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 15 毫秒
1.
Two tomato cDNA libraries were synthesized from poly(A)+ RNAs isolated from unwounded and wounded tomato stems. These cDNA libraries were packaged in gt10 and screened by in situ plaque hybridization with a tomato extensin gene clone (pTom 5.10). Several cDNA clones were identified and isolated from both libraries in this manner and subjected to restriction enzyme digestion, Southern gel blot hybridization, RNA gel blot hybridization, and DNA sequence analyses. From these analyses, the various cDNA clones were found to fall into one of five distinct classes (classes I–V). Class I clones hybridized to a 4.0 kb mRNA which accumulated markedly after wounding and encoded an extensin characterized largely by Ser-(Pro)4-Ser-Pro-Ser-(Pro)4-(Tyr)3-Lys repeats. Class II clones hybridized to a 2.6 kb mRNA which showed no accumulation following wounding and encoded an extensin containing Ser-(Pro)4-Ser-Pro-Ser-(Pro)4-Thr-(Tyr)1–3-Ser repeats. Class III clones hybridized to a 0.6 kb mRNA which greatly accumulated in response to wounding and encoded a glycine-rich protein (GRP) with (Gly)2–6-Tyr-Pro and(Gly)2–6-Arg repeats. Class IV clones contained both class I and class III DNA sequences and consequently hybridized to both the 4.0 kb and the 0.6 kb wound-accumulating mRNAs; these clones encoded a portion of a GRP sequence on one DNA strand and encoded a portion of an extensin sequence on the other DNA strand. Class V clones hybridized to a 2.3 kb mRNA which decreased following wounding and encoded a GRP sequence characterized by (Gly)2–5-Arg repeats.  相似文献   

2.
Rotifer cultures of Brachionus plicatilis (SINTEF-strain, length 250 m) rich in 3 fatty acids were starved for > 5 days at variable temperature (0–18 °C). The net specific loss rate of rotifer numbers were 0.04 day–1 (range 0–0.08 day–1) at 5–18 °C, but reached values up to 0.25 day–1 at 0–3 °C. The loss rate was independent on culture density (range 40–1000 ind ml–1), but was to some extent dependent on the initial physiological state of the rotifers (i.e., egg ratio).The loss rate of lipids was 0.02–0.05 day–1 below 10 °C, where the potential growth rate of the rotifer is low (0–0.09 day–1). The loss rate of lipids increased rapidly for higher temperatures where the rotifer can maintain positive growth, and reached 0.19 day–1 at 18 °C. The Q10 for the lipid loss rate versus temperature was higher than the Q10 for respiration found in other strains. This may suggest that other processes than respiration were involved in lipid catabolism. The content of 3 fatty acids became reduced somewhat faster than the lipids (i.e. in particular 22:6 3), but the fatty acid per cent distribution remained remarkably unaffected by the temperature during starvation.The results showed that rotifer cultures could be starved for up to 4 days at 5–8 °C without essential quantitative losses of lipids, 3 fatty acids, and rotifers. The rotifers exhausted their endogenous lipids through reproduction (anabolism) and respiration (including enhanced locomotion) at higher temperatures. At lower temperatures, the mortality rate became very high.  相似文献   

3.
The field metabolic rates (FMR) and rates of water flux were measured in two species of varanid lizards over five periods of the year in tropical Australia. The energetics of these species were further investigated by directly measuring activity (locomotion) and body temperatures of free-ranging animals by radiotelemetry, and by measuring standard metabolic rate (over a range of body temperatures) and activity metabolism in the laboratory. Seasonal differences in the activity and energetics were found in these goannas despite similar, high daytime temperatures throughout the year in tropical Australia. Periods of inactivity were associated with the dry times of the year, but the onset of this period of inactivity differed with respect to habitat even within the same species. Varanus gouldii, which inhabit woodlands only, were inactive during the dry and late dry seasons. V. panoptes that live in the woodland had a similar seasonal pattern of activity, but V. panoptes living near the floodplain of the South Alligator River had their highest levels of activity during the dry season when they walked long distances to forage at the receding edge of the floodplain. However, during the late dry season, after the floodplain had dried completely, they too became inactive. For V. gouldii, the rates of energy expenditure were 196 kJ kg–1 day–1 for active animals and 66 kJ kg–1 day–1 for inactive animals. The rates of water influx for these groups were respectively 50.7 and 19.5 ml kg–1 day–1. For V. panoptes, the rates of energy expenditure were 143 kJ kg–1 day–1 for active animals and 56 kJ kg–1 day–1 for inactive animals. The rates of water influx for these two groups were respectively 41.4 and 21.0 ml kg–1 day–1. We divided the daily energy expenditure into the proportion of energy that lizards used when in burrows, out of burrows but inactive, and in locomotion for the two species during the different seasons. The time spent in locomotion by V. panoptes during the dry season is extremely high for a reptile (mean of 3.5 h/day spent walking), and these results provide an ecological correlate to the high aerobic capacity found in laboratory measurements of some species of varanids.  相似文献   

