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 609 毫秒
1.
Excised root cultures of Gloriosa superba reached 7.5 g dry wt l–1 and accumulated 240±40 g colchicine g–1 cell dry wt after 4 weeks growth. While all precursors (except trans-cinnamic acid) enhanced colchicine content of root cultures without adversely affecting root growth, treatment with p-coumaric acid + tyramine (each at 20 mg l–1) increased colchicine content to 1.9 mg g–1 cell dry wt.  相似文献   

2.
Summary Optimal growth conditions for Zymomonas mobilis have been established using continuous cultivation methods. Optimal substrate utilization efficiency occurs with 2.5 g l–1 yeast extract, 2.0 g l–1 ammonium sulfate and 6.0 g l–1 magnesium sulfate in the media. Catabolic activity is at its maximum with glucose uptake rates of 16–18 g l–1 h–1 and ethanol production rates of 8–9 g l–1 h–1, Qg values of 22–26 and Qp values between 11 and 13, which results in 40 g l–1 h–1 ethanol yields using a 100 g l–1 substrate feed. Any increase in these parameters goes on cost of substrate utilization efficiency. Calcium pantothenate can not substitute yeast extract.Abbreviations G Glucose (%) - Pant Calcium pantothenate (mg l–1) - D Dilution rate (h–1) - NH4 Ammonium sulfate (%) - Mg Magnesium sulfate (%) - S1 Residual glucose in the fermenter (g l–1) - S0 Glucose feed (g l–1) - Eth Ethanol concentration (g l–1) - GUR Glucose uptake rate (g l–1 h–1) - Qg Specific glucose uptake rate (g g–1 h–1) - Qp Specific ethanol production rate (g g–1 h–1) - EPR Ethanol production rate (g l–1 h–1) - Yg Yield coefficient for glucose (g g–1) - Yp Conversion efficiency (%) - C Biomass concentration (g l–1) Present address: (Until June 1982) Institut für Mikrobiologie, TH Darmstadt, 6100 Darmstdt, Federal Republic of Germany  相似文献   

3.
The influence of salinity, nutrient level and soil aeration on the transpiration coefficient, defined as amount of water transpired/unit biomass produced (transpiration/biomass ratio) of carrots was investigated under non-limiting conditions with respect to water supply.Under optimum conditions and favorable nutrient supply, the transpiration coefficient amounted to 280–310 g H2O g–1 storage root dry weight (RDW). The transpiration coefficient did not change significantly up to salt concentration of 16 mS cm–1 in the soil solution under otherwise optimum conditions. Higher salt concentrations or low nutrient levels increased the transpiration coefficient to values of 390–540 g H2O g–1 RDW. It is suggested that the transpiration coefficient is not affected by salinity as long as toxic effects and nutrient imbalances do not occur. The transpiration coefficient was not increased by impeded soil aeration. Biomass production was more negatively influenced by adverse soil conditions (salinity, low nutrient level, impeded soil aeration) than was the transpiration coefficient.  相似文献   

4.
Zymomonas mobilis ZM4/AcR (pZB5), a mutant recombinant strain with increased acetate resistance, has been isolated following electroporation of Z. mobilis ZM4/AcR. This mutant strain showed enhanced kinetic characteristics in the presence of 12 g sodium acetate l–1 at pH 5 in batch culture on 40 g glucose, 40 g xylose l–1 medium when compared to ZM4 (pZB5). In continuous culture, there was evidence of increased maintenance energy requirements/uncoupling of metabolism for ZM4/AcR (pZB5) in the presence of sodium acetate; a result confirmed by analysis of the effect of acetate on other strains of Z. mobilis. Nomenclature m Cell maintenance energy coefficient (g g–1 h–1)Maximum overall specific growth rate (1 h–1)Maximum specific ethanol production rate (g g–1 h–1)Maximum specific total sugar utilization rate (g g–1 h–1)Biomass yield per mole of ATP (g mole–1 Ethanol yield on total sugars (g g–1)Biomass yield on total sugars (g g–1)True biomass yield on total sugars (g g–1)  相似文献   

