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.
Stable hydrogen isotope ratio measurements of specific plant components are increasingly used in numerous fields of research, including sample origin verification and climate research. A recently suggested method with considerable potential in this context is the D/H isotope ratio (δ2H value) analysis of plant matter methoxyl groups. The method entails ether or ester cleavage with hydriodic acid (HI) to form the gaseous compound methyl iodide (CH3I) and measurement of the δ2H value of this gas. Here we describe a method for the rapid and precise δ2H analysis of plant matter methoxyl groups using gas chromatography/pyrolysis/isotope ratio mass spectrometry (GC/P/IRMS). The conditions for sample preparation were investigated for isotope discrimination effects, the GC conditions were optimized, the reproducibility of the measurement of standards was studied, and the precision of the method was defined. The reproducibility of δ2H values determined for a CH3I standard on 20 consecutive measurements was found to be 2‰. The method was also tested on four methoxyl‐rich plant components: vanillin, lignin, wood and pectin. The analytical precision obtained, when expressed as the average standard deviation for these compounds, was better than 1.6‰. The described method is rapid, allowing preparation and analysis of a sample within 1 h, and produces accurate and reproducible isotopic measurements. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

2.
We have developed a rapid, sensitive, and automated analytical system to simultaneously determine the concentrations and stable isotopic compositions (δ15N, δ18O, and δ13C) of nanomolar quantities of nitrous oxide (N2O) and methane (CH4) in water, by combining continuous‐flow isotope‐ratio mass spectrometry and a helium‐sparging system to extract and purify the dissolved gases. Our system, which is composed of cold traps and a capillary gas chromatograph that use ultra‐pure helium as the carrier gas, achieves complete extraction of N2O and CH4 in a water sample and separation among N2O, CH4, and the other component gases. The flow path following exit from the gas chromatograph was periodically changed to pass the gases through the combustion furnace to convert CH4 and the other hydrocarbons into CO2, or to bypass the combustion furnace for the direct introduction of eluted N2O into the mass spectrometer, for determining the stable isotopic compositions through monitoring the ions of m/z 44, 45, and 46 of CO and N2O+. The analytical system can be operated automatically with sequential software programmed on a personal computer. Analytical precisions better than 0.2‰ and 0.3‰ and better than 1.4‰ and 2.6‰ were obtained for the δ15N and δ18O of N2O, respectively, when more than 6.7 nmol and 0.2 nmol of N2O, respectively, were injected. Simultaneously, analytical precisions better than 0.07‰ and 2.1‰ were obtained for the δ13C of CH4 when more than 5.5 nmol and 0.02 nmol of CH4, respectively, were injected. In this manner, we can simultaneously determine stable isotopic compositions of a 120 mL water sample with concentrations as low as 1.7 nmol/kg for N2O and 0.2 nmol/kg for CH4. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

