Fact-checked by Grok 2 weeks ago

Stable isotope ratio

Stable isotope ratios refer to the relative abundances of non-radioactive isotopes of a chemical element in a sample, typically expressed as the ratio of the heavier isotope to the lighter one, such as ^{18}O/^{16}O for oxygen or ^{13}C/^{12}C for carbon. These ratios arise from natural variations caused by isotopic fractionation, where physical, chemical, or biological processes preferentially concentrate one isotope over another, and they remain constant over geological timescales because stable isotopes do not decay. The study of stable isotope ratios, primarily for light elements like hydrogen (H), carbon (C), nitrogen (N), oxygen (O), and sulfur (S), provides insights into environmental, geological, and biological processes. In analytical practice, stable isotope ratios are quantified using (IRMS) and reported in the (δ) notation, defined as δ(‰) = [(R_sample / R_standard) - 1] × 1000, where R is the isotope ratio (heavy/light) and the standard is an internationally accepted reference material, such as (VSMOW) for and oxygen or Vienna Pee Dee Belemnite (VPDB) for carbon and oxygen. This per mil (‰) scale allows precise comparisons, with typical measurement precisions of 0.05–0.2‰ for oxygen, carbon, , and , and 0.2–2.0‰ for . mechanisms include processes, driven by thermodynamic differences in bond strengths (e.g., heavier isotopes favor stronger bonds in molecules), and kinetic processes, such as or enzymatic reactions where lighter isotopes react faster. Stable isotope ratio analysis has broad applications across sciences, including (e.g., reconstructing past temperatures from δ^{18}O in cores), geothermometry (e.g., estimating formation temperatures of via oxygen equilibrium), and tracing fluid sources in hydrothermal systems. In mineral resource investigations, ratios like δ^{34}S and δ^{18}O help identify the origins of deposits, such as linking in deposits to seawater sulfate or assessing environmental impacts like . Beyond , these ratios inform biogeochemical cycles, such as distinguishing C3 from C4 plant via δ^{13}C values (-20 to -30‰ versus -13‰, respectively), and forensic applications, including microbial identification through carbon, , oxygen, and signatures.

Fundamentals

Definition and notation

A stable isotope ratio refers to the relative abundance of two or more stable isotopes of the same chemical element within a sample, typically expressed as a comparison to an internationally accepted standard material. This ratio provides a measure of isotopic composition that remains constant over time, as stable isotopes do not undergo . The concept is fundamental in , , and for tracing processes that fractionate isotopes without altering the element's chemical identity. The notation for reporting stable ratios uses the (δ) value, defined in parts per thousand (per , ‰) relative to a reference . For example, the carbon ratio is denoted as δ¹³C and calculated as: \delta^{13}\text{C} = \left( \frac{{^{13}\text{C}/^{12}\text{C}}_{\text{sample}} - {^{13}\text{C}/^{12}\text{C}}_{\text{standard}}}{{^{13}\text{C}/^{12}\text{C}}_{\text{standard}}} \right) \times 1000 \, ‰ This formula applies analogously to other elements, where the heavier is typically in the numerator and the lighter in the denominator, emphasizing small variations in natural abundances. The δ value indicates enrichment (positive) or depletion (negative) of the heavier relative to the ; for instance, most terrestrial carbon samples have δ¹³C values between -30‰ and +10‰. Several light elements commonly exhibit stable isotopes used in ratio analyses, including carbon (¹²C and ¹³C), (¹⁴N and ¹⁵N), oxygen (¹⁶O and ¹⁸O), (¹H and ²H, also denoted as ), and (³²S and ³⁴S). These elements are prioritized due to their involvement in key biogeochemical cycles and the measurable effects on their ratios during physical, chemical, and biological processes. For , the deuterium ratio is often written as δ²H or δD, while oxygen may also include δ¹⁷O in advanced studies. The delta notation was first introduced in 1950 by C. R. McKinney, J. M. McCrea, M. L. Murphy, and H. C. Urey in a on improvements in mass spectrometers for measurements, building on Urey's earlier thermodynamic studies of isotopic substances. This standardization enabled precise comparisons across samples and laid the groundwork for modern stable . Alfred O. C. Nier's concurrent advancements in provided the instrumental precision necessary for implementing this notation.