5.
Glucose repressed xylose utilization inCandida tropicalis pre-grown on xylose until glucose reached approximately 0–5 g l–1. In fermentations consisting of xylose (93 g l–1) and glucose (47 g l–1), xylitol was produced with a yield of 0.65 g g–1 and a specific rate of 0.09 g g–1 h–1, and high concentrations of ethanol were also produced (25 g l–1). If the initial glucose was decreased to 8 g l–1, the xylitol yield (0.79 g g–1) and specific rate (0.24 g g–1 h–1) increased with little ethanol formation (<5 g l–1). To minimize glucose repression, batch fermentations were performed using an aerobic, glucose growth phase followed by xylitol production. Xylitol was produced under O2 limited and anaerobic conditions, but the specific production rate was higher under O2 limited conditions (0.1–0.4 vs. 0.03 g g–1 h–1). On-line analysis of the respiratory quotient defined the time of xylose reductase induction.  相似文献   

6.
Summary The transepithelial water permeability in frog urinary bladder is believed to be essentially dependent on the ADH-regulated apical water permeability. To get a better understanding of the transmural water movement, the diffusional water permeability (P d) of the basolateral membrane of urinary bladder was studied. Access to this post-luminal barrier was made possible by perforating the apical membrane with amphotericin B. The addition of this antibiotic increasedP d from 1.12±0.10×10–4 cm/sec (n=7) to 4.08±0.33×10–4 cm/sec (n=7). The effect of mercuric sulfhydryl reagents, which are commonly used to characterize water channels, was tested on amphotericin B-treated bladders. HgCl2 (10–3 m) decreasedP d by 52% andpara-chloromercuribenzoic acid (pCMB) (1.4×10–4 m) by 34%. The activation energy for the diffusional water transport was found to increase from 4.52±0.23 kcal/mol (n=3), in the control situation, to 9.99±0.91 kcal/mol (n=4) in the presence of 1.4×10–4 m pCMB. Our second approach was to measure the kinetics of water efflux, by stop-flow light scattering, on isolated epithelial cells from urinary bladders.pCMB (0.5 or 1.4×10–4 m) was found to inhibit water exit by 91±2%. These data strongly support the existence of proteins responsible for water transport across the basolateral membrane, which are permanently present.  相似文献   

7.
The water status of Fagus sylvatica L. and Quercus petraea (Matt) Liebl. was analysed during a cycle of progressive natural drought in southern Europe. Predawn (Ψpd) and midday water potential were measured in transpiring (Ψleaf) and non-transpiring leaves (Ψxyl). Furthermore, photosynthesis (A), stomatal conductance to water vapour (gs) and sap flow (Fd) were recorded on the same dates. Apparent leaf specific hydraulic conductance in the soil–plant–air continuum (Kh) and whole tree hydraulic conductance (Khsf) were calculated by using the simple analogy of the Ohm’s law. Kh was estimated at different points in the pathway as the ratio between transpiration (E) in the uppermost canopy leaves at midday and the gradient of water potential in the different compartments of the continuum soil–roots–stem–branches–leaves. There was a progressive decrease in water potential measured on non-transpiring leaves at the base of tree crown in both species (Ψlxyl) from the beginning of the growing season to the end of summer. A similar decrease was shown in shoot water potential (Ψuxyl) at the uppermost canopy. Predawn water potential (Ψpd) was high in both species until late July (28 July); afterwards, a significant decrease was registered in F. sylvatica and Q. petraea with minimum values of −0.81±0.03 and −0.75±0.06 MPa, respectively, by 15 September. In both species, leaf specific hydraulic conductance in the overall continuum soil–plant–air (Kh) decreased progressively as water stress increases. Minimum values of Kh and Khsf were recorded when Ψpd was lower. However, Q. petraea showed higher Kh than F. sylvatica for the same Ψpd. The decrease in Kh with water stress was mainly linked to its fall from the soil to the lowermost canopy (Ksrs). Nevertheless, a significant resistance in the petiole–leaf lamina (Kpl) was also recorded because significant differences in all dates were found on Ψ between transpiring and non-transpiring leaves from the same shoot. The decline in Kh was followed by an increase in stomatal control of daily water losses through the decrease of stomatal conductance to water vapour (gs) during the day. It promoted a seasonal increase in the stomatal limitation to carbon dioxide uptake for photosynthesis (A). These facts were more relevant in F. sylvatica, which had concurrently a higher decline in water use at the tree level than Q. petraea. The results showed a strong coupling in F. sylvatica and Q. petraea between processes at leaf and tree level. It may be hypothesised a role of specific hydraulic conductance not only in the regulation of water losses by transpiration but also of carbon uptake.  相似文献   

