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.
Haploids (2n =24) of the common tetraploid (2n=48) potato (SolanumtuberosumL.) provide promising material for attacking many problemsconcerned with the genetics, cytogenetics and breeding of thisspecies. Interspecific 4xx2xcrosses betweenSolanum tuberosumgp.Andigenaorgp.Tuberosumcultivars as pistillate parents andSolanum tuberosumgp.Phurejaassource of pollen (hereafter ‘pollinator’) have beenused to produce maternally derived haploids through parthenogenesis.This paper discusses the nature of the ‘pollinator’effect in haploid extraction. The ‘pollinator’ hada significant effect on haploid frequencies following 4xx2xcrosses.The ‘pollinator’ effect seems to operate via theendosperm, in which haploid (n=2x) embryos are associated withhexaploid endosperm. A superior ‘pollinator’ appearsto have its effect by contributing two haploid (n) gametes tothe central cell. 2n pollen; double fertilization; endosperm; ploidy manipulations; Solanum tuberosum  相似文献   

2.
Increasing ß-amylase activity in wheat (Triticum aestlvum,var. Star) seedling shoot tissues was consistently accompaniedby the development of a characteristic polymorphism of the enzyme,as shown by electrophoresis employing amylopectin-containingpolyacrylamide gels. Very young shoot tissue contained one principalform of the enzyme (ß1), whereas two other major forms(ß2, ß3) appeared complementary to thisupon further growth. In vitro incubation experiments indicatedthat the polymorphism arose via a probably proteolytic conversionof ß1 into ß2 and ß3. The conversioninvolved neither an activation of ß-amylase nor asignificant modification of ß-amylase component plvalues. The electrophoretic ß-amylase patterns ofsubcellular leaf compartments suggested that ß1 issynthesized in the cytoplasm of leaf mesophyfi cells and thatthe other forms arise upon transfer of this ‘primary’form into the vacuole. The development of shoot ß-amylaseactivity did not require light, but appeared to be under thenegative control of the chloroplast and was stimulated by mineralnutrients. No clear relationship between ß-amylaseactivity and starch metabolism was evident, since the leaf activitywas largely absent from mesophyll protoplasts, could not beunequivocally demonstrated in the mesophyll chioroplasts, anddeveloped regardless of whether the tissues contained significantamounts of starch or not. Key words: Wheat, leaves, ß-amylase, polymorphism, compartmentation  相似文献   

3.
We previously showed that plasma membrane Ca2+-ATPase (PMCA) activity accounted for 25–30% of relaxation in bladder smooth muscle (8). Among the four PMCA isoforms only PMCA1 and PMCA4 are expressed in smooth muscle. To address the role of these isoforms, we measured cytosolic Ca2+ ([Ca2+]i) using fura-PE3 and simultaneously measured contractility in bladder smooth muscle from wild-type (WT), Pmca1+/–, Pmca4+/–, Pmca4–/–, and Pmca1+/–Pmca4–/– mice. There were no differences in basal [Ca2+]i values between bladder preparations. KCl (80 mM) elicited both larger forces (150–190%) and increases in [Ca2+]i (130–180%) in smooth muscle from Pmca1+/– and Pmca1+/–Pmca4–/– bladders than those in WT or Pmca4–/–. The responses to carbachol (CCh: 10 µM) were also greater in Pmca1+/– (120–150%) than in WT bladders. In contrast, the responses in Pmca4–/– and Pmca1+/–Pmca4–/– bladders to CCh were significantly smaller (40–50%) than WT. The rise in half-times of force and [Ca2+]i increases in response to KCl and CCh, and the concomitant half-times of their decrease upon washout of agonist were prolonged in Pmca4–/– (130–190%) and Pmca1+/–Pmca4–/– (120–250%) bladders, but not in Pmca1+/– bladders with respect to WT. Our evidence indicates distinct isoform functions with the PMCA1 isoform involved in overall Ca2+ clearance, while PMCA4 is essential for the [Ca2+]i increase and contractile response to the CCh receptor-mediated signal transduction pathway. PMCA; bladder smooth muscle; gene-altered mice  相似文献   

