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 46 毫秒
1.
Pigment combinations are regulated during leaf ontogenesis. To better understand pigment function, alterations in chlorophyll, carotenoid and anthocyanin concentrations were investigated during different leaf development stages in six subtropical landscape plants, namely Ixora chinensis Lam, Camellia japonica Linn, Eugenia oleina Wight, Mangifera indica L., Osmanthus fragrans Lowr and Saraca dives Pierre. High concentrations of anthocyanin were associated with reduced chlorophyll in juvenile leaves. As leaves developed, the photosynthetic pigments (chlorophyll and carotenoid) of all six species increased while anthocyanin concentration declined. Chlorophyll fluorescence imaging of ΦPSII (effective quantum yield of PSII) and of NPQ (non-photochemical fluorescence quenching) and determination of electron transport rate-rapid light curve (RLC) showed that maximum ETR (leaf electron transport rate), ΦPSII and the saturation point in RLC increased during leaf development but declined as they aged. Juvenile leaves displayed higher values of NPQ and Car/Chl ratios than leaves at other developmental stages. Leaf reflectance spectra (400–800 nm) were measured to provide an in vivo non-destructive assessment of pigments in leaves during ontogenesis. Four reflectance indices, related to pigment characters, were compared with data obtained quantitatively from biochemical analysis. The results showed that the ARI (anthocyanin reflectance index) was linearly correlated to anthocyanin concentration in juvenile leaves, while a positive correlation of Chl NDI (chlorophyll normalized difference vegetation index) to chlorophyll a concentration was species dependent. Photosynthetic reflectance index was not closely related to Car/Chl ratio, while a structural-independent pigment index was not greatly altered by leaf development or species. Accordingly, it is suggested that the high concentration of anthocyanin, higher NPQ and Car/Chl ratio in juvenile leaves are important functional responses to cope with high radiation when the photosynthetic apparatus is not fully developed. Another two leaf reflectance indices, ARI and Chl NDI, are valuable for in vivo pigment evaluation during leaf development.  相似文献   

2.
Common hazel (Corylus avellana L., Fusca rubra Dipp.) juvenile leaves from the periphery of the canopy and thus subjected to high fluxes of solar radiation are characterized by red coloration due to anthocyanin accumulation disappearing in mature leaves. To elucidate the physiological role of anthocyanin accumulation, the interrelations between anthocyanin content, a degree of attenuation by the pigments of the light reaching the photosynthetic apparatus (PSA), and PSA tolerance to photoinhibition in C. avellana juvenile leaves were studied. Absorption spectra were calculated taking into account the light losses due to reflection by the leaf. The analysis of the spectra showed that, in red common hazel leaves accumulating high amounts of anthocyanins in the vacuoles of the upper and lower epidermal cells, up to 95% of visible radiation entering the leaf blade was absorbed by these pigments. The rate of the linear electron transport (ETR) in the chloroplast electron transport chain (ETC) was closely correlated with the anthocyanin content (r 2 = 0.87). In red leaves, the saturation of ETR dependence on irradiance was observed at the higher values of PAR than in green leaves. In red juvenile leaves, this value was close to that in mature green leaves tolerant to high light. There were no differences between red and green leaves in the level of non-photochemical quenching, the content of violaxanthin cycle pigments, a degree of their de-epoxidation under natural illumination and at irradiation with high PAR fluxes. Basing on the data obtained, one may conclude that anthocyanins in C. avellana juvenile leaves serve PSA photoprotection, preventing injury of immature PSA with excessive fluxes of PAR.  相似文献   

3.
Net photosynthetic rate, stomatal conductance, ratio of sub-stomatal to atmospheric CO2 concentration, transpiration rate, and water use efficiency changed significantly and assimilation capacity dropped continuously along the salinization and alkalinization process in the afternoon. Assimilation capacity of L. chinensis leaf correlated negatively with the degree of salinization and alkalinization. The photosynthetic characteristics of L. chinensis determined its community formation. By changing the ratio of chlorophyll a/b in leaves and accumulating soluble saccharides in rhizome, L. chinensis could adapt to the saline-alkali condition.  相似文献   

