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.
Measurement of net ecosystem exchange was made using the eddy covariance method above three forests along a north-south climatic gradient in Sweden: Flakaliden in the north, Knottåsen in central and Asa in south Sweden. Data were obtained for 2 years at Flakaliden and Knottåsen and for one year at Asa. The net fluxes (Nep) were separated into their main components, total ecosystem respiration (Rt) and gross primary productivity (Pg). The maximum half-hourly net uptake during the heart of the growing season was highest in the southernmost site with ?0.787 mg COm?2 s?1 followed by Knottåsen with ?0.631 mg COm?2 s?1 and Flakaliden with ?0.429 mg COm?2 s?1. The maximum respiration rates during the summer were highest in Knottåsen with 0.245 mg COm?2 s?1 while it was similar at the two other sites with 0.183 mg COm?2 s?1. The annual Nep ranged between uptake of ?304 g C m?2 year?1 (Asa) and emission of 84 g C m?2 year?1 (Knottåsen). The annual Rt and Pg ranged between 793 to 1253 g C m?2 year?1 and ?875 to ?1317 g C m?2 year?1, respectively. Biomass increment measurements in the footprint area of the towers in combination with the measured net ecosystem productivity were used to estimate the changes in soil carbon and it was found that the soils were losing on average 96–125 g C m?2 year?1. The most plausible explanation for these losses was that the studied years were much warmer than normal causing larger respiratory losses. The comparison of net primary productivity and Pg showed that ca 60% of Pg was utilized for autotrophic respiration.  相似文献   

2.
Withania somnifera is an important medicinal plant that contains withanolides as bioactive compounds. We have investigated the effects of macroelements and nitrogen source in hairy roots of W. somnifera with the aim of optimizing the production of biomass and withanolide A content. The effects of the macroelements NH4NO3, KNO3, CaCl2, MgSO4 and KH2PO4 at concentrations of 0, 0.5, 1.0, 1.5 and 2.0× strengths and of nitrogen source [NH4 +/NO3 ? (0.00/18.80, 7.19/18.80, 14.38/18.80, 21.57/18.80, 28.75/18.80, 14.38/0.00, 14.38/9.40, 14.38/18.80, 14.38/28.20 and 14.38/37.60 mM)] in Murashige and Skoog medium were evaluated for biomass and withanolide A production. The highest accumulation of biomass (139.42 g l?1 FW and 13.11 g l?1 DW) was recorded in the medium with 2.0× concentration of KH2PO4, and the highest production of withanolide A was recorded with 2.0× KNO3 (15.27 mg g?1 DW). The NH4 +/NO3 ? ratio also influenced root growth and withanolide A production, with both parameters being larger when the NO3 ? concentration was higher than that of NH4 +. Maximum biomass growth (148.17 g l?1 FW and 14.79 g l?1 DW) was achieved at NH4 +/NO3 ? ratio of 14.38/37.60 mM, while withanolide A production was greatest (14.68 mg g?1 DW) when the NH4 +/NO3 ? ratio was 0.00/18.80 mM. The results are useful for the large scale cultivation of Withania hairy root culture for the production of withanolide A.  相似文献   

3.
An in vitro system was established for the characterisation of inorganic nitrogen uptake by sugarcane plantlets of variety NCo376. After multiplication and rooting, plantlets (0.27–0.3 g fresh mass) were placed on N-free medium for 4 days, and then supplied with 2–20 mM N as NO3 ?-N only, NH4 +-N only or NO3 ?-N + NH4 +-N (as 1:1). With few exceptions, on all the tested N media, the in vitro plants always had a higher Vmax for NH4 +-N (28.69–66.51 μmol g?1 h?1) than for NO3 ?-N uptake (10.24–30.19 μmol g?1 h?1) and the Km indicated a higher affinity for NO3 ?-N (0.02–7.38 mM) than for NH4 +-N (0.06–9.15 mM). When N was applied as 4 and 20 mM to varieties N12, N19 and N36, the interaction between variety, N form and concentration resulted in differences in the Vmax and Km. The high N-use efficient varieties (N12 and N19), as determined in previous pot and field trials, behaved similarly under all tested conditions and displayed a lower Vmax and Km than the low N-use efficient ones (NCo376 and N36). Based on this finding, it was suggested that the N-use efficient designation (from pot and field trials) may not be ascribed solely to N uptake. Assessment of the relative preference index (RPI) for NO3 ?-N and NH4 +-N uptake revealed that, at present, the RPI has no application in sugarcane due to its preferential uptake of NH4 +-N.  相似文献   

