Hostname: page-component-8448b6f56d-c4f8m Total loading time: 0 Render date: 2024-04-20T00:47:43.990Z Has data issue: false hasContentIssue false

EVOLUTION OF RADIOCARBON CALIBRATION

Published online by Cambridge University Press:  16 August 2021

Paula J Reimer*
Affiliation:
14CHRONO Centre for Climate the Environment and Chronology, Queen’s University Belfast, BelfastBT7 1NN, UK
*
Corresponding author. Email: p.j.reimer@qub.ac.uk.
Rights & Permissions [Opens in a new window]

Abstract

In the late 1950s it was recognized that levels of atmospheric radiocarbon (14C) had not been constant over time. Since then, researchers have sought to document those changes, initially through measurements of known age tree rings and more recently using other archives to create curves to correct or calibrate radiocarbon ages to calendar ages. This paper highlights some, but by no means all, of the efforts to create and extend radiocarbon calibration curves.

Type
Review Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2021. Published by Cambridge University Press for the Arizona Board of Regents on behalf of the University of Arizona

THE EARLY YEARS

Not long after James Arnold and Willard Libby published the “Curve of Knowns” as preliminary proof that radiocarbon (14C) dating worked at least for the past 5000 years (Arnold and Libby Reference Arnold and Libby1949), researchers began to evaluate the validity of two of the fundamental assumptions of the method. Hessel de Vries, who pioneered radiocarbon dating at the University of Groningen, was the first to convincingly demonstrate, using tree rings of known age, that the radiocarbon concentration in the atmosphere was not constant but varied by up to 1% over the last 400 years (de Vries Reference de Vries1958). This study was expanded on by an intercomparison exercise undertaken by the laboratories of Cambridge, Copenhagen, and Heidelberg on 25 sequoia tree rings sampled at 50-year intervals from AD 659 to AD 1859 (Willis et al. Reference Willis, Tauber and Munnich1960). They reported a variation of 1.5% and observed that there appeared to be a periodicity of 150–200 years superimposed on a longer period (Figure 1). That same year the University of Pennsylvania laboratory published results for a number of dendrochronologically dated tree rings (mostly Douglas fir) and some well-constrained archaeological samples relative to the average value of 39 oak tree rings ranging from AD 1747 to 1849 (Ralph and Stuckenrath Reference Ralph and Stuckenrath1960). They found three time periods (AD 1687, 1297 BC, and 1925 BC) with significant deviations from the oak average. Stuiver (Reference Stuiver1961) noted a correlation between solar activity and the 14C of known age tree rings which was strengthened by further measurements (Stuiver Reference Stuiver1965; Suess Reference Suess1965). Stuiver and Suess (Reference Stuiver and Suess1966) analyzed known age tree rings from European oak, North American Douglas fir and sequoia at La Jolla and Yale relative to the oxalic acid standard provided by the U.S. National Bureau of Standards (NBS). They published tables of radiocarbon corrections and what amounts to the first radiocarbon calibration curve extending from AD 1000 to 1800 (Figure 2).

Figure 1 Radiocarbon activities of Sequoia tree rings from AD 1200–1859 relative to the activity of a tree ring from 1859 (from Willis et al. Reference Willis, Tauber and Munnich1960: Fig. 1)

Figure 2 Radiocarbon ages compared to true ages for AD 1000–1800 (from Stuiver and Suess Reference Stuiver and Suess1966: Fig. 1)

Precise 14C measurements on tree rings from Europe, North and South America, South Africa, and India were made at the University of Groningen to investigate the possibility of geographic differences in atmospheric 14C (Lerman et al. Reference Lerman, Mook, Vogel and Olsson1970). They found that the Northern Hemisphere values were similar for all locations, which supported the radiocarbon dating assumption of a well-mixed atmosphere. However, there measurements showed that Southern Hemisphere samples from around AD 1835 were 4.5 ± 1‰ depleted compared to Northern Hemisphere samples. They also provided the first Southern Hemisphere calibration curve from measurements of Argentine trees from AD 1400–1950.

The construction of a bristlecone pine chronology at the University of Arizona Laboratory of Tree Ring Research (Ferguson Reference Ferguson1968) made it possible to extend radiocarbon calibration. Suess published calibration corrections from 4100 BC to 1500 BC based on bristlecone pine measured at La Jolla (Suess Reference Suess1967). The calibration was extended with further measurements of bristlecone pine from 5400 BC to the present (Suess Reference Suess1970) and was widely used by archaeologists (Ottaway and Ottaway Reference Ottaway and Ottaway1972). As noted in comments by Renfrew, calibration had a tremendous impact on the interpretation of the dates of prehistoric Europe (Bermingham and Renfrew Reference Bermingham and Renfrew1972).

Suess’s calibration curve was not overwhelming accepted as a number of researchers did not believe that the fine structure or “wiggles” were real (Burleigh et al. Reference Burleigh, Longworth and Wainwright1972; Switsur Reference Switsur1973). The fact that Suess said he had drawn the curves with “cosmic schwung” to demonstrate the “most probable character” of the curve did not instil confidence (Adams Reference Adams1973; Olsson Reference Olsson2009). Switsur (Reference Switsur1973) compared the calibration tables presented by the University of Arizona (Damon et al. Reference Damon, Ferguson, Long and Wallick1974) and the University of Pennsylvania (Ralph et al. Reference Ralph, Michael and Han1974) at the 8th International Conference on Radiocarbon Dating at Lower Hutt City, New Zealand, and found that the difference in corrected ages was fairly small. He therefore provided calibration corrections from the averages of the Arizona and Pennsylvania measurements. These corrections had what was felt to be an advantage over the Suess calibration corrections as they produced only one calibrated age for each radiocarbon measurement. Suess argued “nature is not necessarily as simple as we would wish and an attempt to draw a smooth curve amongst the measured points is probably the wrong approach” (Switsur Reference Switsur1973).

Renfrew and Clark (Reference Renfrew and Clark1974) reviewed all the curves that had been published to date, as well as a formula for calibration (Wendland and Donley Reference Wendland and Donley1971) and found these were all statistically unsatisfactory. Clark (Reference Clark1975) then examined all the data and excluded duplicates as well as some data that did not use the NBS oxalic acid standard. He found that the variability between replicates of the same tree rings in different laboratories was in excess of the reported errors which generally only included counting statistics. Clark used a spline function to construct a curve from the data with the standard error determined from the replicates and a smoothing parameter determined by a cross-validation technique. The resulting curve was smoother than that of Suess (Reference Suess1970) but not as smooth as Switsur’s (Reference Switsur1973). Indeed, Clark concluded that the “kinks and wobbles” in Suess’s calibration curve are “not justified statistically.” Suess responded that there was no statistical justification necessary for kinks, whereas Clark responded that his method had been tested and would detect “kinks” if they were there and not swamped by measurement errors (Suess and Clark Reference Suess and Clark1976). No international agreement on a calibration curve could be reached at the Ninth International Conference Los Angeles and La Jolla in 1976. It is no wonder that archaeologists began to doubt the validity of radiocarbon dating and calibration (Pilcher and Baillie Reference Pilcher and Baillie1978).

Meanwhile German and Irish oak dendrochronologies, though mostly still floating (i.e., not dendrochronologically linked to known calendar ages), were under development (Becker Reference Becker1979; Pilcher and Baillie Reference Pilcher and Baillie1978). Suess presented measurements of floating German oak and California bristlecone pine to show that the 14C “wiggles” in them were synchronous and of the same amplitude within uncertainty (Suess Reference Suess1979). High-precision gas proportional counters were developed in Groningen with an overall precision of 1.3‰ for modern carbon (Tans and Mook Reference Tans and Mook1978) and in Seattle with a precision of 1.5–2‰ for a Pacific coast tree-ring series dating back to 1200 AD (Stuiver et al. Reference Stuiver, Robinson and Yang1979). A 500-yr-long section of the Southern German Neolithic floating chronology oak was measured in Groningen (de Jong et al. Reference de Jong, Mook and Becker1979) which confirmed the existence of the Suess-type wiggles. Pearson (Reference Pearson1979) developed high-precision liquid scintillation counters and made a thorough study of the source of errors in the data. With careful controls he was able to produce 14C data on samples of Irish oak with an overall uncertainty of 2.5‰ (ca. 20 14C yr).

In 1979, a meeting was held at Tucson, Arizona, with the primary objective being to construct a radiocarbon calibration correction table using only 14C analyses funded by the U.S. National Science Foundation (Damon et al. Reference Damon, Lerman, Long, Bannister, Klein and Linick1980). A method for producing the radiocarbon calibrations was agreed based on multistage linear regression previously presented by Ralph and Klein (Reference Ralph and Klein1978). The workshop report also stated that “the most important accomplishment of the Workshop was the participants’ universal agreement to accept a single calibration curve based on the composite data set.” The participants also recommended establishing an International Calibration Committee to extend the workshop to include all available data worldwide. Statistical analysis results from the workshop were reported at the 10th International Radiocarbon Conference in Bern (Klein et al. Reference Klein, Lerman, Damon and Timothy1980). The calibration curve and tables constructed as agreed were published two years later (Klein et al. Reference Klein, Lerman, Damon and Ralph1982).