3.
In order to generate a reliable and long‐lasting stable isotope ratio standard for CO2 in samples of clean air, CO2 is liberated from well‐characterized carbonate material and mixed with CO2‐free air. For this purpose a dedicated acid reaction and air mixing system (ARAMIS) was designed. In the system, CO2 is generated by a conventional acid digestion of powdered carbonate. Evolved CO2 gas is mixed and equilibrated with a prefabricated gas comprised of N2, O2, Ar, and N2O at close to ambient air concentrations. Distribution into glass flasks is made stepwise in a highly controlled fashion. The isotopic composition, established on automated extraction/measurement systems, varied within very small margins of error appropriate for high‐precision air‐CO2 work (about ±0.015‰ for δ13C and ±0.025‰ for δ18O). To establish a valid δ18O relation to the VPDB scale, the temperature dependence of the reaction between 25 and 47°C has been determined with a high level of precision. Using identical procedures, CO2‐in‐air mixtures were generated from a selection of reference materials; (1) the material defining the VPDB isotope scale (NBS 19, δ13C = +1.95‰ and δ18O = ?2.2‰ exactly); (2) a local calcite similar in isotopic composition to NBS 19 (‘MAR‐J1’, δ13C = +1.97‰ and δ18O = ?2.02‰), and (3) a natural calcite with isotopic compositions closer to atmospheric values (‘OMC‐J1’, δ13C = ?4.24‰ and δ18O = ?8.71‰). To quantitatively control the extent of isotope‐scale contraction in the system during mass spectrometric measurement other available international and local carbonate reference materials (L‐SVEC, IAEA‐CO‐1, IAEA‐CO‐8, CAL‐1 and CAL‐2) were also processed. As a further control pure CO2 reference gases (Narcis I and II, NIST‐RM 8563, GS19 and GS20) were mixed with CO2‐free synthetic air. Independently, the pure CO2 gases were measured on the dual inlet systems of the same mass spectrometers. The isotopic record of a large number of independent batches prepared over the course of several months is presented. In addition, the relationship with other implementations of the VPDB‐scale for CO2‐in‐air (e.g. CG‐99, based on calibration of pure CO2 gas) has been carefully established. The systematic high‐precision comparison of secondary carbonate and CO2 reference materials covering a wide range in isotopic composition revealed that assigned δ‐values may be (slightly) in error. Measurements in this work deviate systematically from assigned values, roughly scaling with isotopic distance from NBS 19. This finding indicates that a scale contraction effect could have biased the consensus results. The observation also underlines the importance of cross‐contamination errors for high‐precision isotope ratio measurements. As a result of the experiments, a new standard reference material (SRM), which consists of two 5‐L glass flasks containing air at 1.6 bar and the CO2 evolved from two different carbonate materials, is available for distribution. These ‘J‐RAS’ SRM flasks (‘Jena‐Reference Air Set’) are designed to serve as a high‐precision link to VPDB for improving inter‐laboratory comparability. a Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

4.
Accurate determinations of stable isotope ratios require a calibration using at least two reference materials with different isotopic compositions to anchor the isotopic scale and compensate for differences in machine slope. Ideally, the δ values of these reference materials should bracket the isotopic range of samples with unknown δ values. While the practice of analyzing two isotopically distinct reference materials is common for water (VSMOW‐SLAP) and carbonates (NBS 19 and L‐SVEC), the lack of widely available organic reference materials with distinct isotopic composition has hindered the practice when analyzing organic materials by elemental analysis/isotope ratio mass spectrometry (EA‐IRMS). At present only L‐glutamic acids USGS40 and USGS41 satisfy these requirements for δ13C and δ15N, with the limitation that L‐glutamic acid is not suitable for analysis by gas chromatography (GC). We describe the development and quality testing of (i) four nicotine laboratory reference materials for on‐line (i.e. continuous flow) hydrogen reductive gas chromatography‐isotope ratio mass‐spectrometry (GC‐IRMS), (ii) five nicotines for oxidative C, N gas chromatography‐combustion‐isotope ratio mass‐spectrometry (GC‐C‐IRMS, or GC‐IRMS), and (iii) also three acetanilide and three urea reference materials for on‐line oxidative EA‐IRMS for C and N. Isotopic off‐line calibration against international stable isotope measurement standards at Indiana University adhered to the ‘principle of identical treatment’. The new reference materials cover the following isotopic ranges: δ2Hnicotine ?162 to ?45‰, δ13Cnicotine ?30.05 to +7.72‰, δ15Nnicotine ?6.03 to +33.62‰; δ15Nacetanilide +1.18 to +40.57‰; δ13Curea ?34.13 to +11.71‰, δ15Nurea +0.26 to +40.61‰ (recommended δ values refer to calibration with NBS 19, L‐SVEC, IAEA‐N‐1, and IAEA‐N‐2). Nicotines fill a gap as the first organic nitrogen stable isotope reference materials for GC‐IRMS that are available with different δ15N values. Comparative δ13C and δ15N on‐line EA‐IRMS data from 14 volunteering laboratories document the usefulness and reliability of acetanilides and ureas as EA‐IRMS reference materials. Published in 2009 by John Wiley & Sons, Ltd.  相似文献   

