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.
To gain information on extended flight energetics, quasi-natural flight conditions imitating steady horizontal flight were set by combining the tetheredflight wind-tunnel method with the exhaustion-flight method. The bees were suspended from a two-component aerodynamic balance at different, near optimum body angle of attack and were allowed to choose their own speed: their body mass and body weight was determined before and after a flight; their speed, lift, wingbeat frequency and total flight time were measured throughout a flight. These values were used to determine thrust, resultant aerodynamic force (magnitude and tilting angle), Reynolds number, total flight distance and total flight impulse. Flights in which lift was body weight were mostly obtained. Bees, flown to complete exhausion, were refed with 5, 10, 15 or 20 l of a 1.28-mol·l-1 glucose solution (energy content w=18.5, 37.0, 55.5 or 74.0 J) and again flown to complete exhaustion at an ambient temperature of 25±1.5°C by a flight of known duration such that the calculation of absolute and relative metabolic power was possible. Mean body mass after exhaustion was 76.49±3.52 mg. During long term flights of 7.47–31.30 min similar changes in flight velocity, lift, thrust, aerodynamic force, wingbeat frequency and tilting angle took place, independent of the volume of feeding solution. After increasing rapidly within 15 s a more or less steady phase of 60–80% of total flight time, showing only a slight decrease, was followed by a steeper, more irregular decrease, finally reaching 0 within 20–30 s. In steady phases lift was nearly equal to resultant aerodynamic force; tilting angle was 79.8±4.0°, thrust to lift radio did not vary, thrust was 18.0±7.4% of lift, lift was somewhat higher/equal/lower than body mass in 61.3%, 16.1%, 22.6% of all totally analysable flights (n=31). The following parameters were varied as functions of volume of feeding solution (5–20 l in steps of 5 l) and energy content. (18.5–74.0 J in steps of 18.5 J): total flight time, velocity, total flight distance, mean lift, thrust, mean resultant aerodynamic force, tilting angle, total flight impulse, wingbeat frequency, metabolic power and metabolic power related to body mass, the latter related to empty, full and mean (=100 mg) body mass. The following positive correlations were found: L=1.069·10-9 f 2.538; R=1.629·10-9 f 2.464; P m=7.079·10-8 f 2.456; P m=0.008v+0.008; P m=18.996L+0.022; P m=19.782R+0.021; P m=82.143T+0.028; P m=1.245·bm f 1.424 ; P mrel e=6.471·bm f 1.040 ; =83.248+0.385. The following negative correlations were found: V=3.939–0.032; T=1.324·10-4–0.038·10-4. Statistically significant correlations were not found in T(f), L(), R(), f(), P m(bm e), P m rel e(bm e), P m rel f(bm e), P m rel f(bm f).Abbreviations A(m2) frontal area - bl(m) body length - bm(mg) body mass - c(mol·1-1) glucose concentration of feeding solution - c D (dimensionless) drag coefficient, related to A - D(N) drag - F w(N) body weight - F wp weight of paper fragment lost at flight start - f wingbeat frequency (s-1) - g(=9.81 m·s-2) gravitational acceleration - I(Ns)=R(t) dt total impulse of a flight - L(N) lift vertical sustaining force component - P m(J·s-1=W) metabolic power - Pm ret (W·g-1) metabolic power, related to body mass - R(N) resultant aerodynamic force - Re v·bl·v -1 (dimensionless) Reynolds number, related to body length - s(m) v(t) dt virtual flight distance of a flight - s(km) total virtual flight distance - T (N) thrust horizontal force component of horizontal flight - T a (°C) ambient temperature - t(s) time - t tot (s or min) total flight time - v(m·s-1) flight velocity - v(l) volume of feeding solution - W (J) energy and energy content of V - ( °) body angle of attack between body longitudinal axis and flow direction - ( °) tilting angle ( 90°) between R and the horizont in horizontal flight v(=1.53·10-5m2·s-1 for air at 25°) kinematic viscosity - (=1.2 kg·m-3 at 25°C) air density  相似文献   