4.
Leaf area expansion, thickness and inclination, gas exchange parameters and relative chlorophyll content were analysed in field‐grown fig (Ficus carica L.) leaves over time, from emergence until after full leaf expansion (FLE). Ficus carica leaves showed a subtle change in shape during the early stages of development, and FLE was reached within ca. 30 days after emergence. Changes in leaf thickness and inclination after FLE demonstrated good adaptation to environmental conditions during summer in areas with a Mediterranean climate. Changes in gas exchange parameters and relative chlorophyll content showed that F. carica is a delayed‐greening species, reaching maximum values 20 days after FLE. Correlation analysis of datasets collected during leaf expansion, confirmed dependence among structural and functional traits in F. carica. Pn was directly correlated with stomatal conductance (Gs), transpiration (E), leaf area (LA) and relative chlorophyll content up to FLE. The effect of pruning on leaf expansion, a cultural technique commonly applied in this fruit tree, was also evaluated. Although leaf development in pruned branches gave a significantly higher relative leaf area growth rate (RGRl) and higher LA than non‐pruned branches, no significant differences were found in other morphological and physiological traits, indicating no pruning effect on leaf development. All studied morphological and physiological characteristics indicate that F. carica is well adapted to semiarid conditions. The delayed greening strategy of this species is discussed.  相似文献   

5.
Leaf developmental patterns were characterized in four rainforest tree species of Syzgium. Leaf optical properties, pigment changes, expansion characteristics, stomatal development, and photosynthetic rates were studied. In both S. luehmannii and S. wilsonii photosynthetic development was delayed until after full leaf expansion. Rates of O2 evolution were negative during expansion of S. luehmannii and S. wilsonii leaves and stomatal conductance was 10-20 mmol m-2 s-1 lower than for corresponding leaves of S. moorei. Stomatal conductance showed that the development of functional stomata was delayed until after full leaf expansion in S. luehmannii and S. wilsonii, however, low stomatal conductance was not responsible for the lack of photosynthetic potential during leaf expansion in these species. Leaves of S. luehmannii and S. wilsonii required less than 10 d for full leaf expansion and contained anthocyanin during expansion. In contrast, leaves of S. moorei and S. corynanthum expanded slowly (20-40 d required for full leaf expansion), exhibited positive rates of O2 evolution and did not accumulate anthocyanin. In S. luehmannii and S. wilsonii anthocyanin was located in the vacuole of distinct cell layers just below the upper epidermis and the possible functions of anthocyanin accumulation are discussed. This is the first report where such variation in leaf development has been characterized in the one genus.Key words: Anthocyanin, leaf expansion, photosynthetic development, delayed leaf greening, stomatal development.   相似文献   

6.
Niu  S.L.  Jiang  G.M.  Li  Y.G.  Gao  L.M.  Liu  M.Z. 《Photosynthetica》2003,41(2):221-226
Net photosynthetic rate (P N), transpiration rate (E), stomatal conductance (g s), leaf water potential (ψleaf), leaf nitrogen content, and photosynthetic nitrogen use efficiency (PNUE) were compared between a typical C4 plant, Agriophyllum squarrosum and a C3 plant, Leymus chinensis, in Hunshandak Sandland, China. The plant species showed different diurnal gas exchange patterns on June 12–14 when photosynthetic photon flux density (PPFD), air temperature (T air), and water potential were moderate. P N, E, and g s of A. squarrosum showed distinct single peak while those of L. chinensis were depressed at noon and had two peaks in their diurnal courses. Gas exchange traits of both species showed midday depression under higher photosynthetic photon flux density (PPFD) and T air when Ψleaf was significantly low down on August 6–8. However, those of A. squarrosum were depressed less seriously. Moreover, A. squarrosum had higher P N, Ψleaf, water use efficiency (WUE), and PNUE than L. chinensis. Thus A. squarrosum was much more tolerant to heat and high irradiance and could utilise the resources on sand area more efficiently than L. chinensis. Hence species like A. squarrosum may be introduced and protected to reconstruct the degraded sand dunes because of their higher tolerance to stress and higher resource use efficiency. This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