5.
Stable isotope ratios of individual plant components have become a valuable tool for the determination of the geographical origin and authenticity of foodstuff. A recently published method with considerable potential in this context is the measurement of the deuterium/hydrogen (D/H) isotope ratios of plant matter methoxyl groups. The method entailed cleavage of methyl ethers or esters with hydriodic acid (HI) to form gaseous methyl iodide (CH3I) and then measurement of the δ2H value of this gas. Here, as a follow up to a previous study, we describe a method for the rapid and precise δ13C analysis of plant matter methoxyl groups using gas chromatography/combustion/isotope ratio mass spectrometry (GC/C/IRMS). Conditions for sample preparation were investigated for isotope discrimination effects, the GC conditions optimized, the reproducibility of the measurement of standards undertaken, and the precision of the method defined. The reproducibility of the δ13C value determined for a CH3I standard on 20 consecutive measurements was found to be 0.17‰. The method was also tested on four methoxyl‐rich plant components: vanillin, lignin, wood and pectin. The analytical precision obtained, expressed as the average standard deviation, for these compounds was found to be better than 0.13‰. The described procedure which is simple and rapid, allowing preparation and analysis of a sample within 1 h, produces accurate and reproducible isotopic measurements. We suggest that this validated δ13C method when employed together with the recently published δ2H method for two‐dimensional stable isotope studies of organic matter containing methoxyl groups will be of considerable value, e.g. for proving the authenticity of foodstuff. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

6.
Developments in continuous‐flow isotope ratio mass spectrometry have made possible the rapid analysis of δ13C in CO2 of small‐volume gas samples with precisions of ≤0.1‰. Prior research has validated the integrity of septum‐capped vials for collection and short‐term storage of gas samples. However, there has been little investigation into the sources of contamination during the preparation and analysis of low‐concentration gas samples. In this study we determined (1) sources of contamination on a Gasbench II, (2) developed an analytical procedure to reduce contamination, and (3) identified an efficient, precise method for introducing sample gas into vials. We investigated three vial‐filling procedures: (1) automated flush‐fill (AFF), (2) vacuum back‐fill (VBF), and (3) hand‐fill (HF). Treatments were evaluated based on the time required for preparation, observed contamination, and multi‐vial precision. The worst‐case observed contamination was 4.5% of sample volume. Our empirical estimate showed that this level of contamination results in an error of 1.7‰ for samples with near‐ambient CO2 concentrations and isotopic values that followed a high‐concentration carbonate reference with an isotope ratio of ?47‰ (IAEA‐CO‐9). This carry‐over contamination on the Gasbench can be reduced by placing a helium‐filled vial between the standard and the succeeding sample or by ignoring the first two of five sample peaks generated by each analysis. High‐precision (SD ≤0.1‰) results with no detectable room‐air contamination were observed for AFF and VBF treatments. In contrast, the precision of HF treatments was lower (SD ≥0.2‰). VBF was optimal for the preparation of gas samples, as it yielded faster throughput at similar precision to AFF. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

7.
We present an advanced system for on‐line position‐specific carbon isotope analysis. The main limitation of on‐line intramolecular isotope ratio measurements has been that optimal pyrolytic fragments are obtained mostly at temperatures where the analyte has not completely reacted. As a result of undetermined isotopic fractionation, the isotopic signatures of the pyrolysis products are not strictly equal to these of the equivalent moieties in the parent molecule. We designed a pyrolytic unit in which both temperature and reaction time are variable parameters, enabling determination of the enrichment factor of the pyrolysis at optimal temperature by construction of a Rayleigh plot. In the case of methyl tert‐butyl ether (MTBE) presented here, a ‘pre‐pyrolysis’ fractionation of MTBE leading to a depletion of 0.9‰ was discovered and the enrichment factor of the optimal pyrolysis reaction was determined at −1.7‰. Absolute δ13C values of two functional groups of MTBE – the methoxy group and the 2‐methylpropane group – could be determined with 95% confidence intervals of 0.4‰ and 0.5‰, respectively. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

