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 375 毫秒
1.
Summary Wood structure and function was investigated in different growth forms of temperate honeysuckles (Lonicera spp.). All three species had many narrow vessels and relatively few wide ones, with the measured K h (flow rate/pressure gradient) approximately 24–55% of the theoretical K h predicted by Poiseuille's law. Only the twiner, Lonicera japonica, had some vessels greater than 50 m in diameter. The twiner also had the narrowest stem xylem diameters, suggesting the greater maximum vessel diameter hydraulically compensated for narrow stems. Conversely, the free-standing shrub, L. maackii, had the greatest annual increments of xylem but the least percent conductive xylem implying that a great portion of the wood was involved with mechanical support. The scrambler, L, sempervirens had low maximum vessel diameter, high Huber values (= xylem area/leaf area), and low specific conductivities (= measured K h/xylem area), much like the shrub. The greatest vessel frequency occurred in the scrambler (901 vessels · mm-2), the highest thus far recorded in vines. The lowest Huber value and highest specific conductivity occurred in the twiner, suggesting little self-support but relatively efficient water conduction. LSC (= measured K h/leaf area) and maximum vessel diameter of Lonicera vines were near the low end of the range for vines in general.  相似文献   

2.
Summary A method for determining the mass flow rate of xylem water in thin stems under natural field conditions is presented. Diurnal courses of xylem water flow and stomatal conductance of the vines Entadopsis polystachya, Cyclanthera multifoliolata, and Serjania brachycarpa were examined in a tropical deciduous forest on the west coast of Mexico. E. polystachya (leaf area 23.6 m2) had a maximum water flow rate of 6.50 kg h-1 or 1.44 kg cm-2 stem basal area h-1; daily water use was 2.00 kg m-2 leaf area day-1. S. brachycarpa (leaf area 4.5 m2) and C. multifoliolata (leaf area 3.6 m2) had a maximum water flow rate of 0.72 and 0.19 kg h-1 or 0.63 and 0.92 kg cm-2 stem basal area h-1. Daily water use was 1.26 and 0.39 kg m-2 leaf area day-1, respectively. The daily courses of xylem water flow were strongly influenced by the orientation of the leaf area to irradiance and its intensity. While leaves of E. polystachya had a constant high stomatal conductance during the day, S. brachycarpa had a maximum stomatal opening in the morning followed by continuous closure during the rest of the day. In contrast to the woody species, the herbaceous C. multifoliolata exhibited a strong midday depression of stomatal conductance and wilting of its leaves. The leaf biomass accounted for 8% (Entadopsis), 16% (Serjania), and 23% (Cyclanthera) of above-ground biomass. The relation of sapwood area to leaf area supplied (Huber value) was 0.19 (Entadopsis), 0.18 (Serjania), and 0.06 (Cyclanthera) cm2 m-2  相似文献   

3.
Morphology, water relations, and xylem anatomy of high-light (sun)- and low-light (shade)-grown Vitis vinifera L. shoots were studied to determine the effects of shading on the hydraulic conductance of the pathway for water flow from the roots to the leaves. Shade shoots developed leaf area ratios (leaf area: plant dry weight) that were nearly threefold greater than sun shoots. Water-potential gradients (·m–1) in the shoot xylem accounted for most of the ·m–1 between soil and shoot apex at low and high transpiration rates in both sun and shade shoots, but the gradients were two- to fourfold greater in shade-grown plants. Low light reduced xylem conduit number in petioles, but had an additional slight effect on conduit diameter in internodes. The hydraulic conductance per unit length (Kh) and the specific hydraulic conductivity (ks, i.e. Kh per xylem cross-sectional area) of internodes, leaf petioles, and leaf laminae at different developmental stages leaf plastochron index was calculated from measurements of water potential and water flow in intact plants, from flow through excised organs, and from vessel and tracheid lumen diameters according to Hagen-Poiseuille's equation. For all methods and conductance parameters, the propensity to transport water to sink leaves was severalfold greater in internodes than in petioles. The Kh and ks increased logarithmically until growth ceased, independent of treatment and measurement method, and increased further in pressurized-flow experiments and Hagen-Poiseuille predictions. However, the increase was less in shade internodes than in sun internodes. Mature internodes of shade-grown plants had a two- to fourfold reduced Kh and significantly lower ks than sun internodes. Except very early in development, leaf lamina conductance and ks from shade-grown plants was also reduced. The strong reduction in Kh with only a slight reduction in leaf area (17% of sun shoots) in the shade shoots indicated a decoupling of water-transport capacity from the transpirational surface supplied by that capacity. This decoupling resulted in strongly reduced leaf specific conductivities and Huber values for both internodes and petioles, which may increase the likelihood of cavitation under conditions of high evaporative demand or soil drought.Abbreviations Ac total cross-sectional area (internodes, petioles, leaf laminae) - Ax xylem cross-sectional area - HV Huber value - Kh hydraulic conductance per unit length - ks specific hydraulic conductivity - LPI leaf plastochron index - LSC leaf specific conductivity - water potential - water-potential gradient - q volume flow of water per unit time Hans R. Schultz was supported in part by the Deutsche Forschungsgemeinschaft (grant Ki-114/8-1). We wish to thank Dr. Thomas Geier, Institut für Biologie, Forschungsanstalt D-6222 Geisenheim, Germany for his advice on sample preparation and microscopy, and two anonomous reviewers for their helpful comments.  相似文献   

