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.
The isotopic compositions of carbon compounds in landfill leachate provide insights into the biodegradation pathways that dominate the different stages of waste decomposition. In this study, the carbon geochemistry of different carbon pools, environmental stable isotopes and compound-specific isotope analysis (CSIA) of leachate dissolved organic carbon (DOC) fractions and gases show distinctions in leachate biogeochemistry and methane production between the young area of active waste emplacement and the old area of historical emplacement at the Trail Road Landfill (TRL).

The active area leachate has low DOC concentrations (<200 mg l?1) dominated by fulvic acid (FA=160 mg l?1), and produces CH4 dominantly by CO2 reduction (D? excess=20.6‰). Leachate generated in the area of older waste has high DOC (>4770 mg l?1) dominated by FA (4482 mg l?1) and simple fatty acids (acetic=1008 mg l?1 and propionic=608 mg l?1), and produces CH4 by the acetate fermentation pathway (D? excess=9.8‰). CSIA shows an advanced degradation and a progressive accumulation of 13C of fatty acids in leachate from the older area. The enriched 13C value of FA (?20 and?26‰ for the older and active parts, respectively,) and of low molecular weight DOC (?8 and?27‰) as well as of the bulk DOC (?21 and?25‰) shows more advanced degradation in the older part of the landfill, which is consistent with the shift in the humic/FA ratios (0.05 and 0.18). The 13C enrichment of acetate (?12‰) above the 13C of DOC (?21‰) and of propionic acid (?19‰), in older leachate, suggests that this acetate has not evolved from the simple degradation of larger organic molecules, but by homoacetogenesis from the enriched dissolved inorganic carbon (DIC) pool (8‰) and H2, which produce a more enriched 13C of acetate. In contrast, the 13C of the minor acetate in the active area (?17‰) indicates that CO2-reducing bacteria must be the primary consumers of H2, which has resulted in enriched 13CDIC (10‰) and depleted 13CCH4 (?58‰).  相似文献   

2.
On the Qinghai–Tibetan Plateau, isotopic signatures in soil–atmosphere CH4 fluxes were investigated in nine grasslands and three wetlands. In the grasslands, the fractionation factor for soil CH4 uptake, αsoil, was much smaller than the usually reported value of 0.9975–1.0095. Stepwise multiple variation analysis indicates that αsoil is higher for higher soil water contents but is lower for higher C/N ratios of soil surface biomass. In the three wetlands, the soil-emitted δ13C–CH4 was similar (?55.3?±?5.5?‰ and ?53.0?±?5.5?‰) in two bogs separated by >1000?km but was lower (?63.4?±?6.3?‰) in a marsh. Environmental factors related to intrasite variations in soil-emitted δ13C–CH4 include the soil C/N ratio, oxidation–reduction potential, soil C concentration and soil water contents. Geographical isotopic surveys revealed environmental constraints on the CH4 consumption pathways in grasslands and the biome type-specific consistency in CH4 production pathways in wetlands.  相似文献   

3.
Abstract

Naturally produced methane shows different δ13C-values with respect to its origin, e.g., geological or biological. Methane-production of ruminants is considered to be the dominant source from the animal kingdom. Isotopic values of rumen methane—given in literature—range between ?80‰ and -50‰ and are related to feed composition and also sampling techniques. Keeping cows, camels and sheep under identical feed conditions and sampling rumen gases via implanted fistulae we compared δPDB 13C-values of methane and CO2 between the species. Referring to mean values obtained from 4 or 5 samples at different times of 11 animals (n = 47) we calculated δPDB 13C-medians resulting in small but not significant differences within and significant differences between the species for CO2 and methane. The δPDB 13C-differences between methane and CO2 were statistically equal within and also between the species. Therefore a linear regression of methane values on CO2 is appropriate and leads to: δPDB 13C(methane)‰ = 1,57 * δPDB 13C(CO2)‰-47‰ with a correlation coefficient of r = 0,87.  相似文献   

4.
Measurement of soil-respired CO2 at high temporal resolution and sample density is necessary to accurately identify sources and quantify effluxes of soil-respired CO2. A portable sampling device for the analysis of δ13C values in the field is described herein.