8.
Gaseous membrane permeation (MP) technologies have been combined with continuous‐flow isotope ratio mass spectrometry for on‐line δ13C measurements. The experimental setup of membrane permeation‐gas chromatography/combustion/isotope ratio mass spectrometry (MP‐GC/C/IRMS) quantitatively traps gas streams in membrane permeation experiments under steady‐state conditions and performs on‐line gas transfer into a GC/C/IRMS system. A commercial polydimethylsiloxane (PDMS) membrane sheet was used for the experiments. Laboratory tests using CO2 demonstrate that the whole process does not fractionate the C isotopes of CO2. Moreover, the δ13C values of CO2 permeated on‐line give the same isotopic results as off‐line static dual‐inlet IRMS δ13C measurements. Formaldehyde generated from aqueous formaldehyde solutions has also been used as the feed gas for permeation experiments and on‐line δ13C determination. The feed‐formaldehyde δ13C value was pre‐determined by sampling the headspace of the thermostated aqueous formaldehyde solution. Comparison of the results obtained by headspace with those from direct aqueous formaldehyde injection confirms that the headspace sampling does not generate isotopic fractionation, but the permeated formaldehyde analyzed on‐line yields a 13C enrichment relative to the feed δ13C value, the isotopic fractionation being 1.0026 ± 0.0003. The δ13C values have been normalized using an adapted two‐point isotopic calibration for δ13C values ranging from ?42 to ?10‰. The MP‐GC/C/IRMS system allows the δ13C determination of formaldehyde without chemical derivatization or additional analytical imprecision. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

9.
Isotope ratio monitoring gas chromatography‐mass spectrometry of the 2H/1H ratio by pyrolysis isotope ratio mass spectrometry (P‐IRMS) was used to analyze benzaldehyde originating from various sources. Based on the δ2HSMOW value of an authentic reference sample determined with an elemental analyzer (EA), the range of reproducibility and linearity was checked. Correct (EA related) and reproducible data were obtained for sample amounts >0.6 μg benzaldehyde (on column). In another series of experiments, the influence of sample preparation, i. e. simultaneous distillation‐extraction (SDE) was found to be negligible. The following ranges of δ2HSMOW values were determined for benzaldehyde using five types of samples, i. e. (i) synthetic (δ2HSMOW –78 to –85‰, ex benzal chloride; +420 to +668‰, ex toluene) and ‘natural’ (including ‘ex‐cassia’) references (δ2HSMOW –83 to –144‰); (ii) bitter almond oils (δ2HSMOW –113 to –148‰); (iii) fruits (δ2HSMOW –111 to –146‰); (iv) kernels (δ2HSMOW –115 to –188‰); and (v) leaves (δ2HSMOW –165 to –189‰).  相似文献   

10.
An analytical line for stable isotope analyses of water recovered from fluid inclusions in minerals was built and successfully tested. The line is based on the principle of continuous‐flow analysis of water via high‐temperature reduction on glassy carbon. It includes a custom‐designed set of high‐efficiency crushers and a cryo‐focusing cell. This paper provides details of the line design and discusses strategies for line conditioning and mitigation of memory effects. The line allows measurements of hydrogen and oxygen isotopes during a single acquisition. The precision of the analyses depends on the amount of water released from the inclusions. The best results are obtained for samples containing at least 0.1–0.2 µL (0.06–0.11 µmol) H2O. For such samples precision is better than 1.5‰ for δD and 0.5‰ for δ18O (1σ). Smaller amounts of water can be measured but at lower precision. Analyses of modern calcite formed under stable conditions in a deep cave allowed assessment of the accuracy of the analyses. The δD values measured in fluid inclusions of this working standard match the δD value of the parent water, and the oxygen isotope values agree within ca. 0.5‰. This indicates that fluid inclusions trapped in calcite at near‐ambient temperatures (e.g. speleothems and low‐temperatures phreatic calcite) faithfully preserve the original isotopic composition of the parent waters. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

11.
Observations made during the 13C isotope analysis of gaseous CO2 in the simultaneous presence of argon in the ion source of the isotope ratio mass spectrometer prompted us to investigate what influence the simultaneous presence of nitrogen would have on both accuracy and precision of bulk 2H isotope analysis of nitrogen‐rich organic compounds. Initially an international reference material, IAEA‐CH7, was mixed with silver nitrate in various ratios to assess the impact that N2 evolved from the pyrolysis of nitrogen‐rich organic compounds would have on measured δ2H‐values of IAEA‐CH7. In a subsequent experiment, benzoic acid was mixed with silver nitrate to mimic the N:H ratio of organic‐rich nitrogen compounds such as cellulose nitrate and RDX. The results of both experiments showed a significant deterioration of both accuracy and precision for the expected δ2H values for IAEA‐CH7 and benzoic acid when model mixtures were converted into hydrogen and nitrogen, and subsequently separated by gas chromatography using standard experimental conditions, namely a 60 cm packed column with molecular sieve 5 Å as stationary phase held at a temperature of 85°C. It was found that bulk 2H stable isotope analysis of nitrogen‐rich organic compounds employing published standard conditions can result in a loss of accuracy and precision yielding δ2H values that are 5 to 25‰ too negative, thus suggesting, for example, that tree‐ring 2H isotope data based on cellulose nitrate may have to be revised. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