Stable isotopes in nature

Stable isotopes occur naturally with specific abundances that reflect the elemental composition of the solar system, primarily derived from stellar nucleosynthesis processes. For carbon, the stable isotopes are ¹²C (98.93% abundance) and ¹³C (1.07% abundance), while for oxygen, the stable isotopes include ¹⁶O (99.757%), ¹⁷O (0.038%), and ¹⁸O (0.205%). These abundances represent the average values across solar system materials and are used as standards for expressing isotopic variations. Similar patterns hold for other elements, such as nitrogen with ¹⁴N at 99.632% and ¹⁵N at 0.368%, though abundances vary across the periodic table based on nuclear stability. The primordial isotopic composition of these stable isotopes originated from nucleosynthesis in previous generations of stars, where light elements like carbon and oxygen were synthesized through fusion reactions in massive stars. Carbon-12 forms primarily via the triple-alpha process (3⁴He → ¹²C) in helium-burning cores, while oxygen-16 results from subsequent carbon capture (¹²C + ⁴He → ¹⁶O). Heavier stable isotopes, such as ¹³C and ¹⁸O, arise from secondary processes like carbon-nitrogen-oxygen (CNO) cycles and alpha captures in stellar envelopes, leading to slight variations in ratios due to differences in stellar masses, metallicities, and evolutionary stages. The solar system's bulk composition homogenized these stellar contributions in the interstellar medium, resulting in near-uniform abundances observed in primitive meteorites. On , stable ratios exhibit variations across major reservoirs due to differences in source materials and mixing. Pre-industrially, the δ¹³C value of atmospheric CO₂ was approximately -6.5‰ relative to the Vienna Pee Dee Belemnite (VPDB) standard, reflecting a balance between volcanic inputs and biospheric exchange; however, it has decreased to about -8.5‰ as of 2020 due to emissions of ¹²C-enriched CO₂. Oceanic dissolved inorganic carbon () shows higher values, averaging around +1‰ in deep waters, with surface waters varying from 0‰ to +2‰ due to biological productivity and . These reservoir-specific ratios establish baselines for tracing material transfers in the . Biogenic processes significantly influence baseline isotope ratios by preferentially incorporating lighter isotopes into . During , discriminate against ¹³C, resulting in organic materials enriched in ¹²C, with typical δ¹³C values of -27‰ for C3 and -13‰ for C4 . This depletion extends to soils and sediments, where accumulated exhibits δ¹³C values around -25‰ on average, lowering the overall ¹³C content in the terrestrial compared to inorganic reservoirs. Such biogenic contributes to the lighter isotopic signatures observed in atmospheric and carbon pools over geological timescales.

Isotopic fractionation

Equilibrium fractionation

Equilibrium fractionation refers to the partial separation of stable isotopes between coexisting phases or compounds that have attained thermodynamic equilibrium, primarily due to differences in the zero-point and vibrational energies of molecules containing different isotopes. These energy differences arise from quantum mechanical effects, where heavier isotopes form slightly stronger bonds and occupy lower energy states, leading to preferential partitioning of heavier isotopes into phases with higher bond stiffness or coordination numbers. In closed systems undergoing progressive while maintaining , the evolution of ratios in the residual phase is described by the fractionation model: R = R_0 f^{(\alpha - 1)}, where R is the in the remaining , R_0 is the initial , f is the of the original remaining, and \alpha is the factor between the phases. This model captures how small differences in \alpha amplify isotopic variations as proceeds, assuming constant \alpha and instantaneous removal of the fractionated product. The fractionation factor \alpha depends strongly on temperature, with a theoretical approximation given by \alpha \approx 1 + \frac{\Delta E}{RT^2}, where \Delta E is the energy difference associated with isotopic substitution in bond strengths, R is the gas constant, and T is the absolute temperature; this form derives from the high-temperature limit of the Bigeleisen-Mayer equation, reflecting the dominance of vibrational contributions to the partition function at elevated temperatures. As temperature increases, \alpha approaches unity, reducing isotopic separation. A key example is the oxygen isotope fractionation between dissolved water and precipitating , where heavier ^{18}\mathrm{O} is preferentially incorporated into the solid phase at lower temperatures. The temperature dependence is expressed as $1000 \ln \alpha = 18.03 \left( \frac{10^3}{T} \right) - 31.98, where T is in , illustrating how \alpha decreases with rising temperature, such as from approximately 1.029 at 25°C to near 1.000 at high temperatures.

