Radiocarbon dating considerations

Last updated

The variation in the 14
C
/12
C
ratio in different parts of the carbon exchange reservoir means that a straightforward calculation of the age of a sample based on the amount of 14
C
it contains will often give an incorrect result. There are several other possible sources of error that need to be considered. The errors are of four general types:

Contents

Atmospheric variation

In the early years of using the technique, it was understood that it depended on the atmospheric 14
C
/12
C
ratio having remained the same over the preceding few thousand years. To verify the accuracy of the method, several artefacts that were datable by other techniques were tested; the results of the testing were in reasonable agreement with the true ages of the objects. However, in 1958, Hessel de Vries was able to demonstrate that the 14
C
/12
C
ratio had changed over time by testing wood samples of known ages and showing there was a significant deviation from the expected ratio. This discrepancy, often called the de Vries effect, was resolved by the study of tree rings. [1] [2] Comparison of overlapping series of tree rings allowed the construction of a continuous sequence of tree-ring data that spanned 8,000 years. [1] (Since that time the tree-ring data series has been extended to 13,900 years.) [3] Carbon-dating the wood from the tree rings themselves provided the check needed on the atmospheric 14
C
/12
C
ratio: with a sample of known date, and a measurement of the value of N (the number of atoms of 14
C
remaining in the sample), the carbon-dating equation allows the calculation of N0 – the number of atoms of 14
C
in the sample at the time the tree ring was formed – and hence the 14
C
/12
C
ratio in the atmosphere at that time. [1] Armed with the results of carbon-dating the tree rings, it became possible to construct calibration curves designed to correct the errors caused by the variation over time in the 14
C
/12
C
ratio. [4] These curves are described in more detail below.

There are three main reasons for these variations in the historical 14
C
/12
C
ratio: fluctuations in the rate at which 14
C
is created, changes caused by glaciation, and changes caused by human activity. [1]

Variations in 14
C
production

Two different trends can be seen in the tree ring series. First, there is a long-term oscillation with a period of about 9,000 years, which causes radiocarbon dates to be older than true dates for the last 2,000 years and too young before that. The known fluctuations in the strength of the Earth's magnetic field match up quite well with this oscillation: cosmic rays are deflected by magnetic fields, so when there is a weaker magnetic field, more 14
C
is produced, leading to a younger apparent age for samples from those periods. Conversely, a stronger magnetic field leads to lower 14
C
production and an older apparent age. A secondary oscillation is thought to be caused by variations in sunspot activity, which has two separate periods: a longer-term, 200-year oscillation, and a shorter 11-year cycle. Sunspots cause changes in the solar system's magnetic field and corresponding changes to the cosmic ray flux, and hence to the production of 14
C
. [1]

There are two kinds of geophysical event which can affect 14
C
production: geomagnetic reversals and polarity excursions. In a geomagnetic reversal, the Earth's geomagnetic field weakens and stays weak for thousands of years during the transition to the opposite magnetic polarity and then regains strength as the reversal completes. A polarity excursion, which can be either global or local, is a shorter-lived version of a geomagnetic reversal. A local excursion would not significantly affect 14C production. During either a geomagnetic reversal or a global polarity excursion, 14
C
production increases during the period when the geomagnetic field is weak. It is fairly certain, though, that in the last 50,000 years there have been no geomagnetic reversals or global polarity excursions. [5]

Since the earth's magnetic field varies with latitude, the rate of 14
C
production changes with latitude, too, but atmospheric mixing is rapid enough that these variations amount to less than 0.5% of the global concentration. [1] This is close to the limit of detectability in most years, [6] but the effect can be seen clearly in tree rings from years such as 1963, when 14
C
from nuclear testing rose sharply through the year. [7] The latitudinal variation in 14
C
was much larger than normal that year, and tree rings from different latitudes show corresponding variations in their 14
C
content. [7]

14
C
can also be produced at ground level, primarily by cosmic rays that penetrate the atmosphere as far as the earth's surface, but also by spontaneous fission of naturally occurring uranium. These sources of neutrons only produce 14
C
at a rate of 1 x 10−4 atoms per gram per second, which is not enough to have a significant effect on dating. [7] [8] At higher altitudes, the neutron flux can be substantially higher, [9] [note 1] and in addition, trees at higher altitude are more likely to be struck by lightning, which produces neutrons. However, experiments in which wood samples have been irradiated with neutrons indicate that the effect on 14
C
content is minor, though for very old trees (such as some bristlecone pines) that grow at altitude some effect can be seen. [9]