CO2 accumulated in a soil chamber was batch sampled sequentially in four gas bags and analysed by Wavelength-Scanned Cavity Ring-down Spectrometry (WS-CRDS). A Keeling plot (1/[CO2] versus δ13C) was used to derive δ13C values of soil-respired CO2. Calibration to the δ13C Vienna Peedee Belemnite scale was by analysis of cylinder CO2 and CO2 derived from dissolved carbonate standards. The performance of gas-bag analysis was compared to continuous analysis where the WS-CRDS analyser was connected directly to the soil chamber.

Although there are inherent difficulties in obtaining absolute accuracy data for δ13C values in soil-respired CO2, the similarity of δ13C values obtained for the same test soil with different analytical configurations indicated that an acceptable accuracy of the δ13C data were obtained by the WS-CRDS techniques presented here. Field testing of a variety of tropical soil/vegetation types, using the batch sampling technique yielded δ13C values for soil-respired CO2 related to the dominance of either C3 (tree, δ13C=?27.8 to?31.9 ‰) or C4 (tropical grass, δ13C=?9.8 to?13.6 ‰) photosynthetic pathways in vegetation at the sampling sites. Standard errors of the Keeling plot intercept δ13C values of soil-respired CO2 were typically<0.4 ‰ for analysis of soils with high CO2 efflux (>7–9 μmol m?2 s?1).  相似文献   

5.
δ13C values of gaseous acetaldehyde were measured by gas chromatograph–combustion–isotope ratio mass spectrometer (GC–C–IRMS) via sodium bisulfite (NaHSO3) adsorption and cysteamine derivatisation. Gaseous acetaldehyde was collected via NaHSO3-coated Sep-Pak® silica gel cartridge, then derivatised with cysteamine, and then the δ13C value of the acetaldehyde–cysteamine derivative was measured by GC–C–IRMS. Using two acetaldehydes with different δ13C values, derivatisation experiments were carried out to cover concentrations between 0.009×10?3 and 1.96×10?3 mg·l?1) of atmospheric acetaldehyde, and then δ13C fractionation was evaluated in the derivatisation of acetaldehyde based on stoichiometric mass balance after measuring the δ13C values of acetaldehyde, cysteamine and the acetaldehyde–cysteamine derivative. δ13C measurements in the derivertisation process showed good reproducibility (<0.5 ‰) for gaseous acetaldehyde. The differences between predicted and measured δ13C values were 0.04–0.31 ‰ for acetaldehyde–cysteamine derivative, indicating that the derivatisation introduces no isotope fractionation for gaseous acetaldehyde, and obtained δ13C values of acetaldehyde in ambient air at the two sites were distinct (?34.00 ‰ at an urban site versus?31.00 ‰ at a forest site), implying potential application of the method to study atmospheric acetaldehyde.  相似文献   

6.
Soil from Free-Air Carbon dioxide Enrichment (FACE) plots (FAL, Braunschweig) under ambient air (375 ppm; δ13C–CO2?9.8‰) and elevated CO2 (550 ppm; for six years; δ13C–CO2?23‰), either under 100% nitrogen (N) (180 kg ha?1) or 50% N (90 kg ha?1) fertilisation treatments, was analysed by thermogravimetry. Soil samples were heated up to the respective temperatures and the remaining soil was analysed for δ13C and δ15N by Isotope Ratio Mass Spectrometry (IRMS). Based on differential weight losses, four temperature intervals were distinguished. Weight losses in the temperature range 20–200 °C were connected mostly with water volatilisation. The maximum weight losses and carbon (C) content were measured in the soil organic matter (SOM) pool decomposed at 200–360 °C. The largest amount of N was detected in SOM pools decomposed at 200–360 °C and 360–500 °C. In all temperature ranges, the δ13C values of SOM pools were significantly more negative under elevated CO2 versus ambient CO2. The incorporation of new C into SOM pools was not inversely proportional to its thermal stability. 50% N fertilisation treatment gained higher C exchange under elevated CO2 in the thermally labile SOM pool (200–360 °C), whereas 100% N treatment induced higher C turnover in the thermally stable SOM pools (360–500 °C, 500–1000 °C). Mean Residence Time of SOM under 100% N and 50% N fertilisation showed no dependence between SOM pools isolated by increasing temperature of heating and the renovation of organic C in those SOM pools. Thus, the separation of SOM based on its thermal stability was not sufficient to reveal pools with contrasting turnover rates of C.  相似文献   

