Hostname: page-component-788cddb947-w95db Total loading time: 0 Render date: 2024-10-19T19:40:49.218Z Has data issue: false hasContentIssue false

RADIOCARBON CONCENTRATION IN SUB-ANNUAL TREE RINGS FROM POLAND AROUND 660 BCE

Published online by Cambridge University Press:  29 September 2023

Andrzej Z Rakowski*
Affiliation:
Silesian University of Technology, Konarskiego 22B str., 44-100 Gliwice, Poland
Jacek Pawlyta
Affiliation:
AGH University of Science and Technology, Mickiewicza Av. 30, 30-059 Krakow, Poland
Hiroko Miyahara
Affiliation:
Humanities and Sciences/Museum Careers, Musashino Art Universally, 1-736 Ogawa-cho, Kodaira, Tokyo, 187-8505, Japan
Marek Krąpiec
Affiliation:
AGH University of Science and Technology, Mickiewicza Av. 30, 30-059 Krakow, Poland
Mihály Molnár
Affiliation:
Hertelendi Laboratory of Environmental Studies, Institute of Nuclear Research of the Hungarian Academy of Sciences (ATOMKI), H-4026 Debrecen, Hungary Isotoptech Zrt., H-4025 Debrecen, Hungary
Damian Wiktorowski
Affiliation:
AGH University of Science and Technology, Mickiewicza Av. 30, 30-059 Krakow, Poland
Masayo Minami
Affiliation:
Institute for Space-Earth Environmental Research, Nagoya University, Furo-cho, Chikusa-ku, Nagoya, 464-8601, Japan
*
*Corresponding author. Email: arakowski@polsl.pl
Rights & Permissions [Opens in a new window]

Abstract

The article presents results of measurements of radiocarbon (14C) concentration in sub-annual dendrochronologically dated tree rings of English oak (Quercus robur L.) from Grabie village near Kraków (southern Poland). Samples of early wood (EW) and late wood (LW) spanning the years 664–658 BCE. α-cellulose was extracted from each sample and their radiocarbon content was measured at the ATOMKI laboratory in Debrecen, Hungary. The EW and LW data confirm a prolonged increase in Δ14C values around 665–663 as was observed by Park et al. (2017), Rakowski et al. (2019), or Sakurai et al. (2020). In addition, we found that this event may consist of two relatively small events, as was proposed by Sakurai et al. (2020). Based on obtained in this and previous study data we estimate that the occurrence of the two events were between 665 and 664 BCE (Rakowski et al. 2019), and in late spring of 663 BCE (May–June, before beginning of LW formation).

Type
Research Article
Copyright
© The Author(s), 2023. Published by Cambridge University Press on behalf of University of Arizona

INTRODUCTION

Radiocarbon (14C) is produced by the nuclear reaction 14N(n, p)14C in the Earth’s atmosphere, and its production depends on the cosmic ray flux reaching the Earth’s atmosphere (Usoskin and Kovaltsov Reference Usoskin and Kovaltsov2012). On average the 14C production rate is of the order of 1.64 and 1.88 atoms·cm–2·s–1 for the modern time and for the preindustrial epoch, respectively (Kovaltsov et al. Reference Kovaltsov, Mishev and Usoskin2012). Modulation of the cosmic ray flux occurs because of the variability of the intensity of the superposition of solar, interplanetary, and terrestrial magnetic fields. On the other hand, changes in the intensity of cosmic flux due to occurrence of events such as solar proton events (SPEs), supernovae explosion (SNe) or gamma-ray burst (GRBs), can increase the production rate for a short time.

Freshly produced atmospheric radiocarbon is oxidized in two steps, first with O2 to carbon oxide, and then to carbon dioxide with hydroxyl radical. The 14CO2 formed in this way becomes part of the global carbon cycle (GCC) and can be exchanged between the atmosphere, the biosphere, and the ocean in various physical and chemical processes. Through photosynthesis radiocarbon is incorporated into plant tissue, including i.e., tree rings, which can be then used in reconstruction of its concentration in the past. Dendrochronologically dated tree rings were used to reconstruct the past changes in atmospheric radiocarbon concentration due to changes in its production, but also changes in the carbon cycle for the last 12,500 yr (van der Plicht Reference van der Plicht2007; Hogg et al. Reference Hogg, Heaton, Hua, Palmer, Turney and Wacker2020; Reimer et al. Reference Reimer, Austin, Bard, Bayliss, Blackwell, Ramsey, Butzin, Cheng, Edwards and Friedrich2020). The development of accelerator mass spectrometry (AMS) measurement techniques enabled the increase the resolution and precision of measurements, while reducing the weight of the sample needed for analysis (i.e., Synal et al. Reference Synal, Stocker and Suter2007; Molnar et al. Reference Molnar, Rinyu, Veres, Seiler, Wacker and Synal2016).