Kinetic fractionation

Kinetic isotope fractionation refers to the separation of stable isotopes during unidirectional physical or chemical processes, where molecules containing isotopes react or move faster than those with heavier isotopes due to differences in strengths arising from lower zero-point energies in lighter isotopologues. This rate-dependent separation is prominent in open systems or irreversible s, contrasting with equilibrium processes by lacking significant back-reaction. The magnitude of fractionation depends on the specific reaction pathway and the difference between isotopes, with larger effects observed for light elements like , carbon, and . The isotope effect in kinetic fractionation is characterized by the fractionation factor \alpha = \frac{k_{\text{light}}}{k_{\text{heavy}}}, where k represents the rate constant for the light or heavy isotopologue, typically greater than 1 since lighter species react faster. This is often expressed as the enrichment factor \epsilon = (\alpha - 1) \times 1000‰, allowing comparison in per mil (‰) units. Key mechanisms include diffusion, evaporation, and biological uptake, each exploiting mass-dependent differences in molecular velocities or reaction barriers. In , lighter isotopes migrate faster according to , with the diffusion coefficient ratio approximated as \frac{D_1}{D_2} \approx \sqrt{\frac{m_2}{m_1}}, where m denotes ; this leads to isotopic gradients in gases or solutions, such as during the of gases through porous media. exemplifies this in the , where lighter H_2^{16}O molecules evaporate more readily than those containing (D) or ^{18}O, fractionating the D/H ratio in vapor relative to liquid water and producing isotopically lighter atmospheric moisture. Biological uptake, such as in microbial or enzymatic processes, further amplifies kinetic effects through selective incorporation; for instance, photosynthetic enzymes favor ^{12}C over ^{13}C during CO_2 fixation. Representative examples illustrate these effects' scale. In , CO_2 diffusion across leaf stomata imparts a kinetic fractionation of approximately -4‰ to -7‰, depleting intracellular CO_2 in ^{13}C and influencing the \delta^{13}C signature of . Similarly, during bacterial sulfate reduction, sulfate-reducing microbes produce highly enriched in ^{32}S, with \epsilon values reaching up to 40‰ due to rate-limiting steps in sulfur bond breakage. These fractionations provide insights into process dynamics without relying on .

Measurement techniques

Mass spectrometry

Mass spectrometry is the primary technique for precise measurement of stable isotope ratios, such as those of carbon, , oxygen, and , in various sample matrices. The basic principle involves ionizing the sample to produce charged particles, accelerating these , separating them based on their using magnetic or electric fields, and detecting the ion currents to calculate the isotopic s. In (IRMS), generates positive from gaseous samples, which are then accelerated to high velocities and directed into a magnetic sector where the curved path of ions depends on their mass, allowing separation of isotopes like ¹³C from ¹²C. detectors collect the separated ion beams, measuring their currents to determine the with high accuracy, often expressed in (δ) notation relative to a . IRMS systems, particularly those with dual-inlet configurations, enable high-precision comparisons by alternating between sample and reference gases, achieving reproducibilities as low as ±0.1‰ for δ¹³C measurements. In dual-inlet IRMS, beams from the sample and reference are switched rapidly into the , minimizing drift and enhancing signal-to-noise ratios for bulk gas analysis. For metal stable isotopes, such as those of iron or , multi-collector (MC-ICP-MS) is employed, where ionizes liquid samples at , producing ions that are skimmed into a high-vacuum magnetic sector for simultaneous multi-isotope detection. MC-ICP-MS offers precisions of 0.01–0.001% for ratios like ⁵⁶Fe/⁵⁴Fe, making it suitable for trace-level analysis in geochemical samples. Sample preparation is crucial for converting analytes into suitable forms for . For organic materials, in an elemental analyzer converts carbon and to CO₂ and N₂ gases, respectively, which are then introduced into the IRMS via a continuous-flow , typically requiring 0.5–2 of sample for C/N analysis. Gaseous samples, such as for oxygen isotopes, undergo equilibration with a reference gas like CO₂ in the presence of a catalyst to exchange isotopes without altering the overall ratio. In MC-ICP-MS, samples are dissolved in acid and purified via ion-exchange to remove interferences before nebulization into the . Calibration relies on international standards to anchor measurements to defined scales. For carbon, the Vienna Pee Dee Belemnite (VPDB) standard is assigned δ¹³C = 0‰ by convention, with working standards like NBS 19 calibrated relative to it for routine use. Similar scales exist for other elements, such as VSMOW for oxygen, ensuring and comparability across laboratories. Error sources include memory effects, where residual ions from previous samples contaminate subsequent analyses, particularly in continuous-flow systems, potentially shifting ratios by up to 1–2‰ if not corrected through blank runs or conditioning. Other artifacts, like fractionation or detector dead time, are mitigated by instrumental design and software corrections to maintain precision.