A number of new radiocarbon measurements were also presented at the 10th International Radiocarbon conference. Pearson (Reference Pearson1980) reported dates on 20-ring blocks of Irish and Scottish oak trees from AD 1510 to AD 1840 using liquid scintillation counters which were in agreement with those from Stuiver (Reference Stuiver1978) on North American Douglas fir measured using gas proportional counters. A visual match of an older floating Irish tree-ring sequence with that of the Suess bristlecone pine provided an approximate age range of 1810–3535 BC. The data exhibited a saw tooth pattern with a period of 110 to 180 years, similar to that seen in the German oak (de Jong et al. Reference de Jong, Mook and Becker1979) with which there was a short overlap. Pearson concluded, however, that the variability in the bristlecone data (Suess Reference Suess1979) was too high to identify wiggles. Bruns et al. (Reference Bruns, Munnich and Becker1980) measured 1–3 ring samples of Southern German Oak from AD 200-800 with gas counters at the University of Heidelberg. Using Fourier analysis, they determined a periodicity of 150 to 180 years similar to that seen by de Jong et al. (Reference de Jong, Mook and Becker1979) and de Jong and Mook (Reference de Jong and Mook1980) in tree rings from several millennia earlier. Likewise, after removing a sine curve with a period of 12,100 years and applying a smoothing spline, Fourier analysis showed a significant 200-yr period in the bristlecone pine data (Suess Reference Suess1980).

The theoretical framework for the Suess “wiggles” was set out by Stuiver and Quay (Reference Stuiver and Quay1980). They used a carbon reservoir model to calculate the 14C production rate from precise 14C measurements of Pacific coast tree rings. The calculated production rates were then compared to those derived from neutron flux measurements and geomagnetic indices and to historical records of sunspot numbers and auroral indices. This work showed that the natural atmospheric 14C changes were related to Earth’s geomagnetic field variation and changes in solar wind shielding properties. In later work, ocean circulation and climate changes were also shown to influence atmospheric 14C (Stuiver et al. Reference Stuiver, Braziunas, Becker and Kromer1991).

The Klein et al. (Reference Klein, Lerman, Damon and Ralph1982) calibration curve was based on dendrochronologically dated trees, primarily bristlecone pine and sequoia of 20 or fewer rings measured at the Universities of Groningen, California at La Jolla, Pennsylvania, and Yale. No other selection criteria were applied and no adjustment was made for systematic differences between laboratories (Klein et al. Reference Klein, Lerman, Damon and Timothy1980), however, the uncertainty from the interlaboratory standardization was included in the calibration uncertainty. The data were weighted by their measurement uncertainty and an additional term to account for increasing uncertainty with sample age and scaled logarithmically. A polynomial regression was done to derive the long period of the curve on the order of a few thousand years. Next a piecewise regression was done on the residuals around the polynomial in intervals of 500 years (“shingles”) overlapping by 250 years and Fourier analysis was done to determine the short period variability. A composite calibration function with uncertainty was determined from these. A section of the curve from AD 600–1950 is shown in Figure 3 with 90% confidence limits for radiocarbon errors of 20, 100, 200, and 300 14C yr.

Figure 3 A section of the calibration curve from AD 600–1950 with 90% confidence limits for radiocarbon errors of 20, 100, 200, and 300 14C yr (from Klein et al. Reference Klein, Lerman, Damon and Ralph1982: Fig. 3A).

Stuiver produced a high-precision decadal calibration curve for AD 1–1950 based on Douglas Fir from the Pacific Northwest and Sequoia from California (Stuiver Reference Stuiver1982). Measurement errors were ±12 14C yr, which were increased to ±18 14C yr with an error multiplier derived from comparison of a set of 30 pairs of contemporaneous samples from different trees. The difference between the Seattle data and that from Belfast (Pearson Reference Pearson1980) for this period was 4.4 14C years with mean age difference of 2 ± 3 14C yr between the complete data sets of 53 sample pairs (AD 955–1840). Offsets varied between 27 for German oak measured and La Jolla and Douglas fir in Seattle, between 37 and 55 14C yr for La Jolla bristlecone pine versus Seattle sequoia and Seattle Douglas fir, respectively. Preliminary measurements from Heidelberg on German oak were offset 58 14C yr from Seattle sequoia.

At the business meeting of the 11th International Radiocarbon Conference in Seattle, 1982, the International Calibration Committee set up by Damon et al. (Reference Damon, Lerman, Long, Bannister, Klein and Linick1980) “was discharged with thanks for their useful work and a smaller committee with W. G. Mook as chairman was re-constituted for the 1982–1985 interval” (Stuiver Reference Stuiver1983).

THE STUIVER YEARS

The 1986 Calibration Curves

The 12th International Radiocarbon Conference, Trondheim, 1985, saw a proliferation of calibration datasets and curves. The bi-decadal curves from 2500 BC to the present (Pearson and Stuiver Reference Pearson and Stuiver1986; Stuiver and Pearson Reference Stuiver and Pearson1986) were officially recommended at the business meeting (Mook Reference Mook1986), the first time the radiocarbon community had come to an agreement on radiocarbon calibration. These curves, which were based on measurements of Irish oak (Pilcher et al. Reference Pilcher, Baillie, Schmidt and Becker1984) and trees from the western U.S., were published in a special calibration issue that also included a decadal curve for the same interval based on the Seattle measurements only (Stuiver and Becker Reference Stuiver and Becker1986). Additional measurements on German oak, both dendrochronologically dated and floating, and bristlecone pine were published in the calibration issue (Stuiver Reference Stuiver1986 and references therein). Stuiver (Reference Stuiver1986) also suggested the possibility of extending the calibration curve with varved sediment data from Lake of the Clouds.

These decadal and bidecadal data, with the exception of the 14C data from varved sediment and the older floating German oak chronology, were combined with a simple weighted average of the data within each bidecade, rather than using more complicated statistical processes, to create the curve atm20 for calibration of terrestrial samples from 7210 cal BC–AD 1950 using the computer program CALIB (Stuiver and Reimer Reference Stuiver and Reimer1986). An ocean-atmosphere box model developed by Oeschger et al. (Reference Oeschger, Siegenthaler, Schotterer and Gugelmann1975) was used to construct a “global” marine radiocarbon calibration curve using the atmospheric curve as input age (Stuiver et al. Reference Stuiver, Pearson and Braziunas1986). The concept of ΔR, the difference between the “global” ocean radiocarbon and that of a specific region was introduced and a table of ΔR values given based on known age, pre-bomb samples for use with the marine curve.

A comparison of different calibration methods and programs, using the 1986 datasets, was undertaken by Aitchinson et al. (1989). The theoretical basis for combining archaeological information and radiocarbon dates through a Bayesian calibration model was developed by Buck et al. (Reference Buck, Kenworthy, Litton and Smith1991) and the first versions of the software packages OxCal (Bronk Ramsey Reference Bronk Ramsey1995) and BCal (Buck et al. Reference Buck, Christen and James1999) were made available in due course. Both BCal and OxCal utilize the ratified IntCal curves as do age-depth modeling programs such as BChron (Parnell et al. Reference Parnell, Haslett, Allen, Buck and Huntley2008) and BACON (Blaauw and Christen Reference Blaauw and Christen2011).

The 1993 Update

The calibration curves were corrected and extended and a probability method of calibration (Stuiver and Reimer Reference Stuiver and Reimer1989) was added to the CALIB program (Stuiver and Reimer Reference Stuiver and Reimer1993). A correction to earlier Seattle data had to be made due to a change in lab procedure that introduced radon into the samples which produced alpha particles during decay (Stuiver and Becker Reference Stuiver and Becker1993). The radon correction was based on the difference between the first and last day of counting for hundreds of samples. The Belfast data was also adjusted to account for “the variation of efficiency with time” in standard counts (Pearson and Qua Reference Pearson and Qua1993). Additional data of German oak measured by Heidelberg and floating German pine with a tentative match to the oak (Kromer and Becker Reference Kromer and Becker1993) were also included in the bidecadal curve. Adjustments for laboratory offsets were also made. For the first time, U-Th and 14C dated coral (Bard et al. Reference Bard, Arnold, Fairbanks and Hamelin1993) were included with a 400-yr reservoir offset to extend the curve with a smoothing spline to nearly 22,000 cal BP. A ratification vote for the bidecadal IntCal93 and the corresponding Marine93 curves was not sought and the curves were not generally accepted in Europe (F. G. McCormac, pers. comm.)

IntCal98 and Marine98

In 1996 a workshop was held at the Wissenschaftsforum, Heidelberg, Germany, to address aspects of high-precision radiocarbon calibration (Kromer et al. Reference Kromer, Ambers, Baillie, Damon, Hesshaimer, Hofmann, Jöris, Levin, Manning, Mccormac, van der Plicht, Spurk, Stuiver and Weninger1996). With regards to calibration data, two main points were made. An intercomparison of the Hohenheim and the Gottingen German oak chronology (Leuschner and Delorme Reference Leuschner and Delorme1988; Becker Reference Becker1993) revealed that 41 years were missing from the published Hohenheim chronology (Spurk et al. Reference Spurk, Friedrich, Hofmann, Remmele, Frenzel, Leuschner and Kromer1998). The previous link between the German oak and Preboreal Pine Chronology (Becker Reference Becker1993; Kromer and Becker Reference Kromer and Becker1993) was also now discarded so the pine chronology was left floating. Gerry McCormac presented evidence to support the original Irish oak 14C data (Pearson et al. Reference Pearson, Pilcher, Baillie, Corbett and Qua1986) as opposed to the revised dates (Pearson et al. Reference Pearson, Becker and Qua1993) used in the 1993 calibration curves. Further potential calibration datasets were presented at the 16th International Radiocarbon Conference in Groningen in 1997. Among these were U-Th dated coral datasets from Barbados, Mururoa and Tahiti (Bard et al. Reference Bard, Arnold, Hamelin, Tisnerat-Laborde and Cabioch1998) and from Vanuatu (Burr et al. Reference Burr, Beck, Taylor, Recy, Edwards, Cabioch, Correge, Donahue and O’Malley1998) and foraminifera from varved Cariaco Basin sediments (Hughen et al. Reference Hughen, Overpeck, Lehman, Kashgarian and Southon1998). The agreement of the high resolution Cariaco Basin data (Hughen et al. Reference Hughen, Overpeck, Lehman, Kashgarian and Southon1998) with the Barbados coral data (Bard et al. Reference Bard, Arnold, Hamelin, Tisnerat-Laborde and Cabioch1998) convinced Minze Stuiver to include varve data in the calibration curve for the first time. At the conference a few scientists expressed their discontent to me about having their data included in the calibration curve without direct involvement. At my suggestion they went to talk to Minze Stuiver about it and he agreed that they would become co-authors on the resulting publication. The international part of the curve name “IntCal” became a reality.