4.
The calanoid copepod, Eudiaplomus graciloides, was reared fromegg to adult on uni-algal diets (0.1. 0.5 and 2.5 mg dry wt1–1) using the green alga, Chlamydomonas reinhardtii,as food, or on a mixed diet consisting of Lake Esrom water filteredthrough a plankton net with pore size 45 µm and supplementedwith C. reinhardtii (2.5 mg dry wt 1–1). On the mixeddiet at 21.0°C growth in body dry wt (W, µg dry wt)was exponential, and the growth constants were 0.21 day–1in the early to mid juvenile stage (N1 - C4) and 0.11 day–1in the late juvenile to early adult stage (C4-A). At 14.5°Cthe corresponding growth rate constants were 0.10 and 0.08 day–1.Similar growth rates were found at uni-algal concentrationsof 0.5 and 2.5 mg dry wt I–1, and it was argued that thethreshold concentration for growth in Eudiaptomus was closeto 0.1 mg dry wt I–1. The clearance (C, ml h–1)of copepodites was measured on the uni-algal diets. The constantsof the regression (C = aWb) were: a = 0.125, b = 0.858 (2000C. reinhardtii ml–1), a = 0.068, b = 0.849 (10 000), a= 0.028, b = 0.875 (50 000). Ingestion rates were calculatedfrom the clearances and the average algal concentrations. Atthe three food levels the average daily rations were 30, 67and 125% of body dry wt. The respiration rate (R, nl O2 h–1)was measured in individuals reared on the mixed diet. The constantsof the regression (R = aWb) were: a = 4.82, b = 1.07 (nauplii,14.5°C), a = 4.17, b = 0.904 (copepodites and adults, 14.5°C),a = 6.87, b = 0.757 (copepodites and adults, 21.0°C). Nosignificant difference in the respiration rate of copepoditesreared on uni-algal diets and the mixed diet could be demonstrated.Energy budgets were calculated. The assimilation efficiencyand the gross growth efficiency of copepodites decreased markedlywith increasing food concentration, the net growth efficiencyvaried from an average of 0.44 at the lowest algal concentrationto 0.60 on the mixed diet. The results are discussed in relationto previous findings with both freshwater and marine copepods.  相似文献   

5.
Ge XJ  Yu Y  Yuan YM  Huang HW  Yan C 《Annals of botany》2005,95(5):843-851
Background and Aims The desert legume genus Ammopiptanthuscomprises two currently endangered species, A. mongolicus andA. nanus. Genetic variability and genetic differentiation betweenthe two species and within each species were examined. • Methods Inter-simple sequence repeat (ISSR) marker datawere obtained and analysed with respect to genetic diversity,structure and gene flow. • Key Results Despite the morphological similarity betweenA. mongolicus and A. nanus, the two species are geneticallydistinct from each other, indicated by 63 % species-specificbands. Low genetic variability was detected for both populationlevel (Shannon indices of diversity Hpop = 0·106, percentageof polymorphic loci P = 18·55 % for A. mongolicus; Hpop= 0·070, P = 12·24 % for A. nanus) and specieslevel (Hsp = 0·1832, P = 39·39 % for A. mongolicus;Hsp = 0·1026, P = 25·89 % for A. nanus). Moderategenetic differentiation was found based on different measures(AMOVA ST and Hickory B) in both A. mongolicus (0·3743–0·3744)and A. nanus (0·2162–0·2369). • Conclusions The significant genetic difference betweenthe two species might be due to a possible vicariant evolutionaryevent from a single common ancestor through the fragmentationof their common ancestor's range. Conservation strategies forthese two endangered species are proposed.  相似文献   