7.
为探讨夏季南亚热带森林演替过程中优势树种幼叶的光保护机制,以演替中期优势树种木荷(Schima superba)、黧蒴(Castanopsis fissa)、锥栗(C.chinensis)和演替后期优势种华润楠(Machilus chinensis)、厚壳桂(Cryptocarya chinensis)、黄果厚壳桂(C.concinna)为材料,分析了2种生长光强(全光照和30%全光照)下6种优势种幼叶和成熟叶的叶片表型、光合色素含量、花色素苷含量、抗氧化能力、类黄酮含量、总酚含量和最大量子产量(Fv/Fm)恢复效率间的差异。结果表明,两个演替阶段幼叶的叶绿素含量(Chl a+b)、Chl a/b比成熟叶低,但光保护物质比成熟叶多;演替中期幼叶的花色素苷含量和总抗氧化能力比演替后期的高,而类黄酮和总酚含量比演替后期的低;全光照下幼叶的总酚、类黄酮、总抗氧化能力及Fv/Fm恢复效率都要比30%全光照的高,并且含有花色素苷的幼叶恢复得更快。因此,植物的光合能力与自身的光保护潜力成反比关系,演替中期优势种幼叶的光保护在很大程度上是因为花色素苷的积累而演替后期优势种是因为自身抗氧化物质(类黄酮、总酚)的共同作用。  相似文献   

8.
The relationship between the microclimate within an Oak-Hickory forest and photosynthetic characters of two resident evergreen herbs with contrasting leaf phenologies was investigated on a monthly basis for 1 full year. Heuchera americana has leaf flushes in the spring and fall, with average leaf life spans of 6–7 months. Hexastylis arifolia produces a single cohort of leaves each spring with a leaf life span of 12–13 months. We predicted that among evergreen plants inhabiting a seasonal habitat, a species for which the frequency of leaf turnover is greater than the frequency of seasonal extremes would have a greater annual range in photosynthetic capacity than a species that only produced a single flush of leaves during the year. Photosynthetic parameters, including apparent quantum yield, maximum photosynthetic capacity (Pmax), temperature of maximum photosynthesis, photochemical efficiency of PSII and leaf nitrogen (N) and chlorophyll concentrations, were periodically measured under laboratory conditions in leaves sampled from natural populations of both species. Mature leaves of both species acclimated to changing understory conditions with the mean seasonal differences being significantly greater for Heuchera than for Hexastylis. Area based maximum photosynthetic rates at 25°C were approximately 250% and 100% greater in winter leaves than summer leaves for Heuchera and Hexastylis respectively. Nitrogen concentrations were highest in winter leaves. Chlorophyll concentrations were highest in summer leaves. Low Pmax/N values for these species suggest preferential allocation of leaf nitrogen into non-photosynthetic pools and/or light-harvesting function at the expense of photosynthetic enzymes and electron transport components. Despite the increase in photosynthetic capacity, there was evidence of chronic winter photoinhibition in Hexastylis, but not in Heuchera. Among these ecologically similar species, there appears to be a trade-off between the frequency of leaf production and the balance of photosynthetic acclimation and photoinhibition.  相似文献   