2.
We quantified metabolic power consumption as a function of wind speed in the presence and absence of simulated solar radiation in rock squirrels, Spermophilus variegatus, a diurnal rodent inhabiting arid regions of Mexico and the western United States. In the absence of solar radiation, metabolic rate increased 2.2-fold as wind speed increased from 0.25 to 4.0 m·s-1. Whole-body thermal resistance declined 56% as wind speed increased over this range, indicating that body insulation in this species is much more sensitive to wind disruption than in other mammals. In the presence of 950 W·m-2 simulated solar radiation, metabolic rate increased 2.3-fold as wind speed was elevated from 0.25 to 4.0 m·s-1. Solar heat gain, calculated as the reduction in metabolic heat production associated with the addition of solar radiation, increased with wind speed from 1.26 mW·g-1 at 0.25 m·s-1 to 2.92 mW·g-1 at 4.0 m·s-1. This increase is opposite to theoretical expectations. Both the unexpected increase in solar heat gain at elevated wind speeds and the large-scale reduction of coat insulation suggests that assumptions often used in heat-transfer analyses of animals can produce important errors.Abbreviations absorptivity of coat to solar radiation - kinematic viscosity of air (mm2·s-1) - reflectivity of coat to solar radiation - a r B expected at zero wind speed (s·m-1) - A P projected surface area of animal on plane perpendicular to solar beam (cm2) - A SKIN skin surface area (cm2) - b Coefficient describing change in r B with change in square-root of wind speed (s1.5·m1.5) - d hair diameter (m) - d characteristic dimension of animal (m) - D H thermal diffusivity of air (m2·s-1) - E evaporative heat loss (W·m-2) - I probability per unit coat depth that photon will strike hair - k constant equalling 1200 J·m-3·°C-1 - l C coat depth m) - l H hair length (m) - M metabolic rate (W·m-2) - n density of hairs of skin (m-2) - Q A solar heat gain to animal (W·m-2) - Q I solar irradiance intercepted by animal (W·m-2) - RQ respiratory quotient - r A thermal resistance of boundary layer (s·m-1) - r B whole-body thermal resistance (s·m-1) - r E thermal resistance between animal surface and environment s·m-1) - r R radiative resistance (s·m-1) - r S sum of r B and r E at 0.25 m·s-1 (s·m-1) - r T tissue thermal resistance s·m-1) - T AIR air temperature (°C) - T B body temperature (°C) - T E operative temperature of environment (°C) - T ES standard operative temperature of environment (°C) - u wind speed (m·s-1)  相似文献   

3.
Summary Drag forces and lift forces acting on honeybee trunks were measured by using specially built sensitive mechanical balances. Measurements were made on prepared bodies in good and in bad flight position, with and without legs, at velocities between 0.5 and 5m·s-1 (Reynolds numbers between 4·102 and 4·103) and at angles of attack between-20° and +20°. From the forces drag coefficients and lift coefficients were calculated. The drag coefficient measured with a zero angle of attack was 0.45 at 3v5m·s-1, 0.6 at 2m·s-1, 0.9 at 1m·s-1 and 1.35 at 0.5m·s-1, thus demonstrating a pronounced effect of Reynolds number on drag. These values are about 2 times lower (better) than those of a drag disc with the same diameter and attacked at the same velocity. The drag coefficient (related to constant minimal frontal area) was minimal at zero angle of attack, rising symmetrically to larger (+) and smaller (-) angles of attack in a non-linear fashion. The absolute value is higher and the rise is steeper at lower speeds or Reynolds numbers, but the incremental factors are independent of Reynolds number. For example, the drag coefficient is 1.44±0.05 times higher at an angle of attack of 20° than at one of 0°. On a double-logarithmic scale the slope of the drag versus Reynolds number plot was 1.5: with decreasing Reynolds number the relationship between drag and velocity changes from quadratic (Newton's law) to linear (viscous flow). Trunk drag was not systematically increased by the legs at any velocity or Reynolds number or any angle of attack. The legs appear to shape the trunk aerodynamically, to form a relatively low-drag trunk-leg system. The body is able to generate dynamic lift. Highly significant positive linear correlations between lift coefficient and angle of attack were determined for the trunk-leg system in the typical flight position. Lift coefficient was +0.05 at zero angle of attack (possibly attained during very fast flight), +0.1 at 5° (attained during fast flight), +0.25 at +20° (attained during slow flight) and +0.55 at 45° (attained whilst changing over to hovering). Average slope cL was 0.66±0.07, and average profile efficiency was 0.10. Non-wing lift contribution due to body form and banking only accounts for a few percent of body weight during fast flight. A non-wing lift contribution due to the legs has been demonstrated. The legs increase trunk lift by 23–24%. Reynolds number lift effects are present but of no biological significance. Force and power calculations do not support maximum flight speeds substantially higher than approximately 7m · s-1 relative to the ambient air. At this speed body drag attains 35% and body lift 8.4% of the body weight, and parasite power is 5% of the maximum metabolic power.Abbreviations angle of attack - A area - c drag coefficient - cL lift coefficient - D drag - F force - L lift - P power - Q quotient - Re Reynolds number - density - dliding number - O2 oxygen consumption - W work - v kinematic viscosity - efficiency - v velocity  相似文献   

4.
Cross-flow filtration (CFF) has been investigated as a method of separating filamentously growing fungal cells and purifying the polysaccharide produced. The effects of transmembrane pressure, module geometry (e.g. channel height or tube diameter), tangential feed velocity and cell as well as polysaccharide concentration are discussed. Apart from these experiments, influences by the recirculation pump used are shown.List of Symbols b f fouling index - b factor refering to the behaviour of the sublayer - C kg · m–3 concentration - C g kg · m–3 solute concentration at the membrane - C b kg · m–3 solute concentration in the bulk phase - D s-1 shear rate - k m · s–1 mass-transfer coefficient - K mPa · sn consistency index - n flow behaviour index - P w m3 · s–1 · m–2 rate of permeation - P w1 m3 · s–1 · m–2 rate of permeation at 1 minute - P w m3 · s–1 · m–2 rate of permeation at the beginning - p Pa pressure - Q m2 largest cross-section of a particle - q m2 smallest cross-section of a particle - Re Reynolds number - R f –1 fouling resistance - R m m–1 membrane resistance - t s time - w m · s–1 tangential feed velocity Greek Symbols friction factor - pTM Pa transmembrane pressure - mPa · s shear viscosity - sp specific viscosity (rel. increase of viscosity sp=rel-1) - [] m3· kg–1 intrinsic viscosity - w m2 · s–1 kinematic viscosity - kg · m–3 density Indices b bulk - cell cells - f fouling - g gelling - PS polysaccharide - rel relative - sp specific - w water  相似文献   