6.
The seasonal variability of phytoplankton in the EquatorialAtlantic was analysed using Sea-viewing Wide Field-of-view Sensor(SeaWiFS)-derived chlorophyll a (Chl a) concentration data from1998 to 2001, together with in situ Chl a and primary productiondata obtained during seven cruises carried out between 1995and 2000. Monthly averaged SeaWiFS Chl a distributions werein agreement with previous observations in the Equatorial Atlantic,showing marked differences between 10° W in the EasternTropical Atlantic (ETRA) and 25° W in the Western TropicalAtlantic (WTRA) provinces (Longhurst et al. 1995. J. PlanktonRes., 17, 1245–1271). The seasonal cycle of SeaWiFS-derivedChl a concentration calculated for 0–10° S, 0–20°W (ETRA) is consistent with in situ Chl a measurements, withvalues ranging from 0.16 mg m–3, from February to April,to 0.52 mg m–3 in August. Lower variability was observedin 10° N–10° S, 20–30° W (WTRA) whereminimum and maximum concentrations occurred in April (0.15 mgm–3) and in August (0.24 mg m–3), respectively.A significant empirical relationship between depth-integratedprimary production and in situ measured sea surface Chl a wasfound for ETRA, allowing us to estimate the seasonal cycle ofdepth-integrated primary production from SeaWiFS-derived Chla. As for Chl a, this model was verified in a small area ofthe Eastern Equatorial Atlantic (0–10° S, 0–20°W), although in this instance it was not completely able todescribe the magnitude and temporal variability of in situ primaryproduction measurements. The annual euphotic depth-integratedprimary production rate estimated for ETRA by our empiricalmodel was 1.4 Gt C year–1, which represents 16% of theopen ocean primary production estimated for the whole AtlanticOcean.  相似文献   

7.
The relationships between photosynthesis and photosyntheticphoton flux densities (PPFD, P-l) were studied during a red-tideof Dinophysis norvegica (July-August 1990) in Bedford Basin.Dinophysis norvegica, together with other dinoflagellates suchas Gonyaulax digitate, Ceratium tripos, contributed {small tilde}50%of the phytoplankton biomass that attained a maximum of 16.7µg Chla 1 and 11.93 106 total cells I–1.The atomic ratios of carbon to nitrogen for D.norvegica rangedfrom 8.7 to 10.0. The photosynthetic characteristics of fractionatedphytoplankton (>30 µm) dominated by D.norvegica weresimilar to natural bloom assemblages: o (the initial slope ofthe P-l curves) ranged between 0.013 and 0.047 µg C [µgChla]–1 h–1 [µmol m s–1]–1the maximum photosynthetic rate, pBm, between 0.66 and 1.85µg C [µghla]–1 h–1; lk (the photoadaptationindex) from 14 to 69 µ,mol m–2 s–1. Carbonuptake rates of the isolated cells of D.norvegica (at 780 µmolm–2 s–1) ranged from 16 to 25 pg C cell–1h and were lower than those for C.tripos, G.digitaleand some other dinoflagellates. The variation in carbon uptakerates of isolated cells of D.norvegica corresponded with PBmof the red-tide phytoplankton assemblages in the P-l experiments.Our study showed that D.norvegica, a toxigenic dinoflagellate,was the main contributor to the primary production in the bloom.  相似文献   

8.
A method for isolating an f-type cytochrome (Chlorella cytochrome554) from Chlorella vulgaris var. viridis (CHODAT) utilizingN, N-diethylaminoethylcellulose is described. The spectrum ofreduced Chlorella cyt. 554 has absorption maxima at 554 (-band), 524 (ß-band), 417 (SORETband), 352 , 319 and 277 (proteinband). The oxidized form has absorption maxima at 554530 , (ß-band), 412 (SORET band),360 322 and 275 (protein band). Thespectral characteristics resembled other f-type cytochromes,e. g. in the high SORET to -extinction ratio (6.8) and an asymmetric-absorption band (especially at liquid N2 temperature) ; butcharacteristic differences were present. Mitochondria from whitelupine seedlings and sweet potato roots reduced Chlorella cyt.554. From the effects of antimycin A and 2-heptyl-4-hydroxyquinolineN-oxide it appears that Chlorella cyt. 554 was reduced sequentiallybefore cytochrome a+a3 and near the level of the cytochromesof the b type. Oxidation was slow using lupine mitochondriaand nil with sweet potato mitochondria. The oxidation-reductionpotential at pH 7.2 and 30? was 0.35 V. Ascorbate, cysteine,glutathione and Na2S2O4 readily reduced Chlorella cyt. 554.The cytochrome was not autoxidizable and was slowly oxidizedby excess potassium ferricyanide. The reduced form did not reactwith CO and was not adsorbed by IRC-50 or Cellex-P cation exchangers. 1 Temporary address until September 1961: Department of HorticulturalScience, University of California, Los Angeles 24, California,U. S. A. 2 Present address: Plant Industry Station, Pioneering ResearchLaboratory, Marketing Quality Research Division, AgriculturalMarketing Service, Beltsville, Maryland, U. S. A. (Received January 16, 1961; )  相似文献   

