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.
Summary This study determined how structural features of the eggshells of coots (Fulica americana) laid at 4150 m in the Peruvian Andes differed from those at sea level in Peru and California and how these features affected exchange of water vapor, O2, and CO2. While barometric pressure at 4150 m was reduced to 60% of that at sea level, the conductance to water vapor, corrected to 760 torr, of montane eggs was 107% of the corresponding lowland value. When the effect of low barometric pressure on the diffusion coefficient of gases was considered, the effective conductance of the montane eggs at altitude was 177% of that at sea level. As a result, daily water loss from the montane eggs was substantially greater than that from lowland ones. The oxygen consumption of montane embryos was lower than that of lowland embryos of all sizes, particularly at larger embryonic masses. Just before pipping, the oxygen consumption of montane embryos was about 60% of the corresponding value for lowland individuals. Air cell oxygen tensions in montane eggs varied between about 65 and 38 torr; these values were about 60–70 torr below those in lowland eggs at equivalent embryonic masses. Just before pipping, the air cell CO2 tension of montane eggs was about 20 torr below levels in sea level eggs. The eggshell conductance to gases of montane eggs appears to have been selected to promote oxygen delivery at the cost of increased losses of water vapor and CO2.  相似文献   

2.
Over the past decade, the concept of isohydry or anisohydry, which describes the link between soil water potential (ΨS), leaf water potential (ΨL), and stomatal conductance (gs), has soared in popularity. However, its utility has recently been questioned, and a surprising lack of coordination between the dynamics of ΨL and gs across biomes has been reported. Here, we offer a more expanded view of the isohydricity concept that considers effects of vapour pressure deficit (VPD) and leaf area index (AL) on the apparent sensitivities of ΨL and gs to drought. After validating the model with tree‐ and ecosystem‐scale data, we find that within a site, isohydricity is a strong predictor of limitations to stomatal function, though variation in VPD and leaf area, among other factors, can challenge its diagnosis. Across sites, the theory predicts that the degree of isohydricity is a good predictor of the sensitivity of gs to declining soil water in the absence of confounding effects from other drivers. However, if VPD effects are significant, they alone are sufficient to decouple the dynamics of ΨL and gs entirely. We conclude with a set of practical recommendations for future applications of the isohydricity framework within and across sites.  相似文献   

3.
Sorrell  B.K.  Brix  H. 《Plant and Soil》2003,253(1):71-79
Internal pressurization and convective gas flow in emergent wetland plants is a function of the water vapour pressure deficit (WPD) and stomatal conductance (G s) separating the external atmosphere from the internal aerenchyma. We have compared the effects of WPD and G s under a range of light intensities on static pressures and convective flows in Phragmites australis, Typha orientalis and Baumea articulata. The capacity of the three species to generate flows per unit leaf area differed, being greatest in P. australisand lowest in B. articulata. In all three species, decreasing light intensity from full sunlight (2200 mol m–2 s–1 photosynthetically active photon flux density (PPFD)) to < 200 and < 10 mol m–2 s–1PPFD caused immediate decreases in photosynthetic assimilation, followed by more gradual decreases in transpiration and G s. However, internal pressures and flows in the two low light intensities remained similar to values recorded in full sunlight. WPD was more significantly related to pressures and flows in P. australis and T. orientalis than G s. In B. articulata, pressures increased at low G s values but flow rates were unaffected, as predicted by earlier models describing pore size effects on pressures and flows. The data suggest that emergent macrophytes can maintain significant internal convection even at low light intensities, and this may be beneficial for nocturnal aeration, particularly in arid climates where the atmospheric humidity at night is low.  相似文献   