Effect of climatic cycles

Because the solubility of CO
2
in water increases with lower temperatures, glacial periods would have led to the faster absorption of atmospheric CO
2
by the oceans. In addition, any carbon stored in the glaciers would be depleted in 14
C
over the life of the glacier; when the glacier melted as the climate warmed, the depleted carbon would be released, reducing the global 14
C
/12
C
ratio. The changes in climate would also cause changes in the biosphere, with warmer periods leading to more plant and animal life. The effect of these factors on radiocarbon dating is not known. [1]

Effects of human activity

caption = Atmospheric
C for the northern and southern hemispheres, showing percentage excess above pre-bomb levels.  The Partial Test Ban Treaty went into effect on 10 October 1963. Hemispheric 14C graphs 1950s to 2010.png
caption = Atmospheric
C
for the northern and southern hemispheres, showing percentage excess above pre-bomb levels.  The Partial Test Ban Treaty went into effect on 10 October 1963.

Coal and oil began to be burned in large quantities during the 1800s. Both coal and oil are sufficiently old that they contain little detectable 14
C
and, as a result, the CO
2
released substantially diluted the atmospheric 14
C
/12
C
ratio. Dating an object from the early 20th century hence gives an apparent date older than the true date. For the same reason, 14
C
concentrations in the neighbourhood of large cities are lower than the atmospheric average. This fossil fuel effect (also known as the Suess effect, after Hans Suess, who first reported it in 1955) would only amount to a reduction of 0.2% in 14
C
activity if the additional carbon from fossil fuels were distributed throughout the carbon exchange reservoir, but because of the long delay in mixing with the deep ocean, the actual effect is a 3% reduction. [1] [11]

A much larger effect comes from above-ground nuclear testing, which released large numbers of neutrons and created 14
C
. From about 1950 until 1963, when atmospheric nuclear testing was banned, it is estimated that several tonnes of 14
C
were created. If all this extra 14
C
had immediately been spread across the entire carbon exchange reservoir, it would have led to an increase in the 14
C
/12
C
ratio of only a few per cent, but the immediate effect was to almost double the amount of 14
C
in the atmosphere, with the peak level occurring in about 1965. The level has since dropped, as the "bomb carbon" (as it is sometimes called) percolates into the rest of the reservoir. [1] [11] [12]

Isotopic fractionation

Photosynthesis is the primary process by which carbon moves from the atmosphere into living things. Two different photosynthetic processes exist: the C3 pathway and the C4 pathway. About 90% of all plant life uses the C3 process; the remaining plants either use C4 or are CAM plants, which can use either C3 or C4 depending on the environmental conditions. Both the C3 and C4 photosynthesis pathways show a preference for lighter carbon, with 12
C
being absorbed slightly more easily than 13
C
, which in turn is more easily absorbed than 14
C
. The differential uptake of the three carbon isotopes leads to 13
C
/12
C
and 14
C
/12
C
ratios in plants that differ from the ratios in the atmosphere. This effect is known as isotopic fractionation. [9] [13]

To determine the degree of fractionation that takes place in a given plant, the amounts of both 12
C
and 13
C
are measured, and the resulting 13
C
/12
C
ratio is then compared to a standard ratio known as PDB. (The 13
C
/12
C
ratio is used because it is much easier to measure than the 14
C
/12
C
ratio, and the 14
C
/12
C
ratio can be easily derived from it.) The resulting value, known as δ13C, is calculated as follows: [9]

where the ‰ (permil) sign indicates parts per thousand. [9] Because the PDB standard contains an unusually high proportion of 13
C
, [note 2] most measured δ13C values are negative. Values for C3 plants typically range from −30‰ to −22‰, with an average of −27‰; for C4 plants the range is −15‰ to −9‰, and the average is −13‰. [13] Atmospheric CO
2
has a δ13C of −8‰. [9]

Sheep on the beach in North Ronaldsay. In the winter, these sheep eat seaweed, which has a higher dC content than grass; samples from these sheep have a dC value of about -13%0, which is much higher than for sheep that feed on grasses. NR sheep.jpg
Sheep on the beach in North Ronaldsay. In the winter, these sheep eat seaweed, which has a higher δC content than grass; samples from these sheep have a δC value of about −13‰, which is much higher than for sheep that feed on grasses.