9.
Background and Aims Neotyphodium lolii is a fungal endophyteof perennial ryegrass (Lolium perenne), improving grass fitnessthrough production of bioactive alkaloids. Neotyphodium speciescan also affect growth and physiology of their host grasses(family Poaceae, sub-family Pooideae), but little is known aboutthe mechanisms. This study examined the effect of N. lolii onnet photosynthesis (Pn) and growth rates in ryegrass genotypesdiffering in endophyte concentration in all leaf tissues. • Methods Plants from two ryegrass genotypes, Nui D andNui UIV, infected with N. lolii (E+) differing approx. 2-foldin endophyte concentration or uninfected clones thereof (E–)were grown in a controlled environment. For each genotype xendophyte treatment, plant growth rates were assessed as tilleringand leaf extension rates, and the light response of Pn, darkrespiration and transpiration measured in leaves of young (30–45d old) and old (>90 d old) plants with a single-chamber openinfrared gas-exchange system. • Key Results Neotyphodium lolii affected CO2-limited ratesof Pn, which were approx. 17 % lower in E+ than E– plants(P < 0·05) in the young plants. Apparent photon yieldand dark respiration were unaffected by the endophyte (P >0·05). Neotyphodium lolii also decreased transpiration(P < 0·05), but only in complete darkness. There wereno endophyte effects on Pn in the old plants (P > 0·05).E+ plants grew faster immediately after replanting (P < 0·05),but had approx. 10 % lower growth rates during mid-log growth(P < 0·05) than E– plants, but there was noeffect on final plant biomass (P > 0·05). The endophyteeffects on Pn and growth tended to be more pronounced in NuiUIV, despite having a lower endophyte concentration than NuiD. • Conclusions Neotyphodium lolii affects CO2 fixation,but not light interception and photochemistry of Pn. The impactof N. lolii on plant growth and photosynthesis is independentof endophyte concentration in the plant, suggesting that theendophyte mycelium is not simply an energy drain to the plant.However, the endophyte effects on Pn and plant growth are stronglydependent on the plant growth phase.  相似文献   

10.
The thermal dependence of enzyme kinetic parameters has beenpresented as an indicator of species’ thermal optima andtolerance limits. Previous studies suggest the relationshipbetween temperature and the apparent Michaelis–Mentenconstant (Km) of an enzyme system can be used to predict wholeplant success at specific temperatures. The apparent Kmfor glutathionereductase (EC 1.6.4.2; GR) (oxidized glutathione as substrate)extracted from leaves of American sloughgrass (Beckmannia syzigachneSteud.), tufted hairgrass (Deschampsia caespitosa L.), tallfescue (Festuca arundinaceae Schreb. ‘Titan’), andmaize (Zea mays L.), was determined over a range of temperatures(1–40 °C). For all species, minimum apparent KmforGR was observed at 1 °C, and Kmvalues increased as temperatureincreased. The apparent Kmvalues differed among all speciesat the lower temperatures (1–15 °C), but were similarat higher temperatures. The enzyme from tufted hairgrass hadthe lowest apparent Kmat low temperatures (<15 °C), followedin increasing order by American sloughgrass, tall fescue andmaize. Our experimental system failed to reproduce thermal kineticwindow profiles similar to those reported elsewhere. With respectto the enzyme systems reported here, results suggest that thesecool-season grasses can be ranked as more to less eurythermicwithin the temperature range from 1 to 15 °C. Copyright0000 American sloughgrass, Beckmannia syzigachne Steud., tufted hairgrass, Deschampsia caespitosa L., tall fescue, Festuca arundinaceae Schreb. ‘Titan’, Zea mays L., plant competition, temperature stress, kinetics, Michaelis–Menten constant (Km), glutathione reductase  相似文献   