4.
The experiments and simulations reported in this paper show that, for stomata sensitive to both CO2 and water vapour concentrations, responses of stomatal conductance (gws) to boundary layer thickness have two components, one resulting from changes in intercellular CO2 concentration (χci) and another from changes in leaf surface water vapour saturation deficit (Dws). The experiments and simulations also show that the boundary layer conductance (gwb) can significantly alter the apparent response of gws to ambient air CO2 mole fraction (χca) and water vapour mole fraction (χwa). Because of the feedback loop involved the responses of gws for χca and χwa each include responses to both χci and Dws. The boundary layer alters the state of the variables sensed by the guard cells—i.e. χci and Dws—and so it is a source of feedback. Thus, when scaling up from responses of stomata to the response of gws for a whole leaf, the effect of the boundary layer must be considered. The results indicate that, for given responses of gws to χci and Dws, the apparent responses of gws to Dwa and χca depend on the size of the leaf and wind speed, showing that this effect of the boundary layer should be considered when comparing data measured under different conditions, or with different methods.  相似文献   

5.
Abstract. Seedlings of Eucalyptus pauciflora Sieb. ex. Spreng. were stressed by withholding water. They were then rewatered, and the time course of recovery of photosynthesis was followed. Recovery always followed a distinct bi-phasic pattern. A first, rapid, stage of recovery commenced between 5 and 60 min after rewatering and was completed by between 30 min and four h after rewatering. Recovery in this stage always involved concurrent increases in stomatal conductance and the leaf's capacity to assimilate CO2 at any intercellular partial pressure of CO2 [ A(pi) relationship]. This stage of rapid recovery was followed either by a constant or gradually declining rate of photosynthesis for the remainder of the light period. In plants kept to a normal diurnal cycle, a second stage of recovery occurred and was completed during the night following rewatering. In this second stage of recovery, the A(Pi) relationship recovered to 90–100% of prestress values. In contrast, the recovery of stomatal conductance was not complete by the first day after rewatering. In darkness, complete recovery of the A(pi) relationship required as little as five h. If plants were kept in continuous high light, then between six and 16 h elapsed after rewatering before the second stage of recovery commenced. After this lag, almost complete recovery of the A(pi) relationship was possible. These results indicate that water stress has two independent and parallel effects on the mesophyll capacity for photosynthesis. The first may be simply reversible when the plant is rewatered, while the second may involve damage to the photosynthetic machinery that requires protein synthesis for its reversal.  相似文献   

6.
Abstract. Stomatal conductance and needle water potential of P. radiata clones were measured after 2, 5 and 8 months on plants grown in controlled environment rooms with markedly different water vapour saturation deficits (D). Conductance was significantly lower at high D, but water potential differences between treatments were not significant. When trees were moved between treatments most of the changes in conductances occurred within 2 h, with residual changes after 24 h. Water potentials were not different 24 h after the trees were moved. The effects were completely reversible.
Transpiration rates of individual trees were highest in the high D treatment and lowest in the low D treatment. They were not linearly related to D because of decreasing conductance with increasing D.
Height growth, diameter growth and foliage areas were not significantly different between treatments. Tracheid lumen diameters tended to be larger in trees grown at higher D although treatment differences were not significant.
There were significant clonal differences in shoot conductance and tracheid dimensions.  相似文献   

7.
We measured the diurnal changes in net photosynthetic rate (P N) and stomatal conductance (g s) of the leaves of a liana, Enkleia malaccensis Griff. (Thymelaeaceae), at the canopy level in the lowland tropical rainforest at Pasoh, Peninsular Malaysia. The measurements were made from a canopy walkway system, 30 m from the ground for 3 d in March 2003. P N increased with increasing photosynthetically active radiation (PAR) before noon, though P N was not enhanced by the strong radiation hit in the afternoon. Plotting g s at saturating PAR (>0.5 mmol m−2 s−1) against the vapour pressure deficit (VPD) failed to reveal a significant correlation between VPD and g s, and g s became very low at VPD >2.5 kPa. The relationship between P N and g s was fitted on the same regression line irrespective of measuring day, indicating that this relationship was not influenced by either VPD or leaf temperature (T L). Therefore, in the liana E. malaccensis, an increase in VPD leads to partial stomatal closure and, subsequently, reductions in P N and the midday depression of P N of this plant.  相似文献   