12.
Isoprene is one of the most important non‐methane hydrocarbons (NMHCs) in the troposphere: it is a significant precursor of O3 and it affects the oxidative state of the atmosphere. The diastereoisomeric 2‐methyltetrols, 2‐methylthreitol and 2‐methylerythritol, are marker compounds of the photooxidation products of atmospheric isoprene. In order to obtain valuable information on the δ13C value of isoprene in the atmosphere, the stable carbon isotopic compositions of the 2‐methyltetrols in ambient aerosols were investigated. The 2‐methyltetrols were extracted from filter samples and derivatized with methylboronic acid, and the δ13C values of the methylboronate derivatives were determined by gas chromatography/combustion/isotope ratio mass spectrometry (GC/C/IRMS). The δ13C values of the 2‐methyltetrols were then calculated through a simple mass balance equation between the 2‐methyltetrols, methylboronic acid and the methylboronates. The δ13C values of the 2‐methyltetrols in aerosol samples collected at the Changbai Mountain Nature Reserves in eastern China were found to be ?24.66 ± 0.90‰ and ?24.53 ± 1.08‰ for 2‐methylerythritol and 2‐methylthreitol, respectively. Based on the measured isotopic composition of the 2‐methyltetrols, the average δ13C value of atmospheric isoprene is inferred to be close to or slightly heavier than ?24.66‰ at the collection site during the sampling period. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

13.

Rationale

The fundamental level of stable isotopic knowledge lies at specific atomic positions within molecules but existing methods of analysis require lengthy off‐line preparation to reveal this information. An automated position‐specific isotope analysis (PSIA) method is presented to determine the stable carbon isotopic compositions of the carboxyl groups of amino acids (δ13CCARBOXYL values). This automation makes PSIA measurements easier and routine.

Methods

An existing high‐performance liquid chromatography (HPLC) gas handling interface/stable isotope ratio mass spectrometry system was modified by the addition of a post‐column derivatisation unit between the HPLC system and the interface. The post‐column reaction was optimised to yield CO2 from the carboxyl groups of amino acids by reaction with ninhydrin.

Results

The methodology described produced δ13CCARBOXYL values with typical standard deviations below ±0.1 ‰ and consistent differences (Δ13CCARBOXYL values) between amino acids over a 1‐year period. First estimates are presented for the δ13CCARBOXYL values of a number of internationally available amino acid reference materials.

Conclusions

The PSIA methodology described provides a further dimension to the stable isotopic characterisation of amino acids at a more detailed level than the bulk or averaged whole‐molecule level. When combined with on‐line chromatographic separation or off‐line fraction collection of protein hydrolysates the technique will offer an automated and routine way to study position‐specific carboxyl carbon isotope information for amino acids, enabling more refined isotopic studies of carbon uptake and metabolism.
  相似文献   

14.
The bromine isotope composition is potentially diagnostic in both degradation monitoring and source apportionment of organobromines in the environment. A method for compound‐specific bromine isotope analysis (δ81Br) based on gas chromatography multiple collector inductively coupled plasma mass spectrometry (GC/ICPMS) was developed for common brominated diaromatic compounds. Brominated diphenyl ethers (BDEs) in Bromkal 70‐5DE, a technical flame‐retardant mixture containing mainly BDEs #47, #99 and #100, were used as test substances, with standard bracketing for the samples achieved through co‐injected monobromobenzene (MBB) with a known δ81Br of ?0.39‰ vs. Standard Mean Ocean Bromine (SMOBr). Three different heated transfer lines were constructed and tested to achieve efficient conduction of the BDEs from the gas chromatograph to the ICPMS instrument. The MBB was analyzed with a precision of 0.4‰ (1 s, n = 18). The precision for BDEs was 1.4–1.8‰ (1 s, n = 10–12 depending on the congener). The lower precision for the BDEs than for MBB may reflect the heat required to prevent condensation of the analytes in ICP torch assembly. The use of an internal standard of similar chemical structure to the analytes alleviates this problem, as illustrated by a difference of 0.3 ± 0.7‰ (1 s, n = 6) between the δ81Br values of co‐injected methoxy BDE‐47 and BDE‐47 extracted from whale blubber. Improvements in precision and accuracy may be achieved by the use of a more efficient heating of the torch assembly in conjunction with a set of internal standards that match the target compounds. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