A few other corrections were made to previously published data. Comparison of tree-ring samples remeasured in the Seattle lab showed that the radon correction made in 1993 needed to be reduced by 50% (Stuiver et al. Reference Stuiver, Reimer and Braziunas1998b). A weak section was found on re-examination of the older part of the German oak chronology and additional trees were found to bridge this section resulting in a shift of 54 yr to the older trees (Spurk et al. Reference Spurk, Friedrich, Hofmann, Remmele, Frenzel, Leuschner and Kromer1998). In addition, a 14C match of the floating German pine chronology to the German oak provided a tentative link and extended the chronology to 11,857 cal BP with an error estimate for the link of 20 cal years (Kromer and Spurk Reference Kromer and Spurk1998). Other new tree-ring data included in the IntCal98 curve were German oak and pine, Sitka spruce and Douglas fir (Stuiver et al. Reference Stuiver, Reimer, Bard, Beck, Burr, Hughen, Kromer, McCormac, van der Plicht and Spurk1998a), Irish oak (McCormac et al. Reference McCormac, Hogg, Higham, Baillie, Palmer, Xiong, Pilcher, Brown and Hoper1998a, Reference McCormac, Hogg, Higham, Lynch-Stieglitz, Broecker, Baillie, Palmer, Xiong, Pilcher, Brown and Hoper1998b) and previously published data (Kromer et al. Reference Kromer, Rhein, Bruns, Schochfischer, Munnich, Stuiver and Becker1986; Pearson et al. Reference Pearson, Becker and Qua1993; Vogel and van der Plicht Reference Vogel and van der Plicht1993).

Previously, 14C ages on dendrodated wood covering only a few years (sub-decadal) were disregarded, but a different approach was used for IntCal98. The “decadal” values were obtained by averaging all full decadal and sub-decadal results. Samples measured on 20-yr blocks of wood were also included by assigning this age to two decades with the standard deviation multiplied by 1.4. Laboratory error multipliers (k values) and offsets were estimated for each tree-ring dataset based on comparison of results from samples of identical calendar age.

Marine reservoir offsets (R) were estimated for the coral data from Bard et al. (Reference Bard, Arnold, Hamelin, Tisnerat-Laborde and Cabioch1998), Burr et al. (Reference Burr, Beck, Taylor, Recy, Edwards, Cabioch, Correge, Donahue and O’Malley1998), and Edwards et al. (Reference Edwards, Beck, Burr, Donahue, Chappell, Bloom, Druffel and Taylor1993) by comparing the marine 14C ages to overlapping tree-ring data. The weighted mean R value was 414 ± 31 14C yr (n =12) for 10,000–8000 cal BP and 509 ± 25 14C yr (n = 21) for 12,000–10,000 cal BP. While the possibility of reservoir age changes over time was acknowledged it was decided to use a value of 500 14C yr for the Late Glacial tropical surface ocean. The coral data were found to have higher 14C variability than the tree rings. To compensate for this the standard deviation of coral 14C ages was based on a 2σ error in the coral 14C measurement and an error multiplier of 1.3 was applied. Foraminifera 14C ages from the Cariaco Basin decreased by 500 and 400 14C yr for, respectively, 12–10 and 10–8 ka cal BP were matched to the IntCal98 tree-ring data to update the floating varve chronology. The observed scatter in the 14C ages was similar to that of the tree rings with an error multiplier of k = 1.3. The IntCal98 curve from 12,000 to 24,000 was generated from the reservoir corrected marine data using a spline with minimum smoothing (Stuiver et al Reference Stuiver, Reimer, Bard, Beck, Burr, Hughen, Kromer, McCormac, van der Plicht and Spurk1998a).

Marine98 curve, which extended from 0 to 12,000 cal BP, was generated from the IntCal98 curve using an atmosphere-ocean box diffusion model (Stuiver et al. Reference Stuiver, Reimer and Braziunas1998b) with model parameters as discussed in Stuiver and Braziunas (Reference Stuiver and Braziunas1993).

THE INTCAL WORKING GROUP

2004 Calibration Curves

When Minze Stuiver retired, I continued the calibration effort with encouragement from Gerry McCormac. Together we received a Leverhulme Trust networking grant to form the IntCal Working Group (IWG) in 2002. The grant funded meetings in Belfast and Woods Hole and a research assistant. Previous and potential data providers, as well as an archaeologist and a statistician, were invited to join the group. At the first meeting in Belfast in 2002 the IWG set out criteria for inclusion of data into the planned IntCal04 calibration curve (Reimer et al. Reference Reimer, Hughen, Guilderson, McCormac, Baillie, Bard, Barratt, Beck, Buck, Damon, Friedrich, Kromer, Ramsey, Reimer, Remmele, Southon, Stuiver and van der Plicht2002). It was agreed that a new statistical method was needed, in part to get rid of what Ron Reimer christened the “pig-in-the-python” around 15,000 cal BP (Figure 4). At the second meeting in Woods Hole in 2003, a random walk model was chosen to account for the uncertainty in the calendar ages of samples as well as the uncertainty in their 14C ages to calculate the curve (Buck and Blackwell Reference Buck and Blackwell2004). Constant regional offsets were used to incorporate the marine data. The IntCal04 and Marine04 curves were extended to 26,000 cal BP (Hughen et al. Reference Hughen, Baillie, Bard, Beck, Bertrand, Blackwell, Buck, Burr, Cutler, Damon, Edwards, Fairbanks, Friedrich, Guilderson, Kromer, Mccormac, Manning, Ramsey, Reimer, Reimer, Remmele, Southon, Stuiver, Talamo, Taylor, van der Plicht and Weyhenmeyer2004; Reimer et al. Reference Reimer, Baillie, Bard, Bayliss, Beck, Bertrand, Blackwell, Buck, Burr, Cutler, Damon, Edwards, Fairbanks, Friedrich, Guilderson, Hogg, Hughen, Kromer, McCormac, Manning, Ramsey, Reimer, Remmele, Southon, Stuiver, Talamo, Taylor, van der Plicht and Weyhenmeyer2004) but no recommendation was given beyond that because of the disparities between datasets. However, a predictive curve called “NotCal04” (to signify that it was not for calibration purposes) was plotted based on a random effects model (van der Plicht et al. Reference van der Plicht, Beck, Bard, Baillie, Blackwell, Buck, Friedrich, Guilderson, Hughen, Kromer, Mccormac, Ramsey, Reimer, Reimer, Remmele, Richards, Southon, Stuiver and Weyhenmeyer2004). This paper also demonstrated the erroneous conclusions that could result from attempts to calibrate radiocarbon ages in this time frame with alternative datasets, such as a smooth curve proposed by van Andel (Reference van Andel1998) based on geomagnetic records, and caution was urged. Nonetheless the NotCal04 curve was used for calibration of Palaeolithic radiocarbon dates such as those from Chauvet cave (e.g., Mellars Reference Mellars2006). van Andel (Reference van Andel2005) argued that alternative datasets and calibration methods such as CalPal (Jöris and Weninger Reference Jöris and Weninger1998) should not be discouraged due to the need to calibrate radiocarbon ages beyond the 26,000 cal BP. He also disputed the importance of the IntCal consensus calibration to the community. Fortunately, this issue was resolved by further updates to the IntCal curves extending to close to the limits of radiocarbon dating and increasing community involvement in the process. In addition to IntCal04 and Marine04, a preliminary calibration curve for the Southern Hemisphere atmosphere, SHCal02 (McCormac et al. Reference McCormac, Reimer, Hogg, Higham, Baillie, Palmer and Stuiver2002) was extended to 11,000 cal BP using the tree-ring portion of the IntCal04 curve with a modeled offset between the Northern and Southern Hemisphere (McCormac et al. Reference McCormac, Hogg, Blackwell, Buck, Higham and Reimer2004).

Figure 4 IntCal98 calibration curve from 12,000 to 18,000 cal BP with 2σ error envelope. The “pig in the python” is around 15,000 cal BP (as published in Reimer et al. Reference Reimer, Hughen, Guilderson, McCormac, Baillie, Bard, Barratt, Beck, Buck, Damon, Friedrich, Kromer, Ramsey, Reimer, Remmele, Southon, Stuiver and van der Plicht2002: Figure 1).

2009 Updates

In 2009 the IntCal09 and Marine09 calibration curves (Reimer et al. Reference Reimer, Baillie, Bard, Bayliss, Beck, Blackwell, Ramsey, Buck, Burr, Edwards, Friedrich, Grootes, Guilderson, Hajdas, Heaton, Hogg, Hughen, Kaiser, Kromer, Mccormac, Manning, Reimer, Richards, Southon, Talamo, Turney, van der Plicht and Weyhenmeyer2009) were extended to 50,000 cal BP using new data from U-Th dated coral (Fairbanks et al. Reference Fairbanks, Mortlock, Chiu, Cao, Kaplan, Guilderson, Fairbanks, Bloom, Grootes and Nadeau2005) and from marine sediment with timescales transferred through climatic correlation to the U-Th dated Hulu Cave speleothem (Wang et al. Reference Wang, Cheng, Edwards, An, Wu, Shen and Dorale2001). These “tie-pointed” records included the non-varved portion of Cariaco Basin (Hughen et al. Reference Hughen, Southon, Lehman, Bertrand and Turnbull2006) and the Iberian margin (Bard et al. Reference Bard, Ménot-Combes and Rostek2004). The Marine09 curve was taken from the Marine04 curve for 0–12,500 cal BP whereas the older portion was derived from the IntCal09 curve with a constant 405 14C yr offset. A Bayesian evaluation of the interhemispheric offset was undertaken but SHCal04 was not updated at this time (Hogg et al. Reference Hogg, Ramsey, Turney and Palmer2009). From a statistical perspective, the random walk model for curve construction was considerably developed for the 2009 update. The IntCal09 extension was the first time a fully Bayesian approach, achieved via Markov Chain Monte Carlo (MCMC), had been used to estimate the calibration curve (Blackwell and Buck Reference Blackwell and Buck2008; Heaton et al. Reference Heaton, Blackwell and Buck2009).