For marine organisms, the details of the photosynthesis reactions are less well understood. Measured δ13C values for marine plankton range from −31‰ to −10‰; most lie between −22‰ and −17‰. The δ13C values for marine photosynthetic organisms also depend on temperature. At higher temperatures, CO
2
has poor solubility in water, which means there is less CO
2
available for the photosynthetic reactions. Under these conditions, fractionation is reduced, and at temperatures above 14 °C the δ13C values are correspondingly higher, reaching −13‰. At lower temperatures, CO
2
becomes more soluble and hence more available to the marine organisms; fractionation increases and δ13C values can be as low as −32‰. [13]

The δ13C value for animals depends on their diet. An animal that eats food with high δ13C values will have a higher δ13C than one that eats food with lower δ13C values. [9] The animal's own biochemical processes can also affect the results: for example, both bone minerals and bone collagen typically have a higher concentration of 13
C
than is found in the animal's diet, though for different biochemical reasons. The enrichment of bone 13
C
also implies that excreted material is depleted in 13
C
relative to the diet. [15]

Since 13
C
makes up about 1% of the carbon in a sample, the 13
C
/12
C
ratio can be accurately measured by mass spectrometry. [16] Typical values of δ13C have been found by experiment for many plants, as well as for different parts of animals such as bone collagen, but when dating a given sample it is better to determine the δ13C value for that sample directly than to rely on the published values. [9] The depletion of 13
C
relative to 12
C
is proportional to the difference in the atomic masses of the two isotopes, so once the δ13C value is known, the depletion for 14
C
can be calculated: it will be twice the depletion of 13
C
. [16]

The carbon exchange between atmospheric CO
2
and carbonate at the ocean surface is also subject to fractionation, with 14
C
in the atmosphere more likely than 12
C
to dissolve in the ocean. The result is an overall increase in the 14
C
/12
C
ratio in the ocean of 1.5%, relative to the 14
C
/12
C
ratio in the atmosphere. This increase in 14
C
concentration almost exactly cancels out the decrease caused by the upwelling of water (containing old, and hence 14
C
depleted, carbon) from the deep ocean, so that direct measurements of 14
C
radiation are similar to measurements for the rest of the biosphere. Correcting for isotopic fractionation, as is done for all radiocarbon dates to allow comparison between results from different parts of the biosphere, gives an apparent age of about 400 years for ocean surface water. [16]

Reservoir effects

Libby's original exchange reservoir hypothesis assumed that the 14
C
/12
C
ratio in the exchange reservoir is constant all over the world, [17] but it has since been discovered that there are several causes of variation in the ratio across the reservoir. [18]

Marine effect

The CO
2
in the atmosphere transfers to the ocean by dissolving in the surface water as carbonate and bicarbonate ions; at the same time the carbonate ions in the water are returning to the air as CO
2
. [17] This exchange process brings14
C
from the atmosphere into the surface waters of the ocean, but the 14
C
thus introduced takes a long time to percolate through the entire volume of the ocean. The deepest parts of the ocean mix very slowly with the surface waters, and the mixing is known to be uneven. The main mechanism that brings deep water to the surface is upwelling. Upwelling is more common in regions closer to the equator; it is also influenced by other factors such as the topography of the local ocean bottom and coastlines, the climate, and wind patterns. Overall, the mixing of deep and surface waters takes far longer than the mixing of atmospheric CO
2
with the surface waters, and as a result water from some deep ocean areas has an apparent radiocarbon age of several thousand years. Upwelling mixes this "old" water with the surface water, giving the surface water an apparent age of about several hundred years (after correcting for fractionation). [18] This effect is not uniform—the average effect is about 440 years, but there are local deviations of several hundred years for areas that are geographically close to each other. [18] [19] The effect also applies to marine organisms such as shells, and marine mammals such as whales and seals, which have radiocarbon ages that appear to be hundreds of years old. [18] These marine reservoir effects vary over time as well as geographically; for example, there is evidence that during the Younger Dryas, a period of cold climatic conditions about 12,000 years ago, the apparent difference between the age of surface water and the contemporary atmosphere increased from between 400 and 600 years to about 900 years until the climate warmed again. [19]