4.
解析植物木质部导水率对逆境的响应和适应对促进植物抗逆性机理研究和受损植被恢复具有重要意义。该文以荒漠河岸林建群种胡杨(Populus euphratica)为研究对象,系统分析了胡杨幼株根、茎、叶水分传输通道对不同浓度盐胁迫的响应和适应。结果表明:(1)胡杨幼株根系对盐胁迫的敏感性高于茎和叶,盐胁迫下根系生长和根尖数显著受到抑制,根木质部易于发生栓塞,导水率明显降低。(2)胡杨幼株茎木质部导水率对盐胁迫的响应依盐浓度而定,轻度(0.05 mol·L–1 Na Cl)和中度(0.15 mol·L–1 Na Cl)盐胁迫下,胡杨可以通过协调导管输水的有效性和安全性来调节木质部的导水率,维持植物正常生长;重度(0.30 mol·L–1 Na Cl)盐胁迫下,胡杨茎木质部导管输水有效性和安全性均明显降低,木质部导水率显著下降,并伴随叶片气孔导度的显著降低,从而严重抑制了胡杨的光合和生长。  相似文献   

5.
We estimated the densities of termites in dead wood using a point-quarter technique in four habitats: pine plantations subject to control burning and unburned, lowland hardwood forest, and turkey oak woodland. The burned pine plantation had the highest estimated termite density, 13×106·ha-1 unburned pine plantation and lowland hardwood had estimated densities of 2.6×106·ha-1 and 2.2×106·ha-1, respectively, and the turkey oak woodland had an estimated densities of 61.9×103·ha-1. There were varying percentages of Reticulitermes flavipes and R. virginicus in the various habitats. There was nearly linear increase in percentage of pine blocks attacked by termites in the pine and hardwood forests and by the end of the growing season, nearly all had been channelized by termites. In the turkey oak habitat 70% of the pine blocks were channelized. Termites removed between 3% and 12% of the original mass of over one-fourth of the pine blocks during the growing season.Wood litter in fall was highest in the lowland hardwood forest: 2869 kg·ha-1·yr-1. Wood litter input in the long leaf pine plantation, 792 kg·ha-1·yr-1 was nearly equivalent to the three year average wood litter in fall in a Danish oak forest.  相似文献   

6.
Because of the different mechanical constraints on vines and self-supporting plants, vines are thought to differ from trees and shrubs in a variety of their growth characteristics. I tested the hypotheses that vines grow faster than shrubs and that supported shoots have delayed leaf expansion relative to stem elongation, using western poison oak, Toxicodendron diversilobum (T. & G.) Greene, a plant that grows as a vine when externally supported but otherwise as a shrub. In the field, supported shoots (vines) had significantly higher aboveground biomass and relative growth rates than did their paired unsupported shoots (shrubs) growing nearby. This was not due to differences in leaf phenology, but may have resulted from vines growing into more favorable habitats for growth. In contrast, whereas 2-yr-old cloned plants in a common garden differed in their stem and internode lengths, they had the same aboveground dry weight, proportion of dry weight that was leaf, and relative rate of increase in primary stem length whether grown with stakes (vines) or without stakes (shrubs). These results suggest that there is no inherent requirement of vines to grow faster than shrubs. As hypothesized, leaf elongation was more delayed relative to stem elongation in staked than unstaked individuals in 19 paired plants (each pair cloned from a different source plant). Thus, physical cues resulting from the presence of support can alter the plant's spatial and temporal patterns of development, but do not necessarily dictate the quantity of biomass that will be produced.  相似文献   

