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.
Using the facultative root hemiparasite Rhinanthus minor and its host Hordeum vulgare several aspects of water relations have been measured in this parasitic association. Extraction of xylem sap by the parasite from the host's roots is facilitated by con siderably higher transpiration per leaf area in the parasite than in the host and by the fact that stomata of attached Rhinanthus were open all day and night despite extremely high ABA concentrations in the leaves. By comparison, another root hemiparasite, Melampyrum arvense, parasitizing various grasses in the field, showed normal diurnal stomatal behaviour. The abnormal behaviour of Rhinanthus stomata was not due to anatomical reasons as closure could be induced by applying high external ABA concentrations. Remarkable differences have been detected between the hydraulic conductance of barley seminal roots showing relatively low values and that of Rhinanthus seminal roots showing very high values. The latter could be related to the observed high ABA concentrations in these roots. Whole plant water uptake, transpirational losses, growth-dependent deposition, and the flows of water within the plants have been measured in singly growing Rhinanthus and Hordeum plants and in the parasitic association between the two. Water uptake, deposition and transpiration in Rhinanthus were dramatically increased after attachment to the barley host; most of the water used by the parasite was extracted as xylem sap from the host, thereby scavenging 20% of the total water taken up by the host's roots. This water uptake by the parasitized host, however, due to a parasite-induced reduction in the host's growth, was decreased by 22% as compared to non-parasitized barley. The overall changes in growth-related water deposition in the host and parasite pointed to decreased shoot growth and relatively favoured root growth in the host and to strongly favoured shoot growth in the parasite. These changes in the host became more severe, when more than one Rhinanthus was parasitizing one barley plant.  相似文献   

2.
The enzyme sucrose: sucrose 1-fructosyltransferase was partially purified from barley leaf growth zones. Four steps (ammonium sulphate precipitation and polyethylene glycol precipitation, followed by chromatography on Concanavalin A-sepharose and hydroxylapatite) yielded a 35-fold purification. The resulting preparation of 1-SST which still contained a number of different activities related to fructan metabolism, was subjected to preparative isoelectric focusing, and sections of the gel were analysed individually for 1-SST and related activities, using sucrose and 1-kestose as substrates. This procedure yielded a 196-fold purification and revealed the presence of two isozymes of 1-SST with pI values of 4.93 and 4.99, as determined by analytical isoelectric focusing of the corresponding fractions. Both isozymes produced glucose and 1-kestose when incubated with sucrose. In addition, small amounts of 6-kestose and tetrasaccharides were formed. In particular, one of the two 1-SST isozymes yielded fructose when incubated with 1-kestose, indicating that it also acts as a fructan exohydrolase. The other isozyme exhibited less fructan exohydrolase activity. Nystose was also degraded by the fructan exohydrolase activity but less than 1-kestose, whereas 6-kestose was not a substrate for the enzyme. Incubation of both 1-SSTs with different concentrations of sucrose showed that the enzyme was not saturated even at 500 mM. As for the barley sucrose: fructan 6-fructosyltransferase, both isozymes of 1-SST yielded two polypeptide bands of molecular weight 50 and 22 kDa upon sodium dodecylsulphate polyacrylamide gel electrophoresis, suggesting their close relationship to invertase (composed of two subunits of similar size), as previously reported for other plants.  相似文献   

3.
Using the facultative root hemiparasite Rhinanthus minor and Hordeum vulgare as a host, the flows, depositions, and metabolism of abscisic acid (ABA) within the host, within the parasite, and between host and parasite have been studied. When the plants were supplied with 5 mM NO(3)(-), there were weak or no effects of parasitism on ABA flows, biosynthesis, and ABA degradation in barley. However, ABA deposition was significantly affected in the leaf laminae (3-fold) and in the leaf sheath (2.4-fold), but not in roots. Dramatic changes in ABA flows, metabolism, and deposition on a per plant basis, however, have been observed in Rhinanthus. Biosynthesis in the roots was 12-fold higher after attachment, resulting in 14-fold higher ABA flows in the xylem. A large portion of this ABA was metabolized, a small portion was deposited. Phloem flows of ABA were increased 13-fold after attachment. The concentrations of ABA in tissues and transport fluids were higher in attached Rhinanthus by an order of magnitude than in host tissues and xylem sap. The same tendency was also found in a comparison between single Rhinanthus and unparasitized barley. As compared with 5 mM NO(3)(-), lower NO(3)(-) or 1 mM NH(4)(+) supply doubled the ABA concentrations in barley leaf laminae, while having only small or non-significant effects in the other organs. The possible function of ABA for the parasite is discussed.  相似文献   