Hard water effect

If the carbon in freshwater is partly acquired from aged carbon, such as rocks, then the result will be a reduction in the 14
C
/12
C
ratio in the water. For example, rivers that pass over limestone, which is mostly composed of calcium carbonate, will acquire carbonate ions. Similarly, groundwater can contain carbon derived from the rocks through which it has passed. These rocks are usually so old that they no longer contain any measurable 14
C
, so this carbon lowers the 14
C
/12
C
ratio of the water it enters, which can lead to apparent ages of thousands of years for both the affected water and the plants and freshwater organisms that live in it. [16] This is known as the hard water effect, because it is often associated with calcium ions, which are characteristic of hard water; however, there can be other sources of carbon that have the same effect, such as humus. The effect is not necessarily confined to freshwater species—at a river mouth, the outflow may affect marine organisms. It can also affect terrestrial snails that feed in areas where there is a high chalk content, though no measurable effect has been found for land plants in soil with a high carbonate content—it appears that almost all the carbon for these plants is derived from photosynthesis and not from the soil. [18]

It is not possible to deduce the effect of the effect by determining the hardness of the water: the aged carbon is not necessarily immediately incorporated into the plants and animals that are affected, and the delay affects their apparent age. The effect is very variable and there is no general offset that can be applied; the usual way to determine the size of the effect is to measure the apparent age offset of a modern sample. [18]

Volcanoes

Volcanic eruptions eject large amounts of carbon into the air. The carbon is of geological origin and has no detectable 14
C
, so the 14
C
/12
C
ratio in the vicinity of the volcano is depressed relative to surrounding areas. Dormant volcanoes can also emit aged carbon. Plants that photosynthesize this carbon also have lower 14
C
/12
C
ratios: for example, plants on the Greek island of Santorini, near the volcano, have apparent ages of up to a thousand years. These effects are hard to predict—the town of Akrotiri, on Santorini, was destroyed in a volcanic eruption thousands of years ago, but radiocarbon dates for objects recovered from the ruins of the town show surprisingly close agreement with dates derived from other means. If the dates for Akrotiri are confirmed, it would indicate that the volcanic effect in this case was minimal. [18]

Hemisphere effect

The northern and southern hemispheres have atmospheric circulation systems that are sufficiently independent of each other that there is a noticeable time lag in mixing between the two. The atmospheric 14
C
/12
C
ratio is lower in the southern hemisphere, with an apparent additional age of 30 years for radiocarbon results from the south as compared to the north. This is probably because the greater surface area of ocean in the southern hemisphere means that there is more carbon exchanged between the ocean and the atmosphere than in the north. Since the surface ocean is depleted in 14
C
because of the marine effect, 14
C
is removed from the southern atmosphere more quickly than in the north. [18]

Island effect

It has been suggested that an "island effect" might exist, by analogy with the mechanism thought to explain the hemisphere effect: since islands are surrounded by water, the carbon exchange between the water and atmosphere might reduce the 14
C
/12
C
ratio on an island. Within a hemisphere, however, atmospheric mixing is apparently rapid enough that no such effect exists: two calibration curves assembled in Seattle and Belfast laboratories, with results from North American trees and Irish trees, respectively, are in close agreement, instead of the Irish samples appearing to be older, as would be the case if there were an island effect. [18]

Contamination

Any addition of carbon to a sample of a different age will cause the measured date to be inaccurate. Contamination with modern carbon causes a sample to appear to be younger than it really is: the effect is greater for older samples. If a sample that is in fact 17,000 years old is contaminated so that 1% of the sample is actually modern carbon, it will appear to be 600 years younger; for a sample that is 34,000 years old the same amount of contamination would cause an error of 4,000 years. Contamination with old carbon, with no remaining 14
C
, causes an error in the other direction, which does not depend on age—a sample that has been contaminated with 1% old carbon will appear to be about 80 years older than it really is, regardless of the date of the sample. [20]

Contamination can occur if the sample is brought into contact with or packed in materials that contain carbon. Cotton wool, cigarette ash, paper labels, cloth bags, and some conservation chemicals such as polyvinyl acetate can all be sources of modern carbon. [21] Labels should be added to the outside of the container, not placed inside the bag or vial with the sample. Glass wool is acceptable as packing material instead of cotton wool. [22] Samples should be packed in glass vials or aluminium foil if possible; [21] [23] polyethylene bags are also acceptable but some plastics, such as PVC, can contaminate the sample. [22] Contamination can also occur before the sample is collected: humic acids or carbonate from the soil can leach into a sample, and for some sample types, such as shells, there is the possibility of carbon exchange between the sample and the environment, depleting the sample's 14
C
content. [21]