4.
Effects of three levels of photosynthetic photon flux (PPF: 60, 160 and 300 μmol m−2s−1) were investigated in one-month-old Phalaenopsis plantlets acclimatised ex vitro. Optimal growth, chlorophyll and carotenoid concentations, and a high carotenoid:chlorophyll a ratio were obtained at 160 μmol m−2s−1, while net CO2 assimilation (A), stomatal conductance (g), transpiration rate (E) and leaf temperature peaked at 300 μmol m−2s−1, indicating the ability of the plants to grow ex vitro. Adverse effects of the highest PPF were reflected in loss of chlorophyll, biomass, non-protein thiol and cysteine, but increased proline. After acclimatisation, glucose-6-phosphate dehydrogenase, shikimate dehydrogenase, phenylalanine ammonia-lyase (PAL) and cinnamyl alcohol dehydrogenase (CAD) increased, as did lignin. Peroxidases (POD), which play an important role in lignin synthesis, were induced in acclimatised plants. Polyphenol oxidase (PPO) and β-glucosidase (β-GS) activities increased to a maximum in acclimatised plants at 300 μmol m−2s−1. A positive correlation between PAL, CAD activity and lignin concentration was observed, especially at 160 and 300 μmol m−2s−1. The study concludes that enhancement of lignin biosynthesis probably not only adds rigidity to plant cell walls but also induces defence against radiation stress. A PPF of 160 μmol m−2s−1was suitable for acclimatisation when plants were transferred from in vitro conditions.  相似文献   

5.
Quasi-continuous DO and pH measurements (total 47 days) were conducted during enclosure experiments (6 enclosures; 5 × 5 × 2.5 m), in which a biomass gradient of silver carp was created. After subtracting the air–water exchanges of O2 and CO2, the chemical and biochemical changes in DO (dissolved oxygen) and DIC (dissolved inorganic carbon) were estimated in order to evaluate MQ (metabolic quotient: DO change divided by DIC change) at intervals of 1 hour. By removing small absolute changes below the threshold value (0.01 mM h?1), the averaged values of the 24 MQ means for the respective 1-hour periods ranged from 0.96 to 1.20 in the six enclosures. Because the MQs in the daytime inversely correlated well with the ratio of NH+ 4–N to (NH+ 4–N + NO? 3–N), not the ecosystems, i.e., density of fish, community structure of zooplankton and phytoplankton, but the form of nitrogen uptaken for primary production principally determined the MQs. The higher MQs observed in the daytime compared with the nighttime (from 14% to 21% except 3% for one enclosure) could not be explained by the denitrification and/or dissolution of CaCO3 in the sediments, therefore suggesting the selectively faster decomposition of part of the organic matter provided through primary production, in other words, an accumulation of another part of the organic matter in the diurnal and/or daily time scale.  相似文献   

6.
Pigments produced by species of Monascus have been used to coloring rice, meat, sauces, wines and beers in East Asian countries. Monascus can produce orange (precursor), yellow and red pigments. Orange pigments have low solubility in culture media and when react with amino groups they become red and largely soluble. The orange pigments are an alternative to industrial pigment production because the low solubility facilitates the downstream operations. The aim of this work was to study the kinetic on the production of orange pigments by Monascus ruber CCT 3802. The shaking frequency of 300 rpm was favorable to production, whereas higher shaking frequencies showed negative effect. Pigment production was partially associated with cell growth, the critical dissolved oxygen concentration was between 0.894 and 1.388 mgO2 L?1 at 30 °C, and limiting conditions of dissolved oxygen decreased the production of orange pigments. The maintenance coefficient (mo) and the conversion factor of oxygen in biomass (Yo) were 18.603 mgO2 g x ?1  h?1 and 3.133 gx gO 2 ?1 and the consideration of these parameters in the oxygen balance to estimate the biomass concentration provided good fits to the experimental data.  相似文献   