8.
Numerous gross morphological attributes are shared among unrelated free‐living bryozoans revealing convergent evolution associated with functional demands of living on soft sediments. Here, we show that the reproductive structures across free‐living groups evolved convergently. The most prominent convergent traits are the collective reduction of external brood chambers (ovicells) and the acquisition of internal brooding. Anatomical studies of four species from the cheilostome genera Cupuladria and Discoporella (Cupuladriidae) show that these species incubate their embryos in internal brooding sacs located in the coelom of the maternal nonpolymorphic autozooids. This sac consists of a main chamber and a narrow neck communicating to the vestibulum. The distal wall of the vestibulum possesses a cuticular thickening, which may further isolate the brood cavity. The presence of this character in all four species strongly supports grouping Cupuladria and Discoporella in one taxon. Further evidence suggests that the Cupuladriidae may be nested within the Calloporidae. Based on the structure of brooding organs, two scenarios are proposed to explain the evolution of the internal brooding in cupuladriids. The evolution of brood chambers and their origin in other free‐living cheilostomes is discussed. Unlike the vast majority of Neocheilostomina, almost all free‐living cheilostomes possess nonprominent chambers for embryonic incubation, either endozooidal and immersed ovicells or internal brooding sacs, supporting the idea that internal embryonic incubation is derived. We speculate that prominent skeletal brood chambers are disadvantageous to a free‐living mode of life that demands easy movement through sediment in instable sea‐floor settings. J. Morphol., 2009. © 2009 Wiley‐Liss, Inc.  相似文献   

9.
Do stomata respond to relative humidity?   总被引:24,自引:12,他引:12  
  相似文献   

10.
Perturbations of the source-sink balances were performed in field-grown coffee (Coffea arabica) trees to investigate the possible role of carbohydrates in feedback regulation of photosynthesis. Four treatments were applied at the whole-plant level: (i) complete defruiting and maintenance of the full leaf area, (ii) the half crop load and full leaf area, (iii) the full crop load and full leaf area and (iv) the full crop load and half leaf area. Sampling and measurements were performed twice during the phase of dry matter accumulation of fruits. Gas exchange, chlorophyll a fluorescence, carbon isotope labelling and steady-state metabolite measurements were assessed in source leaves. The average rate of net photosynthetic rate (A) and stomatal conductance (g(s)) were larger (> 50%), and carbon isotope composition ratio was lower, in trees with a full crop load and half leaf area than in defruited trees, with individuals of the other two treatments showing intermediate values. However, differences in A seem unlikely to have been caused either by photochemical impairments or a direct end-product-mediated feedback down-regulation of photosynthesis. It is proposed that the decreased A in defruited coffee trees was independent of carbon metabolism and was rather directly related to a lower CO(2) availability coupled to lower g(s).  相似文献   

11.
Concurrent measurements of gas exchange, instantaneous isotope discrimination (Δ) against 13CO2 and C18O16O, and extent of 18O enrichment in H2O at the evaporative sites, were followed in a tropical forest pioneer, Piper aduncum, on two different days in Trinidad during February 1995. Δ13CO2 differed from that predicted from measurements of internal:external CO2 concentration (Ci/Ca) and showed a wide range of values which decreased throughout the course of the day. Derivation of Cc (the CO2 concentration at the carboxylation site) was not possible using carbon isotope discrimination under field conditions in situ and was derived assuming a constant value of internal transfer conductance (gw). Under low rates of assimilation the derived Cc/Ca, like Ci/Ca, remained relatively stable over the course of both days and ΔC18O16O followed evaporative demand. Lower values of ΔC18O16O on day 2 occurred in response to the indirect effect of increased leaf-to-air vapour pressure deficits (VPD) and reduced stomatal conductance. For the first time, direct determination of the δH218O of transpired water vapour (δt) allowed derivation of evaporative site enrichment without the prerequisite of isotopic steady state (ISS) defined in the Craig and Gordon model. Generally, δt was less enriched than the source water (δs) in the morning and more enriched in the afternoon, which would be predicted from an increase and decrease in ambient VPD, respectively. On both days, leaves of P. aduncum approached ISS (indicated where δtδs) between 1300 and 1500 h. Evaporative site enrichment was maintained into the late afternoon, despite a decrease in ambient VPD. The data presented provide a greater insight into the natural variation in isotopic discrimination under field conditions, which may help to refine models of terrestrial biome discrimination.  相似文献   