4.
Summary Small birch plants (Betula pendula Roth.) were cultivated in a hydroponic spray solution where the relative addition rate of iron (RFe; g g–1 day–1), was the growth-controlling variable. All other elements were added in free access. An additional treatment was performed where all nutrients, including iron, were in free access (FA). The plants showed deficiency symptoms at steady-state growth and severe limitation of iron, RFe 0.05 and 0.10 day–1. There were few symptoms at RFe of 0.15 or above. Plant relative growth rate (RG; g g–1 day–1), equalled the relative rate of increase in iron supply, RFe. Internal iron concentration of the plants ranged from 40 to 70 g g–1 dry weight (DW) over the range for which iron supply was limiting growth. At FA, the internal concentration was approximately 200 g g–1 DW without further increase in RG, demonstrating that iron may be taken up in excess without affecting growth. Internal concentrations of macronutrients were stable at the different RFe, except for Ca and Mg in shoots which were higher at low iron supply. Uptake rates of iron, calculated per root growth rate (mol g–1 root DW), were approximately twice as high at RFe 0.20 as at 0.05 day–1. The effect of iron limitation on dry matter allocation to leaves was small, with increases in the root fraction being largely at the expense of the stem. Leaf area ratio was constant regardless of RFe and the specific leaf area tended to increase with increasing iron limitation. Net assimilation rate decreased by a factor of 6 from free access to severe iron limitation, largely accounting for the differences in plant RG.  相似文献   

5.
Mathematical model parameters for the methanogenic degradation of propylene glycol were estimated in a sequential manner by means of an optimization technique. Model parameters determined from an initial experimental data set using one bioreactor were then verified with the results from a second bioreactor. The proposed methodology is a useful tool to obtain model parameters for continuous flow reactors with completely mixed regime. Abbrevations: S – substrate concentration (mg COD l–1); S in – influent substrate concentration (mg COD l–1); D L – dilution rate (day–1); – stoichiometric coefficients (ND); nx – number of microbial species (ND); X S – fixed biomass concentration (mg biomass l–1); X L – suspended biomass concentration of (mg biomass l–1); k d – decay rate of biomass (day–1); b S – specific detachment rate of biofilm (day–1); – specific growth rate of biomass (day–1); m – maximum specific growth rate of biomass (day–1); K S – half saturation constant (mg COD l–1); K I – inhibition constant (mg COD l–1).  相似文献   

6.
A model primitive gas containing a mixture of N2, CO and water vapor over a water pool (300 mL, 37 °C) was subjected to electric discharges. The discharge vessel (7 L in volume) was equipped with a CO2 absorber (The CO2 being formed during the discharge), thus simulating possible absorption of CO2 in the primitive ocean. The vessel also has a cold trap ( –15 °C), which protects the primary products against the further decomposition in the discharge phase by enabling these products to adhere to the trap. Since the partial pressures of CO and N2 decreased at rates of 1.5–1.7 cmHg day–1 and 0.1–0.2 cmHg day–1, respectively, the gases were added at regular intervals. The solution was analyzed at regular intervals for HCN, HCHO and urea, and maximum concentrations of about 50, 2, and 140 mM were observed. The discharge phase was continued for 6 months. In the solution, glycine (5.6% yield based on the carbon), glycylglycine (0.64%), orotic acid (0.004%) and small amounts of the other amino acids were found.  相似文献   