15.
A method for online simultaneous δ2H and δ18O analysis in water by high‐temperature conversion is presented. Water is injected by using a syringe into a high‐temperature carbon reactor and converted into H2 and CO, which are separated by gas chromatography (GC) and carried by helium to the isotope ratio mass spectrometer for hydrogen and oxygen isotope analysis. A series of experiments was conducted to evaluate several issues such as sample size, temperature and memory effects. The δ2H and δ18O values in multiple water standards changed consistently as the reactor temperature increased from 1150 to 1480°C. The δ18O in water can be measured at a lower temperature (e.g. 1150°C) although the precision was relatively poor at temperatures <1300°C. Memory effects exist for δ2H and δ18O between two waters, and can be reduced (to <1%) with proper measures. The injection of different amounts of water may affect the isotope ratio results. For example, in contrast to small injections (100 nL or less) from small syringes (e.g. 1.2 µL), large injections (1 µL or more) from larger syringes (e.g. 10 µL) with dilution produced asymmetric peaks and shifts of isotope ratios, e.g. 4‰ for δ2H and 0.4‰ for δ18O, probably resulting from isotope fractionation during dilution via the ConFlo interface. This method can be used to analyze nanoliter samples of water (e.g. 30 nL) with good precision of 0.5‰ for δ2H and 0.1‰ for δ18O. This is important for geosciences; for instance, fluid inclusions in ancient minerals may be analyzed for δ2H and δ18O to help understand the formation environments. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

16.
The most precise method of chlorine isotope analysis described to date is based on the isotope ratio mass spectrometry (IRMS) of chlorine quantitatively converted into chloromethane, CH3Cl. This gas can be produced from several chlorine‐containing compounds and analyzed by IRMS. However, the mass spectrum of chloromethane is rather complicated and the ratio of the most abundant ions (mass‐52/mass‐50) differs from the 37Cl/35Cl isotope ratio. This difference becomes significant when the δ exceeds 10‰. Moreover, the electron ionization source yields approximately 80% of all the ionic species at the useful masses 50 and 52. To overcome these drawbacks, we have devised a negative ion mass spectrometer which retains all the best features of IRMS, including a dual‐inlet system with changeover valve, dual collector assembly and CH3Cl gas as analyte. In the modified ion source we have replaced the ionization chamber with an electron beam by a metal tube with a hot metal filament inside it. Within this tube the 35Cl? and 37Cl? ions are produced with an efficiency dependent on the filament material and its temperature. No other ionic species were found in the mass spectrum except of traces at masses 26 and 28 at ppm levels, probably due to the formation of CN? and CO?. The minimal amount of Cl used in our method is of the order of 5 µmol (3 mg AgCl) and the precision is better than 0.005‰ (1σ). Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

17.
Amino sugars have been used as biomarkers to assess the relative contribution of dead microbial biomass of different functional groups of microorganisms to soil carbon pools. However, little is known about the dynamics of these compounds in soil. The isotopic composition of individual amino sugars can be used as a tool to determine the turnover of these compounds. Methods to determine the δ13C of amino sugars using gas chromatography/combustion/isotope ratio mass spectrometry (GC/C/IRMS) have been proposed in literature. However, due to derivatization, the uncertainty on the obtained δ13C is too high to be used for natural abundance studies. Therefore, a new high‐performance liquid chromatography/isotope ratio mass spectrometry (HPLC/IRMS) methodology, with increased accuracy and precision, has been developed. The repeatability on the obtained δ13C values when pure amino sugars were analyzed were not significantly concentration‐dependent as long as the injected amount was higher than 1.5 nmol. The δ13C value of the same amino sugar spiked to a soil deviated by only 0.3‰ from the theoretical value. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