7.
This study reports an improved method for activating asphalt to produce ultra‐high surface area porous carbons. Pretreatment of asphalt (untreated Gilsonite, uGil ) at 400 °C for 3 h removes the more volatile organic compounds to form pretreated asphalt ( uGil‐P ) material with a larger fraction of higher molecular weight π‐conjugated asphaltenes. Subsequent activation of uGil‐P at 900 °C gives an ultra‐high surface area (4200 m2 g?1) porous carbon material ( uGil‐900 ) with a mixed micro and mesoporous structure. uGil‐900 shows enhanced room temperature CO2 uptake capacity at 54 bar of 154 wt% (35 mmol g?1). The CH4 uptake capacity is 37.5 wt% (24 mmol g?1) at 300 bar. These are relevant pressures in natural gas production. The room temperature working CO2 uptake capacity for uGil‐900 is 19.1 mmol g?1 (84 wt%) at 20 bar and 32.6 mmol g?1 (143 wt%) at 50 bar. In order to further assess the reliability of uGil‐900 for CO2 capture at elevated pressures, the authors study competitive sorption of CO2 and CH4 on uGil‐900 at pressures from 1 to 20 bar at 25 °C. CO2/CH4 displacement constants are measured at 2 to 40 bar, and found to increase significantly with pressure and surface area.  相似文献   

8.
Decomposing leaf litter is a large supply of energy and nutrients for soil microorganisms. How long decaying leaves continue to fuel anaerobic microbial activity in wetland ecosystems is poorly understood. Here, we compare leaf litter from 15 tree species with different growth forms (angiosperms and gymnosperms, deciduous, and longer life span), using litterbags positioned for up to 4 years in a forested peatland in New York State. Periodically, we incubated partially decayed residue per species with fresh soil to assess its ability to fuel microbial methane (CH4) production and concomitant anaerobic carbon dioxide (CO2) production. Decay rates varied by leaf type: deciduous angiosperm > evergreen gymnosperm > deciduous gymnosperm. Decay rates were slower in leaf litter with a large concentration of lignin. Soil with residue of leaves decomposed for 338 days had greater rates of CH4 production (5.8 µmol g?1 dry mass d?1) than less decomposed (<0.42 µmol g?1 dry mass d?1) or more decomposed (2.1 µmol g?1 dry mass d?1) leaf residue. Species-driven differences in their ability to fuel CH4 production were evident throughout the study, whereas concomitant rates of CO2 production were more similar among species and declined with degree of decomposition. Methane production rates exhibited a positive correlation with pectin and the rate of pectin decomposition. This link between leaf litter decay rates, biochemical components in leaves, and microorganisms producing greenhouse gases should improve predictions of CH4 production in wetlands.  相似文献   

9.
Buoyancy of the gas-vacuolate alga Anabaena flosaquae Brébisson was measured under various levels of light, NH4+, and CO2. At high irradiance (50 μE · m?2·?1) the alga was non-buoyant regardless of the availability of CO2 and NH4+. At low irradiance (≤10 μE · m ?2· s?1) buoyancy was controlled by the availability of NH4+ and CO2. When NH4+ was abundant, algal buoyancy was high over a wide range of CO2 concentrations. In the absence of NH4+, algal buoyancy was reduced at high CO2 concentrations, however as the CO2 concentration declined below about 5 μmol · L?1, algal buoyancy increased. These results help explain why gas vacuolate, nitrogen-fixing blue-green algae often form surface blooms in eutrophic lakes.  相似文献   

10.
Sodium polytungstate (SPT) is currently the material of choice for soil density fractionation (DF). We recently detected high levels of N in several types of commercially available SPT (0.74–1.4 mg g?1), raising a concern that undesirable chemical effects on soils may occur during the DF procedure. To address this concern, we conducted two experiments to examine effects of SPT on C and N in the resulting soil fractions. First we suspended A-horizon material from three soil types of greatly differing mineralogy for 24 h in solutions containing three types of commercially obtained SPT and commercial SPT that had been passed through cation exchange resin columns. We compared %C, %N, 15N and 13C values in treated and untreated soils. We also spiked SPT with tracer-level 15NH4 + to measure potential NH4 + absorption by the soil fractions. Results suggest that the N-rich commercially available SPT can have a considerable effect on δ15N values, likely due to the presence of 15N-enriched NH4 + in the SPT. In one of our soils, 15N enrichment of 3‰ was observed associated with overnight soaking in N-rich SPT (0.74 mg g?1). By contrast, when using SPT with low levels of N (0.05 mg g?1), no significant changes in 15N were observed. The remaining soil (after suspension and rinsing) was similar in %C, %N, 15N and 13C to the untreated bulk soil, suggesting that suspension of soil in SPT with low N levels purchased from the manufacturer or else through treatment with cation exchange resins does not greatly alter these variables. Low-N SPT is available commercially although it must be specifically requested from the manufacturer and is currently more expensive to purchase. Our results confirm that SPT tested and known to be low in C and N (<0.06 mg g?1) does not adversely contaminate soils during the soil density fractionation procedure. If using newly purchased or recycled SPT with higher N or C levels than this, we recommend thorough testing for possible contamination effects prior to use. However we caution against using SPT that contains N or C levels >0.5 mg g?1.  相似文献   