11.
Alpha1,6-fucosyltransferase (Fut8) plays important roles inphysiological and pathological conditions. Fut8-deficient (Fut8–/–)mice exhibit growth retardation, earlier postnatal death, andemphysema-like phenotype. To investigate the underlying molecularmechanism by which growth retardation occurs, we examined themRNA expression levels of Fut8–/– embryos (18.5days postcoitum [dpc]) using a cDNA microarray. The DNA microarrayand real-time polymerase chain reaction (PCR) analysis showedthat a group of genes, including trypsinogens 4, 7, 8, 11, 16,and 20, were down-regulated in Fut8–/– embryos.Consistently, the expression of trypsinogen proteins was foundto be lower in Fut8–/– mice in the duodenum, smallintestine, and pancreas. Trypsin, an active form of trypsinogen,regulates cell growth through a G-protein-coupled receptor,the proteinase-activated receptor 2 (PAR-2). In a cell culturesystem, a Fut8 knockdown mouse pancreatic acinar cell carcinoma,TGP49-Fut8-KDs, showed decreased growth rate, similar to thatseen in Fut8–/– mice, and the decreased growth ratewas rescued by the application of the PAR-2-activating peptide(SLIGRL-NH2). Moreover, epidermal growth factor (EGF)-inducedreceptor phosphorylation was attenuated in TGP49-Fut8-KDs, whichwas highly associated with a reduction of trypsinogens mRNAlevels. The addition of exogenous EGF recovered c-fos, c-jun,and trypsinogen mRNA expression in TGP49-Fut8-KDs. Again, theEGF-induced up-regulation of c-fos and c-jun mRNA expressionwas significantly blocked by the protein kinase C (PKC) inhibitor.Our findings clearly demonstrate a relationship between Fut8and the regulation of EGF receptor (EGFR)-trypsin-PAR-2 pathwayin controlling cell growth and that the EGFR-trypsin-PAR-2 pathwayis suppressed in TGP49-Fut8-KDs as well as in Fut8–/–mice.  相似文献   

12.
Niklas  Karl J. 《Annals of botany》1993,72(5):475-483
Perianth MP, gynoecium MG, and androecium MA dry-weight biomass(in g) of 39 species of perfect flowers was measured. Thesedata were pooled with published data from an additional 51 speciesand used to determine size-dependent variations in (MG and MA)in terms of the hypothesis that the quotient of MG and MA exceeds1·0 for out-breeding (xenogamous) species and less than1·0 for in-breeding (autogamous) species. Ordinary leastsquare regression of the pooled data (n = 90) showed MG = 0·118M0·916P (r2 = 0·884) and MA = 0·186 M0·975P(r2 = 0·865), indicating that the biomass of the gynoeciumproportionally decrease as floral size increases. The exponentsof these regressions indicate that the ratio of gynoecial toandroecial biomass decreased with increasing floral size suchthat comparatively small flowers (MP < 0·0021 g) hadMG/MA > 1·0 (predicted for 'out-breeders') while comparativelylarger flowers (MP > 0·0021 g) had MG /MA < 1·0(predicted for 'in-breeders'). Thus, on average, the type ofbreeding system was a size-dependent phenomenon. To test whether the biomass of a floral organ-type is a legitimateindicator of gender reproductive effort, the biomass (in g)of stamen filaments Mm and anther sacs MAS of 39 species wasdetermined. Least square regression of these data showed MAS= 0·188 M0·854fil (r2 = 0·967), indicatingthat species with larger stamen filaments, on the average, boreproportionally smaller anther sacs and thereby cautioning againstthe uncritical use of the allocation of biomass to floral organ-typeas a strict gauge of gender-function investment. To determine whether the loss of one gender-function resultsin proportional reallocation of biomass to the remaining gender-function,the size-dependency of androecial and gynoecial biomass wasdetermined for a total of 33 perfect and imperfect flowers ofCucumis melo. Regression of the data obtained from perfect flowersyielded MA = 0·402 M1·47P (r2 = 0·898)and MG = 4·63 M1·36P (r2 = 0·842). SinceMG/MA M0·11P , the biomass allocation to the gynoeciumrelative to the androecium decreased with increasing floralsize. This result was consistent with the broad interpecificcomparison based on 90 species with perfect flowers . Regressionof the data for imperfect flowers yielded MA = 0·151M1·02P (r2 = 0·675) and MG = 4·68 M1·47P(r2 = 0·996), indicating a near allometric relation forthe androecium and a strong positive anisometry for the gynoecium.Thus, for flowers of comparable size, a loss of female genderobtains a modest to significant again in androecial biomasswhereas the loss of male gender yields only a slight increasein gynoecial biomass. Collectively, the results of these studies indicate that biomassallocation patterns are size-dependent phenomena whose complexitieshave been largely ignored in the literature.Copyright 1993,1999 Academic Press Allometry, floral biomass, reproduction  相似文献   