8.
Summary The relative hydraulic conductivity (k) of xylem and resistance (R) to water flow through trunk, primary roots and branches in Picea abies trees growing under contrasting light conditions were investigated. The xylem permeability to water was measured by forcing 10 mM water solution of KC1 through excised wood specimens. From the values of k, the sapwood transverse area and the length of conducting segments, R of the whole trunk, branches and roots was calculated. The relative conductivity of xylem in open-grown trees exceeded that of shade-grown trees by 1.4–3.1 times, while k was closely correlated with the hydraulically effective radius (R e) of the largest tracheids (R 2 was 0.85–0.94 for open- and 0.51–0.79 for shade-grown trees). Because of both a low k and a smaller sapwood area in shade-grown trees the resistance to water movement through their trunk, roots and branches was many times higher. The distribution of R between single segments of the water-conducting pathway differed considerably in trees from different sites. At high water status the largest share of the total resistance in open- as well as shade-grown trees resides in the apical part of the trunk. The contribution of the branches to total xylem resistance is supposed to increase with developing water deficit.  相似文献   

9.
Summary Water turnover rate (WTR), urine concentration and field metabolic rate (FMR) were examined in house mice, Mus domesticus, permanently inhabiting roadside verge areas and seasonally invading crops in semi-arid wheatlands in South Australia. FMR was approximately proportional to body mass0.5 and mean values varied from 4.8 ml CO2 g–1h–1 (2.9 kJ g–1d–1) in autumn and winter, to 7.0 ml CO2 g–1h–1 (4.2 kJ g–1d–1) in maturing crops during spring. WTR was independent of body mass, indicating that larger mice were selecting a diet containing moister foods. WTR was low in summer and high in winter, and in mice from crops varied from 165 ml l–1 body water d–1 (122 ml kg–1d–1) to 1000 ml l–1d–1 (725 ml kg–1d–1). Seasonal changes in WTR were less extreme on the roadside, where a greater diversity of food was available. In the crops, breeding occurred throughout summer during two of three years, but the population increased only in the one summer when mice had marginally higher WTR. On the roadside breeding and population growth were continuous during summer, except in a drought year. Avcrage urine concentration was inversely related to WTR, and varied from 2.0 to 4.8 Osm l–1. The data indicate that the water conserving abilities of mice equal those of many desert rodents. The water conserving abilities of mice living in crops during summer were fully extended, and in some years aridity limited breeding success and population levels. The degree of moisture stress to which mice are exposed during summer appears to depend not only on rainfall but also on other factors such as availability of food and shelter, and the level of weed infestation in crops.  相似文献   

10.
Salt-induced changes in growth, photosynthetic pigments, various gas exchange characteristics, relative membrane permeability (RMP), relative water content (RWC) and ion accumulation were examined in a greenhouse experiment on eight sunflower (Helianthus annuus L.) cultivars. Sunflower cultivars, namely Hysun-33, Hysun-38, M-3260, S-278, Alstar-Rm, Nstt-160, Mehran-II and Brocar were subjected to non-stress (0 mM NaCl) or salt stress (150 mM NaCl) in sand culture. On the basis of percent reduction in shoot biomass, cvs. Hysun-38 and Nstt-160 were found to be salt tolerant, cvs. Hysun-33, M-3260, S-278 and Mehran-II moderately tolerant and Alstar-Rm and Brocar salt sensitive. Salt stress markedly reduced growth, different gas exchange characteristics such as photosynthetic rate (A), water-use efficiency (WUE) calculated as A/E, transpiration rate (E), internal CO2 concentration (C i) and stomatal conductance (g s) in all cultivars. The effect of 150 mM NaCl stress was non-significant on chlorophyll a and b contents, chlorophyll a/b ratio, RWC, RMP and leaf and root Cl, K+ and P contents; however, salt stress markedly enhanced C i /C a ratio, free proline content and leaf and root Na+ concentrations in all sunflower cultivars. Of all cultivars, cv. Hysun-38 was higher in gas exchange characteristics, RWC and proline contents as compared with the other cultivars. Overall, none of the earlier-mentioned physiological attributes except leaf K+/Na+ ratio was found to be effective in discriminating the eight sunflower cultivars as the response of each cultivar to salt stress appraised using various physiological attributes was cultivar-specific.  相似文献   

