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 62 毫秒
1.
Butterbach-Bahl  K.  Rothe  A.  Papen  H. 《Plant and Soil》2002,240(1):91-103
Complete annual cycles of N2O and CH4 flux in forest soils at a beech and at a spruce site at the Höglwald Forest were followed in 1997 by use of fully automatic measuring systems. In order to test if on a microsite scale differences in the magnitude of trace gas exchange between e.g. areas in direct vicinity of stems and areas in the interstem region at both sites exist, tree chambers and gradient chambers were installed in addition to the already existing interstem chambers at our sites. N2O fluxes were in a range of –4.6–473.3 g N2O-N m–2 h–1 at the beech site and in a range of –3.7–167.2 g N2O-N m–2 h–1 at the spruce site, respectively. Highest N2O emissions were observed during and at the end of a prolonged frost period, thereby further supporting previous findings that frost periods are of crucial importance for controlling annual N2O losses from temperate forests. Fluxes of CH4 were in a range of +10.4––194.0 g CH4 m–2 h–1 at the beech site and in a range of –4.4––83.5 g CH4 m–2 h–1 at the spruce site. In general, both N2O-fluxes as well as CH4-fluxes were higher at the beech site. On a microsite scale, N2O and CH4 fluxes at the beech site were highest within the stem area (annual mean: 49.6±3.3 g N2O-N m–2 h–1; –77.2±3.1 g CH4 m–2 h–1), and significantly lower within interstem areas (18.5±1.4 g N2O-N m–2 h–1; –60.2±1.8 g CH4 m–2 h–1). Significantly higher values of total N, C and pH in the organic layer, as well as increased soil moisture, especially in spring, in the stem areas, are likely to contribute to the higher N2O fluxes within the stem area of the beech. Also for the spruce site, such differences in trace gas fluxes could be demonstrated to exist (mean annual N2O emission within (a) stem areas: 9.7±0.9 g N2O-N m–2 h–1 and (b) interstem areas: 6.2±0.6 g N2O-N m–2 h–1; mean annual CH4 uptake within (a) stem areas: –26.1±0.6 g CH4 m–2 h–1 and (b) interstem areas: –38.4±0.8 g CH4 m–2 h–1), though they were not as pronounced as at the beech site.  相似文献   

2.
Glatzel  Stephan  Stahr  Karl 《Plant and Soil》2001,231(1):21-35
We examined the effect of fertilisation (200 kg cattle slurry-N ha–1 year–1) on the exchange of N2O and CH4 in the soil–plant system of meadow agroecosystems in southern Germany. From 1996 to 1998, we regularly determined the gas fluxes (closed chamber method) and associated environmental parameters. N2O and CH4 fluxes were not significantly affected by fertilisation. N2O fluxes at the unfertilised and fertilised plots were small, generally between 50 and –20 g N m–2 h–1. We identified some incidents of N2O uptake. CH4-C fluxes ranged from 1.3 to –0.2 mg m–2 h–1 and were not significantly different from 0 at both plots. We budgeted an annual net emission of 15.5 and 29.6 mg m–2 N2O-N and an annual CH4-C net emission of 184.2 and 122.7 mg m–2 at the unfertilised and fertilised plots, respectively. Apparently, rapid N mineralization and uptake in the densely rooted topsoil prevents N losses and the inhibition of CH4 oxidation.  相似文献   

3.
Mosier  A.R.  Morgan  J.A.  King  J.Y.  LeCain  D.  Milchunas  D.G. 《Plant and Soil》2002,240(2):201-211
In late March 1997, an open-top-chamber (OTC) CO2 enrichment study was begun in the Colorado shortgrass steppe. The main objectives of the study were to determine the effect of elevated CO2 (720 mol mol–1) on plant production, photosynthesis, and water use of this mixed C3/C4 plant community, soil nitrogen (N) and carbon (C) cycling and the impact of changes induced by CO2 on trace gas exchange. From this study, we report here our weekly measurements of CO2, CH4, NOx and N2O fluxes within control (unchambered), ambient CO2 and elevated CO2 OTCs. Soil water and temperature were measured at each flux measurement time from early April 1997, year round, through October 2000. Even though both C3 and C4 plant biomass increased under elevated CO2 and soil moisture content was typically higher than under ambient CO2 conditions, none of the trace gas fluxes were significantly altered by CO2 enrichment. Over the 43 month period of observation NOx and N2O flux averaged 4.3 and 1.7 in ambient and 4.1 and 1.7 g N m–2 hr –1 in elevated CO2 OTCs, respectively. NOx flux was negatively correlated to plant biomass production. Methane oxidation rates averaged –31 and –34 g C m–2 hr–1 and ecosystem respiration averaged 43 and 44 mg C m–2 hr–1 under ambient and elevated CO2, respectively, over the same time period.  相似文献   