Miyake et al. (Reference Miyake, Nagaya, Masuda and Nakamura2012) were the first to confirm the occurrence of event (henceforth M12) of rapid increase in 14C concentration, occurring within one year. Between 774 and 775 CE, a significant increase of 12‰ in Δ14C value has been reported (Miyake et al. Reference Miyake, Nagaya, Masuda and Nakamura2012). To explain the causes of the M12 event, a four-box carbon cycle model was used, yielding the estimated production rate of 6·108 atoms·cm–2·s–1. According to these authors, this value may suggest a giant solar proton events (SEPs) or supernovae explosion (SNe) as a possible cause of this event. However, considering no evidence of SNe in this time frame, the SEPs is most possible reason of this increase. The production value has been recalculated by Usoskin et al. (Reference Usoskin, Kromer, Ludlow, Beer, Friedrich, Kovaltsov, Solanki and Wacker2013), using a five-box carbon cycle model, and the obtained result is around 4 times smaller than what was presented by Miyake et al. (Reference Miyake, Nagaya, Masuda and Nakamura2012), i.e., 1.3 ± 0.2·108 atoms·cm–2·s–1. In Miyake et al. (Reference Miyake, Masuda, Hakozaki, Nakamura, Tokanai, Kato, Kimura and Mitsutani2014), a new calculation was presented, and the production rate obtained for this event was consistent with that reported by Usoskin et al. (Reference Usoskin, Kromer, Ludlow, Beer, Friedrich, Kovaltsov, Solanki and Wacker2013). The assumption of SEPs as a possible cause of M12 was confirmed by Mekhaldi et al. (Reference Mekhaldi, Muscheler, Adolphi, Aldahan, Beer, McConnell, Possnert, Sigl, Svensson and Synal2015), who recorded a comparable rapid increase in 10Be and 36Cl in ice cores from Antarctica and Greenland around 775 CE and 994 CE. GRBs with typical energy spectra and fluxes could produce significant amounts of 14C and 36Cl, however, the 10Be production rate would not be significantly affected (Pavlov et al. Reference Pavlov, Blinov, Konstantinov, Ostryakov, Vasilyev, Vdovina and Volkov2013; Mekhaldi et al. Reference Mekhaldi, Muscheler, Adolphi, Aldahan, Beer, McConnell, Possnert, Sigl, Svensson and Synal2015). Thus, we conclude a large SEP, or series of SEPs, are a cause of the discussed Miyake events.

The 774–775 CE event has been confirmed independently by several authors (Jull et al. Reference Jull, Panyushkina, Lange, Kukarskih, Myglan, Clark, Salzer, Burr and Leavitt2014; Güttler et al. Reference Güttler, Adolphi, Beer, Bleicher, Boswijk, Christl, Hogg, Palmer, Vockenhuber, Wacker and Wunder2015; Rakowski et al. Reference Rakowski, Krąpiec, Huels, Pawlyta, Dreves and Meadows2015; Büntgen et al. Reference Büntgen, Wacker, Galván, Arnold, Arseneault, Baillie, Beer, Bernabei, Bleicher and Boswijk2018) in dendrochronologically dated annual tree rings from different places around the world, which indicates its global character.