12.
The hydraulic conductance of the leaf lamina (Klamina) substantially constrains whole‐plant water transport, but little is known of its association with leaf structure and function. Klamina was measured for sun and shade leaves of six woody temperate species growing in moist soil, and tested for correlation with the prevailing leaf irradiance, and with 22 other leaf traits. Klamina varied from 7.40 × 10?5 kg m?2 s?1 MPa?1 for Acer saccharum shade leaves to 2.89 × 10?4 kg m?2 s?1 MPa?1 for Vitis labrusca sun leaves. Tree sun leaves had 15–67% higher Klamina than shade leaves. Klamina was co‐ordinated with traits associated with high water flux, including leaf irradiance, petiole hydraulic conductance, guard cell length, and stomatal pore area per lamina area. Klamina was also co‐ordinated with lamina thickness, water storage capacitance, 1/mesophyll water transfer resistance, and, in five of the six species, with lamina perimeter/area. However, for the six species, Klamina was independent of inter‐related leaf traits including leaf dry mass per area, density, modulus of elasticity, osmotic potential, and cuticular conductance. Klamina was thus co‐ordinated with structural and functional traits relating to liquid‐phase water transport and to maximum rates of gas exchange, but independent of other traits relating to drought tolerance and to aspects of carbon economy.  相似文献   

13.
Nogueira  A.  Martinez  C.A.  Ferreira  L.L.  Prado  C.H.B.A. 《Photosynthetica》2004,42(3):351-356
Leaf gas exchange characteristics were measured in twenty woody species that differ in succession status ranging from pioneer species (PS) to late succession species (LS) in a Brazilian rain-reforestation ecosystem. Photon-saturated photosynthetic rate, calculated per either a leaf area (P NA) or a dry mass (P NM) basis, differed among species. P NA and P NM were highest in PS and lowest in LS. Variation among species was 3-fold (from 7 to 23 mol m–2 s–1) for P NA, and 5-fold (from 50 to 275 mol kg–2 s–1) for P NM. The highest P NA (23 mol m–2 s–1) and P NM (275 mol kg–2 s–1) values were recorded in PS Croton urucurana, while the lowest P NA (7 mol m–2 s–1) and P NM (50 mol kg–2 s–1) values were recorded in LS Aspidosperma cylindrocarpon. A considerable overlap was recorded between PS and LS in values of stomatal conductance (g s), transpiration rate (E), and leaf mass to area ratio (ALM). However, C. urucurana also showed highest g s and E. P NM was highly correlated with ALM in both PS and LS (r=–0.75 and –0.90, respectively). The high values of instantaneous transpiration efficiency (ITE) and intrinsic water use efficiency (WUEi) were also observed in the PS when compared with the LS.  相似文献   

14.
Measuring leaf gas exchange from canopy leaves is fundamental for our understanding of photosynthesis and for a realistic representation of carbon uptake in vegetation models. Since canopy leaves are often difficult to reach, especially in tropical forests with emergent trees up to 60 m at remote places, canopy access techniques such as canopy cranes or towers have facilitated photosynthetic measurements. These structures are expensive and therefore not very common. As an alternative, branches are often cut to enable leaf gas exchange measurements. The effect of branch excision on leaf gas exchange rates should be minimized and quantified to evaluate possible bias. We compared light-saturated leaf net photosynthetic rates measured on excised and intact branches. We selected branches positioned at three canopy positions, estimated relative to the top of the canopy: upper sunlit foliage, middle canopy foliage, and lower canopy foliage. We studied the variation of the effects of branch excision and transport among branches at these different heights in the canopy. After excision and transport, light-saturated leaf net photosynthetic rates were close to zero for most leaves due to stomatal closure. However, when the branch had acclimated to its new environmental conditions—which took on average 20 min—light-saturated leaf net photosynthetic rates did not significantly differ between the excised and intact branches. We therefore conclude that branch excision does not affect the measurement of light-saturated leaf net photosynthesis, provided that the branch is recut under water and is allowed sufficient time to acclimate to its new environmental conditions.  相似文献   