5.
Short-term (up to 5 h) transfers of shade-adapted (100 mol · m–2 · s–1) clonal tissue of the marine macroalga Ulva rotundata Blid. (Chlorophyta) to higher irradiances (1700, 850, and 350 mol · m–2 · s–1) led to photoinhibition of room-temperature chlorophyll fluorescence and O2 evolution. The ratio of variable to maximum (Fv/Fm) and variable (Fv) fluorescence, and quantum yield () declined with increasing irradiance and duration of exposure. This decline could be resolved into two components, consistent with the separation of photoinhibition into energy-dissipative processes (photoprotection) and damage to photosystem II (PSII) by excess excitation. The first component, a rapid decrease in Fv/Fm and in Fv, corresponds to an increase in initial (Fo) fluorescence and is highly sensitive to 1 mM chloramphenicol. This component is rapidly reversible under dim (40 mol · m–2 · s–1) light, but is less reversible with increasing duration of exposure, and may reflect damage to PSII. The second (after 1 h exposure) component, a slower decline in Fv/Fm and Fv with declining Fo, appears to be associated with the photoprotective interconversion of violaxanthin to zeaxanthin and is sensitive to dithiothreitol. The accumulation of zeaxanthin in U. rotundata is very slow, and may account for the predominance of increases in Fo at high irradiances.Abbreviations and Symbols CAP chloramphenicol - DTT dithiothreitol - Fo, Fm, Fv initial, maximum, and variable fluorescence - quantum yield - PFD photon flux density - PSII photosystem II To whom correspondence should be addressedWe are grateful to O. Björkman and S. Thayer, Carnegie Institution of Washington, Stanford, Cal., USA, for analysis of xanthophyll pigments reported here. This research was supported by National Science Foundation grant OCE-8812157 to C.B.O. and J.R. Support for G.L. was provided by a NSF-CNRS (Centre National de la Recherche Scientifique) exchange fellowship.  相似文献   

6.
D. H. Greer  W. A. Laing 《Planta》1988,175(3):355-363
Photoinhibition of photosynthesis was induced in intact kiwifruit (Actinidia deliciosa (A. Chev.) C. F. Liang et A. R. Ferguson) leaves grown at two photon flux densities (PFDs) of 700 and 1300 mol·m-2·s-1 in a controlled environment, by exposing the leaves to PFD between 1000 and 2000 mol·m-2·s-1 at temperatures between 10 and 25°C; recovery from photoinhibition was followed at the same range of temperatures and at a PFD between 0 and 500 mol·m-2·s-1. In either case the time-courses of photoinhibition and recovery were followed by measuring chlorophyll fluorescence at 692 nm and 77K and by measuring the photon yield of photosynthetic O2 evolution. The initial rate of photoinhibition was lower in the high-light-grown plants but the long-term extent of photoinhibition was not different from that in low-light-grown plants. The rate constants for recovery after photoinhibition for the plants grown at 700 and 1300 mol·m-2·s-1 or for those grown in shade were similar, indicating that differences between sun and shade leaves in their susceptibility to photoinhibition could not be accounted for by differences in capacity for recovery during photoinhibition. Recovery following photoinhibition was increasingly suppressed by an increasing PFD above 20 mol·m-2·s-1, indicating that recovery in photoinhibitory conditions would, in any case, be very slow. Differences in photosynthetic capacity and in the capacity for dissipation of non-radiative energy seemed more likely to contribute to differences in susceptibility to photoinhibition between sun and shade leaves of kiwifruit.Abbreviations and symbols F o , F m , F v instantaneous, maximum, variable fluorescence - F v /F m fluorescence ratio - F i =F v at t=0 - F F v at t= - K D rate constant for photochemistry - k(F p ) first-order rate constant for photoinhibition - k(F r ) first-order rate constant for recovery - PFD photon flux density - PSII photosystem II - i photon yield of O2 evolution (incident light)  相似文献   