Notes

  1. Even at an altitude of 3 km, the neutron flux is only 3% of the value in the stratosphere where most 14
    C
    is created; at sea level the value is less than 0.5% of the value in the stratosphere. [9]
  2. The PDB value is 11.1‰. [14]

Footnotes

  1. 1 2 3 4 5 6 7 8 9 Bowman (1995), pp. 16–20.
  2. Suess (1970), p. 303.
  3. Reimer, Paula J.; et al. (2013). "IntCal13 and Marine13 radiocarbon age calibration curves 0–50,000 years cal BP". Radiocarbon. 55 (4): 1869–1887. doi: 10.2458/azu_js_rc.55.16947 .
  4. Bowman (1995), pp. 43–49.
  5. Aitken (1990), pp. 68–69.
  6. Rasskazov, Brandt & Brandt (2009), p. 40.
  7. 1 2 3 Grootes, Pieter M. (1992). "Subtle 14
    C
    Signals: The Influence of Atmospheric Mixing, Growing Season and In-Situ Production"
    . Radiocarbon. 34 (2): 219–225. doi: 10.1017/S0033822200013655 .
  8. Ramsey, C.B. (2008). "Radiocarbon dating: revolutions in understanding". Archaeometry. 50 (2): 249–275. doi:10.1111/j.1475-4754.2008.00394.x.
  9. 1 2 3 4 5 6 7 8 9 10 Bowman (1995), pp. 20–23.
  10. Hua, Quan; Barbetti, Mike; Rakowski, Andrzej Z. (2013). "Atmospheric Radiocarbon for the Period 1950–2010". Radiocarbon. 55 (4): 2059–2072. doi: 10.2458/azu_js_rc.v55i2.16177 . ISSN   0033-8222.
  11. 1 2 Aitken (1990), pp. 71–72.
  12. "Limited Test Ban Treaty". Science Magazine. Retrieved July 26, 2013.
  13. 1 2 3 Maslin & Swann (2006), p. 246.
  14. Miller & Wheeler (2012), p. 186.
  15. Schoeninger (2010), p. 446.
  16. 1 2 3 4 Aitken (1990), pp. 61–66.
  17. 1 2 Libby (1965), p. 6.
  18. 1 2 3 4 5 6 7 8 9 Bowman (1995), pp. 24–27.
  19. 1 2 Cronin (2010), p. 35.
  20. Aitken (1990), pp. 85–86.
  21. 1 2 3 Bowman (1995), pp. 27–30.
  22. 1 2 Aitken (1990), p. 89.
  23. Burke, Smith & Zimmerman (2009), p. 175.

Related Research Articles

<span class="mw-page-title-main">Radiocarbon dating</span> Method of determining the age of objects

Radiocarbon dating is a method for determining the age of an object containing organic material by using the properties of radiocarbon, a radioactive isotope of carbon.

Radiometric dating, radioactive dating or radioisotope dating is a technique which is used to date materials such as rocks or carbon, in which trace radioactive impurities were selectively incorporated when they were formed. The method compares the abundance of a naturally occurring radioactive isotope within the material to the abundance of its decay products, which form at a known constant rate of decay. The use of radiometric dating was first published in 1907 by Bertram Boltwood and is now the principal source of information about the absolute age of rocks and other geological features, including the age of fossilized life forms or the age of Earth itself, and can also be used to date a wide range of natural and man-made materials.

Paleoclimatology is the study of climates for which direct measurements were not taken. As instrumental records only span a tiny part of Earth's history, the reconstruction of ancient climate is important to understand natural variation and the evolution of the current climate.

<span class="mw-page-title-main">Carbon-14</span> Isotope of carbon

Carbon-14, C-14, 14
C
or radiocarbon, is a radioactive isotope of carbon with an atomic nucleus containing 6 protons and 8 neutrons. Its presence in organic materials is the basis of the radiocarbon dating method pioneered by Willard Libby and colleagues (1949) to date archaeological, geological and hydrogeological samples. Carbon-14 was discovered on February 27, 1940, by Martin Kamen and Sam Ruben at the University of California Radiation Laboratory in Berkeley, California. Its existence had been suggested by Franz Kurie in 1934.