4.
Summary Subterranean clover plants were grown as swards (about 2000 plants/m2) under controlled conditions with N provided either by N2-fixation (NO 3 withheld) or by assimilation of NO 3 (NO 3 supplied). Crop growth rates were measured by dry matter sampling over periods of up to 70 days at PPFD values of 400–1000 mole quanta/m2/s. When NO 3 was supplied from sowing the swards grew more rapidly than when the swards were not supplied with NO 3 and plants had to establish an N2-fixing apparatus. When inter-plant competition was reduced within the sward, a difference in growth rate in favour of NO 3 -fed plants continued for at least 50 days. When however, a closed canopy was allowed to form, the NO 3 -fed swards had more dry weight than the N2-fed swards at the time of canopy closure but thereafter the two swards grew at similar rates at light flux densities of above about 800 mole quanta/m2/s. At light flux densities of about 400 mole quanta/m2/s N2-fed swards had a growth rate 70–80% of that of NO 3 -fed plants. NO 3 -fed plants had a higher organic N content than did N2-fed plants under all conditions.  相似文献   

5.
Depth distributions of O2 respiration and denitrification activity were studied in 1- to 2-mm thick biofilms from nutrient-rich Danish streams. Acetylene was added to block the reduction of N2O, and micro-profiles of O2 and N2O in the biofilm were measured simultaneously with a polarographic microsensor. The specific activities of the two respiratory processes were calculated from the microprofiles using a one-dimensional diffusion-reaction model. Denitrification only occurred in layers where O2 was absent or present at low concentrations (of a fewM). Introduction of O2 into deeper layers inhibited denitrification, but the process started immediately after anoxic conditions were reestablished. Denitrification activity was present at greater depth in the biofilm when the NO3 concentration in the overlying water was elevated, and the deepest occurrence of denitrification was apparently determined by the depth penetration of NO3 . The denitrification rate within each specific layer was not affected by an increase in NO3 concentration, and the half-saturation concentration (Km) for NO3 therefore considered to be low (<25M). Addition of 0.2% yeast extract stimulated denitrification only in the uppermost 0.2 mm of the denitrification zone indicating a very efficient utilization of the dissolved organic matter within the upper layers of the biofilm.  相似文献   

6.
Summary Nitrogen mineralization, nitrification, denitrification, and microbial biomass were evaluated in four representative ecosystems in east-central Minnesota. The study ecosystems included: old field, swamp forest, savanna, and upland pin oak forest. Due to a high regional water table and permeable soils, the upland and wetland ecosystems were separated by relatively short distances (2 to 5 m). Two randomly selected sites within each ecosystem were sampled for an entire growing season. Soil samples were collected at 5-week intervals to determine rates of N cycling processes and changes in microbial biomass. Mean daily N mineralization rates during five-week in situ soil incubations were significantly different among sampling dates and ecosystems. The highest annual rates were measured in the upland pin oak ecosystem (8.6 g N m–2 yr–1), and the lowest rates in the swamp forest (1.5 g N m–2 yr–1); nitrification followed an identical pattern. Denitrification was relatively high in the swamp forest during early spring (8040 g N2O–N m–2 d–1) and late autumn (2525 g N2O–N m–2 d–1); nitrification occurred at rates sufficient to sustain these losses. In the well-drained uplands, rates of denitrification were generally lower and equivalent to rates of atmospheric N inputs. Microbial C and N were consistently higher in the swamp forest than in the other ecosystems; both were positively correlated with average daily rates of N mineralization. In the subtle landscape of east-central Minnesota, rates of N cycling can differ by an order of magnitude across relatively short distances.  相似文献   