Laser-based methods

Laser-based methods for stable isotope ratio analysis rely on optical spectroscopy techniques that exploit differences in vibrational and rotational transitions between isotopes, leading to distinct or emission spectra. These methods measure the of at specific wavelengths corresponding to isotopologues, such as the slight wavelength shift between ¹⁸O and ¹⁶O in or ¹³C and ¹²C in . The principle is grounded in the Beer-Lambert law, where the intensity of transmitted decreases exponentially with path length and concentration of the absorbing species, allowing precise quantification of isotopic abundances without or mass separation. Key techniques include Cavity Ring-Down Spectroscopy (CRDS), which uses a high-finesse optical cavity to extend the effective path length up to kilometers, measuring the decay time of light intensity after laser pulsing to determine gas concentrations with high sensitivity. CRDS achieves precisions of around ±0.2‰ for δ¹³C in CO₂, enabling faster analysis than traditional isotope ratio mass spectrometry (IRMS) while maintaining comparable accuracy for many applications. Another prominent method is Tunable Diode Laser Absorption Spectroscopy (TDLAS), which employs narrow-linewidth diode lasers to scan and probe specific absorption lines, facilitating real-time monitoring of isotopic ratios in gases like CO₂ and H₂O. TDLAS offers precisions around 0.2‰ for δ¹³C in breath CO₂, supporting non-invasive, continuous measurements. These laser-based approaches provide significant advantages, including portability for field deployment without vacuum systems or cryogens, minimal sample preparation—such as direct analysis of exhaled breath for δ¹³C to assess metabolic processes—and high throughput with measurement frequencies up to 20 Hz. Unlike IRMS, they enable spatially resolved and non-destructive analysis, making them ideal for dynamic environments like ecosystems or biomedical settings. Post-2000 developments have enhanced these methods' practicality, notably through Off-Axis Integrated Cavity Output Spectroscopy (OA-ICOS), which improves beam stability and alignment for robust field-deployable systems measuring isotopes with precisions below 0.5‰ (e.g., 0.08‰ for δ¹⁸O). As of 2025, further advancements include mid-infrared TDLAS systems achieving precisions below 0.1‰ for CO₂ isotopes and integration with liquid chromatography for compound-specific analysis, expanding applications in .