2013 Calibration Curves

For the 2013 curves, three scientists were elected by the community to the IntCal Oversight Committee to provide independent expertise and assessment. New tree-ring measurements were included in the IntCal13 curve most importantly from the floating Late Glacial Pine chronology (Schaub et al. Reference Schaub, Buntgen, Kaiser, Kromer, Talamo, Andersen and Rasmussen2008) anchored by a 14C wiggle-match as described by Hua et al. (Reference Hua, Barbetti, Fink, Kaiser, Friedrich, Kromer, Levchenko, Zoppi, Smith and Bertuch2009) which extended to 14,200 cal BP. The density of data beyond the tree rings was greatly increased compared to IntCal09 due to the incorporation of 14C measurements from speleothems from the Bahamas (Beck et al. Reference Beck, Richards, Edwards, Silverman, Smart, Donahue, Hererra-Osterheld, Burr, Calsoyas, Jull and Biddulph2001; Hoffmann et al. Reference Hoffmann, Beck, Richards, Smart, Singarayer, Ketchmark and Hawkesworth2010) and Hulu Cave, China (Southon et al. Reference Southon, Noronha, Cheng, Edwards and Wang2012) and from terrestrial plant macrofossils from Lake Suigetsu, Japan (Kitagawa and van der Plicht Reference Kitagawa and van der Plicht1998; Bronk Ramsey et al. Reference Bronk Ramsey, Staff, Bryant, Brock, Kitagawa, van der Plicht, Schlolaut, Marshall, Brauer, Lamb, Payne, Tarasov, Haraguchi, Gotanda, Yonenobu, Yokoyama, Tada and Nakagawa2012). There were also new coral data from Tahiti (Durand et al. Reference Durand, Deschamps, Bard, Hamelin, Camoin, Thomas, Henderson, Yokoyama and Matsuzaki2013) and a “tie-pointed” record from the Pakistan margin (Bard et al. Reference Bard, Ménot, Rostek, Licari, Böning, Edwards, Cheng, Wang and Heaton2013; Heaton et al. Reference Heaton, Bard and Hughen2013). Marine reservoir ages were determined for each carbonate dataset from the offset with overlapping tree-ring data, where possible, and kept constant beyond the end of the tree rings. Comparison of Cariaco Basin and Barbados data to overlapping tree-ring data suggested an abrupt drop in reservoir ages during both the Younger Dryas and Heinrich Stadial, therefore those data were not used in IntCal13 and Marine13.

The random walk model for curve construction was developed further (Niu et al. Reference Niu, Heaton, Blackwell and Buck2013) to account for the dependencies in the calendar age estimates of the samples within both the Lake Suigetsu and the “tie-pointed” records; and the wiggle-matched tree-ring measurements. Discrepancies between the diverse datasets used to construct the curve between 13,900–50,000 cal BP were accommodated by allowing for a potential additive increase in the uncertainty of the data. This resulted in a smoother character for the IntCal13 mean curve in this older time period. The increase in the volume of data for 2013, together with the complex dependencies in the calendar age chronologies of the constituent records, meant that the random walk approach had reached its practical limit. It was decided that, for the future, a new method that enabled more flexibility and greater potential to investigate the effect of specific modeling and dataset choices was needed.

The SHCal13 curve was constructed with additional Holocene tree-ring data including a Tasmanian Huon pine floating chronology from 12,679 to 12,073 cal BP (Hua et al. Reference Hua, Barbetti, Fink, Kaiser, Friedrich, Kromer, Levchenko, Zoppi, Smith and Bertuch2009). For portions of the curve without Southern Hemisphere tree-ring measurements, the IntCal13 curve with an atmospheric radiocarbon offset between the Northern and Southern hemispheres (N-S offset) of 43 ± 23 yr was used, which extended SHCal13 to 50,000 cal BP (Hogg et al. Reference Hogg, Hua, Blackwell, Niu, Buck, Guilderson, Heaton, Palmer, Reimer, Reimer, Turney and Zimmerman2013).

2020 Calibration Curves

In 2013 IntCal focus groups were instigated to promote wider engagement with the IWG and provide additional expertise on various topics. As requested by the IWG, the statistical approach to curve construction was completely redesigned for the 2020 calibration curves (Heaton et al. Reference Heaton, Blaauw, Blackwell, Bronk Ramsey, Reimer and Scott2020a). The new methodology utilized Bayesian splines. MCMC was still used to provide the curves but the new Bayesian spline approach allowed much faster implementation and more reliable curve estimation than had been possible with the previous random walk.

The spline method of construction creates a calibration curve based upon a trade-off between the quality of the curve’s fit to the calibration samples, and its roughness/wiggliness. This penalty on roughness aims to prevent over-fitting and the introduction of potential spurious variability. Fit to the calibration data was assessed in F14C where measurement uncertainties are symmetric, while roughness was penalized in Δ14C space. The method was refined to include an additive error term to account for any potential additional variability in observed tree-ring 14C measurements and to explicitly recognize rapid atmospheric 14C change events such as the AD 774–775 event (Miyake et al. Reference Miyake, Nagaya, Masuda and Nakamura2012).

In the older portion of the curve, from 13,900–55,000 cal BP, much work was done by the group to resolve the discrepancies between the various speleothem, ocean, and lacustrine datasets (Butzin et al. Reference Butzin, Köhler and Lohmann2017; Heaton et al. Reference Heaton, Blaauw, Blackwell, Bronk Ramsey, Reimer and Scott2020a; Hughen and Heaton Reference Hughen and Heaton2020). To deal with any discrepancies that may remain, “heavy-tailed” errors were used to reduce the influence of possible outliers.

A total of 220 Holocene tree-ring datasets were considered for possible inclusion in the IntCal20 curve and screened for suitability (Reimer et al. Reference Reimer, Austin, Bard, Bayliss, Blackwell, Bronk Ramsey, Butzin, Cheng, Edwards, Friedrich, Grootes, Guilderson, Hajdas, Heaton, Hogg, Hughen, Kromer, Manning, Muscheler, Palmer, Pearson, van der Plicht, Reimer, Richards, Scott, Southon, Turney, Wacker, Adolphi, Büntgen, Capano, Fahrni, Fogtmann-Schulz, Friedrich, Köhler, Kudsk, Miyake, Olsen, Reinig, Sakamoto, Sookdeo and Talamo2020). Improvements in the linkages of the European Preboreal Pine and the Swiss chronologies were reinforced by the inclusion of single-year subfossil pine trees from the French Alps and allowed a fully atmospheric record to ca. 13,910 cal BP. The Hulu Cave speleothem dataset, which had been extended to 53,900 cal BP (Cheng et al. Reference Cheng, Edwards, Southon, Matsumoto, Feinberg, Sinha, Zhou, Li, Li and Xu2018), provided the backbone for the curve beyond the dendrochronologically dated and 14C wiggle-matched tree-ring measurements ca. 14,000 cal BP. Floating tree rings from Northern Italy and kauri from New Zealand provided higher resolution details to the curves further back in time. The Lake Suigetsu macrofossil data on an extended and improved varve timescale were also included (Bronk Ramsey et al. Reference Bronk Ramsey, Heaton, Schlolaut, Staff, Bryant, Brauer, Lamb, Marshall and Nakagawa2020). Because many of the U-Th coral datasets older than 25,000 cal BP exhibited rather high variability, these data were excluded from IntCal20. For marine data, rather than using a constant offset from the atmospheric curve as had been done in the past, the Hamburg Large Scale Geostrophic model was used to estimate the large variations in the marine reservoir offset over time for each region (Butzin et al. Reference Butzin, Heaton, Köhler and Lohmann2020) using the Hulu 14C record as input. For further details of corrections and adjustments to the datasets, see Reimer et al. (Reference Reimer, Austin, Bard, Bayliss, Blackwell, Bronk Ramsey, Butzin, Cheng, Edwards, Friedrich, Grootes, Guilderson, Hajdas, Heaton, Hogg, Hughen, Kromer, Manning, Muscheler, Palmer, Pearson, van der Plicht, Reimer, Richards, Scott, Southon, Turney, Wacker, Adolphi, Büntgen, Capano, Fahrni, Fogtmann-Schulz, Friedrich, Köhler, Kudsk, Miyake, Olsen, Reinig, Sakamoto, Sookdeo and Talamo2020) and references therein. The IntCal20 curve is more detailed than IntCal13, in part because of increased data coverage and in part due to the efforts made by the group to develop the methods that synthesized the diverse data. In particular, the work to resolve the differences between the various datasets from 13,900–55,000 cal BP has meant we are hopefully better able to discern common signal in this period. While there are some other differences between IntCal20 and IntCal13 in the Holocene, the major changes are from the Late Glacial through to the end of the curve at 55,000 cal BP. For example, radiocarbon ages calibrated with the IntCal20 curve between 34,000 and 42,000 cal BP calibrated ages may be up to 700 years older than with IntCal13 and between 42,000 to 50,000 cal BP these may be more than 1000 years younger. This has major consequences for scientific studies in this time range, for example examining the duration of the overlap between Neanderthals and anatomically modern humans in Europe (Bard et al. Reference Bard, Heaton, Talamo, Kromer, Reimer and Reimer2020).