11.
Lenticel discolouration (LD) has now emerged as a leading postharvest threat in mango, which interferes with the face value of fruits, thereby affecting the trade and causing huge monetary losses to our country. For its management, we designed an experiment using salicylic acid at 200, 400 and 600 ppm concentration along with control fruits, as a dip treatment for 5 min. Our results revealed that salicylic acid at 200 ppm was not only effective in reducing LD significantly but also reduced the activities of polyphenol oxidase (PPO) (0.397 ?A410 O.D min?1 g?1 FW), peroxidase (POD) (0.050 ? A470 O.D min?1 g?1 FW), and lipoxigenase (LOX) (3.227 µmol min?1 g?1 FW) enzymes and helped in increasing the total phenolics (15.46 mg gallic acid equivalent 100 g?1). This treatment also suppressed the rates of ethylene evolution (0.521 µL kg?1 h?1) and respiration (34.46 mL CO2 kg?1 h?1) over untreated mango fruits. With respect to quality parameters, the significant decrease in postharvest decay (23.3%) occurred without any adverse effect on soluble solids concentrates (16° B) and total carotenoids (4.1 mg 100 g?1pulp). Thus, keeping all parameters (physical, physiological, biochemical and quality) in view, salicylic acid at 200 ppm was most effective as a postharvest dip treatment for reducing LD in mango during storage or marketing without adversely affecting the fruit quality.  相似文献   

12.
A Au nanoparticle‐coated Ni nanowire substrate without binder or carbon is used as the electrode (denoted as the Au/Ni electrode) for Li‐oxygen (Li‐O2) batteries. A minimal amount of Au nanoparticles with sizes of <30 nm on a Ni nanowire substrate are coated using a simple electrodeposition method to the extent that maximum capacity can be utilized. This optimized, one body, Au/Ni electrode shows high capacities of 921 mAh g?1Au, 591 mAh g?1Au, and 359 mAh g?1Au, which are obtained at currents of 300 mAg?1Au, 500 mAg?1Au, and 1000 mAg?1Au respectively. More importantly, the Au/Ni electrode exhibits excellent cycle stability over 200 cycles.  相似文献   

13.
The net photosynthetic rate (P N), the sample room CO2 concentration (CO2S) and the intercellular CO2 concentration (C i) in response to PAR, of C3 (wheat and bean) and C4 (maize and three-colored amaranth) plants were measured. Results showed that photorespiration (R p) of wheat and bean could not occur at 2 % O2. At 2 % O2 and 0 μmol mol?1 CO2, P N can be used to estimate the rate of mitochondrial respiration in the light (R d). The R d decreased with increasing PAR, and ranged between 3.20 and 2.09 μmol CO2 m?2 s?1 in wheat. The trend was similar for bean (between 2.95 and 1.70 μmol CO2 m?2 s?1), maize (between 2.27 and 0.62 μmol CO2 m?2 s?1) and three-colored amaranth (between 1.37 and 0.49 μmol CO2 m?2 s?1). The widely observed phenomenon of R d being lower than R n can be attributed to refixation, rather than light inhibition. For all plants tested, CO2 recovery rates increased with increasing light intensity from 32 to 55 % (wheat), 29 to 59 % (bean), 54 to 87 % (maize) and 72 to 90 % (three-colored amaranth) at 50 and 2,000 μmol m?2 s?1, respectively.  相似文献   