Applications

Paleoclimatology and geochemistry

Stable isotope ratios serve as fundamental proxies in and for reconstructing ancient environmental conditions, including temperature, ocean circulation, and atmospheric composition over geological timescales. Oxygen isotope ratios (δ¹⁸O) in ice cores and marine sediments are particularly valuable for inferring past temperatures, as variations reflect both local site effects and global ice volume changes. In the Vostok ice core from , δ¹⁸O records exhibit glacial-interglacial fluctuations of 5–7‰ over the past 420,000 years, with lower values during colder glacial periods indicating cooler precipitation temperatures and expanded ice sheets. These shifts arise from equilibrium fractionation during , where heavier isotopes preferentially precipitate in colder conditions. Carbon isotope ratios (δ¹³C) in sediments provide insights into ancient productivity, carbon cycling, and atmospheric CO₂ levels. During hyperthermal events like the Paleocene-Eocene Thermal Maximum (PETM) approximately 56 million years ago, δ¹³C values in benthic and bulk carbonates show a negative excursion of about -2.5‰, signaling the rapid release of isotopically light carbon from sources such as methane hydrates or wetlands, which perturbed global carbon reservoirs and drove warming. This excursion, preserved in deep-sea records worldwide, highlights how stable isotopes trace major geochemical perturbations without relying on modern analogs. Sulfur isotope ratios (δ³⁴S) in sulfate minerals, such as , help reconstruct ancient chemistry, conditions, and influences from volcanic or microbial processes. In deposits, δ³⁴S values typically mirror contemporaneous sulfate, with excursions indicating changes in bacterial sulfate —which preferentially incorporates lighter ³²S—or volcanic inputs that can introduce distinct isotopic signatures. For instance, Permian-Triassic gypsum beds exhibit δ³⁴S variations of up to 10‰, linked to enhanced microbial activity in anoxic basins and volcanic degassing that altered global sulfur budgets. These records, combined with oxygen isotopes in sulfates, delineate paleo-oceanic sulfate concentrations and biological influences on sulfur cycling. Integrating hydrogen (δD) and oxygen (δ¹⁸O) isotopes enhances reconstructions of past patterns and moisture sources. In polar cores, the deuterium excess (d-excess = δD - 8 × δ¹⁸O) deviates from the during glacial periods, with higher values suggesting drier conditions or shifted evaporation sources over subtropical oceans. For example, East Antarctic cores show d-excess increases of 2–5‰ during the , indicating reduced relative humidity at moisture origins and altered . This combined approach refines paleoclimate models by distinguishing temperature effects from hydrological changes.

Ecology and forensics

Stable isotope ratios play a crucial role in by elucidating trophic interactions within webs. In particular, nitrogen stable isotope ratios (δ¹⁵N) exhibit enrichment of approximately 3–4‰ per , allowing researchers to quantify consumer positions and predator-prey dynamics. This stepwise increase, often standardized at 3.4‰, reflects preferential incorporation of heavier isotopes during metabolic processes and has been widely applied to map energy flow in ecosystems. For instance, in environments, δ¹⁵N analysis of and tissues has revealed omnivory and bottom-up controls in food chains, highlighting how isotopic baselines vary spatially and temporally to influence trophic interpretations. In ecological studies of animal movement, hydrogen stable isotope ratios (δ²H) in metabolically inert tissues like feathers and serve as tracers of geographic origins, mirroring the isotopic composition of at the site of tissue growth. This approach leverages continent-scale isoscapes—maps of spatial isotope variation—to reconstruct routes without direct . For example, δ²H signatures in bird feathers have pinpointed breeding grounds and wintering areas for species like Sharp-shinned Hawks, enabling the delineation of migratory connectivity across hemispheres. Forensic applications of stable isotopes extend to verifying product and tracing substances through their environmental imprints. In wine determination, combined δ¹³C and δ¹⁸O analyses distinguish between and photosynthetic pathways in source , as well as irrigation effects on isotopes, allowing differentiation of regional origins such as versus South American vintages. Similarly, δ¹³C profiling of hydrochloride exploits variations in coca leaf growth environments across , enabling classification of samples to specific countries like or with high accuracy via multivariate statistical models. In archaeological contexts, stable isotopes facilitate the reconstruction of past human diets by analyzing bone collagen and . Carbon isotope ratios (δ¹³C) in human remains from the ancient indicate the adoption of —a C4 with distinct δ¹³C values around -9‰—as a dietary staple, with evidence from sites in and showing increasing reliance from 5000 years ago onward. This method has illuminated shifts in subsistence patterns, such as maize intensification in pre-Columbian societies of the Andean region.