The SHCal20 curve was reinforced with a number of new tree-ring datasets (Hogg et al. Reference Hogg, Heaton, Hua, Palmer, Turney, Southon, Bayliss, Blackwell, Boswijk, Bronk Ramsey, Pearson, Petchey, Reimer, Reimer and Wacker2020). For time periods where direct Southern Hemisphere measurements were not available, the IntCal20 curve was used with an N-S offset model (Heaton et al. Reference Heaton, Blaauw, Blackwell, Bronk Ramsey, Reimer and Scott2020a).

For the Marine20 surface ocean calibration curve the BICYCLE global carbon cycle model, which incorporates a number of components of the carbon cycle (Köhler et al. Reference Köhler, Muscheler and Fischer2006), was revised to allow atmospheric CO2 to be specified by an ice core-based reconstruction and atmospheric Δ14C by the IntCal20 curve (Heaton et al. Reference Heaton, Köhler, Butzin, Bard, Reimer, Austin, Bronk Ramsey, Grootes, Hughen, Kromer, Reimer, Adkins, Burke, Cook, Olsen and Skinner2020b). The resulting model produced a curve with variable marine reservoir offsets similar to the Hamburg Large Scale Geostrophic model. Marine20 is intended to represent the non-polar surface ocean as it has no provision for ice cover, however, Marine20 should be suitable for calibration of polar samples during the Holocene. It is worth noting that the previous model used for marine calibration curves (e.g., Marine86 through Marine13) also did not include ice cover. Because the parameters used in the model differ from those used in the previous marine curves, the average offset between the ocean and the atmosphere is larger. ΔR values used for regional corrections must be calculated with Marine20. Recalculated known age, pre-bomb ΔR values are given in the Marine Radiocarbon Reservoir Database (calib.org/marine) whereas ΔR values for paired marine and terrestrial samples can be calculated with the deltar software (Reimer and Reimer Reference Reimer and Reimer2017) available on the website.

OUTLOOK

The IntCal Working Group, which is now led by Christopher Bronk Ramsey, has already begun discussions on the next update. Kauri measurements through the Laschamp geomagnetic excursion ca. 41–42 ka cal BP have recently been published (Cooper et al. Reference Cooper, Turney, Palmer, Hogg, Mcglone, Wilmshurst, Lorrey, Heaton, Russell, Mccracken, Anet, Rozanov, Friedel, Suter, Peter, Muscheler, Adolphi, Dosseto, Faith, Fenwick, Fogwill, Hughen, Lipson, Liu, Nowaczyk, Rainsley, Bronk Ramsey, Sebastianelli, Souilmi, Stevenson, Thomas, Tobler and Zech2021) and there are new single year tree-ring data coming out soon. It is expected that the statistical methods will not have major modifications for the next update.

In the longer term, methods for calibrating sets of radiocarbon dates with a large subset of the possible realizations generated by the MCMC method are likely to improve “wiggle-matched” dates. As more tree-ring data from different regions with low measurement uncertainty become available it may be possible to provide region-specific terrestrial calibration curves. For the marine curve, the global carbon cycle model could be improved by incorporating time-varying changes to climatic parameters based on palaeoceanographic and palaeoclimatic records, and eventually, with sufficient data, we may be able provide location-specific curves, in particular for the polar regions (Heaton et al. Reference Heaton, Köhler, Butzin, Bard, Reimer, Austin, Bronk Ramsey, Grootes, Hughen, Kromer, Reimer, Adkins, Burke, Cook, Olsen and Skinner2020b).

ACKNOWLEDGMENTS

I would like to thank the Leverhulme Trust for a networking grant which enabled the initial IntCal Working Group meetings, the UK Natural Environment Research Council NE/E018807/1 for funding research leading to IntCal09 and IntCal13, and IGBP PAGES (Past Global Changes) for providing funding for two palaeoscientists on the IntCal Oversight Committee to participate in meetings leading to IntCal13. Pieter Grootes provided valuable suggestions on the manuscript. I am grateful to Tim Heaton for input on the statistical methods used in IntCal09 through IntCal20.