7.
The cyanobacterium Spirulina platensis was used to verify the possibility of employing microalgal biomass to reduce the contents of nitrate and phosphate in wastewaters. Batch tests were carried out in 0.5 dm3 Erlenmeyer flasks under conditions of light limitation (40 mol quanta m–2 s–1) at a starting biomass level of 0.50 g/dm3 and varying temperature in the range 23–40°C. In this way, the best temperature for the growth of this microalga (30°C) was determined and the related thermodynamic parameters were estimated. All removed nitrate was used for biomass growth (biotic removal), whereas phosphate appeared to be removed mainly by chemical precipitation (abiotic removal). The best results in terms of specific and volumetric growth rates ( =0.044 day–1, Q x =33.2 mg dm–3 day–1) as well as volumetric rate and final yield of nitrogen removal ( =3.26 mg dm–3 day–1, =0.739) were obtained at 30°C, whereas phosphorus was more effectively removed at a lower temperature. In order to simulate full-scale studies, batch tests of nitrate and phosphate removal were also performed in 5.0 dm3 vessels (mini-ponds) at the optimum temperature (30°C) but increasing the photon fluence rate to 80 mol quanta m–2 s–1 and varying the initial biomass concentration from 0.25 to 0.86 g/dm3. These additional tests demonstrated that an increase in the inoculum level up to 0.75 g/dm3 enhanced both NO3 and PO4 3– removal, confirming a strict dependence of these processes on biomass activity. In addition, the larger surface area of the ponds and the higher light intensity improved removal yields and kinetics compared to the flasks, particularly concerning phosphorus removal ( =0.032–0.050 day–1, Q x =34.7–42.4 mg dm–3 day–1, =3.24–4.06 mg dm–3 day–1, =0.750–0.879, =0.312–0.623 mg dm–3 day–1, and =0.224–0.440).  相似文献   

8.
Fractional rates (% · day–1) of synthesis and degradation were determined by measuring the output of N-methylhistidine (MeHis) in the excreta at 4 and 8 weeks of age in the chicken. At 4 weeks of age, the fractional rate of synthesis of the meat-type stock was twice that of the egg-type stock (White Leghorn), but the fractional rates of synthesis at 8 weeks of age were similar (4.1–5.1% · day–1) among stocks. The fractional rate of degradation (1.3–1.5% · day–1) of the meat-type stock at 8 weeks of age was less than half the rate of the egg-type stock (2.9% · day–1). The fractional rates of synthesis and degradation at 4 weeks of age in the Satsuma native fowl were relatively high compared with those in the other stocks. In particular, the rate of degradation (8.6% · day–1) at 4 weeks of age was approximately twice that of other stocks. These results show that fractional rates of synthesis and degradation of muscle protein in the chicken differ among genetically diverse groups. The effect of changes in rates of synthesis and degradation on the change in fractional growth rate also differed. From regression coefficients (bK s · FGR and bK d · FGR) of these rates in skeletal muscle protein on the fractional growth rate, it was recognized that the change in growth rate accompanies the changes in both synthesis and degradation in White Leghorn and commercial broilers but only the change in synthesis in White Plymouth Rock (dw) and Satsuma native fowl.  相似文献   

9.
Measurements of the vertical temperature in tropical Lake McIlwaine were used to calculate the time-averaged ( 6 months) vertical diffusivity coefficient (Kz) in the metalimnion and hypolimnion. The mean value of Kz (0.21 cm2 s–1) was correlated with the lake surface area. The mass transport rates of PO4-P and NH4-N, upward from the hypolimnion to the metalimnion, were calculated using Kz and measured values of the nutrient concentration gradients. During a period of 4.5 months when the water was stably stratified, PO4-P was transported upward at a mean rate of 42 kg day–1 and NH4-N at a mean rate of 162 kg day–1 over the entire lake.  相似文献   

10.
Estimates of bacterial production based on total trichloroacetic acid (TCA)-precipitable [methyl-3H]thymidine incorporation and frequency of dividing cell (FDC) techniques were compared to sediment respiration rates in Lake George, New York. Bacterial growth rates based on thymidine incorporation ranged from 0.024 to 0.41 day–1, while rates based on FDC ranged from 1.78 to 2.48 day–1. Respiration rates ranged from 0.11 to 1.8mol O2·hour–1·g dry weight sediment–1. Thymidine incorporation yielded production estimates which were in reasonable agreement with respiration rates. Production estimates based on FDC were 4- to 190-fold higher than those predicted from respiration rates.  相似文献   