Rapid increases in radiocarbon concentration have also been confirmed at different periods. An increase of 11.3‰ was recorded between years 993 and 994 CE in samples of Japanese cedar (Criptomeria japonica; Miyake et al. Reference Miyake, Masuda and Nakamura2013), and Hinoki cypress (Chamaecyparis obtusa) (Miyake et al. Reference Miyake, Masuda, Hakozaki, Nakamura, Tokanai, Kato, Kimura and Mitsutani2014) and have been confirmed by other authors (Fogtman-Schultz et al. Reference Fogtmann-Schulz, Ostbo, Nielsen, Olsen, Karoff and Knudsen2017; Büntgen et al. Reference Büntgen, Wacker, Galván, Arnold, Arseneault, Baillie, Beer, Bernabei, Bleicher and Boswijk2018; Rakowski et al. Reference Rakowski, Krąpiec, Huels, Pawlyta and Boudin2018). Another SEP have been noted around 660 BCE and confirmed by Park et al. (Reference Park, Southon, Fahrni, Creasman and Mewaldt2017), Rakowski et al. (Reference Rakowski, Krąpiec, Huels, Pawlyta, Hamann and Wiktorowski2019), and Sakurai et al. (Reference Sakurai, Tokanai, Miyake, Horiuchi, Masuda, Miyahara, Ohyama, Sakamoto, Mitsurani and Moriya2020). Similarly, to M12, an increase in production rate of 10Be and 36Cl has been noted, confirming the solar origin of this event (O’Hare et al. Reference O’Hare, Mekhaldi, Adolphi, Reisbeck, Aldahan, Anderberg, Beer, Christl, Fahrni and Synal2019).

Other sudden increases in 14C have been observed between 814–815 BCE (Jull et al. Reference Jull, Panyushkina, Miyake, Masuda, Nakamura, Mitsutani, Lange, Cruz, Baisan and Janovics2018), 1054–1055 CE (Terrasi et al. Reference Tarrasi, Marzaioli, Bouompane, Passariello, Capano, Helama, Oinonen, Nojd, Uusitalo and Jull2020), or a series of events between 1261–1262, 1268–1269, and 1279–1280 CE (Miyahara et al. Reference Miyahara, Tokanai, Moriya, Takeyama, Sakurai, Ohyama, Horuchi and Hotta2022). The magnitudes of those events were significantly lower in comparison to the M12 event, and, without information about changes in production rate of 10Be and 36Cl it is difficult to say about the origin, and if those could also be classified as Miyake events. The oldest observed rapid increases in radiocarbon concentration so far identified occurred in mid-Holocene around 5480 BCE (Miyake et al. Reference Miyake, Jull, Panyushkina, Wacker, Salzer, Baisan, Lange, Cruz, Masuda and Nakamura2017), 5259 BCE, and 7176 BCE (Brehm et al. Reference Brehm, Christl, Adolphi, Muscheler, Synal, Mekhaldi, Paleari, Leuschner, Bayliss and Nicolussi2022). The 7176 BCE has magnitude similar to the largest SEP event in 775 CE (Brehm et al. Reference Brehm, Christl, Adolphi, Muscheler, Synal, Mekhaldi, Paleari, Leuschner, Bayliss and Nicolussi2022). The solar origin of the two oldest events has been confirmed by increase in production rate of 10Be and 36Cl (7176 BCE—Paleari et al. Reference Paleari, Mekhaldi, Adolphi, Christl, Vockenhuber, Gautschi, Beer, Brehm, Erhardt and Synal2022; 5480 BCE—Kanzawa et al. Reference Kanzawa, Miyake, Horiuchi, Sasa, Takano and Matsumura2021).

Such sudden and intense increase in the concentration of radiocarbon found application in the precise radiocarbon dating with one-year resolution. This method was used by Wacker et al. (Reference Wacker, Guttler, Goll, Hurni, Synal and Walti2014) to precisely date the cutting year of a timber in the historical Holy Cross chapel of the convent St. John the Baptist in Val Münstair in Switzerland. Also, in the works of Oppenheimer et al. (Reference Oppenheimer, Wacker, Xu, Galván, Stoffel, Guillet, Corona, Sigl, Cosmo and Hajdas2017), Hakozaki et al. (Reference Hakozaki, Miyake, Nakamura, Kimura, Masuda and Okuno2018), Krąpiec et al. (Reference Krąpiec, Rakowski, Pawlyta, Wiktorowski and Bolka2020), Philipsen et al. (Reference Philipsen, Feveile, Olsen and Sindbaek2022), and Kuitems et al. (Reference Kuitems, Wallace and Lindsay2022), these events were used to precisely determine the age based on changes in the concentration of radiocarbon in the sequence of annual tree rings.

The results presented in this article are the continuation of our previous work (Rakowski et al. Reference Rakowski, Krąpiec, Huels, Pawlyta, Hamann and Wiktorowski2019) and include new measurements of the radiocarbon concentration in sub-annual rings. The goal of this study is to use the information of carbon isotopic composition in sub-annual tree rings to estimate beginning of occurrence of this event.