7.
Morphometric and physiological studies were conducted to determine whether the wall ingrowths of transfer cells in the minor-vein phloem of Pisum sativum L. leaves increase the capacity of the cells for solute influx. Size and number of wall ingrowths are positively correlated to the photon flux density (PFD) at which the plants are grown. An analysis of plasmodesmatal frequencies indicated that numerous plasmodesmata are present at all interfaces except those between the sieveelement-transfer-cell complex (SE-TCC) and surrounding cells where plasmodesmata are present but few in number. Flux of exogenous sucrose into the SE-TCC was estimated from kinetic profiles of net sucrose influx into leaf discs, quantitative autoradiography, and measurements of sucrose translocation. Flux based both on the saturable (carrier-mediated) and the linear components of influx was 47% greater in leaves of plants grown at high PFD (1000 mol·m–2·s–1) than those grown in low PFD (200 mol·m–2·s–1) and was paralleled by a 47% increase in SE-TCC plasmalemma surface area. Flux of endogenous photosynthate across the SE-TCC plasmalemma was calculated from carbon balance and morphometric data. The increase in flux in high-light leaves over that in low-light leaves can be explained on the basis of an increase in plasmalemma surface area. In intact leaves, a standing osmotic gradient may facilitate transport of solute into transfer cells with extensive wall elaborations.Abbreviations LPI leaf plastochron index - PCMBS p-chloromercuribenzenesulfonic acid - PFD(s) photon flux density (densities) - SE-TCC sieve-element-transfer-cell complex This research was supported by National Science Foundation Grant DCB-9104159, U.S. Department of Agriculture Competitive Grant 90000854, and Hatch funds.  相似文献   

8.
Plants using the C4 photosynthetic pathway have greater water use efficiency (WUE) than C3 plants of similar ecological function. Consequently, for equivalent rates of photosynthesis in identical climates, C4 plants do not need to acquire and transport as much water as C3 species. Because the structure of xylem tissue reflects hydraulic demand by the leaf canopy, a reduction in water transport requirements due to C4 photosynthesis should affect the evolution of xylem characteristics in C4 plants. In a comparison of stem hydraulic conductivity and vascular anatomy between eight C3 and eight C4 herbaceous species, C4 plants had lower hydraulic conductivity per unit leaf area (KL) than C3 species of similar life form. When averages from all the species were pooled together, the mean KL for the C4 species was 1.60 × 10?4 kg m?1 s?1 MPa?1, which was only one‐third of the mean KL of 4.65 × 10?4 kg m?1 s?1 MPa?1 determined for the C3 species. The differences in KL between C3 and C4 species corresponded to the two‐ to three‐fold differences in WUE observed between C3 and C4 plants. In the C4 species from arid regions, the difference in KL was associated with a lower hydraulic conductivity per xylem area, smaller and shorter vessels, and less vulnerable xylem to cavitation, indicating the C4 species had evolved safer xylem than the C3 species. In the plants from resource‐rich areas, such as the C4 weed Amaranthus retroflexus, hydraulic conductivity per xylem area and xylem anatomy were similar to that of the C3 species, but the C4 plants had greater leaf area per xylem area. The results indicate the WUE advantage of C4 photosynthesis allows for greater flexibility in hydraulic design and potential fitness. In resource‐rich environments in which competition is high, an existing hydraulic design can support greater leaf area, allowing for higher carbon gain, growth and competitive potential. In arid regions, C4 plants evolved safer xylem, which can increase survival and performance during drought events.  相似文献   

