro, Arabidopsis PPH specifically dephytylates the Mg-free chlorophyll pigment, pheophytin (phein), yielding pheophorbide. An Arabidopsis mutant deficient in PPH (pph-1) is unable to degrade chlorophyll during senescence and therefore exhibits a stay-green phenotype. Furthermore, pph-1 accumulates phein during senescence. Therefore, PPH is an important component of the chlorophyll breakdown machinery of senescent leaves, and we propose that the sequence of early chlorophyll catabolic reactions be revised. Removal of Mg most likely precedes dephytylation, resulting in the following order of early breakdown intermediates: chlorophyll → pheophytin → pheophorbide. Chlorophyllide, the last precursor of chlorophyll biosynthesis, is most likely not an intermediate of breakdown. Thus, chlorophyll anabolic and catabolic reactions are metabolically separated.  相似文献   

16.
Summary This study has identified by sequence analysis a single gene in the bchC locus of Rhodobacter sphaeroides and three genes, designated bchX, Y and Z, in the bchA locus, which was previously thought to contain only a single gene. All four genes may reside within the same operon and are transcribed in the order bchC-X-Y-Z. Complementation analysis of eight transposon insertion mutants within these genes suggests that bchX, Y and Z are essential for the reduction of 2-devinyl-2hydroxyethyl chlorophyllide a and that bchC encodes the 2-desacetyl-2-hydroxyethyl bacteriochlorophyllide a dehydrogenase. Similarity between the putative BchX protein and dinitrogenase reductase proteins suggests that BchX may also be a reductase, supplying electrons for reduction of 2-devinyl-2-hydroxyethyl chlorophyllide a.  相似文献   

17.
Glycerol induced a limitation on photosynthetic carbon assimilation by phosphate when supplied to leaves of barley (Hordeum vulgare L.) and spinach (Spinacia oleracea L.). This limitation by phosphate was evidenced by (i) reversibility of the inhibition of photosynthesis by glycerol by feeding orthophosphate (ii) a decrease in light-saturated rates of photosynthesis and saturation at a lower irradiance, (iii) the promotion of oscillations in photosynthetic CO2 assimilation and in chlorophyll fluorescence, (iv) decreases in the pools of hexose monophosphates and triose phosphates and increases in the ratio of glycerate-3-phosphate to triose phosphate, (v) decreased photochemical quenching of chlorophyll fluorescence, and increased non-photochemical quenching, specifically of the component which relaxed rapidly, indicating that thylakoid energisation had increased. In barley there was a massive accumulation of glycerol-3-phosphate and an increase in the period of the oscillations, but in spinach the accumulation of glycerol-3-phosphate was comparatively slight. The mechanism(s) by which glycerol feeding affects photosynthetic carbon assimilation are discussed in the light of these results.Abbreviations Chl chlorophyll - C i intercellular concentration of CO2 - P phosphate - PGA glycerate-3-phosphate - Pi orthophosphate - triose-P sum of glyceraldehyde-3-phosphate and dihydroxyacetone phosphate  相似文献   

18.
Chlorophyll a and chlorophyll b are interconverted in the chlorophyll cycle. The initial step in the conversion of chlorophyll b to chlorophyll a is catalyzed by the chlorophyll b reductases NON‐YELLOW COLORING 1 (NYC1) and NYC1‐like (NOL), which convert chlorophyll b to 7‐hydroxymethyl chlorophyll a. This step is also the first stage in the degradation of the light‐harvesting chlorophyll a/b protein complex (LHC). In this study, we examined the effect of chlorophyll b on the level of NYC1. NYC1 mRNA and NYC1 protein were in low abundance in green leaves, but their levels increased in response to dark‐induced senescence. When the level of chlorophyll b was enhanced by the introduction of a truncated chlorophyllide a oxygenase gene and the leaves were incubated in the dark, the amount of NYC1 was greatly increased compared with that of the wild type; however, the amount of NYC1 mRNA was the same as in the wild type. In contrast, NYC1 did not accumulate in the mutant without chlorophyll b, even though the NYC1 mRNA level was high after incubation in the dark. Quantification of the LHC protein showed no strong correlation between the levels of NYC1 and LHC proteins. However, the level of chlorophyll fluorescence of the dark adapted plant (Fo) was closely related to the accumulation of NYC1, suggesting that the NYC1 level is related to the energetically uncoupled LHC. These results and previous reports on the degradation of chlorophyllide a oxygenase suggest that the a feedforward and feedback network is included in chlorophyll cycle.  相似文献   