13.
The flower-inducing activities in Lemna paucicostata 151 offour major metabolites of benzoic acid (N-benzoyl aspartate,benzyl 6-O-ß-D-apiofuranosyl-O-ß-D-glucopyranoside,O-benzoyl isocitrate and O-benzoyl malate) were measured, andthe effects on the uptake and metabolism of benzoic acid dueto change in the level of the benzoic acid concentration orto the addition of plant hormones were investigated. N-Benzoylaspartate had weak activity, and O-benzoyl isocitrate and malatehad fairly strong activities, while benzyl 6-O-ß-D-apiofuranosyl-ß-D-glucopyranosideshowed no activity. As the concentration of benzoic acid rose,the ratio of N-benzoyl aspartate increased and that of benzyl6-O-ß-D-apiofuranosyl-O-ß-D-glucopyranosidedecreased. GA3 and IAA, inhibitors of flower induction by benzoicacid, seemed to promote conversion to N-benzoyl aspartate insteadof to benzyl 6-O-ß-D-apiofuranosyl-ß-D-glucopyranoside.The conversion to N-benzoyl aspartate was considered to be adetoxification process and that to benzyl 6-O-ß-D-apiofuranosyl-ß-D-glucopyranosidemay be directly related to flower induction in Lemna. (Received November 2, 1987; Accepted January 23, 1988)  相似文献   

14.
The ingestion rates of the copepod, Diacyclops thomasi, on thesoft-bodied rotifer, Synchaeta pectinata, increased 10-fold(0.07–0.77 Synchaeta h–1) over the range 50–250prey l–1. The saturating functional response curve appearedsigmoid but was statistically indistinguishable from a parabola.The response curve was more linear and 10 times lower over thesame range of density when Diacyclops was offered Kerarellacochleans, a species having a stiffened lorica. Diacyclops maximizedits ingestion rate on Synchaeta as a function of the availablegut space. Predation effort, measured as clearance rates, waslinked tentatively to changes in swimming speed of Diacyclopsand was a function of hunger level. Diacyclops, which were starvedfor varying periods of time, increased their ingestion rateson Synchaeta up to a maximum (-3.0 h–1) after 7–10h of food deprivation. The gut passage time of Diacyclops wasestimated to be 7–8 h. Therefore, ingestion rates (andclearance rates) appeared to be strongly correlated to the volumeof food in the gut.  相似文献   

15.
The structure of the zooplankton biotic community and of copepodpopulation in the coastal area of Terra Nova Bay (Ross Sea,Antarctica) was investigated during the 10th Italian AntarcticExpedition (1994/1995). Zooplankton biotic community consistedmainly of pteropods (Limacina helicina and Clione antarctica),Cyclopoid (Oithona similis), Poecilostomatoid (Oncaea curvata)and Calanoid (Ctenocalanus vanus, Paraeuchaeta antarctica, Metridiagerlachei and Stephos longipes) copepods, ostracods, larvalpolychaetes and larval euphausiids. Zooplankton abundance rangedfrom 48.1 ind m–3 to 5968.9 ind m–3, and copepodabundance ranged from 45.2 ind m–3 to 3965.3 ind m–3.The highest peak of zooplankton abundance was observed between25 m and the surface and was mainly due to the contributionof O. similis, O. curvata and C. vanus. Zooplankton biomassranged from 5.28 mg m–3 to 13.04 mg m–3 dry weight;the maximum value was observed between 25 m and the surface.Total lipid content varied from 216.44 to 460.73 mg g–1dry weight.  相似文献   