9.
The hydraulic conductivity of the lateral walls of early metaxylem vessels (Lpx in m · s–1 · MPa–1) was measured in young, excised roots of maize using a root pressure probe. Values for this parameter were determined by comparing the root hydraulic conductivities before and after steam-ringing a short zone on each root. Killing of living tissue virtually canceled its hydraulic resistance. There were no suberin lamellae present in the endodermis of the roots used. The value of Lpx ranged between 3 · 10–7 and 35 · 10–7 m · s–1 · MPa–1 and was larger than the hydraulic conductivity of the untreated root (Lpr = 0.7 · 10–7 to 4.0 · 10–7 m · s–1 · MPa–1) by factor of 3 to 13. Assuming that all flow through the vessel walls was through the pit membranes, which occupied 14% of the total wall area, an upper limit of the hydraulic conductivity of this structure could be given(Lppm=21 · 10–7 to 250 · 10–7 m · s–1 · MPa–1). The specific hydraulic conductivity (Lpcw) of the wall material of the pit membranes (again an upper limit) ranged from 0.3 · 10–12 to 3.8 · 10–12 m2 · s–1 · MPa–1 and was lower than estimates given in the literature for plant cell walls. From the data, we conclude that the majority of the radial resistance to water movement in the root is contributed by living tissue. However, although the lateral walls of the vessels do not limit the rate of water flow in the intact system, they constitute 8–31% of the total resistance, a value which should not be ignored in a detailed analysis of water flow through roots.Abbreviatations and Symbols kwr (T 1 2/W ) rate constant (half-time) of water exchange across root (s–1 or s, respectively) - Lpcw specific hydraulic conductivity of wall material (m2 · s–1 · MPa–1) - Lppm hydraulic conductivity of pit membranes (m · s –1 · MPa–1) - Lpr hydraulic conductivity of root (m · s–1 · MPa–1) - Lpx lateralhydraulic conductivity of walls of root xylem (m · s –1 · MPa–1) This research was supported by a grant from the Bilateral Exchange Program funded jointly by the Natural Sciences and Engineering Research Council of Canada and the Deutsche Forschungsgemeinschaft to C.A.P., and by a grant from the Deutsche Forschungsgemeinschaft, Sonderforschungsbereich 137, to E.S. The expert technical help of Mr. Burkhard Stumpf and the work of Ms. Martina Murrmann and Ms. Hilde Zimmermann in digitizing chart-recorder strips is gratefully acknowledged.  相似文献   

10.
Plants optimize water use and carbon assimilation via transient regulation of stomata resistance and by limiting hydraulic conductivity in a long-term response of xylem anatomy. We postulated that without effective hydraulic regulation plants would permanently restrain water loss and photosynthetic productivity under salt stress conditions. We compared wild-type tomatoes to a transgenic type (TT) with impaired stomatal control. Gas exchange activity, biomass, starch content, leaf area and root traits, mineral composition and main stems xylem anatomy and hydraulic conductivity were analyzed in plants exposed to salinities of 1 and 4 dS m−1 over 60 days. As the xylem cannot easily readjust to different environmental conditions, shifts in its anatomy and the permanent effect on plant hydraulic conductivity kept transpiration at lower levels under unstressed conditions and maintained it under salt-stress, while sustaining higher but inefficient assimilation rates, leading to starch accumulation and decreased plant biomass, leaf and root area and root length. Narrow conduits in unstressed TT plants were related to permanent restrain of hydraulic conductivity and plant transpiration. Under salinity, TT plants followed the atmospheric water demand, sustained similar transpiration rate from unstressed to salt-stressed conditions and possibly maintained hydraulic integrity, due to likely impaired hydraulic regulation, wider conduits and higher hydraulic conductivity. The accumulation of salts and starch in the TT plants was a strong evidence of salinity tolerance via osmotic regulation, also thought to help to maintain the assimilation rates and transpiration flux under salinity, although it was not translated into higher growth.  相似文献   

11.
 Stem xylem features in two evergreen Quercus species (Q. coccifera and Q. ilex) and a deciduous one (Q. faginea) were analysed along an Atlantic-Mediterranean climatic gradient in which rainfall and winter cold experience strong variation. Mean maximum vessel diameter, vessel density, vessel element length, xylem transverse sectional area, Huber value (xylem transverse sectional area per leaf area unit), theoretical leaf specific conductivity (estimated hydraulic conductance per leaf area unit) and total leaf area were determined in 3-year-old branches. Q. faginea presented the widest vessels and the highest theoretical leaf specific conductivity while Q. coccifera showed the lowest total leaf area and the highest Huber value. Studied features did not exhibit significant correlations with mean minimum January temperature in any species but did show significant relationships with rainfall. In Q. coccifera, mean maximum vessel diameter, vessel element length and theoretical leaf specific conductivity increased with higher rainfall while vessel density decreased. Mean maximum vessel diameter and total leaf area in Q. ilex increased with precipitation whereas variables of Q. faginea did not show any significant trend. Results suggest that aridity, rather than minimum winter temperature, controls stem xylem responses in the studied evergreen species. Q. faginea traits did not show any response to precipitation, probably because this species develops deep roots, which in turn makes edaphic and topographic factors more important in the control of soil water availability. In response to aridity Q. coccifera only exhibits adjustment at a xylem level by reducing its water transport capacity through a reduction of vessel diameter without changing the amount of xylem tissue or foliage, whereas Q. ilex adjusts its water transport capacity in parallel to the foliage area. Received: 13 January 1997 / Accepted: 8 April 1997  相似文献   