19.
The concentration of photosynthetic pigments (i.e., chlorophylls a and b, and carotenoids) and chlorophyll degradation enzyme (i.e., chlorophyllase, oxidative bleaching, and Mg-dechelatase) activities on aphid-damaged and non-damaged regions of the infested leaves were determined with two infestation periods (6 and 12 days). Russian wheat aphid [Diuraphis noxia (Mordvilko) (Hemiptera: Aphididae)] feeding caused significant losses of chlorophylls a and b and carotenoids in the damaged regions. However, bird cherry-oat aphid [Rhopalosiphum padi (L.) (Hemiptera: Aphididae)] feeding did not, except a significantly lower level of carotenoids was observed in the damaged regions from the short-infestation (6-day) samples. Interestingly, the non-damaged regions of D. noxia-infested leaves on both sampling dates had a significant increase of chlorophylls a and b and carotenoid concentrations when compared with the uninfested leaves. Although D. noxia feeding did not cause any changes in either chlorophyll a/b or chlorophyll (a+b)/carotenoid ratio between the damaged and non-damaged leaf regions on short-infestation (6-day) samples, a significantly lower chlorophyll a/b ratio was detected in long-infestation (12-day) samples. The assays of chlorophyllase and oxidative bleaching activities showed no significant differences between the damaged and non-damaged regions of the infested leaves on either sampling date. Mg-dechelatase activity, however, was significantly higher in D. noxia-damaged than non-damaged leaf regions from the short-infestation samples, while no differences were detected from the long-infestation samples. Furthermore, the long-infestation samples showed that Mg-dechelatase activity from both D. noxia-damaged and non-damaged regions increased significantly in comparison with the respective regions of either uninfested or R. padi-infested leaves. We infer that non-damaged regions of D. noxia-infested leaves compensate for the pigment losses in the damaged regions, and that Mg-dechelatase activity changed dynamically from a localized response to a systemic response as infestation duration extends. The findings from this study on cereal aphid-elicited chlorosis (or desistance) would help us to elucidate plant resistance mechanisms, in particular plant tolerance to non-defoliating herbivory.  相似文献   

20.
Zeatin allylic phosphate (ZAP) retarded chlorophyll loss in the barleyleaf senescence assay at a concentration 20 times higher than for6-benzyladenine (BA): the effective concentrations for ZAP and BA were 10 and 0.5 , respectively. Sodium molybdate,an inhibitor of phosphatases, decreased the ZAP effective concentration to 0.5 without affecting leaf senescence andtrans-zeatin activity in the control. This demonstrates theimportance of the phosphate group for ZAP activity or its penetration into leafcells. ZAP up-regulated the protein kinase activity of the barley leaf chromatinwith concentration dependence similar to that oftrans-zeatin. Conversely, ZAP was 1000 times less activethan trans-zeatin in the competition with anti-idiotypeantibodies (raised against antibody to zeatin) for binding with atrans-zeatin-binding site oftrans-zeatin-binding protein ZBP67 isolated from barleyleaves. In contrast to trans-zeatin, ZAP did not activateRNA synthesis in the presence of ZBP in the in vitro systemcontaining chromatin and RNA polymerase I isolated from barley leaves. Insummary, data presented show that ZAP possesses cytokinin activity asdemonstrated by the retardation of barley leaf senescence, but moleculartarget(s) for ZAP in barley leaf cells differs, at least partially, from thesefor trans-zeatin. It seems possible that the cytokininactivity of ZAP results from its hydrolysis while producing zeatin.  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