9.
不同种源黄连木秋季色素含量与叶色参数的关系   总被引:2,自引:0,他引:2  
该研究以陕西汉中、河南林州、河北涉县和北京中国科学院植物研究所4个种源黄连木(Pistacia chinensis Bunge)的苗木为对象,用分光光度计和色差仪对其叶绿素、类胡萝卜素、花色素苷含量及叶色参数(L*、a*、b*)进行了测定分析,探讨不同种源苗木秋季叶色变化规律及差异,揭示黄连木叶色呈现与叶片色素含量之间的内在关联,为筛选适合城市绿化的优良黄连木种源提供依据。结果表明:(1)在秋季叶片转色期,随着时间的推移,4个种源黄连木叶片的叶绿素、类胡萝卜素和花色素苷含量的比例呈现不同的变化趋势,其中:河北种源的花色素苷含量较高,叶片呈现红色;陕西种源叶绿素含量较高,叶片呈现绿色的时间较长;河南、北京种源处于两者之间。(2)各个种源黄连木的叶色参数a*值(红/绿)均与花色素苷含量呈正相关关系,与叶绿素含量呈负相关关系,且相关系数均达到显著水平(P0.05),各个种源叶色参数L*值(光泽明亮度)也与叶绿素含量间表现出显著或极显著的正相关性。研究发现,河北种源黄连木秋季的叶色最红,陕西种源黄连木叶片呈现绿色的时间最长;色差仪的应用实现了叶色和各色素含量间量化的关系。  相似文献   

10.
11.
Changes in mesophyll anatomy, gas exchange, and the amounts of nitrogen and cell wall constituents including cellulose, hemicellulose and lignin during leaf development were studied in an evergreen broad‐leaved tree, Quercus glauca, and in an annual herb, Phaseolus vulgaris. The number of chloroplasts per whole leaf in P. vulgaris increased and attained the maximal level around 10 d before full leaf area expansion (FLE), whereas it continued to increase even after FLE in Q. glauca. The increase in the number of palisade tissue cells per whole leaf continued until a few days before FLE in Q. glauca, but it had almost ceased by 10 d before FLE in P. vulgaris. The radius and height of palisade tissue cells in Q. glauca, attained their maximal levels at around FLE whereas the thickness of the mesophyll cell wall and concentrations of the cell wall constituents increased markedly after FLE. These results clearly indicated that, in Q. glauca, chloroplast development proceeded in parallel with the cell wall thickening well after completion of the mesophyll cell division and cell enlargement. The sink–source transition, defined to be the time when the increase in daily carbon exchange rate exceeds the daily increase in leaf carbon content, occurred before FLE in P. vulgaris but after FLE in Q. glauca. During leaf area expansion, the maximum daily increase in nitrogen content on a whole leaf basis (the maximum leaf areas were corrected to be identical for these species) in Q. glauca was similar to that in P. vulgaris. In Q. glauca, however, more than 70% of nitrogen in the mature leaf was invested during its sink phase, whereas in P. vulgaris it was 50%. These results suggest that Q. glauca invests nitrogen for cell division for a considerable period and for chloroplast development during the later stages. We conclude that the competition for nitrogen between cell division and chloroplast development in the area of expanding leaves can explain different greening patterns among plant species.  相似文献   

12.
In the seasonally flooded forest of the Mapire River, a tributary of the Orinoco, seedlings remain totally covered by flood water for over six months. In order to characterize the physiological response to flooding and submergence, seedlings of the tree Pouteria orinocoensis, an important component of the forest vegetation, were subjected experimentally to flooding. Flooding was imposed gradually, the maximum level of flood including submerged and emerged leaves. After 45 d a severe reduction of net photosynthetic rate (P N) and stomatal conductance (g s) was observed in emerged leaves, whereas leaf water potential remained constant. The decrease in P N of emerged leaves was associated to an increase in both relative stomatal and non-stomatal limitations, and the maintenance of the internal/air CO2 concentration (C i/C a) for at least 20 d of flooding. After this time, both P N and g s became almost zero. The decrease in photosynthetic capacity of emerged leaves with flooding was also evidenced by a decrease in carboxylation efficiency; photon-saturated photosynthetic rate, and apparent quantum yield of CO2 fixation. Oxygen evolution rate of submerged leaves measured after three days of treatment was 7 % of the photosynthetic rate of emerged leaves. Submersion determined a chronic photoinhibition of leaves, viewed as a reduction in maximum quantum yield in dark-adapted leaves, whereas the chlorophyll fluorescence analysis of emerged leaves pointed out at the occurrence of dynamic, rather than chronic, photoinhibition. This was evidenced by the absence of photochemical damage, i.e. the maintenance of maximum quantum yield in dark-adapted leaves. Nevertheless, the observed lack of complementarity between photochemical and non-photochemical quenching after 12 d of flooding implies that the capacity for photochemical quenching decreased in a non-co-ordinate manner with the increase in non-photochemical quenching.  相似文献   

13.
Manetas  Y.  Drinia  A.  Petropoulou  Y. 《Photosynthetica》2002,40(3):349-354
We checked the hypothesis that the transient presence of anthocyanins in young leaves serves a photoprotective function. For this purpose, Rosa sp. and Ricinus communis L., whose young leaves are red to become green upon maturation, were used. Thus, young leaves with high and mature leaves with low anthocyanin contents were analysed concerning their carotenoid (Car) composition and susceptibility to photoinhibition. Cars, including the components of the xanthophyll cycle, had similar contents in young and mature leaves, when expressed on a chlorophyll basis. Yet, when expressed on a leaf area basis or on the assumed photon absorptive capacity of leaves, Cars contents were considerably lower in anthocyanic young leaves. Although this may indicate a low photodissipative potential, red young leaves were considerably less susceptible to photoinhibitory damage. The results are compatible with a photoprotective function of anthocyanins, indicating also that their presence may compensate for a low capacity in the xanthophyll cycle-dependent harmless dissipation of excess excitation energy.  相似文献   

14.
Upland cotton (Gossypium hirsutum L.) can move leaves to track the sun throughout the day, so-called leaf diaheliotropic movement. This paper reports an experimental test of the hypothesis that leaf diaheliotropic movement in upland cotton can enhance carbon assimilation and not increase the risk of stress from high energy load. In this experiment, cotton leaves were divided into two groups: one was that leaves could track the sun freely; another was that leaves were retained to the horizontal position. The diaheliotropic leaves recorded higher incident irradiance than the restrained ones, especially in the morning and late afternoon. Compared with restrained leaves, diaheliotropic leaves were generally warmer throughout the day. As expected, diaheliotropic leaves had significantly higher diurnal time courses of net photosynthetic rate (P N) than restrained leaves, except during 14:00–18:00 of the local time. Higher instantaneous water-use efficiency (WUE) was observed in diaheliotropic leaves in the early morning and late afternoon than in the restrained leaves. During the given day, diaheliotropic and restrained leaves had similar diurnal time courses of recovery of maximal quantum yield of PSII photochemistry (Fv/Fm). Diaheliotropic leaves recorded lower or similar photochemical quenching coefficient (qp) than restrained leaves did throughout the day. These results suggest that cotton leaf diaheliotropic movement can improve carbon gain and water use efficiency and not intensify photoinhibition.  相似文献   

15.
Spraying low concentrated (0.5–5.0 mM) solutions of NaHSO3 on Satsuma mandarin (Citrus unshiu Marc.) leaves resulted in enhancement (maximal about 15 % at 1 mM NaHSO3) of net photosynthetic rate (P N) for 6 d. The potential photochemical efficiency of photosystem 2 (PS2, Fv/Fm) and the quantum yield of PS2 electron transport (ΦPS2) were increased under strong photon flux density (PFD). The slow phase of millisecond delayed light emission (ms-DLE) was increased, showing that the transmembrane proton motive force related to photophosphorylation was enhanced. We also observed that low concentrations of NaHSO3 promoted the production of ATP in irradiated leaves. We suggest that the increase in P N in Satsuma mandarin leaves caused by low concentrations of NaHSO3 solution may have been due to the stimulation of photophosphorylation and, hence, the increase in photochemical efficiency through speeding-up of PS2 electron transport. Photoinhibition of photosynthesis in leaves was modified by NaHSO3 treatment under high PFD. Hence the increase in leaf dry mass seems to be associated with the mitigation of photoinhibition caused by strong PFD.  相似文献   

16.
We investigated the effects of leaf color change in the fall on photosynthetic production and nitrogen resorption. Seedlings of Acer platanoides L. and A. saccharum Marsh. were grown in a shade house for 5 months in either 21 % (intermediate light, M) or 4.9 % (low light, L) of incident irradiance. After this period, a subset of the intermediate-light grown seedlings was transferred to a high-light stress treatment (H). Gas exchange, chlorophyll fluorescence, pigments, antioxidant activity, and nitrogen (N) resorption were examined at three leaf senescence stages during September and October. Our results show that plants of both species produce more anthocyanins in the H treatment. In comparison with plants grown in the L and M treatments, plants of both species in the H treatments had lower chlorophyll, carotenoid and chlorophyll fluorescence parameters (F v/F m, Φ PSII, NPQ and ETR) at the third sampling date (October 12–18), and indicating higher levels of photoinhibition in the seedlings exposed to high light. Our results imply that autumn leaf redness is inducible and closely linked to photo-oxidative stress. However, anthocyanins did not enhance antioxidant capacity in red leaves in either species, when exposed to high light. For both species, our results showed a higher N-resorption for high-light stressed plants. We also observed that the number of abscised leaves at the second sampling dates (September 10) was higher than at the third sampling dates. The intra-leaf distribution of anthocyanin, the association between anthocyanin production and the high-light environments, the retention of red leaves, the substantial physiological gain of photosynthetic activity, as well as the links between anthocyanins and increased N resorption led us to assume that one primary role of autumn anthocyanin could be to protect the photosynthetic apparatus from photo-oxidative damage as light filters rather than as antioxidant. Another major role is to extend carbon capture and help supply the energy needed for N resorption from senescing leaves in both A. saccharum and A. Platanoides during high-light stress. Nevertheless, photoprotective capacity of anthocyanins was not able to fully compensate for photoinhibitory stress as the anthocyanins are not optimally located to efficiently reduce light within the leaves.  相似文献   

17.
Flavonoids are thought to participate in protection of the photosynthetic apparatus against photoinhibition under excessive light. Flavone glycoside, scutellarin, is a main active ingredient extracted from Erigeron breviscapus, the plant used in Chinese medicine. Shade-developed leaves of E. breviscapus were transferred from shade to full sunlight to quantify a relationship between the concentration of leaf scutellarin and tolerance to high radiation stress or the recovery from photoinhibition. The maximal quantum yield of PSII photochemistry showed a diurnal fluctuation in both shaded and sunlit leaves throughout the day. It indicated dynamic photoinhibition in the leaves of Erigeron, i.e., higher photoinhibition at solar noon and lower one in the morning and late afternoon. The sun-developed leaves reached the higher scutellarin content and values of nonphotochemical quenching coefficient with a lower degree of photoinhibition than the shade-developed leaves. When the shade-developed leaves were transferred to full sunlight, the content of scutellarin was declining continuously for 10 d and then was increasing for 15 d. After 50 d, all leaves became the sun-developed leaves with their scutellarin contents of about 138.5 ± 5.2 mg g?1(dry mass, DM) which was significantly higher than that of the shade-developed leaves [107.8 ± 9.8 mg g?1(DM)]. During acclimatization, the degree of photoinhibition was negatively correlated with the scutellarin content. Our results demonstrated a synchronous fluctuation between the flavonoid content and degree of protection against photoinhibition.  相似文献   

18.
Changes in net photosynthetic rate on a leaf area basis and anatomical properties during leaf development were studied in an evergreen broad‐leaved tree, Castanopsis sieboldii and an annual herb, Phaseolus vulgaris. In C. sieboldii, surface area of mesophyll cells facing the intercellular air spaces on a leaf area basis (Smes) was already considerable at the time of full leaf area expansion (FLE). However, surface area of chloroplasts facing the intercellular air spaces on a leaf area basis (Sc), and chlorophyll and Rubisco contents on a leaf area basis increased to attain their maximal values 15–40 d after FLE. In contrast, in P. vulgaris, chloroplast number on a leaf area basis, Sc and Smes at 10 d before FLE were two to three times greater than the steady‐state levels attained at around FLE. In C. sieboldii, the internal CO2 transfer conductance (gi) slightly increased for 10 d after FLE but then decreased toward the later stages. Limitation of photosynthesis by gi was only about 10% at FLE, but then increased to about 30% at around 40 d after FLE. The large limitation after FLE by gi was probably due to the decrease in CO2 concentration in the chloroplast caused by the increases in thickness of mesophyll cell walls and in Rubisco content per chloroplast surface area. These results clearly showed that: (1) in C. sieboldii, chloroplast development proceeded more slowly than mesophyll cell expansion and continued well after FLE, whereas in P. vulgaris these processes proceeded synchronously and were completed by FLE; (2) after FLE, photosynthesis in leaves of C. sieboldii was markedly limited by gi. From these results, it is suggested that, in the evergreen broad‐leaved trees, mechanical protection of mesophyll cells has priority over the efficient CO2 transfer and quick construction of the chloroplasts.  相似文献   

19.
Photosynthesis and dark respiration were studied during andafter the spring flush in Brachystegia spiciformis Benth. Variousparameters were examined including anthocyanin and chlorophyllcontent, Hill reaction activity, and gas exchange. Total chlorophyllcontent steadily increased reaching a constant value in fullyexpanded (25-d-old) leaves, whereas anthocyanin concentrationincreased as the chlorophylls but rapidly declined after 23d. Hill reaction measurements confirmed that leaf material fromevery stage of the flush (bud burst to mature leaves) was photosyntheticallyfunctional. The activity was low in flushing leaves with highanthocyanin content and then dramatically increased as leafanthocyanin content declined. Oxygen exchange measurements showedflushing leaves to have lower photosynthetic rates but higherrespiratory activity than mature leaves (60% and 120%, respectively).Gas exchange and in vitro electron transport were also generallycorrelated; Hill reaction activity was 128% of gas exchangein mature leaves and 92% in flushing leaves. It is concludedthat although photosynthetic rates are lower and respirationrates are higher in flushing leaves than in mature, fully expandedleaves, flushing leaves are fully photosynthetically competentand apparently require no net input of carbon for growth anddevelopment. Key words: Brachystegia, Hill reaction, photosynthesis, Zimbabwe  相似文献   

20.
In tropical rainforests, the increased light associated with the formation of treefall gaps can have a critical impact on the growth and survivorship of understory plants. Here we examine both leaf-level and whole-plant responses to simulated light gap formation by two common shade-tolerant shrubs, Hybanthus prunifolius and Ouratea lucens. The species were chosen because they differed in leaf lifespans, a trait that has been correlated with a number of leaf- and plant-level processes. Ouratea leaves typically live about 5 years, while Hybanthus leaves live less than 1 year. Potted plants were placed in the understory shade for 2 years before transfer to a light gap. After 2 days in high light, leaves of both species showed substantial photoinhibition, including reduced CO2 fixation, F v/F m and light use efficiency, although photoinhibition was most severe in Hybanthus. After 17 days in high light, leaves of both species were no longer photoinhibited. In response to increased light, Ouratea made very few new leaves, but retained most of its old leaves which increased photosynthetic capacity by 50%. Within a few weeks of transfer to high light, Hybanthus had dropped nearly all of its shade leaves and made new leaves that had a 2.5-fold greater light-saturated photosynthetic rate. At 80 days after transfer, the number of new leaves was 4.9-fold the initial leaf number. After 80 days in high light, Hybanthus had approximately tenfold greater productivity than Ouratea when leaf area, photosynthetic capacity, and leaf dark respiration rate were all taken into account. Although both species are considered shade tolerant, we found that their growth responses were quite different following transfer from low to high light. The short-lived Hybanthus leaves were quickly dropped, and a new canopy of sun leaves was produced. In contrast, Ouratea showed little growth response at the whole-plant level, but a greater ability to tolerate light stress and acclimate at the leaf level. These differences are consistent with predictions based on leaf lifespan and are discussed within the context of other traits associated with shade-tolerant syndromes. Received: 25 March 1999 / Accepted: 16 August 1999  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