SAMPLE AND METHODS

For this study, sub-fossil oaks (Quercus robur L.) were recovered from a gravel pit near the Vistula River in the village of Grabie, near Kraków (southern Poland; Figure 1). Around 100 slices of oak trunks were taken and after preparation were used for dendrochronological study. Measurements were made with 0.01 mm accuracy using a DENDROLAB 1.0 apparatus, then the ring-width sequence were processed with a set of software TREE-RINGS (Krawczyk and Krąpiec Reference Krawczyk and Krąpiec1995), TSAP (Rinn Reference Rinn2005) and DLP (Holmes Reference Holmes1999).

Figure 1 Location of Grabie village (50.0391 N, 19.992 E).

Based on dendrochronological sequences of subfossil oak trunks from Grabie, a GAA_3U chronology covering the period 994-612 BC was established (Krąpiec Reference Krąpiec2001). A G 58 sample was selected for radiocarbon analysis, and it is the same specimen (G 58) which was used in our previous study Rakowski et al. (Reference Rakowski, Krąpiec, Huels, Pawlyta, Hamann and Wiktorowski2019). Sample G 58, dated to 797-620BC and characterized by relatively large annual increments and high similarity (t=15,5, Gl=79(***)) to the South Polish Oak Chronology C_3000E (Krąpiec Reference Krąpiec2001; Figure S1). The South Polish Oak Chronology C_3000E was established trough teleconnection with Becker’s German oak chronology for southern Germany (Becker Reference Becker1993; t=7.7) and Leuschner’s and Delorme’s oak chronology for central Germany (Leuschner and Delorme Reference Leuschner and Delorme1988; t=6.8).

The tree used in our study belongs to the ring-porous angiosperms species, having clearly developed earlywood (EW) and latewood (LW). The EW consists of large cells, which are formed in the spring just before the leaves appear. To form these, the plant mainly uses stored reserves from the previous growing season. LW, on the other hand, is formed from photosynthesised substances during the growing season, i.e., the year of tree-ring formation (Ermich Reference Ermich1959). This allows a more accurate (half-yearly) determination of the timing of the 14C change.

Well-developed annual increments from the period 663-558 BCE, characterized by adequate thickness, were sliced and separated into LW and EW. Then, the LW was additionally split into 2 parts (except for samples from years 665 BCE and 660 BCE, which were too thin). This gives a total of 19 samples, of which five years (663 BCE, 662 BCE, 661 BCE, 660 BCE, and 658 BCE) were divided into EW, LW1, and LW2, and two years were divided into EW and LW (664 BCE and 659 BCE). From each of the sample, the most stable part of the wood in the form of α-cellulose was designated for study (Santos et. al. Reference Santos, Bird, Pillans, Fifield, Alloway, Chappell, Hausladen and Arneth2001; Nemec et al. Reference Nemec, Wacker, Hajdas and Gaggeler2010). The extraction of α-cellulose from the samples was performed according to the method described by Michczyńska et al. (Reference Michczyńska, Krąpiec, Michczyński, Pawlyta, Goslar, Nawrocka, Piotrowska, Szychowska-Krąpiec, Waliszewska and Zborowska2018). In the first step of this procedure, the wood samples were dried and cut into shavings, then weighed and placed into glass tubes, and subsequently each sample was put into glass tube with deionized water, sodium chlorite, and 1% HCl. The tube was then placed in an ultrasonic bath kept at 70°C for 1 hr, and then sodium chlorite and 1% HCl was added to each sample. This procedure was repeated five times, and the samples were then rinsed with hot deionized water up to neutral pH. This process removes most of the lignin and the remaining sample consists mostly of holocellulose. For the next step, 10% NaOH solution was added to the samples and then put into an ultrasonic bath at 70°C for 45 min. After the solution was removed, the sample was rinsed with cold deionizing water. Next, 17% NaOH solution was added, and samples were put into ultrasonic bath at room temperature for 45 min. After this time deionizing water was added to the samples. This step removes the rest of the remaining lignin and holocellulose. The samples were then rinsed with hot deionized water up to neutral pH. Finally, 1% HCl was added to the samples and the samples were rinsed to neutral pH.