12.
The hydraulic conductivities of excised whole root systems of wheat (Triticum aestivum L. cv. Atou) and of single excised roots of wheat and maize (Zea mays L. cv. Passat) were measured using an osmotically induced back-flow technique. Ninety minutes after excision the values for single excised roots ranged from 1.6·10-8 to 5.5·10-8 m·s-1·MPa-1 in wheat and from 0.9·10-8 to 4.8·10-8 m·s-1·MPa-1 in maize. The main source of variation was a decrease in the value as root length increased. The hydraulic conductivities of whole root systems, but not of single excised roots, were smaller 15 h after excision. This was not caused by occlusion of the xylem at the cut end of the coleoptile. The hydraulic conductivities of epidermal, cortical and endodermal cells were measured using a pressure probe. Epidermal and cortical cells of both wheat and maize roots gave mean values of 1.2·10-7 m·s-1·MPa-1 but in endodermal cells (measured only in wheat) the mean value was 0.5·10-7 m·s-1·MPa-1. The cellular hydraulic conductivities were used to calculate the root hydraulic conductivities expected if water flow across the root was via transcellular (vacuole-to-vacuole), apoplasmic or symplasmic pathways. The results indicate that, in freshly excised roots, the bulk of water flow is unlikely to be via the transcellular pathway. This is in contrast to our previous conclusion (H. Jones, A.D. Tomos, R.A. Leigh and R.G. Wyn Jones 1983, Planta 158, 230–236) which was based on results obtained with whole root systems of wheat measured 14–15 h after excision and which probably gave artefactually low values for root hydraulic conductivity. It is now concluded that, near the root tip, water flow could be through a symplasmic pathway in which the only substantial resistances to water flow are provided by the outer epidermal and the inner endodermal plasma membranes. Further from the tip, the measured hydraulic conductivities of the roots are consistent with flow either through the symplasmic or apoplasmic pathways.Symbols L p, cell cell hydraulic conductivity - L p, root root hydraulic conductivity - L p, root calculated root hydraulic conductivity - root reflection coefficient  相似文献   

13.
  • The fast growth of mulberry depends on high water consumption, but considerable variations in drought tolerance exist across different cultivars. Physiological and anatomical mechanisms are important to plant survival under drought. However, few research efforts have been made to reveal the relationships of these two aspects in relation to drought tolerance.
  • In this study, growth rates, leaf functional physiology and anatomical characteristics of leaf and xylem of 1‐year‐old saplings of seven mulberry cultivars at a common garden were compared. Their relationships were also explored.
  • Growth, leaf physiology and anatomy were significantly different among the tested cultivars. Foliar stable carbon isotope composition (δ13C) was negatively correlated with growth rates, and closely related to several leaf and xylem anatomical traits. Particularly, leaf thickness, predicted hydraulic conductivity and vessel element length jointly contributed 77% of the variability in δ13C. Cultivar Wupu had small stomata, intermediate leaf thickness, the smallest hydraulically weighted vessel diameter and highest vessel number, and higher δ13C; Yunguo1 had high abaxial stomatal density, low specific leaf area, moderate hydraulic conductivity and δ13C; these are beneficial features to reduce leaf water loss and drought‐induced xylem embolism in arid areas. Cultivar Liaolu11 had contrasting physiological and anatomical traits compared with the previous two cultivars, suggesting that it might be sensitive to drought.
  • Our findings indicate that growth and δ13C are closely associated with both leaf and xylem anatomical characteristics in mulberry, which provides fundamental information to assist evaluation of drought tolerance in mulberry cultivars and in other woody trees.
  相似文献   