7.
Relative and absolute line intensities for the ν3 bands of the 12C and 13C isotopic varieties of methane have been measured using a tunable difference-frequency laser spectrometer. From these data the integrated band strength of 13CH4 is calculated to be 0.983 ± 0.007 that of 12CH4, with the uncertainty representing three standard deviations. The absolute ν3 bandstrength for 12CH4 is 266.1 ± 3.0 cm?2 atm ?1 at 294.7 K where the errors are dominated by the pressure measurement. This band strength corresponds to an effective transition moment 〈μ3〉 = 0.0534(3)D for 12CH4 from which the ν4 band dipole moment and the Herman-Wallis F factor can be estimated using a recent force field model for methane.  相似文献   

8.
High and fluctuating salinity is characteristic for coastal salt marshes, which strongly affect the physiology of halophytes consequently resulting in changes in stable isotope distribution. The natural abundance of stable isotopes (δ13C and δ15N) of the halophyte plant Salicornia brachiata and physico-chemical characteristics of soils were analysed in order to investigate the relationship of stable isotope distribution in different populations in a growing period in the coastal area of Gujarat, India. Aboveground and belowground biomass of S. brachiata was collected from six different populations at five times (September 2014, November 2014, January 2015, March 2015 and May 2015). The δ13C values in aboveground (?30.8 to ?23.6?‰, average: ?26.6?±?0.4?‰) and belowground biomass (?30.0 to ?23.1?‰, average: ?26.3?±?0.4?‰) were similar. The δ13C values were positively correlated with soil salinity and Na concentration, and negatively correlated with soil mineral nitrogen. The δ15N values of aboveground (6.7–16.1?‰, average: 9.6?±?0.4?‰) were comparatively higher than belowground biomass (5.4–13.2?‰, average: 7.8?±?0.3?‰). The δ15N values were negatively correlated with soil available P. We conclude that the variation in δ13C values of S. brachiata was possibly caused by soil salinity (associated Na content) and N limitation which demonstrates the potential of δ13C as an indicator of stress in plants.  相似文献   

9.
A combination of C/N ratios, δ13C and δ15N values in suspended matter was used to examine the seasonal (late summer 2004 and spring 2005) relationship with hydrological characteristics of the River Sava watershed in Slovenia. The values of C/N ratios range from 1.2 to 19.1, δ13C values range from?29.2 to?23.0 ‰ and δ15N values from 0.5 to 16.7 ‰ and indicate that the samples are a mixture of two end members: modern soils and plant litter. A simple mixing model was used to indicate that soil organic carbon prevails over plant litter and contributes more than 50% of suspended material. The calculated annual particulate organic carbon flux is estimated as 5.2×1010 g C/year, the annual particulate nitrogen flux 8.5×109 g N/year and the total suspended solid flux is estimated to be 1.3×1012 g/year. Anthropogenic impact was detected only in a tributary stream of the River Sava, which is located in an agriculture–industrial area and is reflected in higher δ15N values in suspended matter and high nitrate concentrations in the late summer season.  相似文献   

10.
Abstract

The isotopic compositions of biogenic carbon dioxide and methane from different sites were investigated. The δ13C values of methane vary mainly between ?55‰ and ?75‰ whereas δ13C values of carbon dioxide were found from about + 11‰ to ?23‰. Especially the latter ones are not so typical for microbial gases. The different sites don't vary over the whole scales but form certain groups. Secondary effects like diffusion change the δ values of both components in an even more negative direction, while oxidation processes near the surface result in more positive δ13C values for methane and very negative δ13C values for carbon dioxide.  相似文献   

11.
We employed tunable diode laser absorption spectroscopy to measure the line strength, the methane (CH4), ethane (C2H6) and the propane (C3H8) broadening coefficients for the 523–422 H2O transition at 3619.61 cm?1. Water amount fractions generated by a stable and accurate humidity transfer standard, traceable to the SI units via the German national humidity standard, were used to calibrate the spectroscopic line strength measurements. We focus on the traceability of the measured line data to the SI and on uncertainty assessments following the guidelines of the Guide to the Expression of Uncertainty in Measurement. We determined the line strength to be (8.42 ± 0.07)×10?20 cm?1/(cm?2 molecule) corresponding to a relative uncertainty of ±0.8%. To the best of our knowledge, we report the first methane, ethane and propane broadening coefficients of (8.037 ± 0.056)×10?5 cm?1/hPa, (9.077 ± 0.064)×10?5 cm?1/hPa and (10.469 ± 0.073)×10?5 cm?1/hPa for the 523–422 H2O transition at 3619.61 cm?1, respectively. The relative combined uncertainties of the stated CH4, C2H6 and C3H8 broadening coefficients are in the ±0.7% range.  相似文献   