15.
Stomatal responsiveness to vapour pressure deficit (VPD) results in continuous regulation of daytime gas‐exchange directly influencing leaf water status and carbon gain. Current models can reasonably predict steady‐state stomatal conductance (gs) to changes in VPD but the gs dynamics between steady‐states are poorly known. Here, we used a diverse sample of conifers and ferns to show that leaf hydraulic architecture, in particular leaf capacitance, has a major role in determining the gs response time to perturbations in VPD. By using simultaneous measurements of liquid and vapour fluxes into and out of leaves, the in situ fluctuations in leaf water balance were calculated and appeared to be closely tracked by changes in gs thus supporting a passive model of stomatal control. Indeed, good agreement was found between observed and predicted gs when using a hydropassive model based on hydraulic traits. We contend that a simple passive hydraulic control of stomata in response to changes in leaf water status provides for efficient stomatal responses to VPD in ferns and conifers, leading to closure rates as fast or faster than those seen in most angiosperms.  相似文献   

16.
The theoretical basis for the link between the leaf exchange of carbonyl sulfide (COS), carbon dioxide (CO(2)) and water vapour (H(2)O) and the assumptions that need to be made in order to use COS as a tracer for canopy net photosynthesis, transpiration and stomatal conductance, are reviewed. The ratios of COS to CO(2) and H(2)O deposition velocities used to this end are shown to vary with the ratio of the internal to ambient CO(2) and H(2)O mole fractions and the relative limitations by boundary layer, stomatal and internal conductance for COS. It is suggested that these deposition velocity ratios exhibit considerable variability, a finding that challenges current parameterizations, which treat these as vegetation-specific constants. COS is shown to represent a better tracer for CO(2) than H(2)O. Using COS as a tracer for stomatal conductance is hampered by our present poor understanding of the leaf internal conductance to COS. Estimating canopy level CO(2) and H(2)O fluxes requires disentangling leaf COS exchange from other ecosystem sources/sinks of COS. We conclude that future priorities for COS research should be to improve the quantitative understanding of the variability in the ratios of COS to CO(2) and H(2)O deposition velocities and the controlling factors, and to develop operational methods for disentangling ecosystem COS exchange into contributions by leaves and other sources/sinks. To this end, integrated studies, which concurrently quantify the ecosystem-scale CO(2), H(2)O and COS exchange and the corresponding component fluxes, are urgently needed.  相似文献   

17.
Close coordination between leaf gas exchange and maximal hydraulic supply has been reported across diverse plant life forms. However, it has also been suggested that this relationship may become weak or break down completely within the angiosperms. We examined coordination between hydraulic, leaf vein, and gas‐exchange traits across a diverse group of 35 evergreen Australian angiosperms, spanning a large range in leaf structure and habitat. Leaf‐specific conductance was calculated from petiole vessel anatomy and was also measured directly using the rehydration technique. Leaf vein density (thought to be a determinant of gas exchange rate), maximal stomatal conductance, and net CO2 assimilation rate were also measured for most species (n = 19–35). Vein density was not correlated with leaf‐specific conductance (either calculated or measured), stomatal conductance, nor maximal net CO2 assimilation, with r2 values ranging from 0.00 to 0.11, P values from 0.909 to 0.102, and n values from 19 to 35 in all cases. Leaf‐specific conductance calculated from petiole anatomy was weakly correlated with maximal stomatal conductance (r2 = 0.16; P = 0.022; n = 32), whereas the direct measurement of leaf‐specific conductance was weakly correlated with net maximal CO2 assimilation (r2 = 0.21; P = 0.005; n = 35). Calculated leaf‐specific conductance, xylem ultrastructure, and leaf vein density do not appear to be reliable proxy traits for assessing differences in rates of gas exchange or growth across diverse sets of evergreen angiosperms.  相似文献   