16.
Three marine phytoplankton species (Skeletonema costatum, Olisthodiscusluteus andGonyaulax tamarensis) were grown in batch culturesat 15°C and a 14:10 L:D cycle at irradiance levels rangingfrom 5 to 450 µEinst m–2 s–1. At each irradiance,during exponential growth, concurrent measurements were madeof cell division, carbon-specific growth rate, photosyntheticperformance (both O2 and POC production), dark respiration,and cellular composition in terms of C, N and chlorophyll a.The results indicate that the three species were similar withrespect to chemical composition, C:N (atomic) = 6.9 ±0.4, photo-synthetic quotient, 1.43 ± 0.09, and photosyntheticefficiency, 2.3 ±0.1 x 10–3 µmol O2 (µgChl a)–1 h–1 (µEinst m–2 s–1)–1.Differences in maximum growth rate varied as the –0.24power of cell carbon. Differences in growth efficiency, werebest explained by a power function of Chl a:C at µ = 0.Compensation intensities, ranged from 1.1 µEinst m–2s–1 for S. costatum to 35 forG. tamarensis and were foundto be a linear function of the maintenance respiration rate.The results indicate that interspecific differences in the µ–Irelationship can be adequately explained in terms of just threeparameters: cell carbon at maximum growth rate, the C:Chl aratio (at the limit as growth approaches zero) and the respirationrate at zero growth rate. A light-limited algal growth modelbased on these results gave an excellent fit to the experimentalµ–I curves and explained 97% of the observed interspecificvariability. 1Present address: Lamont-Doherty Geological Observatory Columbiaof University, Palisades, NY 10964, USA  相似文献   

17.
The cell quotas of microcystin (Qmcyst), protein (Qprot), chlorophylla (Qchloro) and carbohydrate (Qcarbo), as well as the net productionrates of these parameters, were determined during the exponentialand stationary phases in nine batch cultures of Microcystisaeruginosa (CYA 228) at light regimes from 33 to 53 µmolphotons m–2 s–1. The following results were obtained.(i) A parallel pattern was found in the changes of Qmcyst, Qprot,Qchloro and Qcarbo during the entire growth cycle and significantcorrelations were recorded between Qmcyst and Qprot, Qchloroand Qcarbo. (ii) The net microcystin production rate (µmcyst)was positively correlated with the specific cell division rate(µc), the chlorophyll production rate (µchloro)and the protein production rate. (iii) A significant inverselinear relationship was found between µc and Qmcyst, i.e.cultures with a positive µc had a Qmcyst between 110 and400 fg microcystin cell–1, while declining cultures hadQmcyst values >400 fg microcystin cell–1. Maximum variationin Qmcyst within cultures was 3.5-fold. Collectively, the resultsshow that cells produced microcystin at rates approximatingthose needed to replace losses to daughter cells during divisionand that microcystin was produced in a similar way to proteinand chlorophyll, indicating a constitutive microcystin production.  相似文献   

18.
Analysis of the Promoter of the Auxin-Inducible Gene, parC, of Tobacco   总被引:2,自引:0,他引:2  
The auxin-responsive region (AuxRR) in the promoter of the parCgene was analyzed in transgenic tobacco plants in which the5' flanking region of the parC promoter was placed upstreamof the gene for rß-glucuronidase (GUS). The AuxRRwas located between nucleotides (nt) –226 and –54.Detailed dissection of this segment revealed that the presenceof the non-contiguous sequences from nt –226 to –151and from nt –84 to –54 was required for the expressionof the auxin responsiveness of the parC promoter. The sequencefrom nt –226 to –151 was found to contain a sequencewhich resembles the as-1 element in the 35S promoter of cauliflowermosaic virus (CaMV). Although it has been reported that theas-1 element is involved in auxin responsiveness [Liu and Lam(1994) J. Biol. Chem. 269: 668], we showed that introductionof a point mutation into the as-1-like sequence completely eliminatedauxin responsiveness, a result that suggests that the sequenceis indispensable for auxin responsiveness. However, the presenceof the as-1-like sequence alone was not sufficient for auxinresponsiveness, since the segment (nt –226 to –84)that included the as-1-like sequence failed to confer auxinresponsiveness on the core promoter. It is possible that thetwo separately located sequences play specific roles in interactionswith trans-factors that are required for the expression of theauxin responsiveness of the parC promoter. (Received March 11, 1996; Accepted July 9, 1996)  相似文献   