The obtained α-cellulose was dried overnight at 70°C, and then placed in a glass tube together with oxidizing agent (MnO2), pumped out and sealed with torch. Samples of NBS Ox-II standard, IAEA-C3 (cellulose), and fossil wood background samples (IAEA-C9) were prepared and used to control the whole process of sample preparation and measurement (Janovics et al. Reference Janovics, Futó and Molnár2018). After the sealed tube combustion in 550°C (12 hr), the CO2 gas was released and purified in vacuum line and sealed again in tube containing zinc, titanium hydride and iron. A typical number of reagents and catalysts was used according to Rinyu et al. (Reference Rinyu, Molnár, Major, Nagy, Veres, Kimák, Wacker and Synal2013). For sample mass between 0.5 to 1.5 mg, 10 mg of TiH2, 60 mg of Zn, and 4.5 mg of iron powder was used. CO2 consists of 2 steps where the tube is placed in the furnace for 3 hr, initially at 500°C for hydrogen release and iron reduction, then the temperature is raised to 550°C to reduce CO2. This second stage takes approximately 5 hr. The produced graphite is then pressed into the cathode and measured using EnvironMICADAS AMS system at ATOMKI in Debrecen, Hungary (Molnar et al. Reference Molnar, Rinyu, Veres, Seiler, Wacker and Synal2016). The measurement time was extended to obtain more than 400.000 counts giving the overall uncertainties for modern samples better than ±2‰. Data acquisition was performed using BATS software (Wacker et al. Reference Wacker, Christl and Synal2010). The 14C contents are reported as Δ14C in per mil (‰) deviations from the standard sample, 0.7459 activity of NBS oxalic acid (SRM- 4990C). Age correction, isotopic composition correction (δ13C, measured by the AMS system), and Δ14C values were calculated using formulas presented by Stuiver and Polach (Reference Stuiver and Polach1977).

RESULTS AND DISCUSSION

The measured results (pMC, Δ14C) with corresponding uncertainties are presented in Table S1. Figure 2 shows Δ14C values in sub-annual (EW and LW) tree rings for the period 664–658 BCE as well as annual data from Rakowski et al. (Reference Rakowski, Krąpiec, Huels, Pawlyta, Hamann and Wiktorowski2019). EW was shifted 3/12 years from the beginning of the year while LW was shifted by 7.5/12 years. The LW from annual rings have been split into two parts and shifted by 7/12 years (LW1) and 8.5/12 years (LW2) respectively. Earlier results from Rakowski et al. (Reference Rakowski, Krąpiec, Huels, Pawlyta, Hamann and Wiktorowski2019), along with result from Park et al. (Reference Park, Southon, Fahrni, Creasman and Mewaldt2017), Sakurai et al. (Reference Sakurai, Tokanai, Miyake, Horiuchi, Masuda, Miyahara, Ohyama, Sakamoto, Mitsurani and Moriya2020), IntCal20 calibration curve (Reimer et al. Reference Reimer, Austin, Bard, Bayliss, Blackwell, Ramsey, Butzin, Cheng, Edwards and Friedrich2020) and IntCal20 raw datasets: HD, UB, UCIAMS, QL (http://calib.org) are presented in Figure 3.

Figure 2 Results of Δ14C determination in Grabie site tree. New results are shown as stars. Results from Rakowski et al. (Reference Rakowski, Krąpiec, Huels, Pawlyta, Hamann and Wiktorowski2019) are shown as empty crosses.

Figure 3 Results of Δ14C determination in Grabie site tree. New data are shown as red stars. Our results from Rakowski et al. (Reference Rakowski, Krąpiec, Huels, Pawlyta, Hamann and Wiktorowski2019) are shown as empty, red crosses. Sakurai et al. (Reference Sakurai, Tokanai, Miyake, Horiuchi, Masuda, Miyahara, Ohyama, Sakamoto, Mitsurani and Moriya2020) results are shown as black triangles. Park et al. (Reference Park, Southon, Fahrni, Creasman and Mewaldt2017) results are shown as black bow ties. Another data (HD, UB, UCIAMS, and QL) were extracted from IntCal20 raw data sets (http://calib.org). IntCal20 with its uncertainty is shown as a gray-shaded curve. (Please see online version for color figures.)