4.
Boron (B) is an essential nutrient required for plant growth and physiological processes. Long-distance B transport is facilitated by the formation of B–polyol complexes. We investigated B uptake and distribution in response to differing levels of exogenous nitrogen supply in the hemiparasitic association between Rhinanthus minor and Hordeum vulgare (barley) and in unparasitised barley and single Rhinanthus plants. In this system, the polyol mannitol is the major assimilate in Rhinanthus , whereas polyols are not detectable in barley. Furthermore, previous studies have shown that the accumulation of polyols within Rhinanthus is negatively affected by the application of exogenous nitrogen. Within the association, the strongest accumulation of B was detected in lateral buds and inflorescences of Rhinanthus , consistent with the greatest B demand in strong sink organs supplied through the phloem that contain high concentrations of mannitol. In the host, the strongest B accumulation was found in xylem-supported leaf lamellae. Roots and sheaths did not accumulate substantial amounts of B, while re-circulation of B through the phloem vessels accounted for only 10% (unparasitised) and 8% (parasitised) of the xylem sap-imported B in the mannitol-free barley hosts. In contrast, 53% (attached) and 39% (in the absence of a host) of the xylem sap-imported B was re-circulated in the phloem in the mannitol-rich Rhinanthus . We therefore present the first quantitative uptake and flow models of long-distance B transport in polyol-rich and polyol-free plants. Our findings are consistent with a close relationship between B re-translocation and mannitol concentrations in phloem vessels.  相似文献   

5.
Rates of CO2 fixation during the light period and the rates of CO2 release during the night period were measured using mature leaves from 39- to 49-d-old spinach (Spinacia oleracea L., US Hybrid 424; grown in 9 h light, 15 h darkness, daily) and mature leaves from 21-d-old barley (Hordeum vulgare L., cv. Apex; grown in 14 h light, 10 h darkness, daily). At certain times during the light and dark periods leaves were harvested for assay of their contents of soluble carbohydrates, starch, malate and the various amino acids. Evaluation of the results of these measurements shows that in spinach and barley leaves 46% and 26%, respectively, of the carbon assimilated during the light period is deposited in the leaves for export during the night period. Taking into account the carbon consumption in the source leaves by dark respiration, it is evaluated that rates of assimilate export during the light period from spinach and barley leaves [38 and 42 atom C · (mg Chl)–1 · h–1] are reduced in the dark period to 16 atom C · (mg Chl)–1 · h–1 in both species. The calculated C/N ratios of the photoassimilates exported during the dark period were 0.029 and 0.015 for spinach and barley leaves, respectively.This work was supported by the Deutsche Forschungsgemeinschaft. We thank Dr. Dieter Heineke for stimulating discussions and Mrs. Petra Hoferichter and Mrs. Marita Feldkämper for their technical assistance.  相似文献   

6.
P. Scott  R. L. Lyne  T. ap Rees 《Planta》1995,197(3):435-441
The aim of this work was to discover why barley (Hordeum vulgare L.) microspores die when cultured on media containing 40 mM sucrose but undergo embryogenesis on 40 mM maltose. Freshly isolated microspores were cultured for 6–24 h on media containing either [U-14C]maltose or [U-14C]sucrose at 40 mM, and the detailed distribution of 14C was determined. The amounts of glycolytic intermediates, ATP, ADP and AMP, in microspores were also measured. Cultures on sucrose differed from those on maltose in that the initial rate of metabolism was faster but declined rapidly, less 14C was recovered in polymers and more in alanine, there was extensive leakage of assimilated carbon, significant accumulation of ethanol and a lower adenylate energy charge. It is argued that microspores cultured on 40 mM sucrose die because they metabolize the sugar rapidly, become hypoxic and, as a result, accumulate large quantities of ethanol within the cells. Metabolism of maltose is slower and there is sufficient oxygen available to allow cells to survive in culture. Consequently some of the cultured cells undergo embryogenesis.P.S. thanks the Science and Engineering Research Council and Shell Research Ltd., Sittingbourne, for a Cooperative Award in Science and Engineering studentship.  相似文献   