18.
Diurnal depression of leaf hydraulic conductance in a tropical tree species   总被引:10,自引:2,他引:8  
Diurnal patterns of hydraulic conductance of the leaf lamina (Kleaf) were monitored in a field‐grown tropical tree species in an attempt to ascertain whether the dynamics of stomatal conductance (gs) and CO2 uptake (Aleaf) were associated with short‐term changes in Kleaf. On days of high evaporative demand mid‐day depression of Kleaf to between 40 and 50% of pre‐dawn values was followed by a rapid recovery after 1500 h. Leaf water potential during the recovery stage was less than ?1 MPa implying a refilling mechanism, or that loss of Kleaf was not linked to cavitation. Laboratory measurement of the response of Kleaf to Ψleaf confirmed that leaves in the field were operating at water potentials within the depressed region of the leaf ‘vulnerability curve’. Diurnal courses of Kleaf and Ψleaf predicted from measured transpiration, xylem water potential and the Kleaf vulnerability function, yielded good agreement with observed trends in both leaf parameters. Close correlation between depression of Kleaf, gs and Aleaf suggests that xylem dysfunction in the leaf may lead to mid‐day depression of gas exchange in this species.  相似文献   

19.
Three- and four-year-old potted, greenhouse-grown cedar seedlings were subjected to two different watering regimes: half received full water supply and the other half was submitted to moderate drought (50% of the full water supply). Height growth was the greatest for C. atlantica and the most-limited for C. brevifolia in the well-watered set. However, in the dry set, height growth was less affected by drought conditions for C. brevifolia than for C. atlantica. Cedrus libani gave intermediate results for both watering regimes. Moderate drought provoked a decrease in osmotic potential at full leaf turgor and a long-lasting osmotic adjustment. When irrigation was withheld completely to induce severe soil drying, gas exchange decreased and then stopped at predawn water potentials of −3.0 MPa for C. brevifolia, between −2.6 and −2.8 MPa for C. libani, and at −2.4 MPa for C. atlantica, irrespective of watering regime. For all species, the dry set showed lower net photosynthesis (A) and stomatal conductance (g s) than the plants in the well-watered set. A and g s responded to variations in atmospheric water-vapour pressure deficit (VPD). As VPD increased, A and g s decreased, and this trend was proportionate to initial values at low VPD, but remained independent of previous watering treatments, plant water status or species. To conclude, C. brevifolia appears to be a species with limited growth potential but strong soil drought tolerance whereas C. atlantica has strong growth potential when an adequate water supply is available but is more sensitive to soil drought. C. libani shows an intermediate behaviour for growth and drought tolerance.  相似文献   

20.
The effects of the ammonium (NH4+) and nitrate (NO3-) forms of nitrogen and NaCl on the growth, water relations and photosynthesis performance of sunflower (Helianthus annuus L.) were examined under glasshouse conditions. Eight-day-old plants of cv. Hisun 33 were subjected for 21 days to Hoagland's nutrient solution containing 8 mol m-3N as NH4+or NO3-, and salinised with 0, 60, or 120 mol m-3NaCl. Fresh weights of shoots and roots, and leaf area of NO3-supplied non-salinised plants were significantly greater than those of NH4+-supplied non-salinised plants. But addition of NaCl to the rooting medium of these plants had more inhibitory effect on the growth of NO3--supplied plants than on NH4+-supplied plants. Both leaf water and osmotic potentials of plants grown with NH4+were lower than those of plants given NO3-under both non-saline and saline conditions. Chlorophylls a and b concentrations were higher in plants grown with NH4+than N03--supplied plants at the lower two levels of salinisation. The rate of photosynthesis in plants was considerably higher in non-salinised plants grown with NO3-than with NH4+, but with increase in salinisation the photosynthesis rate decreased in NO3--supplied plants, but not in those given NH4+. The rate of transpiration was increased significantly by salinisation in NO3--supplied plants, but not consistently so in NH4+-supplied plants. The stomatal conductances were much higher in plants given NO3-than with NH4+when grown under non-saline conditions, but not when salinised. As a consequence, water-use efficiency in NO3--supplied control plants was better than in NH4+-supplied under non-saline conditions, but worse under saline conditions. The different forms of nitrogen and the addition of NaCl to the growing medium did not affect the relative intercellular concentrations of CO2 (Ci/Ca). Overall, the NH4+form of nitrogen inhibited the growth of sunflowers under non-saline conditions, but NO3-and NaCl interacted to inhibit growth more than did NH4+under saline conditions.  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