11.
Aerobic respiration with oxygen and anaerobic respiration with nitrate (denitrification) and sulfate (sulfate reduction) were measured during winter and summer in two coastal marine sediments (Denmark). Both aerobic respiration and denitrification took place in the oxidized surface layer, whereas sulfate reduction was most significant in the deeper, reduced sediment. The low availability of nitrate apparently limited the activity of denitrification during summer to less than 0.2 mmoles NO 3 m–2 day–1, whereas activities of 1.0–3.0 mmoles NO 3 m–2 day–1 were measured during winter. Sulfate reduction, on the contrary, increased from 2.6–7.6 mmoles SO 4 2– m–2 day–1 during winter to 9.8–15.1 mmoles SO 4 2– m–2 day–1 during summer. The aerobic respiration was high during summer, 135–140 mmoles O2 m–2 day–1, as compared to estimated winter activities of about 30 mmoles O2 m–2 day–1. The little importance of denitrification relative to aerobic respiration and sulfate reduction is discussed in relation to the availability and distribution of oxygen, nitrate, and sulfate in the sediments and to the detritus mineralization.  相似文献   

12.
Removal of inorganic nitrogen sources by cells of the aerial microalga Trentepohlia aurea grown on the surface of substrate, such as filter paper, has been investigated in a batch system. When the alga grew on the paper dampened with medium, it actively ingested inorganic nitrogenous compounds in the medium. Immobilized cells on the filter papers were called algal biofilm in this study. When the algal biofilms were soaked in modified Bold's Basal medium (using 1 g NH4Cl l–1 as a N source), the removal rate was 4.25 mg ammonium-N l–1 day–1 in 40 days. In modified medium with added 26 mg nitrite-N, the removal rate of the total inorganic N ion by the biofilms reached 5.11 mg N l–1 day–1. This removal rate of total N ion was higher than that in the medium by addition of 26 mg nitrate-N. In addition, we tried to examine simultaneous removal of ammonium, nitrate, and nitrite ions and growth inhibition of cyanobacteria in the medium by using the algal biofilms. Consequently, it was demonstrated that the algal biofilms of T. aurea could be utilized as a biofunctional material for the purification of wastewater.  相似文献   

13.
Submerged macrophytes are a major component of freshwater ecosystems, yet their net effect on water column phosphorus (P), algae, and bacterioplankton is not well understood. A 4-month mass-balance study during the summer quantified the net effect of a large (5.5 ha) undisturbed macrophyte bed on these water-column properties. The bed is located in a slow-flowing (0.05–0.1 cm s–1) channel between two lakes, allowing for the quantification of inputs and outputs. The P budget for the study period showed that, despite considerable short-term variation, the macrophyte bed was a negligible net sink for P (0.06 mg m–2 day–1, range from –0.76 to +0.79 mg m–2 day–1), demonstrating that loading and uptake processes in the weedbed roughly balance over the summer. Chlorophyll a was disproportionately retained relative to particulate organic carbon (POC), indicating that the algal component of the POC was preferentially trapped. However, the principal contribution of the weedbed to the open water was a consistent positive influence on bacterioplankton production over the summer. Conservative extrapolations based on measured August specific exports (m–2 day–1) of P and bacterial production exiting the weedbed applied to five regional lakes varying in lake morphometry and macrophyte cover suggest that even in the most macrophyte dominated of lakes (66% cover), P loading from submerged weedbeds never exceeds 1% day–1 of standing epilimnetic P levels, whereas subsidization of bacterioplankton production can reach upward of 20% day–1. The presence of submerged macrophytes therefore differentially modifies algae and bacteria in the water column, while modestly altering P dynamics over the summer.  相似文献   

14.
Ammonia-nitrogen excretion in Daphnia pulex   总被引:3,自引:2,他引:1  
Ammonia-nitrogen excretion rates were measured in natural summer and cultured populations of Daphnia pulex from Silver Lake, Clay County, Minnesota, USA during 1973. The mean rate of ammonia-nitrogen excretion for the summer populations was 0.20 µg N animal–1 day–1 or 5.11 µg N mg–1 dry body weight day–1 (N = 80) measured at 15°, 20°, and 25°C. These rates appear to be temperature and weight dependent, but they are probably affected by factors other than temperature and dry body weight. Ammonia-nitrogen excretion rates of Daphnia pulex cultured on Chlamydomonas reinhardi yielded the following relationship with temperature: Log10E = (0.061) T 1.773, where E is µg N animal–1 day–1 and T is temperature °C. The ammonia-nitrogen excretion on a mg–1 dry body weight day–1 basis was related to temperature according to the following similar expression Log10E = (0.043) T + 0.153, where E is µg N mg–1 dry body weight day–1, and T is temperature °C. The length-weight relationship of Daphnia pulex for the summer populations (N = 1583) was log10W = (0.526) Log10L + 1.357, where W is weight in µg and L is length in mm.  相似文献   