12.
A survey study was conducted on man-made plantations located at two different areas in the arid region of Syria to determine the variations in natural abundances of the 15N and 13C isotopes in leaves of several woody legume and non-legume species, and to better understand the consequence of such variations on nitrogen fixation and carbon assimilation. In the first study area (non-saline soil), the δ15N values in four legume species (Acacia cyanophylla,?1.73 ‰ Acacia farnesiana,?0.55 ‰ Prosopis juliflora,?1.64 ‰; and Medicago arborea,+1.6 \textperthousand) and one actinorhizal plant (Elaeagnus angustifolia,?0.46 to?2.1 ‰) were found to be close to that of the atmospheric value pointing to a major contribution of N2 fixing in these species; whereas, δ15N values of the non-fixing plant species were highly positive. δ13C ‰; in leaves of the C3 plants were found to be affected by plant species, ranging from a minimum of?28.67 ‰; to a maximum of?23 ‰. However, they were relatively similar within each plant species although they were grown at different sites. In the second study area (salt affected soil), a higher carbon discrimination value (Δ13C ‰) was exhibited by P. juliflora, indicating that the latter is a salt tolerant species; however, its δ15N was highly positive (+7.03 ‰) suggesting a negligible contribution of the fixed N2. Hence, it was concluded that the enhancement of N2 fixation might be achieved by selection of salt-tolerant Rhizobium strains.  相似文献   

13.
In order to investigate fractionation of calcium (Ca) isotopes in vertebrates as a diagnostic tool to detect Ca metabolism dysfunction we analyzed the Ca isotopic composition (δ44/40Ca?=?[(44Ca/40Ca)sample/(44Ca/40Ca)reference]?1) of diet, faeces, blood, bones and urine from Göttingen minipigs, an animal model for human physiology. Samples of three groups were investigated: 1. control group (Con), 2. group with glucocorticosteroid induced osteoporosis (GIO) and 3. group with Ca and vitamin D deficiency induced osteomalacia (?CaD). In contrast to Con and GIO whose average δ44/40Cafaeces values (0.39?±?0.13‰ and 0.28?±?0.08‰, respectively) tend to be lower than their diet (0.47?±?0.02‰), δ44/40Cafaeces of ?CaD (?0.27?±?0.21‰) was significantly lower than their δ44/40Cadiet (0.37?±?0.03‰), but also lower than δ44/40Cafaeces of Con and GIO. We suggest that the low δ44/40Cafaeces of ?CaD might be due to the contribution of isotopically light Ca from gastrointestinal fluids during gut passage. Assuming that this endogenous Ca source is a common physiologic feature, a fractionation during Ca absorption is also required for explaining δ44/40Cafaeces of Con and GIO. The δ44/40Caurine of all groups are high (>2.0‰) reflecting preferential renal reabsorption of light Ca isotopes. In Göttingen minipigs we found a Ca isotope fractionation between blood and bones (Δ44/40Cablood-bone) of 0.68?±?0.15‰.  相似文献   

14.
Recovery of gold from arsenopyrite-hosted ore in the Giant Mine camp, Yellowknife, NWT, Canada, has left a legacy of arsenic contamination that poses challenges for mine closure planning. Seepage from underground chambers storing some 237,000 tonnes of arsenic trioxide dust, has As concentrations exceeding 4000 ppm. Other potential sources and sinks of As also exist. Sources and movement of water and arsenic are traced using the isotopes of water and sulphate. Mine waters (16 ppm As; AsV/AsIII ≈ 150) are a mixture of two principal water sources – locally recharged, low As groundwaters (0.5 ppm As) and Great Slave Lake (GSL; 0.004 ppm As) water, formerly used in ore processing and discharged to the northwest tailings impoundment (NWTP). Mass balance with δ18O shows that recirculation of NWTP water to the underground through faults and unsealed drillholes contributes about 60% of the mine water. ;[emsp]>Sulphate serves to trace direct infiltration to the As2O3 chambers. Sulphate in local, low As groundwaters (0.3–0.6 ppm As; δ34SSO4  ~ 4 ‰ and δ18OSO4  ~ ? 10 ‰) originates from low-temperature aqueous oxidation of sulphide-rich waste rock. The high As waters gain a component of 18O-enriched sulphate derived from roaster gases (δ18OSO4  = + 3.5 ‰), consistent with their arsenic source from the As2O3 chambers. High arsenic in NWTP water (~ 8 ppm As; δ18OSO4  = ? 2 ‰) derived from mine water, is attenuated to close to 1 ppm during infiltration back to the underground, probably by oxidation and sorption by ferrihydrite.  相似文献   