The sub-annual data from the Grabie core G 58 EW and LW rings show a prolonged increase in the Δ14C values of 20.4 ± 1.0‰ between 665 (data from Rakowski et al. Reference Rakowski, Krąpiec, Huels, Pawlyta, Hamann and Wiktorowski2019) and EW 661 BCE. The radiocarbon concentration (Δ14C) in this period raised from –1.3 ± 2.3‰ in 665 BCE to 19.1±2.2‰ in EW from 661 BCE showing high value of radiocarbon concentration during the rest of measured period. Similar to the literature the Δ14C curve shows first the rapid increase of about 8.3 ± 2.8‰ between 665 and 664 BCE, and then gradual increase to the maximum value of 19.1 ± 2.2‰ in EW in 661 BCE. Park et al. (Reference Park, Southon, Fahrni, Creasman and Mewaldt2017) and Sakurai et al. (Reference Sakurai, Tokanai, Miyake, Horiuchi, Masuda, Miyahara, Ohyama, Sakamoto, Mitsurani and Moriya2020) reported increase in Δ14C value of 14.3 ± 1.5‰ in German oak series within 4 yr, and 13.3 ± 2.1‰ in Choukai-Jindai cedar within 6 yr, respectively. The difference in radiocarbon concentration (Δ14C) in LW from 665 BCE and 664 BCE presented in literature is 8.9 ± 0.4‰ (Park et al. Reference Park, Southon, Fahrni, Creasman and Mewaldt2017), and 9.8 ± 2.2‰ (Sakurai et al. Reference Sakurai, Tokanai, Miyake, Horiuchi, Masuda, Miyahara, Ohyama, Sakamoto, Mitsurani and Moriya2020) and is similar to the value obtained in this study.

The Δ14C values measured in LW from 664 BCE (5.6 ± 2.2‰), and in EW from 663 BCE (6.0 ± 2.1‰), show no significant differences. This is because, to a significant extent, the EW is formed from the material accumulated in the previous year and a much smaller amount of it derives from the material created by photosynthesis in the year of growth (Speer Reference Speer2010). Olsson and Possnert (Reference Olsson and Possnert1992) observed that there is no delay between the radiocarbon concentration in the atmospheric sample in comparison with its concentration in the EW and LW during the maximum bomb effect (1962–1963). According Ermich (Reference Ermich1959), EW formation in this climatic zone occurs from April to mid-June, and LW formations ends in mid-September. This limits the beginning of the phenomenon to the period before formation of the LW, as there is no significant difference between Δ14C values in LW1 and LW2 in year 664 BCE. The most likely period of occurrence is May–June of year 663 BCE.

O’Hare et al. (2019) confirmed the origin of the cause of those changes. Based on the high-resolution data sets of 14C (tree rings), and 10Be and 36Cl (ice core) the production rate for those isotopes during this period was estimated, which led to the conclusion that SEP with a very hard energy spectrum triggered this phenomenon. The 660 BCE event was an order of magnitude stronger than any solar events observed during the industrial era and was comparable to the 774/775 CE event.

Sakurai et al. (Reference Sakurai, Tokanai, Miyake, Horiuchi, Masuda, Miyahara, Ohyama, Sakamoto, Mitsurani and Moriya2020) estimated the net production of 14C to (1.3–1.5) × 108 atoms/cm2 and is around 40% smaller than the net production during the event 774/775 CE (2.2 × 108 atoms/cm2), but around 15% higher than the 993/994 CE event, when the net production was 1.2 × 108 atoms/cm2. To compare, the net production rate for the solar events SPE56 and SEP72 were 2.6 × 106 atoms/cm2 and 6.2 × 105 atoms/cm2.

The production rate of 14C for the increases in those two events of 664 BCE and 663 BCE estimated using an 11-box carbon cycle model (Güttler et al. Reference Güttler, Adolphi, Beer, Bleicher, Boswijk, Christl, Hogg, Palmer, Vockenhuber, Wacker and Wunder2015) was ∼1.2 × 108 atoms/cm2 and ∼1.3 × 108 atoms/cm2, respectively. For this estimation, LW data for 664–660 BCE were used for the estimation on the latter event. Please note we took weighted means for the LW1 and LW2 when available (Table S1, S2). For the first event, we used the yearly data of 665–664 BCE obtained by Rakowski et al. (Reference Rakowski, Krąpiec, Huels, Pawlyta, Hamann and Wiktorowski2019; