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 31 毫秒
1.
Phytoplankton growth rate in response to irradiance can be approximated by a hyperbola defined by three coefficients: i) initial slope (α); ii) asymptote (μm); and, iii) X-axis intercept or compensation irradiance (Ic). To mathematically represent the interaction of temperature and irradiance on growth rate, one must describe the relationship between these constants and temperature. The marine diatom, Skeletonema costatum (Greville) Cleve, was grown in unialgal culture at different levels of irradiance and 2-3 photoperiods at 0, 5, 10, 16 and 22 C. The value of Ic is ca. 1.0 ly·day?1 or less at all temperatures. The initial slope (div·ly?1) is a “u-shaped” function of temperature described by the second degree polynomial, α= 0.25–0.02T+0.001T2. Within the range 0–10 C, μm (div·day?1) is an exponential function of temperature described by the equation, μm= 0.48 exp (0.126T). At each temperature and selected levels of irradiance, cell size and cellular content of C, N and chl a were determined. The C:chl a and N:chl a ratios increased with irradiance because of increases in C and decreases in chl a. At lower temperatures (0, 5, 10 C), the rate of increase of both ratios with irradiance was greater than at the higher temperatures (16, 22 C). Cellular content of N was independent of irradiance and temperature, and the C:N ratio ranged from 5 to 8 with a slight tendency to lower values at low irradiance. Cell volume was not influenced by either temperature or irradiance.  相似文献   

2.
3.
Consider the model yijk=u ± ai ± bi ± cij ± eijk i=1, 2,…, t; j=1, 2,…b; k=1, 2,…,nij where μ is a constant and ai, bi, cij are distributed independently and normally with zero means and variances Δ2 Δ2/bdij and δ2 respectively. It is assumed that di's, and dij's are known (positive) constants (for all i and j). In this paper procedures for estimating the variance components (Δ2, Δ2b and Δ2a) and for testing the hypothesis Hoc2c2 = y3 and Hoa2b2 = y4 (where y2, y3, and y4, are specified constants) are presented. A generalization for the mixed model case is discussed in the last section.  相似文献   

4.
Thermoregulatory sweating [total body (m sw,b), chest (m sw,c) and thigh (m sw,t) sweating], body temperatures [oesophageal (T oes) and mean skin temperature (T sk)] and heart rate were investigated in five sleep-deprived subjects (kept awake for 27 h) while exercising on a cycle (45 min at approximately 50% maximal oxygen consumption) in moderate heat (T air andT wall at 35° C. Them sw,c andm sw,t were measured under local thermal clamp (T sk,1), set at 35.5° C. After sleep deprivation, neither the levels of body temperatures (T oes,T sk) nor the levels ofm sw, b,m sw, c orm sw, t differed from control at rest or during exercise steady state. During the transient phase of exercise (whenT sk andT sk,1 were unvarying), them sw, c andm sw, t changes were positively correlated with those ofT oes. The slopes of them sw, c versusT oes, orm sw, t versusT oes relationships remained unchanged between control and sleep-loss experiments. Thus the slopes of the local sweating versusT oes, relationships (m sw, c andm sw, t sweating data pooled which reached 1.05 (SEM 0.14) mg·cm–2·min–1°C–1 and 1.14 (SEM 0.18) mg·cm–2·min–1·°C–1 before and after sleep deprivation) respectively did not differ. However, in our experiment, sleep deprivation significantly increased theT oes threshold for the onset of bothm sw, c andm sw, t (+0.3° C,P<0.001). From our investigations it would seem that the delayed core temperature for sweating onset in sleep-deprived humans, while exercising moderately in the heat, is likely to have been due to alterations occurring at the central level.  相似文献   

5.
We examined the in situ CO2 gas-exchange of fruits of a tropical tree, Durio zibethinus Murray, growing in an experimental field station of the Universiti Pertanian Malaysia. Day and night dark respiration rates were exponentially related to air temperature. The temperature dependent dark respiration rate showed a clockwise loop as time progressed from morning to night, and the rate was higher in the daytime than at night. The gross photosynthetic rate was estimated by summing the rates of daytime dark respiration and net photosynthesis. Photosynthetic CO2 refixation, which is defined as the ratio of gross photosynthetic rate to dark respiration rate in the daytime, ranged between 15 and 45%. The photosynthetic CO2 refixation increased rapidly as the temperature increased in the lower range of air temperature T c (T c <28.5 °C), while it decreased gradually as the temperature increased in the higher range (T c 28.5 °C). Light dependence of photosynthetic CO2 refixation was approximated by a hyperbolic formula, where light saturation was achieved at 100 mol m–2 s–1 and the asymptotic CO2 refixation was determined to be 37.4%. The estimated gross photosynthesis and dark respiration per day were 1.15 and 4.90 g CO2 fruit–1, respectively. Thus the CO2 refixation reduced the respiration loss per day by 23%. The effect of fruit size on night respiration rate satisfied a power function, where the exponent was larger than unity.  相似文献   

6.
A low molecular weight acid phosphatase was purified to homogeneity from chicken heart with a specific activity of 42 U/mg and a recovery of about 1%. Nearly 800 fold purification was achieved. The molecular weight was estimated to be 18 kDa by SDS-polyacrylamide gel electrophoresis. Para-nitrophenyl phosphate, phenyl phosphate and flavin mononucleotide were efficiently hydrolysed by the enzyme and found to be good substrates. Fluoride and tartrate had no inhibitory effect while phosphate, vanadate and molybdate strongly inhibited the enzyme. The acid phosphatase was stimulated in the presence of glycerol, ethylene glycol, methanol, ethanol and acetone, which reflected the phosphotransferase activity. When phosphate acceptors such as ethylene glycol concentrations were increased, the ratio of phosphate transfer to hydrolysis was also increased, demonstrating the presence of a transphosphorylation reaction where an acceptor can compete with water in the rate limiting step involving hydrolysis of a covalent phospho enzyme intermediate. Partition experiments carried out with two substrates, para-nitrophenyl phosphate and phenyl phosphate, revealed a constant product ratio of 1.7 for phosphotransfer to ethylene glycol versus hydrolysis, strongly supporting the existence of common covalent phospho enzyme intermediate. A constant ratio of K cat/K m, 4.3×104, found at different ethylene glycol concentrations, also supported the idea that the rate limiting step was the hydrolysis of the phospho enzyme intermediate.  相似文献   

7.
2013年5月至2014年6月,对干旱河谷区云南松(Pinus yunnanensis)人工林进行增加降水试验,试验设置对照(CK,0 mm m~(-2)a~(-1))、增水10%(A1,80 mm m~(-2)a~(-1))、增水20%(A2,160 mm m~(-2)a~(-1))和增水30%(A3,240 mm m~(-2)a~(-1))4个处理水平。采用LI-8100开路式土壤碳通量测量系统测定每月土壤呼吸速率。结果表明,4个处理云南松人工林土壤呼吸速率均呈明显的季节变化,7月最高,2月最低。与CK相比,A1年均土壤呼吸速率无显著性差异(P0.05),A2显著增加了12.88%(P0.05),而A3明显减少了17.71%(P0.05)。3个增水处理均提高了土壤呼吸的温度敏感性,减弱了土壤呼吸与土壤湿度的关系。与土壤温度相比,土壤湿度对土壤呼吸的影响相对较小。增水增加了湿季土壤微生物碳、氮含量,干季对微生物碳含量无影响,但明显降低了微生物氮含量。这说明,降水增加对干旱河谷区云南松人工林土壤呼吸的影响是不尽相同的,适当的增水会促进土壤呼吸,而过量的增水会抑制土壤呼吸。  相似文献   

8.
Pigeon flight in a wind tunnel   总被引:2,自引:0,他引:2  
Summary Core temperatureT c, breast temperatureT s–br and leg temperatureT s–1 were measured simultaneously in pigeons during rest and flight in a wind tunnel, using thermistors.MeanT c at rest is 39.8±0.7°C and is independent of ambient temperatureT a (10–30°C). In the first minutes of flight,T c increases to 1.5–3.0°C above resting level and remains at this higher level. This hyperthermia increases withT a (v=const.). It is±constant in the lowT a range (10.6–13.9°C) at flight speeds v ranging from 10–18 m s–1 and normal body mass, but increases with v and elevated body mass in the highT a range (23.7–28.8°C). T s–1 is adapted toT a at rest and increases in flight up to 3–4°C belowT c. This increase inT s–1 is linear toT a. T s–br is always lower thanT c, in extreme cases reaching restingT c in flight.Supported by the Deutsche Forschungsgemeinschaft  相似文献   

9.
Abstract

This research is focussed on kinetic, thermodynamic and thermal inactivation of a novel thermostable recombinant α-amylase (Tp-AmyS) from Thermotoga petrophila. The amylase gene was cloned in pHIS-parallel1 expression vector and overexpressed in Escherichia coli. The steady-state kinetic parameters (Vmax, Km, kcat and kcat/Km) for the hydrolysis of amylose (1.39?mg/min, 0.57?mg, 148.6?s?1, 260.7), amylopectin (2.3?mg/min, 1.09?mg, 247.1?s?1, 226.7), soluble starch (2.67?mg/min, 2.98?mg, 284.2?s?1, 95.4) and raw starch (2.1?mg/min, 3.6?mg, 224.7?s?1, 61.9) were determined. The activation energy (Ea), free energy (ΔG), enthalpy (ΔH) and entropy of activation (ΔS) at 98?°C were 42.9?kJ mol?1, 74?kJ mol?1, 39.9?kJ mol?1 and ?92.3 J mol?1 K?1, respectively, for soluble starch hydrolysis. While ΔG of substrate binding (ΔGE-S) and ΔG of transition state binding (ΔGE-T) were 3.38 and ?14.1?kJ mol?1, respectively. Whereas, EaD, Gibbs free energy (ΔG*), increase in the enthalpy (ΔH*) and activation entropy (ΔS*) for activation of the unfolding of transition state were 108, 107, 105?kJ mol?1 and ?4.1 J mol?1 K?1. The thermodynamics of irreversible thermal inactivation of Tp-AmyS revealed that at high temperature the process involves the aggregation of the protein.  相似文献   

10.
In short-term field trials at combinations of ambient temperature (°C) and insolation (W·m−2), larval Colorado potato beetles (Leptinotarsa decemlineata [Say] [Coleoptera: Chrysomelidae]) were observed after their release on the adaxial surface of leaflets on potato plants (Solanum tuberosum L. Solanaceae). The larvae either began feeding or moved under the leaflet; mean interval from release to expression of these behaviors (2.9±0.05 min [n =358]) was independent of air temperature and insolation. Proportion of larvae moving under the leaflet increased logistically with both air temperature and insolation. A 1 W·m−2 change in insolation (P) evoked the same effect on this proportion as a 0.0838 °C change in air temperature (T a ), so the two quantities were combined as T*=T a +P·0.0838 °C/(W·m−2), which has units of °C. The proportion of larvae moving under the leaflet increased logistically with T*. In 1-day field trials we monitored air temperature, insolation and proportion of larvae under the leaflet, and compared the latter to predictions from the logistic regression derived from the short-term trials. Consistently more larvae occurred under leaflets than predicted from the logistic regression; this bias diminished as T* increased until at T*≥40 °C, observed and predicted proportions were equal. This pattern of deviation from the predictions of the logistic regression is consistent with a thermoregulatory strategy in which larvae move away from hostile conditions, rather than seek optimal conditions.  相似文献   

11.
Multi-proton spin-echo images were collected from cold-acclimated winter wheat crowns (Triticum aestivum L.) cv. Cappelle Desprez at 400 MHz between 4 and ?4 °C. Water proton relaxation by the spin-spin (T2) mechanism from individual voxels in image slices was found to be mono-exponential. The temperature dependence of these relaxation rates was found to obey Arrhenius or absolute rate theory expressions relating temperature, activation energies and relaxation rates, Images whose contrast is proportional to the Arrhenius activation energy (Ea), Gibb's free energy of activation (ΔG?), and the entropy of activation (ΔS?) for water relaxation on a voxel basis were constructed by post-image processing. These new images exhibit contrast based on activation energies rather than rules of proton relaxation. The temperature dependence of water proton T2 relaxation rates permits prediction of changes in the physical state of water in this tissue over modest temperature ranges. A simple model is proposed to predict the freezing temperature kof various tissue in wheat crowns. The average Ea and ΔH? for water proton T2 relaxation over the above temperature range in winter wheat tissue were ?6.4 ± 14.8 and ?8.6 ± 14.8kj mol?1, respectively. This barrier is considerably lower than the Ea for proton translation in ice at 0°C, which is reported to be between 46.0 and 56.5 kj mol?1  相似文献   

12.
《Free radical research》2013,47(5):393-399
The one-electron reduction potential of 3-amino-l, 2, 4-benzotriazine 1, 4-dioxide, tirapazamine (SR 4233) in aqueous solution has been determined by pulse radiol-ysis. Reversible electron transfer was achieved between radiolytically-generated one-electron reduced radicals of tirapazamine (T), and quinones or benzyl viologen as redox standards. The reduction potential Em7(T/T±) was -0.45 ± 0.01 V vs. NHE at pH 7. From the pH dependence of the reduction potential, pKa = 5.6 ± 0.2 was estimated for the tirapazamine radical, a value similar to the pKa determined by other methods.  相似文献   

13.
The purpose of this research was to investigate why and how mechanical milling results in an unexpected shift in differential scanning calorimetry (DSC) measured fusion enthalpy (∆fus H) and melting point (T m) of adipic acid, a pharmaceutical excipient. Hyper differential scanning calorimetry (hyper-DSC) was used to characterize adipic acid before and after ball-milling. An experimental study was conducted to evaluate previous postulations such as electrostatic charging using the Faraday cage method, crystallinity loss using powder X-ray diffraction (PXRD), thermal annealing using DSC, impurities removal using thermal gravimetric analysis (TGA) and Karl Fischer titration. DSC thermograms showed that after milling, the values of ∆fus H and T m were increased by approximately 9% and 5 K, respectively. Previous suggestions of increased electrostatic attraction, change in particle size distribution, and thermal annealing during measurements did not explain the differences. Instead, theoretical analysis and experimental findings suggested that the residual solvent (water) plays a key role. Water entrapped as inclusions inside adipic acid during solution crystallization was partially evaporated by localized heating at the cleaved surfaces during milling. The correlation between the removal of water and melting properties measured was shown via drying and crystallization experiments. These findings show that milling can reduce residual solvent content and causes a shift in DSC results.  相似文献   

14.
Šantrůček  J.  Schreiber  L.  Macková  J.  Vráblová  M.  Květoň  J.  Macek  P.  Neuwirthová  J. 《Photosynthesis research》2019,141(1):33-51

We suggest a new technique for estimating the relative drawdown of CO2 concentration (c) in the intercellular air space (IAS) across hypostomatous leaves (expressed as the ratio cd/cb, where the indexes d and b denote the adaxial and abaxial edges, respectively, of IAS), based on the carbon isotope composition (δ13C) of leaf cuticular membranes (CMs), cuticular waxes (WXs) or epicuticular waxes (EWXs) isolated from opposite leaf sides. The relative drawdown in the intracellular liquid phase (i.e., the ratio cc/cbd, where cc and cbd stand for mean CO2 concentrations in chloroplasts and in the IAS), the fraction of intercellular resistance in the total mesophyll resistance (rIAS/rm), leaf thickness, and leaf mass per area (LMA) were also assessed. We show in a conceptual model that the upper (adaxial) side of a hypostomatous leaf should be enriched in 13C compared to the lower (abaxial) side. CM, WX, and/or EWX isolated from 40 hypostomatous C3 species were 13C depleted relative to bulk leaf tissue by 2.01–2.85‰. The difference in δ13C between the abaxial and adaxial leaf sides (δ13CAB ? 13CAD, Δb–d), ranged from ??2.22 to +?0.71‰ (??0.09 ± 0.54‰, mean ± SD) in CM and from ??7.95 to 0.89‰ (??1.17 ± 1.40‰) in WX. In contrast, two tested amphistomatous species showed no significant Δb–d difference in WX. Δb–d correlated negatively with LMA and leaf thickness of hypostomatous leaves, which indicates that the mesophyll air space imposes a non-negligible resistance to CO2 diffusion. δ13C of EWX and 30-C aldehyde in WX reveal a stronger CO2 drawdown than bulk WX or CM. Mean values of cd/cb and cc/cbd were 0.90 ± 0.12 and 0.66 ± 0.11, respectively, across 14 investigated species in which wax was isolated and analyzed. The diffusion resistance of IAS contributed 20 ± 14% to total mesophyll resistance and reflects species-specific and environmentally-induced differences in leaf functional anatomy.

  相似文献   

15.
Internal conductances to CO2 transfer from the stomatal cavity to sites of carboxylation (gi) in hypostomatous sun-and shade-grown leaves of citrus, peach and Macadamia trees (Lloyd et al. 1992) were related to anatomical characteristics of mesophyll tissues. There was a consistent relationship between absorptance of photosynthetically active radiation and chlorophyll concentration (mmol m?2) for all leaves, including sclerophyllous Macadamia, whose transmittance was high despite its relatively thick leaves. In thin peach leaves, which had high gi, the chloro-plast volume and mesophyll surface area exposed to intercellular air spaces (ias) per unit leaf area were similar to those in the thicker leaves of the evergreen species. Peach leaves, however, had the lowest leaf dry weight per area (D/a), the lowest tissue density (Td) and the highest chloro-plast surface area (Sc) exposed to ias. There were negative correlations between gi and leaf thickness or D/a, but positive correlations between gi and Sc or Sc/Td. We developed a one-dimensional diffusion model which partitioned gi into a gaseous diffusion conductance through the ias (gias) plus a liquid-phase conductance through mesophyll cell walls (gcw). The model accounted for a significant amount of variation (r2=0.80) in measured gi by incorporating both components. The gias component was related to the one-dimensional path-length for diffusion across the mesophyll and so was greater in thinner peach leaves than in leaves of evergreen species. The gcw component was related to tissue density and to the degree of chloroplast exposure to the ias. Thus the negative correlations between gi and leaf thickness or D/a related to gias whereas positive correlations between gi and Sc or Sc/Td, related to gcw. The gcw was consistently lower than gias, and thus represented a greater constraint on CO2 diffusion in the mesophylls of these hypostomatous species.  相似文献   

16.
We investigated the tree growth and physiological response of five pine forest stands in relation to changes in atmospheric CO2 concentration (ca) and climate in the Iberian Peninsula using annually resolved width and δ13C tree‐ring chronologies since ad 1600. 13C discrimination (Δ≈ci/ca), leaf intercellular CO2 concentration (ci) and intrinsic water‐use efficiency (iWUE) were inferred from δ13C values. The most pronounced changes were observed during the second half of the 20th century, and differed between stands. Three sites kept a constant ci/ca ratio, leading to significant ci and iWUE increases (active response to ca); whereas a significant increase in ci/ca resulted in the lowest iWUE increase of all stands at a relict Pinus uncinata forest site (passive response to ca). A significant decrease in ci/ca led to the greatest iWUE improvement at the northwestern site. We tested the climatic signal strength registered in the δ13C series after removing the low‐frequency trends due to the physiological responses to increasing ca. We found stronger correlations with temperature during the growing season, demonstrating that the physiological response to ca changes modulated δ13C and masked the climate signal. Since 1970 higher δ13C values revealed iWUE improvements at all the sites exceeding values expected by an active response to the ca increase alone. These patterns were related to upward trends in temperatures, indicating that other factors are reinforcing stomatal closure in these forests. Narrower rings during the second half of the 20th century than in previous centuries were observed at four sites and after 1970 at all sites, providing no evidence for a possible CO2‘fertilization’ effect on growth. The iWUE improvements found for all the forests, reflecting both a ca rise and warmer conditions, seem to be insufficient to compensate for the negative effects of the increasing water limitation on growth.  相似文献   

17.
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   

18.
Leaf‐level measurements have shown that mesophyll conductance (gm) can vary rapidly in response to CO2 and other environmental factors, but similar studies at the canopy‐scale are missing. Here, we report the effect of short‐term variation of CO2 concentration on canopy‐scale gm and other CO2 exchange parameters of sunflower (Helianthus annuus L.) stands in the presence and absence of abscisic acid (ABA) in their nutrient solution. gm was estimated from gas exchange and on‐line carbon isotope discrimination (Δobs) in a 13CO2/12CO2 gas exchange mesocosm. The isotopic contribution of (photo)respiration to stand‐scale Δobs was determined with the experimental approach of Tcherkez et al. Without ABA, short‐term exposures to different CO2 concentrations (Ca 100 to 900 µmol mol?1) had little effect on canopy‐scale gm. But, addition of ABA strongly altered the CO2‐response: gm was high (approx. 0.5 mol CO2 m?2 s?1) at Ca < 200 µmol mol?1 and decreased to <0.1 mol CO2 m?2 s?1 at Ca >400 µmol mol?1. In the absence of ABA, the contribution of (photo)respiration to stand‐scale Δobs was high at low Ca (7.2‰) and decreased to <2‰ at Ca > 400 µmol mol?1. Treatment with ABA halved this effect at all Ca.  相似文献   

19.
In a previous study we have shown that triiodothyronine (T3) added to a serum-free medium supplemented with insulin, transferrin, and selenous acid (ITS) can stimulate Caco-2 cell differentiation. In this study we have focused on the effects of T3on sucrase activity. The results obtained demonstrate that T3(50 nM) does not change Caco-2 cell proliferation but enhances sucrase activity from 50 to 80%. Similar increases were observed whether or not insulin was present in the culture medium, showing that there was no synergistic effect between T3and insulin on sucrase activity. Moreover, T3acts specifically during the differentiation period since addition of T3to the defined TS medium before confluency is reached does not stimulate sucrase activity. Sucrase kinetic parameters were evaluated for the first time in Caco-2 cells under various culture conditions. The presence of a single enzyme was verified, with aKmof about 7 mMand aVmaxaround 20 nmol of substrate hydrolyzed min−1mg−1of protein. Our results showed that T3did not change the enzyme's affinity for sucrose but doubled theVmax. Moreover, immunoblotting using anti-sucrase–isomaltase (SI) antibodies revealed an approximately twofold increase in the relative amount of SI immunoreactive protein in T3-stimulated cells compared to untreated cells. Results obtained by both Northern hybridization and RT-PCR amplification showed a significant increase in SI mRNA contents. These results suggest that T3acts primarily on sucrase expression at the mRNA level.  相似文献   

20.
Tiina Nõges 《Hydrobiologia》1996,338(1-3):91-103
The material for pigment analysis was collected 1–3 times a year from Lake Peipsi-Pihkva in 1983, 1987, 1988, 1991 and 1992–1995. Concentrations of chlorophyll a, b and c (Chla, Chlb, Chlc), pheopigment (Pheo) and adenosine triphosphate (ATP) were measured biweekly in 1985–1986. The mean of all Chla values was 20.2 mg m–1 (median 13.3 mg m–1) indicating the eutrophic state of the lake. Average Chlb, Chlc, Pheo and carotenoid (Car) contents were 3.7 mg m–3, 4.1 mg m–3, 3.0 mg m–3 and 4.8 mg m–3, respectively. The average Chlb/Chla ratio was 22.9%, Chlc/Chla 23.4%, Pheo/Chla 38%, Car/Chla 37% and ATP/Chla 3%, the medians being 14.3, 13.6, 17.5, 39.4 and 1.9%, respectively. The proportion of Chla in phytoplankton biomass was 0.41%, median 0.32%. There were no significant differences in temperature, oxygen concentration, Chla, and ATP between the surface and bottom water; the lake was polymictic during the vegetation period. The Chla concentration had its first peak in May followed by a decrease in June and July. In late summer Chla increased again achieving its seasonal maximum in late autumn. The ATP concentration was the highest during spring and early summer, decreasing drastically in autumn together with the decline of primary production. ATP/Chla was the highest during the clear water period in June and early July, which coincided also with the high proportion of Chla in phytoplankton biomass. The highest Chla occurred in November (average 37.2 mg m–3) when Secchi transparency was the lowest (1.05 m). Concentrations of Chlb, Chlc and carotenoids were the highest in August, that of Pheo in June. Concentrations of Chla and other pigments were the lowest in the northern part of Lake Peipsi (mean 14.7 mg m–3, median 12.5 mg m–3) and the highest in the southern part of Lake Pihkva (mean 47.9 mg m–3, median 16.3 mg m–3). An increase of Chla and decrease of Secchi depth could be noticed in 1983–1988, while in 1988–1994 the tendency was opposite.  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