15.
ABSTRACT

Karst springs in the Main Range of the Crimean Mountains and the Crimean Piedmont show a restricted range of values (δ18O?=?–10.5 to –8.0 ‰, δ2H?=?–72 to –58 ‰), somewhat more negative than the weighted mean of meteoric precipitation. This suggests preferential recharge at higher elevations during winter months. Groundwater tapped by boreholes splits in three groups. A first group has isotopic properties similar to those of the springs. The second group shows significantly lower values (δ18O?=?–13.3 to –12.0 ‰, δ2H?=?–95 to –82 ‰), suggesting recharge during colder Pleistocene times. The third group has high isotope values (δ18O?=?–2.5 to +1.0 ‰, δ2H?=?–24 to –22 ‰); the data points are shifted to the right of the Local Meteoric Water Line, suggesting water–rock exchange processes in the aquifer. These boreholes are located in the Crimean Plains and discharge mineralized (ca. 25 g L?1) thermal (65°C) water from a depth of 1600–1800 m. Groundwater associated with mud volcanoes on the Kerch peninsula have distinct isotope characteristics (δ18O?=?–1.6 to +9.4 ‰, δ2H?=?–30 to –18 ‰). Restricted δ2H variability along with variable and high δ18O values suggest water–rock interactions at temperatures exceeding 95 °C.  相似文献   

16.
Abstract

Much uncertainty still exists regarding spatial and temporal variability of stable isotope ratios (13C/12C and D/H) in different CH4-emission sources. Such variability is especially prevalent in freshwater wetlands where a range of processes can influence stable isotope compositions, resulting in variations of up to ~50‰ for δ13C-CH4 and ~150‰ for δD-CH4 values. Within a temperate-zone bog and marsh situated in southwestern Ontario, Canada, gas bubbles in pond sediments exhibit only minor seasonal and spatial variation in δ13C-CH4, δD-CH4 and δ13C-CO2 values. In pond sediments, CO2 appears to be the main source of carbon during methanogenesis either directly via CO2 reduction or indirectly through dissimilation of autotrophic acetate. In contrast, CH4 production occurs primarily via acetate fermentation at shallow depths in peat soils adjacent to ponds at each wetland. At greater depths within soils, σCO2 and H2O increasingly exert an influence on δ13C- and δD-CH4 values. Secondary alteration processes (e.g., methanotrophy or diffusive transport) are unlikely to be responsible for depth-related changes in stable isotope values of CH4. Recent models that attempt to predict δD-CH4 values in freshwater environments from D/H ratios in local precipitation do not adequately account for such changes with depth. Subenvironments (i.e., soil-forming and open water areas) in wetlands should be considered separately with respect to stable isotope signatures in CH4 emission models.  相似文献   

17.
The coupling between cavity ring-down spectroscopy (CRDS) and an environmental chamber in the investigation of photo-induced reaction mechanisms is demonstrated for the first time. The development of the CRDS device and the corresponding analytical performances are presented. The first application is devoted to the investigation of the branching ratio of the ?OH radical reaction of CH3C(O)OH and CH3C(O)OD under tropospheric conditions. An environmental chamber coupled to two complementary detection systems is used:
  • gas chromatography with FTIR spectroscopy for quantitative detection of acetic acid;
  • CRDS for quantitative detection of CO2.
  • Investigation of the reaction kinetics of ?OH+CH3C(O)OH gives a rate constant of (6.5±0.5)×10-13 cm3?molecule-1?s-1 (296 K) and shows good agreement with literature data. The product study indicates that the H-abstraction channel from the acid group is the dominant pathway with a branching ratio of (78±13)%, whereas the corresponding D-abstraction channel in the ?OH+CH3C(O)OD reaction represents only (36±7)%. This result could be attributed to a strong kinetic isotope effect. Glyoxylic acid has also been detected for the first time as by-product. These results illustrate the high interest of the CRDS technique in the investigation of atmospheric relevant problems.  相似文献   