7.
Summary The influence of the concentration of oxygen on lipase production by the fungus Rhizopus delemar was studied in different fermenters. The effect of oxygen limitation ( 47 mol/l) on lipase production by R. delemar is large as could be demonstrated in pellet and filamentous cultures. A model is proposed to describe the extent of oxygen limitation in pellet cultures. Model estimates indicate that oxygen is the limiting substrate in shake flask cultures and that an optimal inoculum size for oxygen-dependent processes can occur.Low oxygen concentrations greatly negatively affect the metabolism of R. delemar, which could be shown by cultivation in continuous cultures in filamentous growth form (Doptimal=0.086 h-1). Continuous cultivations of R. delemar at constant, low-oxygen concentrations are a useful tool to scale down fermentation processes in cases where a transient or local oxygen limitation occurs.Symbols and Abbreviations CO Oxygen concentration in the gas phase at time = 0 (kg·m-3) - CO 2i Oxygen concentration at the pellet liquid interface (kg·m-3) - CO 2i Oxygen concentration in the bulk (kg·m-3) - D Dilution rate (h-1) - IDO 2 Diffusion coefficient for oxygen (m2·s-1) - dw Dry weight of biomass (kg) - f Conversion factor (rs O 2 to oxygen consumption rate per m3) (-) - k Radial growth rate (m·s-1) - K Constant - kla Volumetric mass transfer coefficient (s-1) - klA Oxygen transfer rate (m-3·s-1) - kl Mass transfer coefficient (m·s-1) - K O 2 Affinity constant for oxygen (mol·m-3) - K w Cotton plug resistance (m-3·s-1) - M Henry coefficient (-) - NV Number of pellets per volume (m-3) - R Radius (m) - RO Radius of oxygen-deficient core (m) - RQ Respiration quotient (mol CO2/mol O2) - rs O 2 Specific oxygen consumption rate per dry weight biomass (kg O2·s-1[kg dw]-1) - rX Biomass production rate (kg·m-3·s-1) - SG Soytone glucose medium (for shake flask experiments) - SG 4 Soytone glucose medium (for tower fermenter and continuous culture experiments) - V Volume of medium (m-3) - X Biomass (dry weight) concentration (kg·m-3) - XR o Biomass concentration within RO for a given X (kg·m-3) - Y O 2 Biomass yield calculated on oxygen (kg dw/kg O2) - Thiele modulus - Efficiency factor =1-(RO/R)3 (-) - Growth rate (m-1·s-1·kg1/3) - Dry weight per volume of pellet (kg·m-3)  相似文献   

8.
Husen  Jia  Dequan  Li 《Photosynthetica》2002,40(1):139-144
The responses to irradiance of photosynthetic CO2 assimilation and photosystem 2 (PS2) electron transport were simultaneously studied by gas exchange and chlorophyll (Chl) fluorescence measurement in two-year-old apple tree leaves (Malus pumila Mill. cv. Tengmu No.1/Malus hupehensis Rehd). Net photosynthetic rate (P N) was saturated at photosynthetic photon flux density (PPFD) 600-1 100 (mol m-2 s-1, while the PS2 non-cyclic electron transport (P-rate) showed a maximum at PPFD 800 mol m-2 s-1. With PPFD increasing, either leaf potential photosynthetic CO2 assimilation activity (Fd/Fs) and PS2 maximal photochemical activity (Fv/Fm) decreased or the ratio of the inactive PS2 reaction centres (RC) [(Fi – Fo)/(Fm – Fo)] and the slow relaxing non-photochemical Chl fluorescence quenching (qs) increased from PPFD 1 200 mol m-2 s-1, but cyclic electron transport around photosystem 1 (RFp), irradiance induced PS2 RC closure [(Fs – Fo)/Fm – Fo)], and the fast and medium relaxing non-photochemical Chl fluorescence quenching (qf and qm) increased remarkably from PPFD 900 (mol m-2 s-1. Hence leaf photosynthesis of young apple leaves saturated at PPFD 800 mol m-2 s-1 and photoinhibition occurred above PPFD 900 mol m-2 s-1. During the photoinhibition at different irradiances, young apple tree leaves could dissipate excess photons mainly by energy quenching and state transition mechanisms at PPFD 900-1 100 mol m-2 s-1, but photosynthetic apparatus damage was unavoidable from PPFD 1 200 mol m-2 s-1. We propose that Chl fluorescence parameter P-rate is superior to the gas exchange parameter P N and the Chl fluorescence parameter Fv/Fm as a definition of saturation irradiance and photoinhibition of plant leaves.  相似文献   

9.
D. H. Greer  W. A. Laing  T. Kipnis 《Planta》1988,174(2):152-158
Photoinhibition of photosynthesis was induced in attached leaves of kiwifruit grown in natural light not exceeding a photon flux density (PFD) of 300 mol·m-2·s-1, by exposing them to a PFD of 1500 mol·m-2·s-1. The temperature was held constant, between 5 and 35° C, during the exposure to high light. The kinetics of photoinhibition were measured by chlorophyll fluorescence at 77K and the photon yield of photosynthetic O2 evolution. Photoinhibition occurred at all temperatures but was greatest at low temperatures. Photoinhibition followed pseudo first-order kinetics, as determined by the variable fluorescence (F v) and photon yield, with the long-term steady-state of photoinhibition strongly dependent on temperature wheareas the observed rate constant was only weakly temperature-dependent. Temperature had little effect on the decrease in the maximum fluorescence (F m) but the increase in the instantaneous fluorescence (F o) was significantly affected by low temperatures in particular. These changes in fluorescence indicate that kiwifruit leaves have some capacity to dissipate excessive excitation energy by increasing the rate constant for non-radiative (thermal) energy dissipation although temperature apparently had little effect on this. Direct photoinhibitory damage to the photosystem II reaction centres was evident by the increases in F o and extreme, irreversible damage occurred at the lower temperatures. This indicates that kiwifruit leaves were most susceptible to photoinhibition at low temperatures because direct damage to the reaction centres was greatest at these temperatures. The results also imply that mechanisms to dissipate excess energy were inadequate to afford any protection from photoinhibition over a wide temperature range in these shade-grown leaves.Abbreviations and symbols fluorescence yield correction coefficient - F o, F m, F v instantaneous, maximum, variable fluorescence - K D, K F, K P, K T rate constants for non-radiative energy dissipation, fluorescence, photochemistry, energy transfer to photosystem I - PFD photon flux density - PSI, II photosystem I, II - i photon yield of photosynthesis (incident light)  相似文献   