7.
Net rates of carbon assimilation per unit leaf area by fully expanded, vegetative leaves of diploid (2x) and autotetraploid (4x) barley (Hordeum vulgare L. cultivars OAC-21 and Brant) were not significantly different (90% level) when measured under controlled environment conditions with air levels of CO2 and either 2 or 20% O2. Leaf thickness increased with ploidy so that net photosynthetic rates measured on single leaves were lower for 4x than 2x barley varieties when compared on a dry or fresh weight basis. Rates of 14CO2 fixation by isolated mesophyll protoplasts prepared from seedlings were also lower for 4x than 2x varieties [about 108 and 125 μmol (mg ChI)?1 h?1, respectively]. Carbohydrate accumulation in leaves of 5-weekold plants averaged 28% (2x) and 47% (4x) of the total photosynthetic weight gain during the first 9 h of the light period. Estimated photoassimilate export from leaves was 15% (OAC-21) and 38% (Brant) lower for 4x compared to 2x isolines. The sucrose and oligofructan content of 4x compared to 2x leaves increased as a result of decreased photosynthate transport. Total tiller dry weight of plants raised in a glasshouse was greater for 4x than 2x barley varieties at ear emergence, but tiller height decreased with increasing ploidy. The nonstructural carbohydrate content of the inflorescence, leaves and lower stem organs was significantly (P≤ 0.01) higher in 4x than in 2x lines at this sampling. During the first 15 days of grain development total tiller dry weight increased by 46% (2x) compared to 8% (4x) when the results of both varieties were averaged together. The dry weight gain of the ear during this period was about 60 to 80% lower for 4x compared to 2x isolines. The nonstructural carbohydrate content of the inflorescence was also about 24% (Brant) and 51% (OAC-21) lower for 4x as compared to 2x plants 15 days post ear emergence. The above results suggested that photosynthate partitioning in autotetraploid barley was sink-limited.  相似文献   

8.
Hordoindolines, the tryptophan-rich polypeptides affecting grain hardness in barley, appeared as three pairs of polypeptides in the acidic polyacrylamide gel electrophoresis (A-PAGE) and two-dimensional A-PAGE?×?SDS-PAGE patterns of starch-granule proteins from 18 barley cultivars. On capillary RP-HPLC/nESI-MS/MS spectrometry, one pair of polypeptides was found to correspond to hordoindoline A (HINA), one to hordoindoline B1 (HINB1) and one to hordoindoline B2 (HINB2), the two polypeptides of each pair deriving from post-translational cleavage of a native hordoindoline at different positions at the N-terminus and/or C-terminus. Amongst the barley cultivars analyzed, cvs Hart and Sundance, which were claimed to be unique in lacking the Hina gene coding for HINA, revealed similar Hina coding sequences and accumulated hordoindoline HINA on their starch granules. The amount of total hordoindolines (HINA?+?HINB1?+?HINB2) on the starch granules, as quantified by densitometric scanning of A-PAGE gels, was comparable with that of puroindolines (PINA?+?PINB) in soft-textured wheat. By contrast, the amount of B-type hordoindolines (HINB1 and HINB2 combined) was 50?% lower than that of PINB, suggesting that the absence of barley cultivars with soft kernels is likely due to the reduced amount of B-type hordoindolines accumulated on the starch granules. Approximately 22 and 27?% of the phenotypic variation for kernel hardness in 56 barley cultivars analyzed by the Single Kernel Characterization System (SKCS) were explained by differences in kernel weight and B-type hordoindoline level, respectively. By contrast, the outer husk of barley grain showed no effect on the SKCS index.  相似文献   

9.
Remains of barley (Hordeum vulgare) grains found at archaeological sites in the Fertile Crescent indicate that about 10,000 years ago the crop was domesticated there from its wild relative Hordeum spontaneum. The domestication history of barley is revisited based on the assumptions that DNA markers effectively measure genetic distances and that wild populations are genetically different and they have not undergone significant change since domestication. The monophyletic nature of barley domestication is demonstrated based on allelic frequencies at 400 AFLP polymorphic loci studied in 317 wild and 57 cultivated lines. The wild populations from Israel-Jordan are molecularly more similar than are any others to the cultivated gene pool. The results provided support for the hypothesis that the Israel-Jordan area is the region in which barley was brought into culture. Moreover, the diagnostic allele I of the homeobox gene BKn-3, rarely but almost exclusively found in Israel H. spontaneum, is pervasive in western landraces and modern cultivated varieties. In landraces from the Himalayas and India, the BKn-3 allele IIIa prevails, indicating that an allelic substitution has taken place during the migration of barley from the Near East to South Asia. Thus, the Himalayas can be considered a region of domesticated barley diversification.  相似文献   