11.
Can elevated CO(2) improve salt tolerance in olive trees?   总被引:2,自引:0,他引:2  
We compared growth, leaf gas exchange characteristics, water relations, chlorophyll fluorescence, and Na+ and Cl concentration of two cultivars (‘Koroneiki’ and ‘Picual’) of olive (Olea europaea L.) trees in response to high salinity (NaCl 100 mM) and elevated CO2 (eCO2) concentration (700 μL L−1). The cultivar ‘Koroneiki’ is considered to be more salt sensitive than the relatively salt-tolerant ‘Picual’. After 3 months of treatment, the 9-month-old cuttings of ‘Koroneiki’ had significantly greater shoot growth, and net CO2 assimilation (ACO2) at eCO2 than at ambient CO2, but this difference disappeared under salt stress. Growth and ACO2 of ‘Picual’ did not respond to eCO2 regardless of salinity treatment. Stomatal conductance (gs) and leaf transpiration were decreased at eCO2 such that leaf water use efficiency (WUE) increased in both cultivars regardless of saline treatment. Salt stress increased leaf Na+ and Cl concentration, reduced growth and leaf osmotic potential, but increased leaf turgor compared with non-salinized control plants of both cultivars. Salinity decreased ACO2, gs, and WUE, but internal CO2 concentrations in the mesophyll were not affected. eCO2 increased the sensitivity of PSII and chlorophyll concentration to salinity. eCO2 did not affect leaf or root Na+ or Cl concentrations in salt-tolerant ‘Picual’, but eCO2 decreased leaf and root Na+ concentration and root Cl concentration in the more salt-sensitive ‘Koroneiki’. Na+ and Cl accumulation was associated with the lower water use in ‘Koroneiki’ but not in ‘Picual’. Although eCO2 increased WUE in salinized leaves and decreased salt ion uptake in the relatively salt-tolerant ‘Koroneiki’, growth of these young olive trees was not affected by eCO2.  相似文献   

12.
Miniature heat balance-sap flow gauges were used to measure water flows in small-diameter roots (3–4 mm) in the undisturbed soil of a mature beech–oak–spruce mixed stand. By relating sap flow to the surface area of all branch fine roots distal to the gauge, we were able to calculate real time water uptake rates per root surface area (Js) for individual fine root systems of 0.5–1.0 m in length. Study aims were (i) to quantify root water uptake of mature trees under field conditions with respect to average rates, and diurnal and seasonal changes of Js, and (ii) to investigate the relationship between uptake and soil moisture θ, atmospheric saturation deficit D, and radiation I. On most days, water uptake followed the diurnal course of D with a mid-day peak and low night flow. Neighbouring roots of the same species differed up to 10-fold in their daily totals of Js (<100–2000 g m−2 d−1) indicating a large spatial heterogeneity in uptake. Beech, oak and spruce roots revealed different seasonal patterns of water uptake although they were extracting water from the same soil volume. Multiple regression analyses on the influence of D, I and θ on root water uptake showed that D was the single most influential environmental factor in beech and oak (variable selection in 77% and 79% of the investigated roots), whereas D was less important in spruce roots (50% variable selection). A comparison of root water uptake with synchronous leaf transpiration (porometer data) indicated that average water fluxes per surface area in the beech and oak trees were about 2.5 and 5.5 times smaller on the uptake side (roots) than on the loss side (leaves) given that all branch roots <2 mm were equally participating in uptake. Beech fine roots showed maximal uptake rates on mid-summer days in the range of 48–205 g m−2 h−1 (i.e. 0.7–3.2 mmol m−2 s−1), oak of 12–160 g m−2 h−1 (0.2–2.5 mmol m−2 s−1). Maximal transpiration rates ranged from 3 to 5 and from 5 to 6 mmol m−2 s−1 for sun canopy leaves of beech and oak, respectively. We conclude that instantaneous rates of root water uptake in beech, oak and spruce trees are above all controlled by atmospheric factors. The effects of different root conductivities, soil moisture, and soil hydraulic properties become increasingly important if time spans longer than a week are considered.  相似文献   

13.
S-Adenosyl-l-methionine (AdoMet) was produced by a mutant strain Kluyveromyces lactis AM-65 grown on whey. A full factorial design method of three factors – (NH4)2SO4 (factor x 1), corn steep liquor (factor x 2) and l-methionine (factor x 3) on three levels – was used to determine the optimal medium conditions for the production of AdoMet. A time course shake-flask experiment in optimal whey medium (x 1=3.1 g l–1, x 2=12.7 g l–1, x 3=4.6 g l–1) was also carried out and the results confirmed the results of the factorial design and subsequent quadratic modelling and optimization of AdoMet production which reached 90 mg g–1 cell dry wt.  相似文献   