References

REFERENCES

Adams, RM. 1973. Review of: Radiocarbon Variations and Absolute Chronology by Ingrid U. Olsson. Journal of Near Eastern Studies 32:253256.Google Scholar
Aitchison, TC, Leese, M, Michczynska, DJ, Mook, WG, Otlet, RL, Ottaway, BS, Pazdur, MF, van der Plicht, J, Reimer, PJ, Robinson, SW, Scott, EM, Stuiver, M, Weninger, B. 1989. A comparison of methods used for the calibration of radiocarbon-dates. Radiocarbon 31:846864.Google Scholar
Arnold, JR, Libby, WF. 1949. Age determinations by radiocarbon content: checks with samples of known age. Science 110:678680.CrossRefGoogle ScholarPubMed
Bard, E, Arnold, M, Fairbanks, RG, Hamelin, B. 1993. 230Th, 234U and 14C ages obtained by mass-spectrometry on corals. Radiocarbon 35:191199.CrossRefGoogle Scholar
Bard, E, Arnold, M, Hamelin, B, Tisnerat-Laborde, N, Cabioch, G. 1998. Radiocarbon calibration by means of mass spectrometric 230Th/234U and 14C ages of corals: an updated database including samples from Barbados, Mururoa and Tahiti. Radiocarbon 40:10851092.CrossRefGoogle Scholar
Bard, E, Heaton, TJ, Talamo, S, Kromer, B, Reimer, RW, Reimer, PJ, 2020. Extended dilation of the radiocarbon time scale between 40,000 and 48,000 y BP and the overlap between Neanderthals and Homo sapiens. Proceedings of the National Academy of Sciences: 202012307.CrossRefGoogle Scholar
Bard, E, Ménot-Combes, G, Rostek, F. 2004. Present status of radiocarbon calibration and comparison records based on Polynesian corals and Iberian Margin sediments. Radiocarbon 46:11891202.10.1017/S0033822200033087CrossRefGoogle Scholar
Bard, E, Ménot, G, Rostek, F, Licari, L, Böning, P, Edwards, RL, Cheng, H, Wang, Y, Heaton, TJ. 2013. Radiocarbon calibration/comparison records based on marine sediments from the Pakistan and Iberian margins. Radiocarbon 55:19992019.CrossRefGoogle Scholar
Beck, JW, Richards, DA, Edwards, RL, Silverman, BW, Smart, PL, Donahue, DJ, Hererra-Osterheld, S, Burr, GS, Calsoyas, L, Jull, AJT, Biddulph, D. 2001. Extremely large variations of atmospheric 14C concentration during the last glacial period. Science 292:24532458.CrossRefGoogle ScholarPubMed
Becker, B. 1979. Holocene tree ring series from southern central Europe for archaeologic dating, radiocarbon calibration, and stable isotope analysis. In: Berger R, Suess HE, editors. Proceedings of the 9th International Radiocarbon Dating Conference. Berkeley/Los Angeles: University of California Press. p. 554–565.CrossRefGoogle Scholar
Becker, B. 1993. An 11,000-year German oak and pine dendrochronology for radiocarbon calibration. Radiocarbon 35:201213.CrossRefGoogle Scholar
Bermingham, A, Renfrew, C. 1972. The history of C14 calibration. Antiquity 46:151.Google Scholar
Blaauw, M, Christen, JA. 2011. Flexible paleoclimate age-depth models using an autoregressive gamma process. Bayesian Analysis 6:457474.CrossRefGoogle Scholar
Blackwell, PG, Buck, CE. 2008. Estimating radiocarbon calibration curves. Bayesian Analysis 3(2):225268.CrossRefGoogle Scholar
Bronk Ramsey, C. 1995. Radiocarbon calibration and analysis of stratigraphy: the OxCal program. Radiocarbon 37:425430.CrossRefGoogle Scholar
Bronk Ramsey, C, Heaton, TJ, Schlolaut, G, Staff, RA, Bryant, CL, Brauer, A, Lamb, HF, Marshall, MH, Nakagawa, T. 2020. Reanalysis of the atmospheric radiocarbon calibration record from Lake Suigetsu, Japan. Radiocarbon 62:989999.CrossRefGoogle Scholar
Bronk Ramsey, C, Staff, RA, Bryant, CL, Brock, F, Kitagawa, H, van der Plicht, J, Schlolaut, G, Marshall, MH, Brauer, A, Lamb, HF, Payne, RL, Tarasov, PE, Haraguchi, T, Gotanda, K, Yonenobu, H, Yokoyama, Y, Tada, R, Nakagawa, T. 2012. A complete terrestrial radiocarbon record for 11.2 to 52.8 kyr BP. Science 338:370374.CrossRefGoogle Scholar
Bruns, M, Munnich, KO, Becker, B. 1980. Natural-radiocarbon variations from AD 200 to 800. Radiocarbon 22:273277.CrossRefGoogle Scholar
Buck, CE, Blackwell, PG. 2004. Formal statistical models for estimating radiocarbon calibration curves. Radiocarbon 46:1093–102.CrossRefGoogle Scholar
Buck, CE, Christen, JA, James, GN. 1999. BCal: an on-line Bayesian radiocarbon calibration tool. Internet Archaeology 7.Google Scholar
Buck, CE, Kenworthy, JB, Litton, CD, Smith, AFM. 1991. Combining archaeological and radiocarbon information: a Bayesian approach to calibration. Antiquity 65:808821.CrossRefGoogle Scholar
Burleigh, R, Longworth, IH, Wainwright, GJ. 1972. Relative and absolute dating of four late Neolithic enclosures: an exercise in the interpretation of radiocarbon determinations. Proceedings of the Prehistoric Society 38:389407.CrossRefGoogle Scholar
Burr, GS, Beck, JW, Taylor, FW, Recy, J, Edwards, RL, Cabioch, G, Correge, T, Donahue, DJ, O’Malley, JM. 1998. A high-resolution radiocarbon calibration between 11,700 and 12,400 calendar years BP derived from 230Th ages of corals from Espiritu Santo Island, Vanuatu. Radiocarbon 40:10931105.CrossRefGoogle Scholar
Butzin, M, Köhler, P, Lohmann, G. 2017. Marine radiocarbon reservoir age simulations for the past 50,000 years. Geophysical Research Letters 44:84738480.CrossRefGoogle Scholar
Butzin, M, Heaton, T, Köhler, P, Lohmann, G. 2020. A short note on marine reservoir age simulations used in intCal20. Radiocarbon 62:865871. doi: 10.1017/RDC.2020.9.CrossRefGoogle Scholar
Cheng, H, Edwards, RL, Southon, J, Matsumoto, K, Feinberg, JM, Sinha, A, Zhou, W, Li, H, Li, X, Xu, Y. 2018. Atmospheric 14C/12C changes during the last glacial period from Hulu Cave. Science 362:12931297.CrossRefGoogle ScholarPubMed
Clark, RM. 1975. A calibration curve for radiocarbon dates. Antiquity 49:251266.CrossRefGoogle Scholar
Cooper, A, Turney, CSM, Palmer, J, Hogg, A, Mcglone, M, Wilmshurst, J, Lorrey, AM, Heaton, TJ, Russell, JM, Mccracken, K, Anet, JG, Rozanov, E, Friedel, M, Suter, I, Peter, T, Muscheler, R, Adolphi, F, Dosseto, A, Faith, JT, Fenwick, P, Fogwill, CJ, Hughen, K, Lipson, M, Liu, J, Nowaczyk, N, Rainsley, E, Bronk Ramsey, C, Sebastianelli, P, Souilmi, Y, Stevenson, J, Thomas, Z, Tobler, R, Zech, R. 2021. A global environmental crisis 42,000 years ago. Science 371:811818.CrossRefGoogle ScholarPubMed
Damon, PE, Ferguson, CW, Long, A, Wallick, ET. 1974. Dendrochronologic calibration of the radiocarbon time scale. American Antiquity 39(2):350366.CrossRefGoogle Scholar
Damon, PE, Lerman, JC, Long, A, Bannister, B, Klein, J, Linick, T. 1980. Report on the workshop on the calibration of the radiocarbon dating time scale. Radiocarbon 22:947949.CrossRefGoogle Scholar
de Jong, A, Mook, W. 1980. Medium-term atmospheric 14C variations. Radiocarbon 22:267272.CrossRefGoogle Scholar
de Jong, A, Mook, W, Becker, B. 1979. Confirmation of the Suess wiggles: 3200–3700 BC. Nature 280:4849.CrossRefGoogle Scholar
de Vries, H. 1958. Variation in concentration of radiocarbon with time and location on earth. Proceedings of the Koninklijke Nederlandse Akademie Van Wetenschappen Series B-Palaeontology Geology Physics Chemistry Anthropology B61:94102.Google Scholar
Durand, N, Deschamps, P, Bard, E, Hamelin, B, Camoin, G, Thomas, AL, Henderson, GM, Yokoyama, Y, Matsuzaki, H. 2013. Comparison of 14C and U-Th ages in corals from IODP# 310 cores offshore Tahiti. Radiocarbon 55:19471974.CrossRefGoogle Scholar
Edwards, RL, Beck, JW, Burr, GS, Donahue, DJ, Chappell, JMA, Bloom, AL, Druffel, ERM, Taylor, FW. 1993. A large drop in atmospheric 14C/12C and reduced melting in the Younger Dryas, documented with Th-230 ages of corals. Science 260:962968.CrossRefGoogle Scholar
Fairbanks, RG, Mortlock, RA, Chiu, TC, Cao, L, Kaplan, A, Guilderson, TP, Fairbanks, TW, Bloom, AL, Grootes, PM, Nadeau, MJ. 2005. Radiocarbon calibration curve spanning 0 to 50,000 years BP based on paired 230Th/234U/238U and 14C dates on pristine corals. Quaternary Science Reviews 24:17811796.CrossRefGoogle Scholar
Ferguson, CW. 1968. Bristlecone pine: science and esthetics: a 7100-year tree-ring chronology aids scientists; old trees draw visitors to California mountains. Science 159:839846.Google ScholarPubMed
Heaton, T, Bard, E, Hughen, K. 2013. Elastic tie-pointing—transferring chronologies between records via a Gaussian Process. Radiocarbon 55:19751997. doi: 10.2458/azu_js_rc.55.17777.CrossRefGoogle Scholar
Heaton, T, Blackwell, P, Buck, C. 2009. A Bayesian approach to the estimation of radiocarbon calibration curves: the IntCal09 methodology. Radiocarbon 51:11511164. doi: 10.1017/S0033822200034214.CrossRefGoogle Scholar
Heaton, TJ, Blaauw, M, Blackwell, PG, Bronk Ramsey, C, Reimer, PJ, Scott, EM. 2020a. The IntCal20 approach to radiocarbon calibration curve construction: a new methodology using Bayesian splines and errors-in-variables. Radiocarbon 62:821863. doi: 10.1017/RDC.2020.46.CrossRefGoogle Scholar
Heaton, TJ, Köhler, P, Butzin, M, Bard, E, Reimer, RW, Austin, WEN, Bronk Ramsey, C, Grootes, PM, Hughen, KA, Kromer, B, Reimer, PJ, Adkins, J, Burke, A, Cook, MS, Olsen, J, Skinner, LC. 2020b. Marine20—the marine radiocarbon age calibration curve (0–55,000 cal BP). Radiocarbon 62:779820. doi: 10.1017/RDC.2020.68.CrossRefGoogle Scholar
Hoffmann, DL, Beck, JW, Richards, DA, Smart, PL, Singarayer, JS, Ketchmark, T, Hawkesworth, CJ. 2010. Towards radiocarbon calibration beyond 28 ka using speleothems from the Bahamas. Earth and Planetary Science Letters 289:110.CrossRefGoogle Scholar