10.
G-banding karyotypes of three cultivars in barley were analyzed. Multiple closely adjacent G-bands were able to be observed in each early metaphase or metaphase chromosome treatted by an ASG method. The more concentrated the chromosome, the less was the number of G-bands during mitosis. The position of band distribution, staining degree and band numbers between homologous chromosomes were basically identical. Chromosome pairing for karyotype analysis could be carried out more accurately. G-banding patterns of different chromosome pairs were not the same, they could be used as the markers to distinguish one from another chromosome pair. During the same mitotic stage the banding patterns including number, relative position and staining degree of the bands between different cultivars were basically the same, but they had differences in the size and staining degree of some bands near centromeres. G-banding technique and G-banding of metaphase chromosomes were discussed.  相似文献   

11.
Studies in peculiarities of the DNA secondary structure in barley by means of thermal denaturation and renaturation shows that there are three types of the nucleotide sequences organization in DNA. More than 95% of the genome composition contain distributed repetitive sequences, in one part of the concentration of the repetitive sequences being higher as compared to bulk of them. About 3.5% of DNA is enriched with A-T pairs and contains no repetitive sequences. There is no "unique" part in the barley genome, which is natural for animals. Slowly renaturation sequences repeat 4 times.  相似文献   

12.
We previously reported that cleistogamy/chasmogamy (CL/CH) of barley ( Hordeum vulgare L.) is controlled by either two tightly linked genes or one gene with multiple alleles. To clarify the morphological and physiological mechanisms of barley CL, we analysed the lodicule size and auxin response of two cultivars whose CL/CH was controlled by two different genes, cly1 and Cly2 . In both cases, lodicules of the CL parent were smaller than those of the CH parent. Analyses of lodicule size and flowering phenotype of f 1 plants and segregation analyses of the mapping population indicated that lodicule size co-segregated with the flowering phenotype. Indole-3-acetic acid (IAA) and other synthetic auxins, such as 2,4-dichlorophenoxyacetic acid, induced abnormal flowering in CH ears, in which florets remained open for a few days instead of the normal hour or so, but not in CL ears. This auxin effect also co-segregated with the flowering phenotype. Analyses of auxin-related compounds in the floret organs revealed that the anther contained high levels of IAA, whereas indole-3-carboxylic acid, a putative decarboxylated metabolite of IAA, accumulated only in lodicules of CH plants just at flowering. These results indicate that lodicule size and auxin response are pleiotropic effects of the CL gene, which may play a role in auxin response or metabolism.  相似文献   

13.
The least ambiguous genetic markers are those based on completely characterized DNA sequence polymorphisms. Unfortunately, assaying allele states by allele sequencing is slow and cumbersome. The most desirable type of genetic marker would be unambiguous, inexpensive to assay and would be assayable singly or in parallel with hundreds of other markers (multiplexable). In this report we sequenced alleles at 54 barley (Hordeum vulgare ssp. vulgare) loci, 38 of which contained single-nucleotide polymorphisms (SNPs). Many of these 38 loci contained multiple polymorphisms, and a total of 112 polymorphisms were scored in five barley genotypes. The polymorphism data set was analyzed both by using the individual mutations as cladistic characters and by reducing data for each locus to haplotypes. We compared the informativeness of these two approaches by consensus tree construction and bootstrap analysis. Both approaches provided similar results. Since some of the loci sequenced contained insertion/deletion events and multiple point mutations, we thought that these multiple-mutated loci might represent old alleles that predated the divergence of barley from H. spontaneum. We evaluated sequences from a sample of H. spontaneum accessions from the Eastern Mediterranean, and observed similar alleles present in both cultivated barley and H. spontaneum, suggesting either multiple domestication events or multiple transfers of genes between barley and its wild ancestor.  相似文献   

14.
The cold-regulated (COR14) protein of 14 kDa is a polypeptide accumulated under low-temperature conditions in the chloroplasts of barley leaves. In H. vulgare the COR14 antibody cross-reacts with two proteins, with a slightly different relative molecular weight around the marker of 14.4 kDa, referred to as COR14a and COR14b (high and low relative molecular weight, respectively). In a collection of H. spontaneum genotypes a clear polymorphism was found for the corresponding COR proteins. While some accessions showed the same COR pattern as cultivated barley, in 38 out of 61 accessions examined the COR14 antibody cross-reacted with an additional coldregulated protein with a relative molecular weight of about 24 kDa (COR24). The accumulation of COR24 was often associated with the absence of COR14b; the relationship between the COR14b/COR24 polymorphism and the adaptation of H. spontaneum to different environments is discussed. By studying COR14 accumulation in cultivated barley we have found that the threshold induction-temperature of COR14a is associated with the loci controlling winter hardiness. This association was demonstrated by using either a set of 30 cultivars of different origin, or two sets of frost-tolerant and frost-sensitive F1 doubled-haploid lines derived from the cross Dicktoo (winter type) x Morex (spring type). These results suggest that the threshold induction-temperature of COR14a can be a potential biochemical marker for the identification of superior frostresistant barley genotypes.  相似文献   