14.
Zair  I.  Chlyah  A.  Sabounji  K.  Tittahsen  M.  Chlyah  H. 《Plant Cell, Tissue and Organ Culture》2003,73(3):237-244
Somatic embryogenesis through callus initiation has been quantified under salt stress conditions for 8 wheat cultivars currently cultivated in Morocco. The cultivars were classed according to the mean number of somatic embryos formed per immature embryo half and regenerated plants per 100 explants under saline conditions. Regenerated plants from control callus (R0–0) and callus initiated on 10 g l–1 NaCl (R0–10) did not show significant differences concerning plant height, spike length and grain number per ear but, the R0 plants remained less developed than parent plants. When watered with a solution containing more than 20 g l–1 NaCl, the seeds of cultivar Te derived from R0–10 regenerated plants exhibited the best elongation of roots and coleoptiles. Furthermore, a chlorophyll fluorescence test showed a clear improvement in salt tolerance of R0–10 plants at four to five-leaf stage, compared to R0–0 plants. It is concluded that plant regeneration from callus initiated on high NaCl levels may be a valid method of selection for salt tolerance.  相似文献   

15.
In a seasonally dry tropical region the water use efficiency (WUE) of three grasses (C3 winter annualPolypogon monspeliensis, C4 perennialDichanthium annulatum and C4 warm seasonal annualEchinochloa colonum) was evaluated during summer and winter under nine experimental conditions (3 soil moisture×3 herbage removal). Generally leaf water status and transpiration rate decreased with soil moisture stress and increased with clipping intensity. During winter the transpiration rate of Dichanthium was much lower than that of Polypogon and its own rate in summer. Both soil moisture stress and clipping intensity increased the WUE in all instances. Despite differences in photosynthetic type, growing season and life form, these grasses exhibited broadly similar positive relationships, across nine treatments for WUE: soil moisture stress, and water consumption: production. The range of WUE (g. mm–1) calculated on TNP through the nine treatments was: summer—Dichanthium 2.9–10.0, Echinochloa 2.0–6.7; winter—Dichanthium 4.3–36.3, Polypogon 1.9–12.0.  相似文献   

16.
Synopsis Arsenic persists in Chautauqua Lake, New York waters 13 years after cessation of herbicide (sodium arsenite) application and continues to cycle within the lake. Arsenic concentrations in lake water ranged from 22.4–114.81 g l–1, = 49.0 ag l–1. Well water samples generally contained less than 10 g l–1 arsenic. Arsenic concentrations in lake water exceeded U.S. Public Health Service recommended maximum concentrations (10 g l–1) and many samples exceeded the maximum permissible limit (50 g l–1). Fish accumulated arsenic from water but did not magnify it. Fish to water arsenic ratios ranged from 0.4–41.6. Black crappie (Pomoxis nigromaculatus) contained the highest arsenic concentrations (0.14–2.04 g g–1 ), X = 0.7 g g–1) while perch (Perca flavescens), muskellunge (Esox masquinongy) and largemouth bass (Micropterus salmoides) contained the lowest concentrations (0.02–0.13 g g–1). Arsenic concentrations in fish do not appear to pose a health hazard for human consumers.  相似文献   

17.
Single-nodal cuttings of Solanum tuberosum (four cultivars) and Solanum chacoense were induced to produce in vitro microtubers on Murashige & Skoog (MS) medium supplemented with 8 g l–1 sucrose and various concentrations of kinetin and paclobutrazol. The cultures were kept 10 days in darkness and then transferred to a 14 h daylength with 100 µE m–2 sec–1 light intensity at 21 °C. Kinetin (2.5 mg l–1) had no significant influence on tuber formation. However, its addition together with paclobutrazol (0.001 mg l–1) significantly enhanced tuberization. Paclobutrazol alone stimulated early tuber initiation and inhibited stem growth. Despite some genotype × treatment interactions, all genotypes (from very early to late and wild type) formed the maximum proportion of explants bearing microtubers on the media containing both plant growth regulators.  相似文献   