References

  1. [1]
    Fundamentals of Stable Isotope Geochemistry
    The so-called stable isotopes are nuclei that do not appear to decay to other isotopes on geologic timescales, but may themselves be produced by the decay of ...
  2. [2]
    [PDF] Stable Isotope Geochemistry I: Theory
    Stable isotope geochemistry is concerned with variations of the isotopic compositions of light ele- ments arising from chemical fractionations rather than ...
  3. [3]
    Stable Isotopes and Mineral Resource Investigations in the United ...
    Dec 21, 1999 · Hence, they are called stable isotopes. Stable isotope geochemistry investigates variations in the ratios of stable isotopes such as 2H/1H ...<|control11|><|separator|>
  4. [4]
    [PDF] Guidelines and recommended terms for expression of stable-isotope ...
    The delta value (d) is used to express the relative difference of a ratio of the numbers (or the amounts) of two isotopes in a specimen compared with that of a ...
  5. [5]
    Stable Isotope Ratios and Forensic Analysis of Microorganisms - NIH
    Here we discuss the use of stable isotope ratios in microbial forensics, using as a database the carbon, nitrogen, oxygen, and hydrogen stable isotope ratios.
  6. [6]
    Determining Delta Values - Education - Stable Isotopes NOAA GML
    Isotope ratio mass spectrometers measure relative isotopic ratios much better than actual ratios. By comparing to a standard, the precision of the data values ...
  7. [7]
  8. [8]
    Stable isotopes - an overview | ScienceDirect Topics
    For convenience, the δ (delta) notation is used to describe isotope ratios, which is defined as: δ × [ sample ] = ( R [ sample ] ) − ( R [ standard ] ) ...<|separator|>
  9. [9]
    Stable Isotopes - Biotechnology Center
    The most common stable isotopes are variants of elements like hydrogen, carbon, nitrogen, and oxygen (Figure 1).
  10. [10]
    Stable Isotopes H, C, N, O, S | Czech Geological Survey
    Stable Isotopes H, C, N, O, S ... The laboratory specializes in the stable isotope geochemistry of light elements, blending analytical method development with the ...
  11. [11]
    Introduction | G-WADI
    Delta Notation. By convention, the heavier isotope is written first to give a heavy/light isotope ratio. For example, δ 2H or δ D = 2H/1H. δ 18O = 18O/16O ...
  12. [12]
    Process Patent Protection via Analysis of Stable Isotope Ratios
    ... Urey, (14-16) who established the “delta notation” approach to isotopic reporting that still is used today. This notation takes an analogous form for all ...<|separator|>
  13. [13]
    [PDF] Urey : The Thermodynumic Properties of Isotopic Substances.
    By HAROLD C. UREY. (Institute of Nuclear Studies, University of Chicago.) THE discovery of the isotopes of the elements resulted from the careful study of ...
  14. [14]
    [PDF] W. M. White Geochemistry Chapter 9: Stable Isotopes
    Nier in the 1930's and 1940's were the precursors to this development. The real history of stable isotope geochemistry begins in 1947 with the publication of ...
  15. [15]
    [PDF] Table of Isotopic Masses and Natural Abundances
    This table lists the mass and percent natural abundance for the stable nuclides. The mass of the longest lived isotope is given for elements without a ...<|control11|><|separator|>
  16. [16]
    The origin of the elements: a century of progress - Journals
    Aug 17, 2020 · This review assesses the current state of knowledge of how the elements were produced in the Big Bang, in stellar lives and deaths, and by interactions in ...<|control11|><|separator|>
  17. [17]
    Changes to Carbon Isotopes in Atmospheric CO2 Over the Industrial ...
    Oct 23, 2020 · Annual mean, global mean values are provided for δ13C. From 1850 to 2015 atmospheric δ13CO2 decreased by 1.8‰, with 1.5% of this drop ...Introduction · Atmospheric Changes Over... · Projected Future Changes in...
  18. [18]
    An international intercomparison of stable carbon isotope ... - ASLO
    et al. 2017), δ13C-DIC values in ocean waters range typically from −6.56‰ to +3.10‰, and based on experience and rec- ommendation of leading researchers, an “ ...<|control11|><|separator|>