15.
To investigate if latent manganese (Mn) deficiency leads to increased transpiration, barley plants were grown for 10 weeks in hydroponics with daily additions of Mn in the low n M range. The Mn-starved plants did not exhibit visual leaf symptoms of Mn deficiency, but Chl a fluorescence measurements revealed that the quantum yield efficiency of PSII (Fv/Fm) was reduced from 0.83 in Mn-sufficient control plants to below 0.5 in Mn-starved plants. Leaf Mn concentrations declined from 30 to 7 μg Mn g−1 dry weight in control and Mn-starved plants, respectively. Mn-starved plants had up to four-fold higher transpiration than control plants. Stomatal closure and opening upon light/dark transitions took place at the same rate in both Mn treatments, but the nocturnal leaf conductance for water vapour was still twice as high in Mn-starved plants compared with the control. The observed increase in transpiration was substantiated by 13C-isotope discrimination analysis and gravimetric measurement of the water consumption, showing significantly lower water use efficiency in Mn-starved plants. The extractable wax content of leaves of Mn-starved plants was approximately 40% lower than that in control plants, and it is concluded that the increased leaf conductance and higher transpirational water loss are correlated with a reduction in the epicuticular wax layer under Mn deficiency.  相似文献   

16.
Temperate cereals, such as wheat (Triticum spp.) and barley (Hordeum vulgare), respond to prolonged cold by becoming more tolerant of freezing (cold acclimation) and by becoming competent to flower (vernalization). These responses occur concomitantly during winter, but vernalization continues to influence development during spring. Previous studies identified VERNALIZATION1 (VRN1) as a master regulator of the vernalization response in cereals. The extent to which other genes contribute to this process is unclear. In this study the Barley1 Affymetrix chip was used to assay gene expression in barley seedlings during short or prolonged cold treatment. Gene expression was also assayed in the leaves of plants after prolonged cold treatment, in order to identify genes that show lasting responses to prolonged cold, which might contribute to vernalization-induced flowering. Many genes showed altered expression in response to short or prolonged cold treatment, but these responses differed markedly. A limited number of genes showed lasting responses to prolonged cold treatment. These include genes known to be regulated by vernalization, such as VRN1 and ODDSOC2, and also contigs encoding a calcium binding protein, 23-KD jasmonate induced proteins, an RNase S-like protein, a PR17d secretory protein and a serine acetyltransferase. Some contigs that were up-regulated by short term cold also showed lasting changes in expression after prolonged cold treatment. These include COLD REGULATED 14B (COR14B) and the barley homologue of WHEAT COLD SPECIFIC 19 (WSC19), which were expressed at elevated levels after prolonged cold. Conversely, two C-REPEAT BINDING FACTOR (CBF) genes showed reduced expression after prolonged cold. Overall, these data show that a limited number of barley genes exhibit lasting changes in expression after prolonged cold treatment, highlighting the central role of VRN1 in the vernalization response in cereals.  相似文献   

17.
The nature and composition of the external, mucilageneous layer of barley roots was studied by extraction methods and electron microscopy. Barley roots were extracted with chloramphenicol-supplemented water at 35°C, with NH4Cl at various concentrations and with pectinase solutions. The kinetics of transfer of bacteria, total and reducing sugars, proteins, Ca++ and K+ was studied, and the removal of the mucigel from the extracted roots was followed under the electron microscope. Within 2 to 3 hours of treatment with water, the rate of release of sugars, ions, proteinaceous material and bacteria, was reduced to almost zero. Increasing concentrations of ammonium chloride enhanced transfer of ions to the extracting solution but affected sugar extraction to a lesser extent. Electron micrographs of ammonium chloride-extracted roots revealed that the amorphous, rather than the fibrillar fraction of the mucigel was removed. At 103 meq of NH4Cl, distortion of the epidermal layer of the extracted roots was observed. With pectinase as an extractant, there was some enhancement of sugars and ions transfer from the roots to the extracting solution. Electron micrographs showed that the main site of extraction of pectinase was the boundary layer between the root surface and the mucigel. Paper chromatography of the acid hydrolyzate of the water extracted, ethanol-precipitated fraction showed the presence of compounds identical in Rf values to D-glucose, D-arabinose, D-glucuronic acid and D-galacturonic acid. Present methods available for the extraction of the mucigel do not allow the differentiation between extracted pectic compounds which originate from the internal root tissue, and the mucigel. re]19751128 Dept. of Field Crops Dept. of Soil and Water Sciences Dept. of Microbiology and Phytopathology  相似文献   