Hogg, A, Ramsey, CB, Turney, C, Palmer, J. 2009. Bayesian evaluation of the Southern Hemisphere radiocarbon offset during the Holocene. Radiocarbon 51:11651176.CrossRefGoogle Scholar
Hogg, AG, Heaton, TJ, Hua, Q, Palmer, JG, Turney, CSM, Southon, J, Bayliss, A, Blackwell, PG, Boswijk, G, Bronk Ramsey, C, Pearson, C, Petchey, F, Reimer, P, Reimer, R, Wacker, L. 2020. SHCal20 Southern Hemisphere calibration, 0–55,000 years cal BP. Radiocarbon 62:759778.CrossRefGoogle Scholar
Hogg, AG, Hua, Q, Blackwell, PG, Niu, M, Buck, CE, Guilderson, TP, Heaton, TJ, Palmer, JG, Reimer, PJ, Reimer, RW, Turney, CSM, Zimmerman, SRH. 2013. SHCal13 Southern Hemisphere calibration, 0–50,000 years cal BP. Radiocarbon 55:18891903.CrossRefGoogle Scholar
Hua, Q, Barbetti, M, Fink, D, Kaiser, KF, Friedrich, M, Kromer, B, Levchenko, VA, Zoppi, U, Smith, AM, Bertuch, F. 2009. Atmospheric 14C variations derived from tree rings during the early Younger Dryas. Quaternary Science Reviews 28:29822990.CrossRefGoogle Scholar
Hughen, K, Heaton, T. 2020. Updated Cariaco Basin 14C calibration dataset from 0–60 cal kyr BP. Radiocarbon 62:10011043. doi: 10.1017/RDC.2020.53.CrossRefGoogle Scholar
Hughen, K, Southon, J, Lehman, S, Bertrand, C, Turnbull, J. 2006. Marine-derived 14C calibration and activity record for the past 50,000 years updated from the Cariaco Basin. Quaternary Science Reviews 25:32163227.CrossRefGoogle Scholar
Hughen, KA, Baillie, MGL, Bard, E, Beck, JW, Bertrand, CJH, Blackwell, PG, Buck, CE, Burr, GS, Cutler, KB, Damon, PE, Edwards, RL, Fairbanks, RG, Friedrich, M, Guilderson, TP, Kromer, B, Mccormac, G, Manning, S, Ramsey, CB, Reimer, PJ, Reimer, RW, Remmele, S, Southon, JR, Stuiver, M, Talamo, S, Taylor, FW, van der Plicht, J, Weyhenmeyer, CE. 2004. Marine04 marine radiocarbon age calibration, 0–26 cal kyr BP. Radiocarbon 46:10591086.CrossRefGoogle Scholar
Hughen, KA, Overpeck, JT, Lehman, SJ, Kashgarian, M, Southon, JR. 1998. A new 14C calibration data set for the last deglaciation based on marine varves. Radiocarbon 40: 483494.CrossRefGoogle Scholar
Jöris, O, Weninger, B. 1998. Extension of the 14C calibration curve to ca. 40,000 cal BC by synchronizing Greenland 18O/16O ice core records and North Atlantic foraminifera profiles: a comparison with U/Th coral data. Radiocarbon 40:495504.CrossRefGoogle Scholar
Kitagawa, H, van der Plicht, J. 1998. A 40,000-year varve chronology from Lake Suigetsu, Japan: extension of the 14C calibration curve. Radiocarbon 40:505515.CrossRefGoogle Scholar
Klein, J, Lerman, JC, Damon, PE, Timothy, L. 1980. Radiocarbon concentration in the atmosphere; 8000-year record of variations in tree rings; first results of a USA workshop. Radiocarbon 22:950961.CrossRefGoogle Scholar
Klein, J, Lerman, JC, Damon, PE, Ralph, EK. 1982. Calibration of radiocarbon-dates—tables based on the consensus data of the workshop on calibrating the radiocarbon time scale. Radiocarbon 24:103150.CrossRefGoogle Scholar
Köhler, P, Muscheler, R, Fischer, H. 2006. A model-based interpretation of low-frequency changes in the carbon cycle during the last 120,000 years and its implications for the reconstruction of atmospheric Δ14C. Geochemistry, Geophysics, Geosystems 7.Google Scholar
Kromer, B, Ambers, J, Baillie, MGL, Damon, PE, Hesshaimer, V, Hofmann, J, Jöris, O, Levin, I, Manning, SW, Mccormac, FG, van der Plicht, J, Spurk, M, Stuiver, M, Weninger, B. 1996. Report: summary of the workshop “Aspects of high-precision radiocarbon calibration”. Radiocarbon 38:607610.CrossRefGoogle Scholar
Kromer, B, Becker, B. 1993. German Oak and pine 14C calibration, 7200–9439 BC. Radiocarbon 35:125135.CrossRefGoogle Scholar
Kromer, B, Rhein, M, Bruns, M, Schochfischer, H, Munnich, KO, Stuiver, M, Becker, B. 1986. Radiocarbon calibration data for the 6th to the 8th millennia BC. Radiocarbon 28:954960.CrossRefGoogle Scholar
Kromer, B, Spurk, M. 1998. Revision and tentative extension of the tree-ring based 14C calibration, 9200–11,855 cal BP. Radiocarbon 40:11171125.CrossRefGoogle Scholar
Lerman, JC, Mook, WG, Vogel, JC. 1970. 14C in tree rings from different localities. In: Olsson, IU, ed. Radiocarbon variations and absolute chronology. New York: John Wiley and Sons. p. 257299.Google Scholar
Leuschner, H-H, Delorme, A. 1988. Tree-ring work in Göttingen: absolute oak chronologies back to 6255 BC. Proceedings of the Wood and archaeology: acts of the European Symposium held at Louvain-la-Neuve, October 1987. Bois et archéologie: actes du Symposium Européen tenu à Louvain-la-Neuve, Octobre 1987, 1988. p. 123–132.Google Scholar
McCormac, FG, Hogg, AG, Blackwell, PG, Buck, CE, Higham, TFG, Reimer, PJ. 2004. SHCal04 Southern Hemisphere calibration, 0–11.0 cal kyr BP. Radiocarbon 46:10871092.CrossRefGoogle Scholar
McCormac, FG, Hogg, AG, Higham, TFG, Baillie, MGL, Palmer, JG, Xiong, LM, Pilcher, JR, Brown, D, Hoper, ST. 1998a. Variations of radiocarbon in tree rings: Southern Hemisphere offset preliminary results. Radiocarbon 40:11531159.CrossRefGoogle Scholar
McCormac, FG, Hogg, AG, Higham, TFG, Lynch-Stieglitz, J, Broecker, WS, Baillie, MGL, Palmer, J, Xiong, L, Pilcher, JR, Brown, D, Hoper, ST. 1998b. Temporal variation in the interhemispheric 14C offset. Geophysical Research Letters 25:13211324.CrossRefGoogle Scholar
McCormac, FG, Reimer, PJ, Hogg, AG, Higham, TFG, Baillie, MGL, Palmer, J, Stuiver, M. 2002. Calibration of the radiocarbon time scale for the Southern Hemisphere: AD 1850–950. Radiocarbon 44:641651.CrossRefGoogle Scholar
Mellars, P. 2006. A new radiocarbon revolution and the dispersal of modern humans in Eurasia. Nature 439:931935.CrossRefGoogle ScholarPubMed
Miyake, F, Nagaya, K, Masuda, K, Nakamura, T. 2012. A signature of cosmic-ray increase in AD 774–775 from tree rings in Japan. Nature 486:240242.CrossRefGoogle ScholarPubMed
Mook, WG. 1986. Business meeting: recommendations/resolutions adopted by the Twelfth International Radiocarbon Conference. Radiocarbon 28:799.CrossRefGoogle Scholar
Niu, M, Heaton, T, Blackwell, P, Buck, C. 2013. The Bayesian approach to radiocarbon calibration curve estimation: the IntCal13, Marine13, and SHCal13 methodologies. Radiocarbon 55:19051922.CrossRefGoogle Scholar
Oeschger, H, Siegenthaler, U, Schotterer, U, Gugelmann, A. 1975. A box diffusion model to study the carbon dioxide exhange in nature. Tellus 27:168192.CrossRefGoogle Scholar
Olsson, IU. 2009. Radiocarbon dating history: early days, questions, and problems met. Radiocarbon 51:143.10.1017/S0033822200033695CrossRefGoogle Scholar
Ottaway, B, Ottaway, J. 1972. The Suess calibration curve and archaeological dating. Nature 239:512513.CrossRefGoogle Scholar
Parnell, AC, Haslett, J, Allen, JR, Buck, CE, Huntley, B. 2008. A flexible approach to assessing synchroneity of past events using Bayesian reconstructions of sedimentation history. Quaternary Science Reviews 27:18721885.CrossRefGoogle Scholar
Pearson, G, Pilcher, J, Baillie, M, Hillam, J. 1977. Absolute radiocarbon dating using a low altitude European tree-ring calibration. Nature 270:2528.CrossRefGoogle Scholar
Pearson, GW. 1979. Precise 14C measurement by liquid scintillation counting. Radiocarbon 21(1):121.CrossRefGoogle Scholar
Pearson, GW. 1980. High-precision radiocarbon dating by liquid scintillation-counting applied to radiocarbon time-scale calibration. Radiocarbon 22(2):337345.CrossRefGoogle Scholar
Pearson, GW, Becker, B, Qua, F. 1993. High-precision 14C measurement of German and Irish Oaks to show the natural 14C variations from 7890 to 5000 BC. Radiocarbon 35:93104.CrossRefGoogle Scholar
Pearson, GW, Pilcher, JR, Baillie, MGL, Corbett, DM, Qua, F. 1986. High-precision 14C measurement of Irish oaks to show the natural 14C variations from AD 1840 to 5210 BC. Radiocarbon 28:911934.CrossRefGoogle Scholar
Pearson, GW, Qua, F. 1993. High-precision 14C measurement of Irish oaks to show the natural 14C variations from AD 1840–5000 BC: a correction. Radiocarbon 35:105123.CrossRefGoogle Scholar
Pearson, GW, Stuiver, M. 1986. High-precision calibration of the radiocarbon time scale, 500–2500 BC. Radiocarbon 28:839862.CrossRefGoogle Scholar
Pilcher, JR, Baillie, MGL. 1978, Implications of a European radiocarbon calibration. Antiquity 52(206):217222.CrossRefGoogle Scholar
Pilcher, JR, Baillie, MGL, Schmidt, B, Becker, B. 1984. A 7,272-year tree-ring chronology for Western-Europe. Nature 312:150152.10.1038/312150a0CrossRefGoogle Scholar
Ralph, EK, Klein, J. 1978. Composite computer plots of 14C-dates for tree-ring dated bristlecone pine and Sequoia. In: Berger R, Suess HE, editors. Radiocarbon dating, the 9th International Radiocarbon Confernece. Proceedings. Berkeley/Los Angeles: University of California Press. p. 545–553.Google Scholar
Ralph, EK, Michael, HN, Han, MC. 1974. Radiocarbon dates and reality. Archaeology of Eastern North America 2(1):120.Google Scholar
Ralph, EK, Stuckenrath, R. 1960. Carbon-14 measurements of known age samples. Nature 188: 4746:185187.CrossRefGoogle Scholar