10.
A pressure-clamp technique was devised for the direct measurement of cell-to-cell and apoplasmic components of root hydraulic conductance; the experimental results were analyzed in terms of a theoretical model of water and solute flow, based on a composite membrane model of the root. When water is forced under a constant pressure into a cut root system, an exponential decay of flow is observed, until a constant value is attained; when pressure is released, a reverse water flow out of the root system is observed which shows a similar exponential behavour. The model assumes that the transient flow occurs through a cell-to-cell pathway and the observed decrease is the result of accumulation of solutes in front of the root semi-permeable membrane, whilst the steady-state component results from the movement of water through the parallel apoplasmic pathway. Root conductance components are estimated by fitting the model to experimental data. The technique was applied to the root systems of potted cherry (Prunus avium L.) seedlings; average apoplasmic conductance was 15.5 × 10–9m3· s–1· MPa–1, with values ranging from 12.0 × 10–9 to 18.5 × 10–9m3· s–1· MPa–1; average cell-to-cell conductance was 11.7 × 109 m3· s–1· MPa–1, with values ranging from 8.5 × 10–9 to 15.3 × 10–9 m3 · s–1·MPa–1. Cell-to-cell conductance amounted on average to 43% of total root conductance, with values between 41 and 45%. Leaf specific conductance (conductance per unit of leaf area supported) of the root systems ranged from 2.7 × 10–8 to 5.6 × 10–8 m· s–1·MPa–1, with an average of 3.7 × 10–8 m · s–1·MPa–1. The newly developed technique allows the interaction of mass flow of water and of solutes to be explored in the roots of soil-grown plants.Abbreviations and Symbols A Lp root hydraulic conductance - AaL p a root apoplasmic conductance - AccL p cc root cell-to-cell conductance - Cs(t) concentration of solutes in apical root compartment at time t - Jv flow of water through the root - J v a apoplasmic flow of water - Jv/cc cell-to-cell flow of water - LSC leaf specific conductance of the root system - P root hydrostatic pressure - Pappl applied pressure - s(t) root osmotic pressure at time t - m osmotic pressure of rooting medium - reflection coefficient of root membrane - time constant of cell-to-cell flow decay This research was funded within the EC Project Long-term effects of CO2-increase and climate change on European forests (LTEEF) (EV5V-CT94-0468); F.M. was supported by a Ministero dell' Universitá e della Ricerca Scientifica e Tecnologica — British Council agreement (Project The ecological significance of cavitation in woody plants); M.C. was supported by a Consiglio Nazionale delle Ricerche — British Council agreement. We gratefully thank Prof. P.G. Jarvis (University of Edinburgh, UK) for revising an earlier version of this paper and Prof. E. Steudle (University of Bayreuth, Germany) for helpful comments.  相似文献   

11.
Leaves of Populus balsamifera grown under full natural sunlight were treated with 0, 1, or 2 l SO2·1-1 air under one of four different photon flux densities (PFD). When the SO2 exposures took place in darkness or at 300 mol photons·m-2·s-1, sulfate accumulated to the levels predicted by measurements of stomatal conductance during SO2 exposure. Under conditions of higher PFD (750 and 1550 mol·m-2·s-1), however, the predicted levels of accumulated sulfate were substantially higher than those obtained from anion chromatography of the leaf extracts. Light-and CO2-saturated capacity as well as the photon yield of photosynthetic O2 evolution were reduced with increasing concentration of SO2. At 2 l SO2·1-1 air, the greatest reductions in both photosynthetic, capacity and photon yield occurred when the leaves were exposed to SO2 in the dark, and increasingly smaller reductions in each occurred with increasing PFD during SO2 exposure. This indicates that the inhibition of photosynthesis resulting from SO2 exposure was reduced when the exposure occurred under conditions of higher light. The ratio F v/F M (variable/maximum fluorescence emission) for photosyntem II (PSII), a measure of the photochemical efficiency of PSII, remained unaffected by exposure of leaves to SO2 in the dark and exhibited only moderate reductions with increasing PFD during the exposure, indicating that PSII was not a primary site of damage by SO2. Pretreatment of leaves with SO2 in the dark, however, increased the susceptibility of PSII to photoinhibition, as such pretreated leaves exhibited much greater reductions inF V/F M when transferred to moderate or high light in air than comparable control leaves.Abbreviations and symbols A1200 photosynthetic capacity (CO2-saturated rate of O2 evolution at 1200 mol photons·m-2·s-1) - Fo instantaneous fluorescence emission - FM maximum fluorescence emission - FV variable fluorescence emission - PFD photon flux density (400–700 nm) - PSII photosystem II  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 biogenic amine octopamine was injected into the haemolymph of 20-days old male locusts,Locusta migratoria, and the content of fructose 2,6-bisphosphate, a potent activator of glycolysis, was measured in the flight muscle after various time. Octopamine brought about a transient increase in fructose 2,6-bisphosphate. After the injection of 10 l of 10 mmol·l-1 d, l-octopamine fructose 2,6-bisphosphate was increased by 61% within 2 min. Ten minutes after the injection fructose 2,6-bisphosphate was increased to 6.71±0.89 nmol·g-1 flight muscle, almost 300% over the control value. Flight caused fructose 2,6-bisphosphate in flight muscle to decrease, but this decrease was counteracted by octopamine injected into the haemolymph of flying locusts. Octopamine and fructose 2,6-bisphosphate may act as signals to stimulate the oxidation of carbohydrate and to integrate muscle performance and metabolism. This mechanism appears particularly significant in the initial stage of flight when carbohydrates are the main fuel.Abbreviations F2,6P2 fructose 2,6-bisphosphate - F6P fructose 6-phosphate - PFK1 6-phosphofructokinase (EC 2.7.1.11) - P i inorganic phosphate - PP i -PFK pyrophosphate dependent fructose 6-phosphate phosphotransferase (EC 2.7.1.90)  相似文献   