14.
Incorporating plasmonic nanoparticles (NPs) in an organic solar cell (OSC) can improve device performance. In our simulation studies, at NP resonance, absorption in poly(3-hexythiophene)/phenyl-C61-butyric acid methyl ester (P3HT/PCBM) can be increased by encapsulating 50 nm Ag NPs with Al2O3, HfO2, MoO3, and SiO2. At Ag NP resonance, when the oxide thickness is significant enough, oxides with high relative permittivity induces a higher electric field enhancement at the metal/dielectric interface. This is translated to improved absorption in the polymer layer. By integrating against AM1.5G, overall absorption in P3HT/PCBM is improved when incorporating Ag NPs encapsulated with a thin oxide shell into the polymer film. However, polymeric absorption loss is induced for oxide-encapsulated Ag NPs if MoO3 and SiO2 shells are more than 5 nm. For Al2O3 and HfO2, Ag NPs should not be encapsulated with shells thicker than 10 nm. Modeling studies are also extended to absorption in a CH3NH3PbI3 perovskite layer. It is revealed that both Al2O3 and HfO2 have an optimal shell thickness of about 20 nm to ensure maximum absorption in CH3NH3PbI3. The results can be utilized as a useful guideline when designing photovoltaics from an optical point of view.  相似文献   

15.
To reduce CO2 emissions from alcoholic fermentation, Arthrospira platensis was cultivated in tubular photobioreactor using either urea or nitrate as nitrogen sources at different light intensities (60 μmol m?2 s?1?≤?I?≤?240 μmol m?2 s?1). The type of carbon source (pure CO2 or CO2 from fermentation) did not show any appreciable influence on the main cultivation parameters, whereas substitution of nitrate for urea increased the nitrogen-to-cell conversion factor (Y X/N ), and the maximum cell concentration (X m ) and productivity (P X ) increased with I. As a result, the best performance using gaseous emissions from alcoholic fermentation (X m ?=?2,960?±?35 g m?3, P X ?=?425?±?5.9 g m?3 day?1 and Y X/N ?=?15?±?0.2 g g?1) was obtained at I?=?120 μmol m?2 s?1 using urea as nitrogen source. The results obtained in this work demonstrate that the combined use of effluents rich in urea and carbon dioxide could be exploited in large-scale cyanobacteria cultivations to reduce not only the production costs of these photosynthetic microorganisms but also the environmental impact associated to the release of greenhouse emissions.  相似文献   

16.
The present research is focused on cultivation of microalgae strain Chlorella vulgaris for bio-fixation of CO2 coupled with biomass production. In this regard, a single semi-batch vertical tubular photobioreactor and four similar photobioreactors in series have been employed. The concentration of CO2 in the feed stream was varied from 2 to 12 % (v/v) by adjusting CO2 to air ratio. The amount of CO2 capture and algae growth were monitored by measuring decrease of CO2 concentration in the gas phase, microalgal cell density, and algal biomass production rate. The results show that 4 % CO2 gives maximum amount of biomass (0.9 g L?1) and productivity (0.118 g L?1 day?1) of C. vulgaris in a single reactor. In series reactors, average productivity per reactor found to be 0.078 g L?1 day?1. The maximum CO2 uptake for single reactor also found with 4 % CO2, and it is around 0.2 g L?1 day?1. In series reactors, average CO2 uptake is 0.13 g L?1 day?1 per reactor. TOC analysis shows that the carbon content of the produced biomass is around 40.67 % of total weight. The thermochemical characteristics of the cultivated C. vulgaris samples were analyzed in the presence of air. All samples burn above 200 °C and the combustion rate become faster at around 600 °C. Almost 98 wt% of the produced biomass is combustible in this range.  相似文献   

17.
We evaluated the nighttime CO2 flux (ecosystem respiration) on Rishiri Island, located at the northern tip of Hokkaido, Japan, from 2009 to 2011, by using the relationship between atmospheric 222Rn and CO2 concentrations. The annual mean CO2 flux was 1.8 μmol m?2 s?1, with a maximum monthly mean in July (4.6 ± 2.6 μmol m?2 s?1) and a broad minimum from December to March (0.33 ± 0.29 μmol m?2 s?1). The annual mean was comparable to fluxes at the JapanFlux sites in northern Japan. During the season of snow cover (mid-December to early April), the CO2 flux was low (0.45 ± 0.43 μmol m?2 s?1). Total annual respiration was estimated at 679 ± 174 g cm?2, about 8 % of which occurred during the season of snow cover.  相似文献   