Reimer, PJ, Austin, WEN, Bard, E, Bayliss, A, Blackwell, PG, Bronk Ramsey, C, Butzin, M, Cheng, H, Edwards, RL, Friedrich, M, Grootes, PM, Guilderson, TP, Hajdas, I, Heaton, TJ, Hogg, AG, Hughen, KA, Kromer, B, Manning, SW, Muscheler, R, Palmer, JG, Pearson, C, van der Plicht, J, Reimer, RW, Richards, DA, Scott, EM, Southon, JR, Turney, CSM, Wacker, L, Adolphi, F, Büntgen, U, Capano, M, Fahrni, SM, Fogtmann-Schulz, A, Friedrich, R, Köhler, P, Kudsk, S, Miyake, F, Olsen, J, Reinig, F, Sakamoto, M, Sookdeo, A, Talamo, S. 2020. The IntCal20 Northern Hemisphere radiocarbon age calibration curve (0–55 cal kBP). Radiocarbon 62:725757.CrossRefGoogle Scholar
Reimer, PJ, Baillie, MGL, Bard, E, Bayliss, A, Beck, JW, Bertrand, CJH, Blackwell, PG, Buck, CE, Burr, GS, Cutler, KB, Damon, PE, Edwards, RL, Fairbanks, RG, Friedrich, M, Guilderson, TP, Hogg, AG, Hughen, KA, Kromer, B, McCormac, G, Manning, S, Ramsey, CB, Reimer, RW, Remmele, S, Southon, JR, Stuiver, M, Talamo, S, Taylor, FW, van der Plicht, J, Weyhenmeyer, CE. 2004. IntCal04 terrestrial radiocarbon age calibration, 0–26 cal kyr BP. Radiocarbon 46:10291058.Google Scholar
Reimer, PJ, Baillie, MGL, Bard, E, Bayliss, A, Beck, JW, Blackwell, PG, Ramsey, CB, Buck, CE, Burr, GS, Edwards, RL, Friedrich, M, Grootes, PM, Guilderson, TP, Hajdas, I, Heaton, TJ, Hogg, AG, Hughen, KA, Kaiser, KF, Kromer, B, Mccormac, FG, Manning, SW, Reimer, RW, Richards, DA, Southon, JR, Talamo, S, Turney, CSM, van der Plicht, J, Weyhenmeyer, CE. 2009. IntCal09 and Marine09 radiocarbon age calibration curves, 0–50,000 years cal BP. Radiocarbon 51:11111150.10.1017/S0033822200034202CrossRefGoogle Scholar
Reimer, PJ, Hughen, KA, Guilderson, TP, McCormac, G, Baillie, MGL, Bard, E, Barratt, P, Beck, JW, Buck, CE, Damon, PE, Friedrich, M, Kromer, B, Ramsey, CB, Reimer, RW, Remmele, S, Southon, JR, Stuiver, M, van der Plicht, J. 2002. Preliminary report of the first workshop of the IntCal04 radiocarbon calibration/comparison working group. Radiocarbon 44:653661.CrossRefGoogle Scholar
Reimer, RW, Reimer, PJ. 2017. An online application for ΔR calculation. Radiocarbon 59:16231627.CrossRefGoogle Scholar
Renfrew, C, Clark, RM. 1974. Problems of the radiocarbon calendar and its calibration. Archaeometry 16:518.CrossRefGoogle Scholar
Schaub, M, Buntgen, U, Kaiser, KF, Kromer, B, Talamo, S, Andersen, KK, Rasmussen, SO. 2008. Lateglacial environmental variability from Swiss tree rings. Quaternary Science Reviews 27:2941.CrossRefGoogle Scholar
Southon, J, Noronha, AL, Cheng, H, Edwards, RL, Wang, YJ. 2012. A high-resolution record of atmospheric C-14 based on Hulu Cave speleothem H82. Quaternary Science Reviews 33:3241.CrossRefGoogle Scholar
Spurk, M, Friedrich, M, Hofmann, J, Remmele, S, Frenzel, B, Leuschner, HH, Kromer, B. 1998. Revisions and extension of the Hohenheim oak and pine chronologies: new evidence about the timing of the Younger Dryas/Preboreal transition. Radiocarbon 40:11071116.10.1017/S0033822200019159CrossRefGoogle Scholar
Stuiver, M. 1961. Variations in radiocarbon concentration and sunspot activity. Journal of Geophysical Research 66:273276.CrossRefGoogle Scholar
Stuiver, M. 1965. Carbon-14 content of 18th-and 19th-century wood: variations correlated with sunspot activity. Science 149:533534.CrossRefGoogle Scholar
Stuiver, M. 1978. Radiocarbon timescale tested against magnetic and other dating methods. Nature 273(5660):271274.CrossRefGoogle Scholar
Stuiver, M. 1982. A high-precision calibration of the AD radiocarbon time scale. Radiocarbon 24(1):126.CrossRefGoogle Scholar
Stuiver, M. 1983. International agreements and the use of the new oxalic-acid standard. Radiocarbon 25:793795.CrossRefGoogle Scholar
Stuiver, M. 1986. Editorial comment. Radiocarbon 28(2B):iiiii.Google Scholar
Stuiver, M, Becker, B. 1986. High-precision decadal calibration of the radiocarbon time scale, AD 1950–2500 BC. Radiocarbon 28:863910.CrossRefGoogle Scholar
Stuiver, M, Becker, B. 1993. High-precision decadal calibration of the radiocarbon time scale, AD 1950–6000 BC. Radiocarbon 35:3565.CrossRefGoogle Scholar
Stuiver, M, Braziunas, TF. 1993. Modeling atmospheric 14C influences and 14C ages of marine samples to 10,000 BC. Radiocarbon 35:137189.CrossRefGoogle Scholar
Stuiver, M, Braziunas, TF, Becker, B, Kromer, B. 1991. Climatic, solar, oceanic, and geomagnetic influences on late-glacial and Holocene atmospheric 14C/12C change. Quaternary Research 35:124.CrossRefGoogle Scholar
Stuiver, M, Pearson, GW. 1986. High-precision calibration of the radiocarbon time scale, AD 1950–500 BC. Radiocarbon 28(2B):805838.CrossRefGoogle Scholar
Stuiver, M, Pearson, GW, Braziunas, T. 1986. Radiocarbon age calibration of marine samples back to 9000 cal yr BP. Radiocarbon 28:9801021.CrossRefGoogle Scholar
Stuiver, M, Quay, PD. 1980. Changes in atmospheric 14C attributed to a variable sun. Science 207:1119.CrossRefGoogle Scholar
Stuiver, M, Reimer, P. 1989. Histograms obtained from computerized radiocarbon age calibration. Radiocarbon 31:817–23.CrossRefGoogle Scholar
Stuiver, M, Reimer, PJ. 1986. A computer-program for radiocarbon age calibration. Radiocarbon 28:10221030.CrossRefGoogle Scholar
Stuiver, M, Reimer, PJ. 1993. Extended 14C data-base and revised Calib 3.0 14C age calibration program. Radiocarbon 35:215230.CrossRefGoogle Scholar
Stuiver, M, Reimer, PJ, Bard, E, Beck, JW, Burr, GS, Hughen, KA, Kromer, B, McCormac, G, van der Plicht, J, Spurk, M. 1998a. INTCAL98 radiocarbon age calibration, 24,000–0 cal BP. Radiocarbon 40:10411083.CrossRefGoogle Scholar
Stuiver, M, Reimer, PJ, Braziunas, TF. 1998b. High-precision radiocarbon age calibration for terrestrial and marine samples. Radiocarbon 40:11271151.CrossRefGoogle Scholar
Stuiver, M, Robinson, S, Yang, I. 1979. 14C dating up to 60,000 years BP with high efficiency proportional counters. In: Berger R, Suess HE, eds. Proceedings of the 9th International Radiocarbon Dating Conference. Berkeley/Los Angeles: University of California Press. p. 202–215.Google Scholar
Stuiver, M, Suess, HE. 1966. On the relationship between radiocarbon dates and true sample ages. Radiocarbon 8:534540.CrossRefGoogle Scholar
Suess, HE. 1965. Secular variations of the cosmic-ray produced carbon 14 in the atmosphere and their interpretations. Journal of Geophysical Research 70:5937–52.CrossRefGoogle Scholar
Suess, HE. 1967. Bristlecone pine calibration of the radiocarbon time scale from 4100 BC to 1500 BC. Proceedings of the Radioactive dating and methods of low-level counting. Monaco: International Atomic Energy Agency.Google Scholar
Suess, HE. 1970. Bristlecone pine calibration of the radiocarbon time-scale 5200 B.C. to the present. In: Olsson IU, editor. Proceedings of the Twelfth Nobel Symposium, Uppsala, Sweden. p. 303–311.Google Scholar
Suess, HE. 1979. The 14C level during the fourth and second half of the fifth millennium B.C. and the 14C calibration curve. In: Berger R, Suess HE, eds. Proceedings of the 9th International Radiocarbon Dating Conference. Berkeley/Los Angeles. University of California Press. p. 538–544.Google Scholar
Suess, HE. 1980. The radiocarbon record in tree rings of the last 8000 years. Radiocarbon 22(2):200209.CrossRefGoogle Scholar
Suess, H, Clark, R. 1976. A calibration curve for radiocarbon dates. Antiquity 50(197):61.Google Scholar
Switsur, V. 1973. The radiocarbon calendar recalibrated. Antiquity 47(186):131.Google Scholar
Tans, P, Mook, W. 1978. Design, construction and calibration of a high accuracy carbon-14 counting set up. Radiocarbon 21(1):2240.CrossRefGoogle Scholar
van Andel, TH. 1998. Middle and Upper Palaeolithic environments and the calibration of 14C dates beyond 10,000 BP. Antiquity 72:26.CrossRefGoogle Scholar
van Andel, TH. 2005. The ownership of time: approved 14C calibration or freedom of choice? Antiquity 79:944948.CrossRefGoogle Scholar
van der Plicht, J, Beck, JW, Bard, E, Baillie, MGL, Blackwell, PG, Buck, CE, Friedrich, M, Guilderson, TP, Hughen, KA, Kromer, B, Mccormac, FG, Ramsey, CB, Reimer, PJ, Reimer, RW, Remmele, S, Richards, DA, Southon, JR, Stuiver, M, Weyhenmeyer, CE. 2004. NotCal04—comparison/calibration 14C records 26–50 cal kyr BP. Radiocarbon 46:12251238.CrossRefGoogle Scholar
Vogel, JC, van der Plicht, J. 1993. Calibration curve for short-lived samples, 1900–3900 BC. Radiocarbon 35:8791.CrossRefGoogle Scholar
Wang, YJ, Cheng, H, Edwards, RL, An, ZS, Wu, JY, Shen, CC, Dorale, JA. 2001. A high-resolution absolute-dated Late Pleistocene monsoon record from Hulu Cave, China. Science 294:23452348.CrossRefGoogle ScholarPubMed
Wendland, WM, Donley, DL. 1971. Radiocarbon-calendar age relationship. Earth and Planetary Science Letters 11:135139.CrossRefGoogle Scholar
Willis, EH, Tauber, H, Munnich, KO. 1960. Variations in the atmospheric radiocarbon concentration over the past 1300 years. American Journal of Science Radiocarbon Supplement 2:14.CrossRefGoogle Scholar
Figure 0

Figure 1 Radiocarbon activities of Sequoia tree rings from AD 1200–1859 relative to the activity of a tree ring from 1859 (from Willis et al. 1960: Fig. 1)

Figure 1

Figure 2 Radiocarbon ages compared to true ages for AD 1000–1800 (from Stuiver and Suess 1966: Fig. 1)

Figure 2

Figure 3 A section of the calibration curve from AD 600–1950 with 90% confidence limits for radiocarbon errors of 20, 100, 200, and 300 14C yr (from Klein et al. 1982: Fig. 3A).

Figure 3

Figure 4 IntCal98 calibration curve from 12,000 to 18,000 cal BP with 2σ error envelope. The “pig in the python” is around 15,000 cal BP (as published in Reimer et al. 2002: Figure 1).