18.
Summary Ultrastructural changes associated with osmotically-induced water transport and water permeability were examined in two flatworm species,Schistosoma mansoni andHymenolepis diminuta. The structure of the surface layer of these parasites is unusual in that it is a syncytial epithelial layer that lacks tight junctions and lateral extracellular spaces. The permeability coefficients observed in this study are therefore necessarily associated only with the transcellular route of transepithelial transport. The ultrastructural changes associated with volume transport across the epithelial syncytium were also unusual in that the basally located channels extending distally from the inward-facing membrane into the syncytial layer remained open regardless of the direction of water flow.Despite the structural differences, most of the features of diffusive (P d ) and osmotic (P osm ) water fluxes across the syncytium resembled those observed in other epithelia: (i) Low water permeability with maximum values of 4.1×10–5 forP d and 9.6×10–5 forP osm.(ii)P osm>P d by 2.0- to 3.2-fold. (iii) Outward water permeability less than inward water permeability. This asymmetry could not be attributed to collapsing channels when net volume transport was directed outward since channels in the syncytium remained open regardless of the direction of water flow. The asymmetry could be explained by tissue contraction or swelling when bathed in anisotonic fluids. (iv)P osm values were not significantly altered by tissue unstirred layers but bothP osm andP d values were underestimated when the bulk fluid was not vigorously stirred.The lower permeability inS. mansoni relative toH. diminuta may be attributed to the membranous surface coat of the former species.  相似文献   

19.
Environmental and physiological regulation of transpiration were examined in several gap-colonizing shrub and tree species during two consecutive dry seasons in a moist, lowland tropical forest on Barro Colorado Island, Panama. Whole plant transpiration, stomatal and total vapor phase (stomatal + boundary layer) conductance, plant water potential and environmental variables were measured concurrently. This allowed control of transpiration (E) to be partitioned quantitatively between stomatal (g s) and boundary layer (g b) conductance and permitted the impact of invividual environmental and physiological variables on stomatal behavior and E to be assessed. Wind speed in treefall gap sites was often below the 0.25 m s–1 stalling speed of the anemometer used and was rarely above 0.5 m s–1, resulting in uniformly low g b (c. 200–300 mmol m–2 s–1) among all species studied regardless of leaf size. Stomatal conductance was typically equal to or somewhat greater than g b. This strongly decoupled E from control by stomata, so that in Miconia argentea a 10% change in g s when g s was near its mean value was predicted to yield only a 2.5% change in E. Porometric estimates of E, obtained as the product of g s and the leaf-bulk air vapor pressure difference (VPD) without taking g b into account, were up to 300% higher than actual E determined from sap flow measurements. Porometry was thus inadequate as a means of assessing the physiological consequences of stomatal behavior in different gap colonizing species. Stomatal responses to humidity strongly limited the increase in E with increasing evaporative demand. Stomata of all species studied appeared to respond to increasing evaporative demand in the same manner when the leaf surface was selected as the reference point for determination of external vapor pressure and when simultaneous variation of light and leaf-air VPD was taken into account. This result suggests that contrasting stomatal responses to similar leaf-bulk air VPD may be governed as much by the external boundary layer as by intrinsic physiological differences among species. Both E and g s initially increased sharply with increasing leaf area-specific total hydraulic conductance of the soil/root/leaf pathway (G t), becoming asymptotic at higher values of G t. For both E and g s a unique relationship appeared to describe the response of all species to variations in G t. The relatively weak correlation observed between g s and midday leaf water potential suggested that stomatal adjustment to variations in water availability coordinated E with water transport efficiency rather than bulk leaf water status.  相似文献   

20.
Khan  W.M.  Prithiviraj  B.  Smith  D.L. 《Photosynthetica》2002,40(4):621-624
On the first day after foliar application, chitosan pentamer (CH5) and chitin pentamer (CHIT5) decreased net photosynthetic rate (P N) of soybean and maize, however, on subsequent days there was an increase in P N in some treatments. CH5 caused an increase in maize P N on day 3 at 10–5 and 10–7 M; the increases were 18 and 10 % over the control plants. This increase was correlated with increases in stomatal conductance (g s) and transpiration rate (E), while the intercellular CO2 concentration (C i) was not different from the control plants. P N of soybean plants did not differ from the control plants except for treatment CH5 (10–7 M) which caused an 8 % increase on day 2, along with increased g s, E, and C i. On days 5 and 6 the CHIT5 treatment caused a 6–8 % increase in P N of maize, which was accompanied by increases in g s, E, and C i. However, there was no such increase for soybean plants treated with CHIT5. In general, foliar application of high molecular mass chitin (CHH) resulted in decreased P N, particularly for 0.010 % treated plants, both in maize and soybean. Foliar applications of chitosan and chitin oligomers did not affect (p > 0.05) maize or soybean height, root length, leaf area, shoot or root or total dry mass.  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