14.
The water relations and hydraulic architecture of a tropical tree (Schefflera morototoni) and of two temperate species (Acer saccharum and Thuja occidentalis) are reported. Among the water relations parameters measured were leaf and stem water storage capacity, leaf water potential, transpiration, and vulnerability of stems to cavitation and loss of hydraulic conductivity by embolisms. Among the hydraulic architecture parameters measured were hydraulic conductivity per unit pressure gradient, specific conductivity, leaf-specific conductivity, and Huber value. In terms of vulnerability of stems to cavitation, stem and leaf capacitances, and leaf-specific conductivity, all three species followed the same sequence: Schefflera > Acer > Thuja. It is argued here that the high stem capacitance and high leaf-specific conductivity of Schefflera are necessary to compensate for its high vulnerability to cavitation. Extractable water storage per unit leaf area in Schefflera stems is >100 times that of Acer and may permit the species to survive unusually long, dry seasons in Panama. Although Schefflera frequently grows >20 meters, the biggest resistance to water flow in the shoots resides in the leaves.  相似文献   

15.
Leaf bearing stems ofCuratella americana L. were very susceptible to induced cavitation: embolisms began at a pressure of 0.5 MPa (15 %) and at 2.0 MPa most of the conductivity was lost (85 %). Nevertheless, in nature similar leaf specific conductivities, of about 90 x 10-5 kg m-2 s-1 MPa-1 during both wet and dry seasons indicated absence of drought induced embolisms. Leaf water potentials were neither very negative or considerably different between seasons but stomatal conductance decreased from 236 mmol m-2 s-1 measured during wet period to 100 mmol m-2 s-1 during drought season. Therefore, it was concluded thatCuratella had an accurate homeostatic balance of leaf water status to keep up xylem integrity. Acknowledgements: Financial support was provided by Decanato de Investigaciones - USB (S1-CB-811). Dr John Sperry (University of Utah) allowed me to learn how to use and to build the equipment used in this study. Mr. M. Edreida and Mr. T. Perez helped me in the field and in the laboratory, respectively. Dr D. HenrIquez corrected the English grammar.  相似文献   

16.
Background and Aims: The branch-base xylem structure of the endangered Wollemia nobiliswas anatomically investigated. Wollemia nobilis is probablythe only extant tree species that produces only first-orderbranches and where all branches are cleanly abscised. An investigationwas carried out to see if these unusual features might influencebranch-base xylem structure and water supply to the foliage. Methods: The xylem was sectioned at various distances along the branchbases of 6-year-old saplings. Huber values and relative theoreticalhydraulic conductivities were calculated for various regionsof the branch base. Key Results: The most proximal branch base featured a pronounced xylem constriction.The constriction had only 14–31 % (average 21 %)of the cross-sectional area and 20–42 % (average28 %) of the theoretical hydraulic conductivity of themore distal branch xylem. Wollemia nobilis had extremely lowHuber values for a conifer. Conclusions: The branch-base xylem constriction would appear to facilitatebranch abscission, while the associated Huber values show thatW. nobilis supplies a relatively large leaf area through a relativelysmall diameter ‘pipe’. It is tempting to suggestthat the pronounced decline of W. nobilis in the Tertiary isrelated to its unusual branch-base structure but physiologicalstudies of whole plant conductance are still needed.  相似文献   