7.
Direct measurements of net production rates and pore water profiles of solutes in the fine-grained sediments of Saginaw Bay, imply corresponding steady-state fluxes to the overlying water of 1.1–1.3 (I), 450–1010 (NH4 +), 1250–2650 (Si(OH)4), 3000–3400 (Ca2+), 440–1330 (Mg2+), 1.5–728 (Fe2+), and 179–281 (Mn2+) moles/m2/day and 11.0–11.8 (alkalinity) meq/m2/day at 17.5 °C. Silica production rates in sediments apparently follow first order kinetics with a rate coefficient of 0.09/day and a steady-state silica concentration of 1.2 mM at 23.5°C. The remaining solutes follow kinetics approximately independent of solute concentration over the range of concentrations observed. Measured solute production rates are consistent with observed solute profiles only if lateral diffusion gradients are maintained in the sediments by the burrowing and irrigation activity of benthic organisms such asChironomous, the dominant burrower in Saginaw Bay. Assuming that solute fluxes from Saginaw Bay are representative of all of the post-glacial sediments of Lake Huron, the iodine flux from sediments is comparable to the total fluvial input of iodine. The extrapolated silica fluxes from Lake Huron sediments balance the estimated biogenic silica flux to the sediments.  相似文献   

8.
Although denitrification has the potential to reduce nitrate (NO 3 ) pollution of surface waters, the quantification of denitrification rates is complex because it requires differentiation from other mechanisms and is highly variable in both space and time. This study first measured potential denitrification rates at a wetland forest site in south Louisiana before receipt of secondary wastewater effluent, and then, following 30 months of effluent application, landscape gradients of dissolved nitrate (NO 3 ) and nitrous oxide (N2O) were measured. A computer model was developed to quantify N transformations. Floodwater NO 3 and N2O concentrations were higher in the forest receiving effluent than in the adjacent control forest. Denitrification rates of NO 3 -amended soil cores ranged from 0.03–0.45 g N m–2 d–1 with an overall mean of 0.10 g N m–2 d–1. Effluent N is being applied at a rate of approximately 0.034 g N m–2 d–1, with approximately 95% disappearing along a 1 km transect. In the treatment forest, floodwater NO 3 concentrations decreased from 1000 M at the inflow point to 50 M along the 1 km transect. Nitrous oxide concentrations increased from 0.25 M to 1.2 M within the first 100 m, but decreased to 0.1 M over the next 900 m. The initial increase in N2O was presumably a result ofin situ denitrification. Model analyses indicated that denitrification was directly associated with nitrification and was limited by the availability of NO 3 produced by nitrification. Due to different redox potential optima, coupling of nitrification and denitrification was a function of a balance of environmental conditions that was moderately favorable to both processes. N removal efficiency was largely dependent on the proportion of effluent NH 4 + to NO 3 . When NH 4 + /NO 3 was 1, average N removal efficiency ranged from 95–100%, but ratios that were >1 reduced average efficiencies to as low as 57%. Actual effluent NH 4 + /NO 3 loading ratios at this site are approximately 0.2 and are consistently <1.  相似文献   

9.
Denitrification potentials of epilithic microbial populations were assessed using the acetylene inhibition method, in which acetylene is used to block the reduction of nitrous oxide (N2O) to nitrogen (N2). Samples of the epilithic community were incubated in filtered river water containing modified Bushnell-Haas salts, glycerol, and yeast extract—under aerobic (0.2 atm O2) and anaerobic (0.2 atm He) acetylene atmospheres. N2O was produced under both atmospheres only if exogenous nitrate of nitrite was added. Denitrification potentials were typically higher when nitrite was the added electron acceptor. The rates of denitrification were temperature-and carbon-dependent and the maximum rate, 8.53 g N2O–N per cm2 per day occurred at 23°C when nitrite was the electron acceptor.  相似文献   