18.
This study was carried out to compare grains of barley (Hordeum vulgare L. distichon) regarding contents and compositions of 5-n-alkylresorcinols. Mixtures of resorcinol homologues were isolated from acetone extracts from five barley cultivars. These polyketide metabolites were identified by chromatographic and spectroscopic means. The content and homologue patterns among different varieties were similar. The predominant compounds were 1,3-dihydroxy-5-n-heneicosylbenzene (C21:0), 1,3-dihydroxy-5-n-nonadecylbenzene (C19:0) and 1,3-dihydroxy-5-n-pentacosylbenzene (C25:0). The alkylresorcinol concentrations, in contrast to their compositions, depended on environmental and agricultural factors.  相似文献   

19.
A cytological study of barley microspores during pretreatment of the uninucleate stage to the early culture stage was conducted utilizing six genotypes. Among the three main pretreatments investigated, microspores completed the first mitotic division during 28 d cold pretreatment of spikes, with or without leaf sheath attached, and during 0.3 M mannitol pretreatment of anthers at 25 degrees C. However, during a 4 d pretreatment in 0.3 M mannitol at 4 degrees C this first mitotic division was blocked or delayed and subsequently most often occurred during the first day on culture medium. The first mitotic division of most microspores pretreated in 0.3 M mannitol was mostly symmetrical (55-60%), whereas it was asymmetric (94%) during the 28 d cold pretreatment of spikes. Following the first mitotic division during the mannitol pretreatment at 25 degrees C, closely associated daughter nuclei often appeared to fuse via membrane coalescence, leading to a high frequency of large uninucleate microspores. Based upon nuclear size, the frequencies of fused uninucleate microspores in genotypes GBC 778, GBC 777 and Igri were estimated to be 87%, 54% and 75%, respectively, after a 4 d mannitol pretreatment at 25 degrees C. Chromosome numbers in dividing nuclei and relative densitometry measurements of nuclear DNA in microspores from cv. Igri confirmed the apparent fused nature of large nuclei in uninucleate microspores. The high frequency of fused nuclei indicates that nuclear fusion occurred between both symmetric and asymmetric nuclei. Microspores of cv. Igri cultured on filter paper following three different pretreatments provided an average of about 12 000 embryo-like structures (ELS) per plate. In samples, 85-97% of these ELS regenerated green shoots. The frequency of doubled haploids (74-83%) following all pretreatments was similar to the frequencies of fused nuclei. The pretreatment of spikes in 0.3 M mannitol at 4 degrees C for 4 d is preferred as it appears to provide genotype independent induction and suspension of nuclear division, as well as regenerating green plants in a shorter time than cold alone.  相似文献   

20.
We report here that calmodulin isolated from the monocotyledon barley is indistinguishable by a variety of criteria from calmodulin isolated from the dicotyledon spinach. In contrast to previous reports, we find that barley (Hordeum vulgare) calmodulin has an amino acid composition similar to that of vertebrate and spinach calmodulins, including the presence of a single trimethyllysinyl residue, and that barley calmodulin quantitatively activates cyclic nucleotide phosphodiesterase. Furthermore, spinach and barley calmodulins are similar in terms of tryptic peptide maps and immunoreactivity with various antisera that differ in their molecular specificities for calmodulins. Limited amino acid sequence analysis demonstrates that the region around the single histidinyl and trimethyllysinyl residues is identical among barley, spinach, and vertebrate calmodulins and that barley calmodulin, like spinach calmodulin, has a novel glutamine residue at position 96. We conclude that calmodulin is highly conserved among higher plants and that detailed sequence analysis is required before significant differences, if any, can be assigned to barley or other higher plant calmodulins. These studies suggest that calmodulin's fundamental importance to the eukaryotic cell may have been established prior to the evolutionary emergence of higher plants.  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