18.
We measured CO2 concentration and determined evasion rate and piston velocity across the water–air interface in flow-through chambers at eight stations along two 20 km long streams in agricultural landscapes in Zealand, Denmark. Both streams were 9–18-fold supersaturated in CO2 with daily means of 240 and 340 μM in January–March and 130 and 180 μM in June–August. Annual CO2 medians were 212 μM in six other streams and 460 μM in four groundwater wells, while seven lakes were weakly supersaturated (29 μM). Air concentrations immediately above stream surfaces were close to mean atmospheric conditions except during calm summer nights. Piston velocity from 0.4 to 21.6 cm h?1 was closely related to current velocity permitting calculation of evasion rates for entire streams. CO2 evasion rates were highest in midstream reaches (170–1,200 mmol m?2 day?1) where CO2-rich soil water entered fast stream flow, while rates were tenfold lower (25–100 mmol m?2 day?1) in slow-flowing lower reaches. CO2 evasion mainly derived from the input of CO2 in soil water. The variability of CO2 evasion along the two lowland streams covered much of the range in sub-Arctic and temperate streams reported previously. In budgets for the two stream catchments, loss of carbon from soils via the hydrological cycle was substantial (3.2–5.7 mmol m?2 day?1) and dominated by CO2 consumed to form HCO3 ? by mineral dissolution (69–76%) and export of organic carbon (15–23%) relative to dissolved CO2 export (7–9%).  相似文献   

19.
Earthworms are among the world’s most important ecosystem engineers because of their effects on soil fertility and plant productivity. Their dependence on plants for carbon, however, means that any changes in plant community structure or function caused by rising atmospheric CO2 or loss of plant species diversity could affect earthworm activity, which may feed back on plant communities. Production of surface casts measured during three consecutive years in field experimental plots (n = 24, 1.2 m2) planted with local calcareous grassland species that varied in plant species richness (diversity levels: high, 31 species; medium, 12; low, 5) and were exposed to ambient (356 μl CO2 l?1) or elevated (600 μl CO2 l?1) CO2 was only consistently stimulated in high diversity plots exposed to elevated CO2 (+120 %, 31 spp: 603 ± 52 under ambient CO2 vs. 1,325 ± 204 g cast dwt. m?2 year?1 under elevated CO2 in 1996; +77 %, 940 ± 44 vs. 1,663 ± 204 g cast dwt. m?2 year?1 in 1998). Reductions in plant diversity had little effect on cast production in ecosystems maintained at ambient CO2, but the stimulatory effect of elevated CO2 on cast production disappeared when plant species diversity was decreased to 12 and 5 species. High diversity plots were also the only communities that included plant species that an earlier field study showed to be among the most responsive to elevated CO2 and to be most preferred by earthworms to deposit casts near. Further, the +87 % CO2-induced increase in cast production measured over the 3 years corresponded to a parallel increase in cumulative total nitrogen of 5.7 g N m?2 and would help explain the large stimulation of aboveground plant biomass production observed in high-diversity communities under elevated CO2. The results of this study demonstrate how the loss of plant species from communities can alter responses of major soil heterotrophs and consequently ecosystem biogeochemistry.  相似文献   

20.
Forest soils and canopies are major components of ecosystem CO2 and CH4 fluxes. In contrast, less is known about coarse woody debris and living tree stems, both of which function as active surfaces for CO2 and CH4 fluxes. We measured CO2 and CH4 fluxes from soils, coarse woody debris, and tree stems over the growing season in an upland temperate forest. Soils were CO2 sources (4.58 ± 2.46 µmol m?2 s?1, mean ± 1 SD) and net sinks of CH4 (?2.17 ± 1.60 nmol m?2 s?1). Coarse woody debris was a CO2 source (4.23 ± 3.42 µmol m?2 s?1) and net CH4 sink, but with large uncertainty (?0.27 ± 1.04 nmol m?2 s?1) and with substantial differences depending on wood decay status. Stems were CO2 sources (1.93 ± 1.63 µmol m?2 s?1), but also net CH4 sources (up to 0.98 nmol m?2 s?1), with a mean of 0.11 ± 0.21 nmol m?2 s?1 and significant differences depending on tree species. Stems of N. sylvatica, F. grandifolia, and L. tulipifera consistently emitted CH4, whereas stems of A. rubrum, B. lenta, and Q. spp. were intermittent sources. Coarse woody debris and stems accounted for 35% of total measured CO2 fluxes, whereas CH4 emissions from living stems offset net soil and CWD CH4 uptake by 3.5%. Our results demonstrate the importance of CH4 emissions from living stems in upland forests and the need to consider multiple forest components to understand and interpret ecosystem CO2 and CH4 dynamics.  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