18.
We present a method for in situ sulphur (S) isotopic analysis of significantly small areas (1.5 µm in diameter) in pyrite using secondary ion mass spectrometry (NanoSIMS) to interpret microbial sulphur metabolism in the early earth. We evaluated the precision and accuracy of S isotopic ratios obtained by this method using hydrothermal pyrite samples with homogeneous S isotopic ratios. The internal precision of the δ34S value was 1.5‰ at the level of 1 sigma of standard error (named 1SE) for a single spot, while the external reproducibility was estimated to be 1.6‰ at the level of 1 sigma of standard deviation (named 1SD, n = 25). For each separate sample, the average δ34S value was comparable with that measured by a conventional method, and the accuracy was better than 2.3‰. Consequently, the in situ method is sufficiently accurate and precise to detect the S isotopic variations of small sample of the pyrite (less than 20 µm) that occurs ubiquitously in ancient sedimentary rocks. This method was applied to measure the S isotopic distribution of pyrite within black chert fragments in early Archean sandstone. The pyrite had isotopic zoning with a 34S‐depleted core and 34S‐enriched rim, suggesting isotopic evolution of the source H2S from ?15 to ?5‰. Production of H2S by microbial sulphate reduction (MSR) in a closed system provides a possible explanation for both the 34S‐depleted initial H2S and the progressive increase in the δ34SH2S value. Although more extensive data are necessary to strengthen the explanation for the origin of the MSR, the results show that the S isotopic distribution within pyrite crystals may be a key tracer for MSR activity in the early earth. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

19.
The stable carbon isotope compositions of tetrols, erythritol and threitol were determined by gas chromatography/combustion/isotope ratio mass spectrometry (GC/C/IRMS). Using four tetrols with various δ13C values derivatized by methylboronic acid, the carbon isotope analysis method achieved excellent reproducibility and high accuracy. There was no carbon isotopic fractionation during the derivatization processes. The differences in the carbon isotopic compositions of methylboronates between the measured and calculated ranged from ?0.20 to 0.12‰, within the specification of the GC/C/IRMS system. It was demonstrated that δ13C values of tetrols could be calculated by a simple mass balance equation between tetrols, methylboronic acid, and methylboronates. The analogous 2‐methyltetrols, marker compounds of photooxidation products of atmospheric isoprene, should have similar behavior using the same derivatization reagent. This method may provide insight on sources and sinks of atmospheric isoprene. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

20.
Carbon isotope analysis by bulk elemental analysis coupled with isotope ratio mass spectrometry has been the mainstay of δ13C analyses both at natural abundance and in tracer studies. More recently, compound‐specific isotope analysis (CSIA) has become established, whereby organic constituents are separated online by gas or liquid chromatography before oxidation and analysis of CO2 for constituent δ13C. Theoretically, there should be concordance between bulk δ13C measurements and carbon‐weighted δ13C measurements of carbon‐containing constituents. To test the concordance between the bulk and CSIA, fish oil was chosen because the majority of carbon in fish oil is in the triacylglycerol form and ~95% of this carbon is amenable to CSIA in the form of fatty acids. Bulk isotope analysis was carried out on aliquots of oil extracted from 55 fish samples and δ13C values were obtained. Free fatty acids (FFAs) were produced from the oil samples by saponification and derivatised to fatty acid methyl esters (FAMEs) for CSIA by gas chromatography/combustion/isotope ratio mass spectrometry. A known amount of an internal standard (C15:0 FAME) was added to allow analyte quantitation. This internal standard was also isotopically calibrated in both its FFA (δ13C = ?34.30‰) and FAME (δ13C = ?34.94‰) form. This allowed reporting of FFA δ13C from measured FAME δ13C values. The bulk δ13C was reconstructed from CSIA data based on each FFA δ13C and the relative amount of CO2 produced by each analyte. The measured bulk mean δ13C (SD) was ?23.75‰ (1.57‰) compared with the reconstructed bulk mean δ13C of ?23.76 (1.44‰) from CSIA and was not significantly different. Further analysis of the data by the Bland‐Altman method did not show particular bias in the data relative to the magnitude of the measurement. Good agreement between the methods was observed with the mean difference between methods (range) of 0.01‰ (?1.50 to 1.30). Copyright © 2010 John Wiley & Sons, Ltd.  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