15.
Microbial activity under alpine snowpacks, Niwot Ridge, Colorado   总被引:19,自引:9,他引:10  
Experiments were conducted during 1993 at Niwot Ridge in the Colorado Front Range to determine if the insulating effect of winter snow cover allows soil microbial activity to significantly affect nitrogen inputs and outputs in alpine systems. Soil surface temperatures under seasonal snowpacks warmed from –14 °C in January to 0 °C by May 4th. Snowmelt began in mid-May and the sites were snow free by mid June. Heterotrophic microbial activity in snow-covered soils, measured as C02 production, was first identified on March 4, 1993. Net C02 flux increased from 55 mg CO2-C m–2 day–1 in early March to greater than 824 mg CO2-C m-2 day–1 by the middle of May. Carbon dioxide production decreased in late May as soils became saturated during snowmelt. Soil inorganic N concentrations increased before snowmelt, peaking between 101 and 276 mg kg–1 soil in May, and then decreasing as soils became saturated with melt water. Net N mineralization for the period of March 3 to May 4 ranged from 2.23 to 6.63 g N m–2, and were approximately two orders of magnitude greater than snowmelt inputs of 50.4 mg N m–2 for NH4 + and 97.2 mg N m–2 for NO3 . Both NO3 and NH4 + concentrations remained at or below detection limits in surface water during snowmelt, indicating the only export of inorganic N from the system was through gaseous losses. Nitrous oxide production under snow was first observed in early April. Production increased as soils warned, peaking at 75 g N2O-N m–2 day–1 in soils saturated with melt water one week before the sites were snow free. These data suggest that microbial activity in snow-covered soils may play a key role in alpine N cycling before plants become active.  相似文献   

16.
Metaseiulus occidentalis females from the carbaryl-organophosphate-sulfur resistant strain (COS) lived longer (25.3 days versus 19.7 days), had a higher total fecundity (43.8 eggs female–1 versus 33.6 eggs female–1) and a higher daily fecundity rate (2.4 eggs female–1 day–1 versus 2.0 eggs female–1 day–1), and exhibited a higher intrinsic rate of increase (0.243 individuals female–1 day–1 versus 0.182 individuals female–1 day–1) and shorter generation time (13.9 days versus 17.0 days), at 24–28°C, 47–56%rh under continuous fluorescent light, when reared on a diet of 0–48-h-old eggs rather than a diet of mixed actives ofTetranychus pacificus McGregor on bean leaf disks. The sexratio of the progeny was female-biased for both diets, 2.1 females to 1 male forM. occidentalis reared on eggs and 2.0:1 : forM. occidentalis reared on mixed actives, suggesting that diet influences sex-ratio in some unknown way.There was no significant difference in oviposition rates for repeatedly-mated and once-matedM. occidentalis females reared on a diet of younger (0–24-h old) eggs compared to a diet of older (72–96-h old) eggs ofT. pacificus.The COS strain ofM. occidentalis exhibited life-table parameters comparable to the other strains reported in the literature, suggesting that the reproductive attributes of this acarine predator were not reduced as a result of artificial laboratory selection. Diet, a biotic factor, produced substantial differences in life-table parameters, suggesting that this factor can influence conclusions regarding the potential efficacy of biological control agents.  相似文献   

17.
Cell cultures of chili pepper (Capsicum annuum L.) were established from callus tissue inoculated in MS liquid medium supplemented with 6.25 M 2,4-d and 0.44 M BA. Cell clones were isolated by plating the cell suspension on filter paper discs supported by polyurethane foam that were bathed with culture medium containing 15% PEG. The cell clones T6 and T7 were chosen based on their characteristics of growth and friability. These cell clones were established as cell suspensions in the presence of 15% PEG and subsequently subcultured in increasing concentrations of osmoticum. By this approach the cell clones T7 and T6 were capable of growing in the presence of 20 and 25% PEG, respectively. The cell clone T7 was found to grow better in the presence of 5–10% PEG after a period of subculturing in the absence of osmoticum indicating that the tolerance trait was stable. The tolerant cell clones exhibited a 3 to 3.5-fold decrease in the osmotic potentials in comparison with the nonselected cells suggesting that osmotic adjustment occurred. K+ was the major contributing solute to the osmotic potential in all the cell cultures among those tested and was found to be higher in concentration in the PEG-tolerant clones (1.3–3 times higher than nonselected cells). Proline and glycine betaine levels showed a positive correlation with the degree of tolerance to water deficit in the PEG-tolerant cell clones. The levels of proline in the cell clone T7 subcultured in the absence of PEG in the culture medium decreased to values similar to those of nonselected cells, whereas the contents of glycine betaine in the same conditions were maintained at high levels.Abbreviations BA benzyladenine - 2,4-d 2,4-dichlorophenoxyacetic acid - MS Murashige and Skoog medium - PEG polyethylene glycol  相似文献   