  19. [19]
    [PDF] Paleovegetation reconstruction using δ 13C of Soil Organic Matter
    Abstract. The relative contributions of C3 and C4 plants to vegetation at a given locality may be estimated by means of δ13C of soil organic matter.
  20. [20]
    [PDF] Equilibrium stable isotope fractionation
    The effect of isotopic substitution on rotational and translational energies can be expressed in terms of vibrational frequencies! αClO-Cl = 1.0096 at 298 K.
  21. [21]
    [PDF] Kinetic and equilibrium mass-dependent isotope fractionation laws ...
    Equilibrium partitioning of isotopes between compounds depends upon zero-point energy differences that reflect the net effect of numerous vibrational modes.<|control11|><|separator|>
  22. [22]
    [PDF] Stable Isotope Theory : Equilibrium Fractionation - geo
    Apr 7, 2009 · Vibrational frequency and heat capacity are closely related because thermal energy in a crystal is stored as vibrational energy of the atoms ...
  23. [23]
    [PDF] Data of Geochemistry Sixth Edition - USGS Publications Warehouse
    ... oxygen isotope fractionation between calcium carbonate and water: Geochim. ... partition function ratios of calcite, and the experimental calcite-water ...
  24. [24]
    Chapter 2: Fundamentals of Isotope Geochemistry
    Mar 3, 1999 · The following section presents a very brief discussion of the fundamentals of stable and radio-isotope geochemistry.
  25. [25]
    [PDF] Kinetic Isotope Effects in Organic Chemistry - Macmillan Group
    Sep 14, 2005 · 1. Isotopic substitution does not affect the potential energy surface of the reaction or the energies of the electronic states. 2.
  26. [26]
    [PDF] An Introduction to Isotopic Calculations
    Isotope ratios are also measures of the absolute abun- dance of isotopes; they are usually arranged so that the. more abundant isotope appears in the ...
  27. [27]
    4.1 Kinetic Fractionation – Stable Isotope Hydrology - GW Books
    For diffusion, the diffusive velocity is inversely proportional to the mass of the molecule and therefore molecules with lighter isotopes will diffuse faster.Missing: formula | Show results with:formula
  28. [28]
    HESS - Technical note: Isotopic fractionation of evaporating waters
    Jan 16, 2024 · Di/D is the diffusivity ratio between light and heavy isotopes. Following Merlivat (1978), Di/D is 0.9755 and 0.9723 for 2H and 18O, ...
  29. [29]
    CARBON ISOTOPE DISCRIMINATION AND PHOTOSYNTHESIS
    FARQUHAR ET AL. Genetic Control of Discrimination. Genetic studies of W, p;lPa, and a are in their infancy. These traits are most likely to be polygenic ...
  30. [30]
    [PDF] Isotope Ratio Mass Spectrometry - CalTech GPS
    May 19, 2014 · Isotope Ratio Mass Spectrometry (IRMS) is a specialized technique used to provide information about the geographic, chemical, and biological ...
  31. [31]
    Isotope Ratio Mass Spectrometry
    The stable isotope analyses provided by IRMS assist the Terrestrial-Atmospheric Processes Integrated Research Platform in understanding and quantifying nutrient ...
  32. [32]
    Multicollector-Inductively Coupled Plasma Mass Spectrometer (MC ...
    May 17, 2007 · MC-ICPMS is an instrument that measures isotopic ratios that are used in geochemistry, geochronology, and cosmochemistry.
  33. [33]
    [PDF] Geol. 655 Isotope Geochemistry
    We will focus exclusively, however, on mass spectrometers used for isotope ratio determination. Most isotope ratio mass spectrometers are of a similar design, ...
  34. [34]
    [PDF] Determination of δ13C, δ15N, or δ34S by Isotope-Ratio-Monitoring ...
    As a precaution against memory in analysis of unknowns, samples with extreme δ34S values are reana- lyzed in duplicate or triplicate in subsequent autoruns so.
  35. [35]
    [PDF] Report of Investigation - Reference Material 8562
    12. The VPDB scale is determined by assigning a δ13C value of +1.95 ‰ to RM 8544 NBS 19 Limestone [1,4]. The differences in measured isotope-number ratios of ...
  36. [36]
    [PDF] Value Assignment of Light Stable Isotope Reference Materials