17.
18.
A. Rieger  R. Hampp 《Planta》1991,184(3):415-421
The activities of enzymes which catalyze one step in each of the five major carbon pathways in green plants were measured in secondary pulvini and other tissues of Phaseolus coccineus L. leaves. We were able to detect activities of fumarase (EC 4.2.1.2; tricarboxylic-acid pathway), NAD-glyceraldehyde-phosphate dehydrogenase (NAD-GAPDH, EC 1.2.1.12; glycolysis), 6-phosphogluconate dehydrogenase (6-PGDH, EC 1.1.1.44; oxidative pentose-phosphate pathway), ribulose-1, 5-bisphosphate carboxylase (Rubisco, EC 4.1.1.39; photosynthetic carbon-reduction pathway), and of hydroxypyruvate reductase (HP-R, EC 1.1.1.81; photosynthetic carbon-oxidation pathway). On a protein basis the activities of Rubisco and HP-R in pulvinar regions were very low (below 1 and 2 mol · (kg protein) –-1 · h–-1, respectively), but the activities of fumarase and NAD-GAPDH were between 10- and 5-fold higher compared with the laminar tissue (up to 7 and 50 mol · (kg protein)–-1 · h–-1, respectively). Similarly, the protein specific activities of 6-PGDH were increased in the pulvinus (3–4 compared with approx. 1 mol · (kg protein)–-1 · h–-1 in the leaf blade). No differences in specific activities were detected between day and night positions of the leaves. By applying quantitative histochemical techniques we determined the longitudinal and transversal compartmentation of the activities of fumarase, NAD-GAPDH, and 6-PGDH in pulvinar tissues. Levels of activity of all three enzymes increased towards the middle part of the pulvinus. Here, expressed on a dry-weight (DW) basis, the analysis of cross sections showed highest activities in the outer parts of the extensor in the order given, approx. 0.6, 5, and 0.25 mol · (kg DW)–-1 · h–-1 for fumarase, NAD-GAPDH and 6-PGDH. When related to protein, levels of activity were comparably high within the inner parts of extensor and flexor, and partly also in the abaxial part of the bundle (fumarase, 6-PGDH). The tissue-specific compartmentation of the respective activities is discussed in relation to leaf movement and shows parallels with guard-cell function.Abbreviations Chl chlorophyll - DW dry weight - GAPDH glyceraldehyde-phosphate dehydrogenase - HP-R hydroxypyruvate reductase - Rubisco ribulose-1,5-bisphosphate carboxylase - 6-PGDH 6-phosphogluconate dehydrogenase This investigation was supported by a grant from the Deutsche Forschungsgemeinschaft.  相似文献   

19.
Bai E  Boutton TW  Liu F  Wu XB  Archer SR 《Oecologia》2008,156(3):479-489
δ13C values of C3 plants are indicators of plant carbon–water relations that integrate plant responses to environmental conditions. However, few studies have quantified spatial variation in plant δ13C at the landscape scale. We determined variation in leaf δ13C, leaf nitrogen per leaf area (Narea), and specific leaf area (SLA) in April and August 2005 for all individuals of three common woody species within a 308 × 12-m belt transect spanning an upland–lowland topoedaphic gradient in a subtropical savanna in southern Texas. Clay content, available soil moisture, and soil total N were all negatively correlated with elevation. The δ13C values of Prosopis glandulosa (deciduous N2-fixing tree legume), Condalia hookeri (evergreen shrub), and Zanthoxylum fagara (evergreen shrub) leaves increased 1–4‰ with decreasing elevation, with the δ13C value of P. glandulosa leaves being 1–3‰ higher than those of the two shrub species. Contrary to theory and results from previous studies, δ13C values were highest where soil water was most available, suggesting that some other variable was overriding or interacting with water availability. Leaf Narea was positively correlated with leaf δ13C of all species (p < 0.01) and appeared to exert the strongest control over δ13C along this topoedaphic gradient. Since leaf Narea is positively related to photosynthetic capacity, plants with high leaf Narea are likely to have low p I/p a ratios and therefore higher δ13C values, assuming stomatal conductance is constant. Specific leaf area was not correlated significantly with leaf δ13C. Following a progressive growing season drought in July/August, leaf δ13C decreased. The lower δ13C in August may reflect the accumulation of 13C-depleted epicuticular leaf wax. We suggest control of leaf δ13C along this topoedaphic gradient is mediated by leaf Narea rather than by stomatal conductance limitations associated with water availability.  相似文献   

20.
Measurements of xylem conduit length and width and the distribution of xylem conduit ends were made in inter-nodes (I), nodes (N) and twig junctions (J) of 1-, 2- and 3-year-old twigs of plants of Quercus cerris L. Parallel measurements were also made of the loss of hydraulic conductivity of twigs subjected to pressure differentials across conduit pit membranes, equalling the leaf water potential at the turgor loss point. The loss of theoretical hydraulic conductivity was calculated as the ratio of i esivr4 (where r is the conduit radius) of the non-conducting conduits to that of all the conduits in the outermost wood ring of I, N and J. Stem zones such as 1-year-old nodes and junctions were localized with narrower and shorter xylem conduits and with higher percentages of conduit ends than internodes. Such ‘constricted zonesrsquo; were less vulnerable to embolism than internodes. Latewood conduits were consistently narrower, shorter and less vulnerable to embolism than earlywood ones. A positive relation therefore existed between conduit diameter and length and vulnerability to embolism. The overall vulnerability to embolism of Q. cerris plants is discussed in terms of xylem conduit width and length and of the distribution of conduit ends.  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