19.
A series of experiments was conducted to assess net CO2assimilationand growth responses to waterlogging of grafted and seedlingtrees in the genus Annona. Seedlings of A. glabra, A. muricataandA. squamosa L., and scions of ‘Gefner’ atemoya(A. squamosaxA. cherimola Mill.), ‘49-11’ (‘Gefner’atemoyaxA. reticulata L.), ‘4-5’ (‘Priestley’atemoyaxA. reticulata), A. reticulata grafted onto either A.glabra, A. reticulata orA. squamosa rootstocks were floodedfor up to 60 d. Soil anaerobiosis occurred on the third dayof flooding. Seedlings ofA. glabra and A. muricata, and thescions ‘49-11’, ‘Gefner’ atemoya, andA. reticulata grafted onto A. glabra rootstock were consideredflood tolerant based on their ability to survive and grow inflooded conditions. Scions of the normally flood-sensitive A.reticulata, ‘Gefner’ atemoya, and ‘49-11’tolerated root waterlogging when grafted onto the flood-tolerantspecies, A. glabra. In contrast, flooding of A. squamosa seedlingsand rootstocks, and A. reticulata rootstocks greatly reducedgrowth and net CO2assimilation rates, and resulted in 20–80%tree mortality. Stem anatomical responses to long-term flooding(12 continuous months) were assessed in seedlings of A. glabraand A. muricata, and trees of ‘49-11’ grafted ontoA. glabra. Flooded trees developed hypertrophied stem lenticels,particularly in A. glabra, and enlarged xylem cells resultingin thicker stems with reduced xylem density. Flooding did notincrease air spaces in pre-existing xylem near the pith or inxylem tissue that was formed during flooding. Thus, flood tolerancedid not involve aerenchyma formation in the stem. Copyright1999 Annals of Botany Company Flood tolerance, net CO2assimilation, photosynthesis, stem anatomy, shoot growth, anaerobiosis, Annonaceae.  相似文献   

20.
Factorial combinations of four photoperiods (10, 11·33,12·66 and 16 h d-1) and three mean diurnal temperatures(20·2, 24·1 and 28·1°C) were imposedon nodulated plants of three Nigerian bambara groundnut genotypes[Vigna subterranea (L.) Verdc., syn. Voandzeia subterranea (L.)Thouars] grown in glasshouses in The Netherlands. The photothermalresponse of the onset of flowering and the onset of poddingwere determined. The time from sowing to first flower (f) wasdetermined by noting the day on which the first open flowerappeared. The time from sowing to the onset of podding (p) wasestimated from linear regressions of pod dry weight againsttime from sowing. Developmental rates were derived from thereciprocals of f and p. In two genotypes, 'Ankpa 2' and 'Yola',flowering occurred irrespective of photoperiod and 1/f was controlledby temperature only, occurring sooner at 28·1 than at20·2°C. The third genotype, 'Ankpa 4', was sensitiveto temperature and photoperiod and f was increased by coolertemperatures and photoperiods > 12·66 h d-1 at 20·2°Cand > 11·33 h d-1 at 24·1 and 28·1°C.In contrast, p was affected by temperature and photoperiod inall three genotypes. In bambara groundnut photoperiod-sensitivitytherefore increases between the onset of flowering and the onsetof podding. The most photoperiod-sensitive genotype with respectto p was 'Ankpa 4', followed by 'Yola' and 'Ankpa 2'. Therewas also variation in temperature-sensitivity between the genotypesinvestigated. Evaluation of bambara groundnut genotypes foradaptation to different photothermal environments will thereforerequire screening for flowering and podding responses.Copyright1994, 1999 Academic Press Vigna subterranea (L.) Verdc., Voandzeia subterranea (L.) Thouars, bambara groundnut, phenology, photoperiod, daylength, temperature, flowering, podding  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