<span class="mw-page-title-main">Isotope analysis</span> Analytical technique used to study isotopes

Isotope analysis is the identification of isotopic signature, abundance of certain stable isotopes of chemical elements within organic and inorganic compounds. Isotopic analysis can be used to understand the flow of energy through a food web, to reconstruct past environmental and climatic conditions, to investigate human and animal diets, for food authentification, and a variety of other physical, geological, palaeontological and chemical processes. Stable isotope ratios are measured using mass spectrometry, which separates the different isotopes of an element on the basis of their mass-to-charge ratio.

<span class="mw-page-title-main">Proxy (climate)</span> Preserved physical characteristics allowing reconstruction of past climatic conditions

In the study of past climates ("paleoclimatology"), climate proxies are preserved physical characteristics of the past that stand in for direct meteorological measurements and enable scientists to reconstruct the climatic conditions over a longer fraction of the Earth's history. Reliable global records of climate only began in the 1880s, and proxies provide the only means for scientists to determine climatic patterns before record-keeping began.

Isotope geochemistry is an aspect of geology based upon the study of natural variations in the relative abundances of isotopes of various elements. Variations in isotopic abundance are measured by isotope ratio mass spectrometry, and can reveal information about the ages and origins of rock, air or water bodies, or processes of mixing between them.

<span class="mw-page-title-main">Sulfur cycle</span> Biogeochemical cycle of sulfur

The sulfur cycle is a biogeochemical cycle in which the sulfur moves between rocks, waterways and living systems. It is important in geology as it affects many minerals and in life because sulfur is an essential element (CHNOPS), being a constituent of many proteins and cofactors, and sulfur compounds can be used as oxidants or reductants in microbial respiration. The global sulfur cycle involves the transformations of sulfur species through different oxidation states, which play an important role in both geological and biological processes. Steps of the sulfur cycle are:

Carbon (6C) has 15 known isotopes, from 8
C
to 22
C
, of which 12
C
and 13
C
are stable. The longest-lived radioisotope is 14
C
, with a half-life of 5.70(3)×103 years. This is also the only carbon radioisotope found in nature, as trace quantities are formed cosmogenically by the reaction 14
N
+
n
14
C
+ 1
H
. The most stable artificial radioisotope is 11
C
, which has a half-life of 20.3402(53) min. All other radioisotopes have half-lives under 20 seconds, most less than 200 milliseconds. The least stable isotope is 8
C
, with a half-life of 3.5(1.4)×10−21 s. Light isotopes tend to decay into isotopes of boron and heavy ones tend to decay into isotopes of nitrogen.

An isotopic signature is a ratio of non-radiogenic 'stable isotopes', stable radiogenic isotopes, or unstable radioactive isotopes of particular elements in an investigated material. The ratios of isotopes in a sample material are measured by isotope-ratio mass spectrometry against an isotopic reference material. This process is called isotope analysis.

Absolute dating is the process of determining an age on a specified chronology in archaeology and geology. Some scientists prefer the terms chronometric or calendar dating, as use of the word "absolute" implies an unwarranted certainty of accuracy. Absolute dating provides a numerical age or range, in contrast with relative dating, which places events in order without any measure of the age between events.

Paleoceanography is the study of the history of the oceans in the geologic past with regard to circulation, chemistry, biology, geology and patterns of sedimentation and biological productivity. Paleoceanographic studies using environment models and different proxies enable the scientific community to assess the role of the oceanic processes in the global climate by the re-construction of past climate at various intervals. Paleoceanographic research is also intimately tied to paleoclimatology.

The environmental isotopes are a subset of isotopes, both stable and radioactive, which are the object of isotope geochemistry. They are primarily used as tracers to see how things move around within the ocean-atmosphere system, within terrestrial biomes, within the Earth's surface, and between these broad domains.

The Suess effect, also referred to as the 13C Suess effect, is a change in the ratio of the atmospheric concentrations of heavy isotopes of carbon (13C and 14C) by the admixture of large amounts of fossil-fuel derived CO2, which is depleted in 13CO2 and contains no 14CO2. It is named for the Austrian chemist Hans Suess, who noted the influence of this effect on the accuracy of radiocarbon dating. More recently, the Suess effect has been used in studies of climate change. The term originally referred only to dilution of atmospheric 14CO2. The concept was later extended to dilution of 13CO2 and to other reservoirs of carbon such as the oceans and soils.