18.
Among 150 strains, including marine cyanobacteria isolated from coastal areas of Japan and a freshwater cyanobacterium from the IAM collection, Spirulina platensis IAM M-135, the marine cyanobacterium Synechococcus sp. NKBG 042902 contained the highest amount of phycocyanin (102 mg/g dry cell weight). We have proposed that the cyanobacterium could be an alternative producer for phycocyanin. The effects of light intensity and light quality on the phycocyanin content in cells of Synechococcus sp. NKBG 042902 were investigated. When the cyanobacterium was cultured under illumination of 25 mol m–2 s–1 using a cool-white fluorescent lamp, the phycocyanin content was highest, and the phycocyanin and biomass productivities were 21 mg 1–1 day–1 and 100 mg 1–1 day–1 respectively. Red light was essential for phycocyanin production by this cyanobacterium. Phycocyanin and biomass production were carried out by the cyanobacterium cultures grown under only red light (peak wavelength at 660 nm) supplied from light-emitting diodes (LED). Maximum phycocyanin and biomass productivities were 24 mg 1–1 day–1 and 130 mg 1–1 day–1 when the light intensity of the LED was 55 mol m–2 s–1.  相似文献   

19.
Phytoplankton production was measured in situ in Kainji lake from December 1970 to September 1972 using the oxygen light and dark bottle technique. Seasonal variations in solar radiation, transparency, temperature, and composition of subsurface light were also measured. Oxygen production per unit area varied from 220 to 4500 mg O2 m–2 day–1, the maximum production rate from 95 to 400 mg O2 m–3 h–1. Seasonal mixing of lake water and river water of varying turbidity changed the optical properties of the lake water and consequently affected phytoplankton production. The annual flood pattern was found to be an important factor regulating phytoplankton production in the lake.  相似文献   

20.
The Catskill Mountains of southeastern New York receive among the highest rates of atmospheric nitrogen (N) deposition in eastern North America, and ecosystems in the region may be sensitive to human disturbances that affect the N cycle. We studied the effects of a clearcut in a northern hardwood forest within a 24-ha Catskill watershed on the net rates of N mineralization and nitrification in soil plots during 6 years (1994–1999) that encompassed 3-year pre- and post-harvesting periods. Despite stream NO3 concentrations that increased by more than 1400 mol l–1 within 5 months after the clearcut, and three measures of NO3 availability in soil that increased 6- to 8-fold during the 1st year after harvest, the net rates of N mineralization and nitrification as measured by in situ incubation in the soil remained unchanged. The net N-mineralization rate in O-horizon soil was 1– 2 mg N kg–1 day–1 and the net nitrification rate was about 1 mg N kg–1 day–1, and rates in B-horizon soil were only one-fifth to one-tenth those of the O-horizon. These rates were obtained in single 625 m2 plots in the clearcut watershed and reference area, and were confirmed by rate measurements at 6 plots in 1999 that showed little difference in N-mineralization and nitrification rates between the treatment and reference areas. Soil temperature increased 1 ± 0.8 °C in a clearcut study plot relative to a reference plot during the post-harvest period, and soil moisture in the clearcut plot was indistinguishable from that in the reference plot. These results are contrary to the initial hypothesis that the clearcut would cause net rates of these N-cycling processes to increase sharply. The in situ incubation method used in this study isolated the samples from ambient roots and thereby prevented plant N uptake; therefore, the increases in stream NO3 concentrations and export following harvest largely reflect diminished uptake. Changes in temperature and moisture after the clearcut were insufficient to measurably affect the net rates of N mineralization and nitrification in the absence of plant uptake. Soil acidification resulting from the harvest may have acted in part to inhibit the rates of these processes. The US Governments right to retain a non-exclusive, royalty-free license in and to any copyright is acknowledged.  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