    Memory and conditioning effects in IRMS systems can have half-lives of a few minutes. (Verkouteren et al., 2003a), which are not fully compensated by ...
  37. [37]
    Breath Analysis Using Laser Spectroscopic Techniques
    This review presents an update on the latest developments in laser-based breath analysis. Keywords: breath analysis; biomarkers, laser spectroscopic detection ...<|control11|><|separator|>
  38. [38]
  39. [39]
  40. [40]
    [PDF] A review of optical isotope techniques
    The development of stable isotope techniques has provided scientists an important set of tools that can be used to gain new insights into the coupled carbon ...
  41. [41]
    Using laser spectroscopy for measuring the ratios of stable isotopes
    The advantage of using stable isotope ratios is that water molecules serve as ubiquitous and already present natural tracers. Within recent years, these ...
  42. [42]
  43. [43]
    Record of δ18O and 17O‐excess in ice from Vostok Antarctica ...
    Jan 24, 2008 · The observed shifts suggest higher normalized relative humidity and/or wind speeds over the source oceanic regions in glacial times. 1.
  44. [44]
    Magnitude of the carbon isotope excursion at the Paleocene ...
    The PETM is marked by a large negative carbon isotope excursion (CIE) in terrestrial and marine carbonates and organic matter (Kennett and Stott, 1991, Koch et ...
  45. [45]
    Sulfur cycle in buried evaporites | Geology - GeoScienceWorld
    Jun 2, 2017 · The δ34S in dissolved marine sulfate increased by 10‰ from the Late Permian to the Early Triassic. Despite precipitation of gypsum from Permian ...Missing: paleoceanography | Show results with:paleoceanography
  46. [46]
    Experimental determination of carbonate-associated sulfate δ 34 S ...
    Apr 4, 2014 · The earliest records of marine δ34S were based on gypsum, in which the isotopic composition of sulfur reflects that of the brine from which it ...
  47. [47]
    Deuterium excess in an East Antarctic ice core suggests higher ...
    Oct 21, 1982 · Deuterium excess in an East Antarctic ice core suggests higher relative humidity at the oceanic surface during the last glacial maximum. Jean ...Missing: patterns | Show results with:patterns
  48. [48]
    Deuterium and oxygen 18 in precipitation: Modeling of the isotopic ...
    Dec 20, 1984 · Jouzel, J., L. Merlivat, C. Lorius, Deuterium excess in an East Antarctic ice core suggests higher relative humidity at the oceanic surface ...
  49. [49]
    USING STABLE ISOTOPES TO ESTIMATE TROPHIC POSITION ...
    Mar 1, 2002 · The stable isotopes of nitrogen (δ 15 N) and carbon (δ 13 C) provide powerful tools for estimating the trophic positions of and carbon flow to consumers in ...
  50. [50]
    Body size predicts ontogenetic nitrogen stable-isotope (δ 15 N ...
    Jun 19, 2024 · δ15N values of a consumer are typically enriched by 3–4‰ relative to its diet, though the value most frequently used is 3.4‰. This is commonly ...
  51. [51]
    Revisiting the use of δ 15 N in meso-scale studies of marine food ...
    This study aimed to assess the importance of considering spatio-temporal variations in isotopic signatures of consumers when using δ 13 C and especially δ 15 N ...
  52. [52]
    Using oxygen and hydrogen stable isotopes to track the migratory ...
    We applied stable isotope analysis of hydrogen (H) and oxygen (O) of feathers, and large scale models of spatial isotopic variation (isoscapes) to formulate ...
  53. [53]
    Advancing Stable Isotope Analysis for Alcoholic Beverages ...
    Mar 10, 2025 · Stable isotopes such as hydrogen (δ2H), oxygen (δ18O), carbon (δ13C), and strontium (87Sr/86Sr) play critical roles in identifying wine ...
  54. [54]
    Geographically Sourcing Cocaine's Origin - PubMed Central - NIH
    Mar 23, 2016 · We show how the combination of trace cocaine alkaloids, stable isotopes, and multivariate statistical analyses can be used to classify illicit cocaine as ...
  55. [55]
    Early isotopic evidence for maize as a staple grain in the Americas
    Jun 3, 2020 · We use carbon isotopes in human bone as the earliest direct evidence for maize as a staple grain in the Americas.