<span class="mw-page-title-main">Stable isotope ratio</span> Ratio of two stable isotopes

The term stable isotope has a meaning similar to stable nuclide, but is preferably used when speaking of nuclides of a specific element. Hence, the plural form stable isotopes usually refers to isotopes of the same element. The relative abundance of such stable isotopes can be measured experimentally, yielding an isotope ratio that can be used as a research tool. Theoretically, such stable isotopes could include the radiogenic daughter products of radioactive decay, used in radiometric dating. However, the expression stable-isotope ratio is preferably used to refer to isotopes whose relative abundances are affected by isotope fractionation in nature. This field is termed stable isotope geochemistry.

The calculation of radiocarbon dates determines the age of an object containing organic material by using the properties of radiocarbon, a radioactive isotope of carbon.

Minze Stuiver was a geochemist who was at the forefront of geoscience research from the 1960s until his retirement in 1998. He helped transform radiocarbon dating from a simple tool for archaeology and geology to a precise technique with applications in solar physics, oceanography, geochemistry, and carbon dynamics. Minze Stuiver's research encompassed the use of radiocarbon (14C) to understand solar cycles and radiocarbon production, ocean circulation, lake carbon dynamics and archaeology as well as the use of stable isotopes to document past climate changes.

<span class="mw-page-title-main">Bomb pulse</span> Sudden increase of carbon-14 in the Earths atmosphere due to nuclear bomb tests

The bomb pulse is the sudden increase of carbon-14 (14C) in the Earth's atmosphere due to the hundreds of aboveground nuclear bombs tests that started in 1945 and intensified after 1950 until 1963, when the Limited Test Ban Treaty was signed by the United States, the Soviet Union and the United Kingdom. These hundreds of blasts were followed by a doubling of the relative concentration of 14C in the atmosphere. We discuss “relative concentration”, because measurements of 14C levels by mass spectrometers are most accurately made by comparison to another carbon isotope, often the common isotope 12C. Isotope abundance ratios are not only more easily measured, they are what 14C carbon daters want, since it is the fraction of carbon in a sample that is 14C, not the absolute concentration, that is of interest in dating measurements. The figure shows how the fraction of carbon in the atmosphere that is 14C, of order only a part per trillion, has changed over the past several decades following the bomb tests. Because 12C concentration has increased by about 30% over the past fifty years, the fact that “pMC”, measuring the isotope ratio, has returned (almost) to its 1955 value, means that 14C concentration in the atmosphere remains some 30% higher than it once was. Carbon-14, the radioisotope of carbon, is naturally developed in trace amounts in the atmosphere and it can be detected in all living organisms. Carbon of all types is continually used to form the molecules of the cells of organisms. Doubling of the concentration of 14C in the atmosphere is reflected in the tissues and cells of all organisms that lived around the period of nuclear testing. This property has many applications in the fields of biology and forensics.

Hydrogen isotope biogeochemistry is the scientific study of biological, geological, and chemical processes in the environment using the distribution and relative abundance of hydrogen isotopes. There are two stable isotopes of hydrogen, protium 1H and deuterium 2H, which vary in relative abundance on the order of hundreds of permil. The ratio between these two species can be considered the hydrogen isotopic fingerprint of a substance. Understanding isotopic fingerprints and the sources of fractionation that lead to variation between them can be applied to address a diverse array of questions ranging from ecology and hydrology to geochemistry and paleoclimate reconstructions. Since specialized techniques are required to measure natural hydrogen isotope abundance ratios, the field of hydrogen isotope biogeochemistry provides uniquely specialized tools to more traditional fields like ecology and geochemistry.

<span class="mw-page-title-main">Fractionation of carbon isotopes in oxygenic photosynthesis</span>

Photosynthesis converts carbon dioxide to carbohydrates via several metabolic pathways that provide energy to an organism and preferentially react with certain stable isotopes of carbon. The selective enrichment of one stable isotope over another creates distinct isotopic fractionations that can be measured and correlated among oxygenic phototrophs. The degree of carbon isotope fractionation is influenced by several factors, including the metabolism, anatomy, growth rate, and environmental conditions of the organism. Understanding these variations in carbon fractionation across species is useful for biogeochemical studies, including the reconstruction of paleoecology, plant evolution, and the characterization of food chains.

References