14.
Data for the maximum carboxylation velocity of ribulose-1,5-biosphosphate carboxylase, Vm, and the maximum rate of whole-chain electron transport, Jm, were calculated according to a photosynthesis model from the CO2 response and the light response of CO2 uptake measured on ears of wheat (Triticum aestivum L. cv. Arkas), oat (Avena sativa L. cv. Lorenz), and barley (Hordeum vulgare L. cv. Aramir). The ratio Jm/Vm is lower in glumes of oat and awns of barley than it is in the bracts of wheat and in the lemmas and paleae of oat and barley. Light-microscopy studies revealed, in glumes and lemmas of wheat and in the lemmas of oat and barley, a second type of photosynthesizing cell which, in analogy to the Kranz anatomy of C4 plants, can be designated as a bundle-sheath cell. In wheat ears, the CO2-compensation point (in the absence of dissimilative respiration) is between those that are typical for C3 and C4 plants.A model of the CO2 uptake in C3–C4 intermediate plants proposed by Peisker (1986, Plant Cell Environ. 9, 627–635) is applied to recalculate the initial slopes of the A(pc) curves (net photosynthesis rate versus intercellular partial pressure of CO2) under the assumptions that the Jm/Vm ratio for all organs investigated equals the value found in glumes of oat and awns of barley, and that ribulose-1,5-bisphosphate carboxylase is redistributed from mesophyll to bundle-sheath cells. The results closely match the measured values. As a consequence, all bracts of wheat ears and the inner bracts of oat and barley ears are likely to represent a C3–C4 intermediate type, while glumes of oat and awns of barley represent the C3 type.Abbreviations A net photosynthesis rate (mol·m-2·s-1) - Jm maximum rate of whole-chain electron transport (mol·e-·m-2·s-1) - pc (bar) intercellular partial pressure of CO2 - PEP phosphoenolpyruvate - PPFD photosynthetic photon flux density (mol quanta·m-2·s-1) - RuBPCase ribulose bisphosphate carboxylase/oxygenase - RuBP ribulose bisphosphate - Vm maximum carboxylation velocity of RuBPCase (mol·m-2·s-1) - T* CO2 compensation point in the absence of dissimilative respiration (bar)  相似文献   

15.
Measurement of the light response of photosynthetic CO2 uptake is often used as an implement in ecophysiological studies. A method is described to calculate photosynthetic parameters, such as the maximum rate of whole electron transport and dissimilative respiration in the light, from the light response of CO2 uptake. Examples of the light-response curves of flag leaves and ears of wheat (Triticum aestivum cv. ARKAS) are shown.Abbreviations and symbols A net photosynthesis rate - D 1 rate of dissimilative respiration occurring in the light - f loss factor - I incident PPFD - I effective absorbed PPFD - J rate of whole electron transport - J m maximum rate of whole electron transport - p c intercellular CO2 partial pressure - PPFD photosynthetic photon flux density - q effectivity factor for the use of light (electrons/quanta) - absorption coefficient - I * CO2 compensation point in the absence of dissimilative respiration (bar) - II conversion factor for calculation of CO2 uptake from the rate of whole electron transport - convexity factor Gas-exchange rates relate to the projective area and are given in mol·m-2·s-1. Electron-transport rates are given in mol electrons·m-2·s-1; PPFD is given in mol quanta·m-2·s-1.  相似文献   

16.
A modified Rotating Biological Contactor (RBC) was used for the treatability studies of synthetic tapioca wastewaters. The RBC used was a four stage laboratory model and the discs were modified by attaching porous nechlon sheets to enhance biofilm area. Synthetic tapioca wastewaters were prepared with influent concentrations from 927 to 3600 mg/l of COD. Three hydraulic loads were used in the range of 0.03 to 0.09 m3·m–2·d–1 and the organic loads used were in the range of 28 to 306 g COD· m–2·d–1. The percentage COD removal were in the range from 97.4 to 68. RBC was operated at a rotating speed of 18 rpm which was found to be the optimal rotating speed. Biokinetic coefficients based on Kornegay and Hudson models were obtained using linear analysis. Also, a mathematical model was proposed using regression analysis.List of Symbols A m2 total surface area of discs - d m active depth of microbial film onany rotating disc - K s mg ·l–1 saturation constant - P mg·m–2·–1 area capacity - Q l·d–1 hydraulic flow rate - q m3·m–2·d–1 hydraulic loading rate - S 0 mg·l–1 influent substrate concentration - S e mg·l–1 effluent substrate concentration - w rpm rotational speed - V m3 volume of the reactor - X f mg·l–1 active biomass per unit volume ofattached growth - X s mg·l–1 active biomass per unit volume ofsuspended growth - X mg·l–1 active biomass per unit volume - Y s yield coefficient for attachedgrowth - Y A yield coefficient for suspendedgrowth - Y yield coefficient, mass of biomass/mass of substrate removed Greek Symbols hr mean hydraulic detention time - (max)A d–1 maximum specific growth rate forattached growth - (max)s d–1 maximum specific growth rate forsuspended growth - max d–1 maximum specific growth rate - d–1 specific growth rate - v mg·l–1·hr–1 maximum volumetric substrateutilization rate coefficient  相似文献   

17.
Dennis H. Greer 《Planta》1995,197(1):31-38
Bean (Phaseolus vulgaris L.) plants were grown at two light periods of 8 and 13 h with a similar photon flux density (PFD) giving a daily photon receipt (DPR) of 17.9 and 38.2 mol · m–2, respectively. Shoot growth and leaf area development were followed at regular intervals and diurnal whole-plant photosynthesis measured. Single mature trifoliate leaves were exposed to photoinhibitory treatments at PFDs of 800 and 1400 mol · m–2 · s–1 and at temperatures of 12 and 20°C. Chlorophyll fluorescence and photon yields were measured at regular intervals throughout each treatment. Plants grown in 13 h had significantly greater leaf areas than those grown in 8 h. There were no differences in maximum rates of photosynthesis, photon yields and only minor but significant differences in Fv/Fm for plants in the two treatments, showing photosynthetic characteristics were dependent on PFD but not DPR. A significant decline in photosynthesis and Fv/Fm occurred over the 13-h but little change in photosynthesis for plants in the 8 h, indicating some feedback inhibition of photosynthesis was occurring. Plants grown in 8 h were consistently more susceptible to photoinhibition of photosynthesis at all treatments than 13-h plants. Nevertheless, photoinhibition was exacerbated by increases in PFD, and by decreases in temperature for leaves from both treatments. However, for plants from the 8-h day, exposing leaves to 12°C and 1400 mol · m–2 · s–1 caused photo-oxidation and severe bleaching but no visible damage on leaves from 13-h-grown plants. Closure of the photosystem II reaction-centre pool was partially correlated with increasing extents of photoinhibition but the relationship was similar for plants from both treatments. There remains no clear explanation for their wide differences in susceptibility to photoinhibition.Abbreviations and Symbols DPR daily photon receipt - F0 and Fm initial and maximal fluorescence - Fv/Fm fluorescence ratio in dark-treated leaves - F/Fm intrinsic efficiency of PSII during illumination - PFD photon flux density - i photon yield (incident basis) - psi quantum yield of PSII electron transport - Pmax maximum rate of photosynthesis - qN non-photochemical quenching coefficient - qP photochemical quenching coefficient Many thanks to my colleague William Laing who spent a considerable effort in developing the programme to run the photosynthesis apparatus. I am also indebted to one reviewer with whom I corresponded to resolve some issues in the paper. This project was funded by the New Zealand Foundation for Research, Science and Technology.  相似文献   