10.
Methane distribution in European tidal estuaries   总被引:16,自引:3,他引:13  
Methane concentrations have been measured along salinity profilesin nine tidal estuaries in Europe (Elbe, Ems, Thames, Rhine,Scheldt, Loire, Gironde, Douro and Sado). The Rhine, Scheldt andGironde estuaries have been studied seasonally. A number ofdifferent methodologies have been used and they yieldedconsistent results. Surface water concentrations ranged from0.002 to 3.6 M, corresponding to saturation ratios of 0.7 to1580 with a median of 25. Methane concentrations in thefresh-water end-members varied from 0.01 to 1.4 M. Methaneconcentrations in the marine end-members were close to saturationoffshore and on the order of 0.1 M in estuarine plumes. Methaneversus salinity profiles in river-dominated, stratified estuaries(Rhine and Douro) appeared rather erratic whereas those in thewell mixed, long-residence time estuaries (Elbe, Ems, Thames,Scheldt, Loire, Gironde and Sado) revealed consistent trends. Inthese systems dissolved methane initially decreases withincreasing salinity, then increases to a maximum at intermediateto high salinities before decreasing again going offshore. Tidalflats and creeks were identified as a methane source to estuarinewaters. The global estuarine flux of methane to the atmospherehas been calculated by combining the median water-air methanegradient (68.2 nmol dm–3) with a global area weighted transfercoefficient and the global area of estuaries. Estuaries emit 1.1to 3.0 Tg CH4 yr–1, which is less than 9% of the global marinemethane emission.  相似文献   

11.
Summary Biotransformation of benzaldehyde to benzyl alcohol bySaccharomyces cerevisiae immobilized in different support matrices was investigated. Polymers with intrinsic hydrophobic and/or hydrophilic nature as well as mixed hydrophobic and hydrophilic supports were examined both in aqueous and bisphasic aqueous-organic systems. The hydrophobic support material ENTP-2000 or mixed silicone:alginate (50-2550-75) proved to be most suitable not only for nonconventional media but also for conventional aqueous media for production of benzyl alcohol. With ENTP-2000, catalytic activity and maximum yield were 159 mol h–1 g–1 dry weight catalyst and 0.89 mM, respectively, in hexane containing 2% moisture. Corresponding values in aqueous media were 246 mol h–1 g–1 dry weight catalyst and 1.53 mM. With 5050 silicone:alginate, catalytic activity and maximum yield were 177 mol h–1 g–1 dry weight catalyst and 1.18 mM, respectively, in hexane containing 2% moisture. Corresponding values in aqueous media were 192 mol h–1 g–1 dry weight catalyst and 0.8 mM.  相似文献   

12.
Gasche  R.  Papen  H. 《Plant and Soil》2002,240(1):67-76
In order to evaluate differences in the magnitude of NO and NO2 flux rates between soil areas in direct vicinity to tree stems and areas of increasing distance to tree stems, we followed in 1997 at the Höglwald Forest site with a fully automated measuring system a complete annual cycle of NO and NO2 fluxes from soils of an untreated spruce stand, a limed spruce strand, and a beech stand using at each stand measuring chambers which were installed onto the soils in such a way that they formed a stem to stem gradient. Flux data obtained since the end of 1993 from measuring chambers placed at the interstem areas of the stands, which had been used for the calculation of the long year annual mean of NO and NO2 flux rates from soils of the stands, are compared to both (a) those obtained from the interstem chambers in 1997 and (b) those from the stem to stem gradient chambers. Daily mean NO fluxes obtained in 1997 were in a range of 0.3 – 280.1 g NO-N m–2 h–1 at the untreated spruce stand, 0.5 – 273.2 g NO-N m–2 h–1 at the limed spruce stand and 0.5 - 368.8 g NO-N m–2 h–1 at the beech stand, respectively. Highest NO emission rates were observed during summer, lowest during winter. Daily mean NO2 fluxes were in a range of –83.1 – 7.6 g NO2-N m–2 h–1 at the untreated spruce stand, -85.1 – 2.1 g NO2-N m–2 h–1 at the limed spruce stand and –77.9 to –2.0 g NO2-N m–2 h–1 at the beech site, respectively. As had been observed for the years 1994–1996, also in 1997 NO emission rates were highest at the untreated spruce stand and lowest at the beech stand and liming of a spruce stand resulted in a significant decrease in NO emission rates. For NO2 no marked differences in the magnitude of flux rates were found between the three different stands. Results obtained from the stem to stem gradient experiments revealed that at all stands studied NO emission rates were significantly higher (between 1.6- and 2.6-fold) from soil areas close to the tree stems and decreased – except at the beech stand - with increasing distance from the stems, while for NO2 deposition no marked differences were found. Including the contribution of soil areas in direct vicinity to the beech stems in the estimation of the annual mean NO source strength revealed that the source strength has been underestimated by 40% in the past.  相似文献   

13.
Egorova  E.A.  Bukhov  N.G. 《Photosynthetica》2002,40(3):343-347
Photosystem 2 (PS2)-driven electron transfer was studied in primary leaves of barley (Hordeum vulgare L.) seedlings grown under various photon fluxes (0.3–170.0 mol m–2 s–1) of blue (BR) or red (RR) radiation using modulated chlorophyll fluorescence. The Fv/Fm ratio was 0.78–0.79 in leaves of all radiation variants, except in seedlings grown under BR or RR of 0.3 mol m–2 s–1. The extent of the photochemical phase of the polyphasic Fv rise induced by very strong white light was similar in leaves of all radiation treatments. Neither radiation quality nor photon flux under plant cultivation influenced the amount of non QB-transferring centres of PS2 except in leaves of seedlings grown under BR of 0.3 mol m–2 s–1, in which the amount of such centres increased threefold. Both BR and RR stimulated the development of photochemically competent PS2 at photon fluxes as low as 3 mol m–2 s–1. Three exponential components with highly different half times were distinguished in the kinetics of Fv dark decay. This indicates different pathways of electron transfer from QA , the reduced primary acceptor of PS2, to other acceptors. Relative magnitudes of the individual decay components did not depend on the radiation quality or the photon flux during plant cultivation. Significant differences were found, however, between plants grown under BR or RR in the rate of the middle and fast components of Fv dark decay, which showed 1.5-times faster intersystem linear electron transport in BR-grown leaves.  相似文献   

14.
Clough  T.J.  Ledgard  S.F.  Sprosen  M.S.  Kear  M.J. 《Plant and Soil》1998,199(2):195-203
A field lysimeter experiment was conducted over a 406 day period to determine the effect of different soil types on the fate of synthetic urinary nitrogen (N). Soil types included a sandy loam, silty loam, clay and peat. Synthetic urine was applied at 1000 kg N ha-1, during a winter season, to intact soil cores in lysimeters. Leaching losses, nitrous oxide (N2O) emissions, and plant uptake of N were monitored, with soil 15N content determined upon destructive sampling of the lysimeters. Plant uptake of urine-N ranged from 21.6 to 31.4%. Soil type influenced timing and form of inorganic-N leaching. Macropore flow occurred in the structured silt and clay soils resulting in the leaching of urea. Ammonium (NH 4 + –N), nitrite (NO 2 - –N) and nitrate (NO3 -–N) all occurred in the leachates with maximum concentrations, varying with soil type and ranging from 2.3–31.4 g NH 4 + –N mL-1, 2.4–35.6 g NO 2 - –N mL-1, and 62–102 g NO 3 - –N mL-1, respectively. Leachates from the peat and clay soils contained high concentrations of NO 2 - –N. Gaseous losses of N2O were low (<2% of N applied) over a 112 day measurement period. An associated experiment showed the ratio of N2–N:N2O–N ranged from 6.2 to 33.2. Unrecovered 15N was presumed to have been lost predominantly as gaseous N2. It is postulated that the high levels of NO 2 - –N could have contributed to chemodenitrification mechanisms in the peat soil.  相似文献   

15.
Denitrification processes were measured by the acetylene-blockage technique under changing flood conditions along the aquatic/terrestrial transition zone on the Amazon floodplain at Lago Camaleão, near Manaus, Brazil. In flooded sediments, denitrification was recorded after the amendment with NO 3 (100 mol liter–1) throughout the whole study period from August 1992 to February 1993. It ranged from 192.3 to 640.7 mol N m–2 h–1 in the 0- to 5-cm sediment layer. Without substrate amendment, denitrification was detected only during low water in November and December 1992, when it occurred at a rate of up to 12.2 mol N m–2 h–1 Higher rates of denitrification at an average rate of 73.3 mol N m–2 h–1 were measured in sediments from the shallow lake basin that were exposed to air at low water. N2O evolution was never detected in flooded sediments, but in exposed sediments, it was detected at an average rate of 28.3 mol N m–2 h–1 during the low-water period. The results indicate that under natural conditions there is denitrification and hence a loss in nitrogen from the Amazon floodplain to the atmosphere. Rates of denitrification in flooded sediments were one to two orders of magnitude smaller than in temperate regions. However, the nitrogen removal of exposed sediments exceeded that of undisturbed wetland soils of temperate regions, indicating a considerable impact of the flood pulse on the gaseous turnover of nitrogen in the Amazon floodplain.  相似文献   

16.
Nitrification and denitrification rates were estimated simultaneously in soil-floodwater columns of a Crowley silt loam (Typic Albaqualfs) rice soil by an15N isotopic dilution technique. Labeled NO 3 was added to the floodwater of soil-water columns, half were treated with urea fertilizer. The (NO 3 +NO 2 )–N and (NO 3 +NO 2 )–N concentrations in the floodwater were measured over time and production and reduction rates for NO 3 calculated. Nitrate reduction in the urea amended columns averaged 515 mol N m–2h–1 and nitrification averaged 395 mol N m–2h–1 over the 35–153 d incubation. The nitrification rate for 4–19 d sampling period (1,560 mol N m–2h–1) in the urea amended columns was almost 9 times greater than the reduction rate (175 mol N m–2h–1) over the same period. Without the addition of urea the NO 3 production rate averaged 32 mol N m–2h–1 and reduction 101 mol N m–2h–1.  相似文献   

17.
Summary The function of the caecal bulb, and its adaptation to chronic high- or low-Na+ intake, was investigated by in vivo perfusion of anaesthetised birds. Effects of acute aldosterone injection (125 g·kg–1 body mass) were also measured.Evidence was found for primary active net absorption of Na+, inducing parallel Na-linked absorption of water and Cl and secretion of K+. Around 20–35% of total Cl absorption and K+ secretion were independent of Na+ fluxes, and these components appear to be driven by passive processes with apparent conductances of 6.3×10–3 (G Cl) and 1.1×10–3 (G K) S·cm–2.Acetate (40mM) stimulated Na+ fluxes (8.5–9.9 Eq·cm–2·h–1) and Na-linked water fluxes (27–44 l·cm–2·h–1). Increased coupling ratios (2.9–4.6 l·Eq–1) and other data indicate that these effects may be due to increased osmotic permeabilities of barriers involved in the Na-linked water transfer pathway.Low-Na+ maintenance enhanced EPD (49–69 mV, serosa positive) and all net fluxes:J Na (6.8–11.6);J K (–3.2––4.3);J Cl (4.3–5.6 Eq·cm serosal area–2·h–1);J v (28–43 l·cm–2·h–1) (mucosal-serosal fluxes positive).Acute aldosterone enhancedJ Na (10.8–14.0 Eq·cm–2·h–1) and EPD (54–66 mV) by 3 h after injection, but had no effect on the Na-linked components ofJ K orJ Cl.Abbreviations ECPD, EPD Electrochemical or electrical potential difference - G Cl ,G K ionic conductances (Cl, K+) - J v ,J ion net volume or ion flux rate, mucosa-serosa positive;P d (Cl) diffusive permeability coefficient (of Cl) - SEDM standard error of difference between means  相似文献   

18.
Jia  Yinsuo  Gray  V.M. 《Photosynthetica》2003,41(4):605-610
We determined for Vicia faba L the influence of nitrogen uptake and accumulation on the values of photon saturated net photosynthetic rate (P Nmax), quantum yield efficiency (), intercellular CO2 concentration (C i), and carboxylation efficiency (C e). As leaf nitrogen content (NL) increased, the converged onto a maximum asymptotic value of 0.0664±0.0049 mol(CO2) mol(quantum)–1. Also, as NL increased the C i value fell to an asymptotic minimum of 115.80±1.59 mol mol–1, and C e converged onto a maximum asymptotic value of 1.645±0.054 mol(CO2) m–2 s–1 Pa–1 and declined to zero at a NL-intercept equal to 0.596±0.096 g(N) m–2. fell to zero for an NL-intercept of 0.660±0.052 g(N) m–2. As NL increased, the value of P Nmax converged onto a maximum asymptotic value of 33.400±2.563 mol(CO2) m–2 s–1. P N fell to zero for an NL-intercept of 0.710±0.035 g(N) m–2. Under variable daily meteorological conditions the values for NL, specific leaf area (L), root mass fraction (Rf), P Nmax, and remained constant for a given N supply. A monotonic decline in the steady-state value of Rf occurred with increasing N supply. L increased with increasing N supply or with increasing NL.  相似文献   

19.
García-Núñez  C.  Rada  F.  Boero  C.  González  J.  Gallardo  M.  Azócar  A.  Liberman-Cruz  M.  Hilal  M.  Prado  F. 《Photosynthetica》2004,42(1):133-138
Stress-induced restrictions to carbon balance, growth, and reproduction are the causes of tree-line formation at a global scale. We studied gas exchange and water relations of Polylepis tarapacana in the field, considering the possible effects of water stress limitations imposed on net photosynthetic rate (P N). Daily courses of microclimatic variables, gas exchange, and leaf water potential were measured in both dry-cold and wet-warm seasons at an altitude of 4 300 m. Marked differences in environmental conditions between seasons resulted in differences for the dry-cold and wet-warm seasons in mean leaf water potentials (–1.67 and –1.02 MPa, respectively) and mean leaf conductances (33.5 and 58.9 mmol m–2 s–1, respectively), while differences in mean P N (2.5 and 2.8 mol m–2 s–1, respectively) were not as evident. This may be related to limitations imposed by water deficit and lower photon flux densities during dry and wet seasons, respectively. Hence P. tarapacana has coupled its gas exchange characteristics to the extreme daily and seasonal variations in temperature and water availability of high elevations.  相似文献   

20.
We analysed the stable isotope composition of emitted N2O in a one-year field experiment (June 1998 to April 1999) in unfertilized controls, and after adding nitrogen by applying slurry or mineral N (calcium ammonium nitrate). Emitted N2O was analysed every 2–4 weeks, with additional daily sampling for 10 days after each fertilizer application. In supplementary soil incubations, the isotopic composition of N2O was measured under defined conditions, favouring either denitrification or nitrification. Soil incubated for 48 h under conditions favouring nitrification emitted very little N2O (0.024 mol gdw –1) and still produced N2O from denitrification. Under denitrifying incubation conditions, much more N2O was formed (0.91 mol gdw –1 after 48 h). The isotope ratios of N2O emitted from denitrification stabilized at 15N = –40.8 ± 5.7 and 18O = 2.7 ± 6.3. In the field experiment, the N2O isotope data showed no clear seasonal trends or treatment effects. Annual means weighted by time and emission rate were 15N = –8.6 and 18O = 34.7 after slurry application, 15N = –4.6 and 18O = 24.0 after mineral fertilizer application and 15N = –6.4 and 18O = 35.6 in the control plots, respectively. So, in all treatments the emitted N2O was 15N-depleted compared to ambient air N2O (15N = 11.4 ± 11.6, 18O = 36.9 ± 10.7). Isotope analyses of the emitted N2O under field conditions per se allowed no unequivocal identification of the main N2O producing process. However, additional data on soil conditions and from laboratory experiments point to denitrification as the predominant N2O source. We concluded (1) that the isotope ratios of N2O emitted from the field soil were not only influenced by the source processes, but also by microbial reduction of N2O to N2 and (2) that N2O emission rates had to exceed 3.4 mol N2O m–2 h–1 to obtain reliable N2O isotope data.  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