    18.
    Variations in the relative isotopic abundance of C and N (δ13C and δ15N) were measured during the composting of different agricultural wastes using bench-scale bioreactors. Different mixtures of agricultural wastes (horse bedding manure?+?legume residues; dairy manure?+?jatropha mill cake; dairy manure?+?sugarcane residues; dairy manure alone) were used for aerobic–thermophilic composting. No significant differences were found between the δ13C values of the feedstock and the final compost, except for dairy manure?+?sugarcane residues (from initial ratio of ?13.6?±?0.2?‰ to final ratio of ?14.4?±?0.2?‰). δ15N values increased significantly in composts of horse bedding manure?+?legumes residues (from initial ratio of +5.9?±?0.1?‰ to final ratio of +8.2?±?0.5?‰) and dairy manure?+?jatropha mill cake (from initial ratio of +9.5?±?0.2?‰ to final ratio of +12.8?±?0.7?‰) and was related to the total N loss (mass balance). δ13C can be used to differentiate composts from different feedstock (e.g. C3 or C4 sources). The quantitative relationship between N loss and δ15N variation should be determined.  相似文献   

    19.
    A facile two-step approach has been used for the synthesis of porous SnO2 rods: the initial room-temperature precipitation of precursor SnC2O4 and its subsequent thermal decomposition at 550 °C. Both the as-obtained porous SnO2 microrods (length ~10.0?±?3.5 μm, diameter ~1.1?±?0.4 μm) and submicrorods (length ~5.8?±?1.9 μm, diameter ~0.4?±?0.1 μm) are the crystalline mixtures of major tetragonal and minor orthorhombic crystal phases, showing a tetragonal fraction of 84.7 and 87.0 %, respectively. When applied as a lithium-ion battery anode, the porous submicrorods (specific surface area ~13.6 m2 g?1) can deliver an initial discharge capacity of 1,730.7 mAh g?1 with a high coulombic efficiency of 61.6 % and show the 50th discharge capacity of 662.8 mAh g?1 at 160 mA g?1 within a narrow potential range of 10.0 mV to 2.0 V. Similarly, even the anode of porous microrods (specific surface area ~11.8 m2 g?1) can still exhibit an initial discharge capacity of 1,661.1 mAh g?1 at 160 mA g?1 with a coulombic efficiency of 60.9 %. Regardless of the polymorphic nature, the acquired porosity may only alleviate the huge volume change of anodes for the first cycle; thus, the structural parameters of average size and specific surface area can be feasibly associated with the enhanced lithium storage capability. Anyway, these indicate a facile oxalate precursor method for the controlling synthesis and high performance of rodlike SnO2 for lithium-ion batteries.  相似文献   

    20.
    Abstract

    Carbon and nitrogen stable isotope compositions of organic matter, TOC/TN ratio, and manganese concentration in a sediment core that was collected in northern part of Lake Baikal (VER92ST10-GC2, water depth at 922 m, about 3 m long) were investigated to elucidate the origin of the sedimentary organic matter and its associated environmental factors.

    The sediment core was composed of mainly two parts: turbidite sections and other sections. Constant δ13C and δ15N values of the turbidite sections were observed (- 26.8 ±0.2 ‰ for δ13C and 3.2 ± 0.1 ‰ for δ15N) throughout the core. The higher δ13C in turbidite sections (about - 27 ‰) than that of the other sections (- 31 to - 29 ‰) was clearly observed, and δ15N was different between turbidite sections (about 3‰) and other sections (3 to 5 ‰). δ13C of other sections was close to that of pelagic phytoplankton, indicating that sediment other than turbidite sections is composed of autochthonous components. The variation of stable isotopes in other sections may be possibly caused by the changes in either phytoplankton growth rate or contribution ratios of terrestrial to aquatic plants for δ13C. Either denitrification or fluctuation of δ15N in pelagic phytoplankton can be the cause of variable δ15N in other sections.  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