18.
The problem of optimising agitation and aeration in a given fermenter is addressed. The objective function is total electric power consumed for agitation, compression and refrigeration. The major constraint considered is to ensure that the dissolved oxygen concentration is above the critical value. It is shown that it is possible to analytically calculate the optimal pair (air flowrate, stirrer speed) and that, at least for the industrial antibiotics fermentation used as case-study, the optimum lies within a window for satisfactory operation, limited by other possible constraints to the problem. Savings achievable by optimal operation as compared with current industrial procedure were found to be around 10% at pilot plant scale (0.26 m3) and 20% at full scale (85 m3).List of Symbols A fermenter cross sectional area (m2) - C dissolved oxygen concentration (mole m–3) - C * DO concentration in equilibrium with the gas (mole m–3) - C crit critical DO concentration (mole m–3) - C p specific heat of air at constant pressure (J kg–1 K–1) - C sp dissolved oxygen set point (mole m–3) - C v specific heat of air at constant volume (J kg–1 K–1) - D agitator diameter (m) - f pressure correction of air flow-rate - (Fl g)F aeration number at flooding - (Fr g)F froude number at flooding - k coefficient in expression for mass transfer coefficient - K La volumetric oxygen transfer coefficient (s–1) - m power exponent in expression for mass transfer coefficient - n gas flow rate exponent in expression for mass transfer coefficient - n * number of impellers - N rotation speed (s–1) - N F rotation speed at flooding (s–1) - N p unaerated power number - N pg aerated power number - OUR Oxygen Uptake Rate (mole m–3 s–1) - p 0 atmospheric pressure (N m–2) - p 1 compressor exit pressure (N m–2) - p 2 pressure at the bottom of the fermenter (N m–2) - p 3 pressure at the top of the fermenter (N m–2) - P c compression power (W) - P d power added by expansion (W) - P ev power removed by evaporation (W) - P g agitation power (W) - P m power added by metabolism (W) - P r power removed by refrigeration (W) - P t total power (W) - Q air flow-rate at atmospheric conditions (m3 s–1) - Q f air flow-rate at average fermenter conditions (m3 s–1) - s 0 absolute humidity at atmospheric conditions - s 3 absolute humidity at fermenter exit - T tank diameter (m) - V liquid volume (m3) - v s gas superficial velocity (m s–1) - i parameter defined in the text - safety margin for dissolved oxygen (mole m–3) - ratio of specific heats of air - g agitation efficiency - c compression efficiency - r refrigeration efficiency - liquid density (kg m–3) - g air density (kg m–3) - latent heat of vaporisation of water (J kg–1) The authors are grateful to Elsa Silva, Carlos Lopes, Carlos Aguiar, Fernando Mendes, and Alexandre Cardoso, who helped with parts of this work, and to CIPAN for permission to publish these data.  相似文献   

19.
The stability and, consequently, the lifetime of immobilized enzymes (IME) are important factors in practical applications of IME, especially so far as design and operation of the enzyme reactors are concerned. In this paper a model is presented which describes the effect of intraparticle diffusion on time stability behaviour of IME, and which has been verified experimentally by the two-substrate enzymic reaction. As a model reaction the ethanol oxidation catalysed by immobilized yeast alcohol dehydrogenase was chosen. The reaction was performed in the batch-recycle reactor at 303 K and pH-value 8.9, under the conditions of high ethanol concentration and low coenzyme (NAD+) concentration, so that NAD+ was the limiting substrate. The values of the apparent and intrinsic deactivation constant as well as the apparent relative lifetime of the enzyme were calculated.The results show that the diffusional resistance influences the time stability of the IME catalyst and that IME appears to be more stabilized under the larger diffusion resistance.List of Symbols C A, CB, CE mol · m–3 concentration of coenzyme NAD+, ethanol and enzyme, respectively - C p mol · m3 concentration of reaction product NADH - d p mm particle diameter - D eff m2 · s–1 effective volume diffusivity of NAD+ within porous matrix - k d s–1 intrinsic deactivation constant - K A, KA, KB mol · m–3 kinetic constant defined by Eq. (1) - K A x mol · m–3 kinetic constant defined by Eq. (5) - r A mol · m–3 · s–1 intrinsic reaction rate - R m particle radius - R v mol · m–3 · s–1 observed reaction rate per unit volume of immobilized enzyme - t E s enzyme deactivation time - t r s reaction time - V mol · m–3 · s–1 maximum reaction rate in Eq. (1) - V x mol · m–3 · s–1 parameter defined by Eq. (4) - V f m3 total volume of fluid in reactor - w s kg mass of immobilized enzyme bed - factor defined by Eqs. (19) and (20) - kg · m–3 density of immobilized enzyme bed - unstableness factor - effectiveness factor - Thiele modulus - relative half-lifetime of immobilized enzyme Index o values obtained with fresh immobilized enzyme  相似文献   

20.
Biochemical and biophysical parameters, including D1-protein turnover, chlorophyll fluorescence, oxygen evolution activity and zeaxanthin formation were measured in the marine seagrassZostera capricorni (Aschers) in response to limiting (100 mol·m–2·–1), saturating (350 mol·m–2·s–1) or photoinhibitory (1100 mol·m–2·s–1) irradiances. Synthesis of D1 was maximal at 350 mol·m–2·s–1 which was also the irradiance at which the rate of photosynthetic O2 evolution was maximal. Degradation of D1 was saturated at 350 mol·m–2·s–1. The rate of D1 synthesis at 1100 mol·m–2·s–1 was very similar to that at 350 mol·m–2·s–1 for the first 90 min but then declined. At limiting or saturating irradiance little change was observed in the ratio of variable to maximal fluorescence (Fv/Fm) measured after dark adaptation of the leaves, while significant photoinhibition occurred at 1100 mol·m–2·s–1. The proportion of zeaxanthin in the total xanthophyll pool increased with increasing irradiance, indicative of the presence of a photoprotective xanthophyll cycle in this seagrass. These results are consistent with a high level of regulatory D1 turnover inZostera under non-photoinhibitory irradiance conditions, as has been found previously for terrestrial plants.We would like to thank Professor Peter Böger (Department of Plant Biochemistry, University of Konstanz, Germany) for the kind gift of D1 antibodies. This work was partly supported by a University of Queensland Enabling Grant to CC.  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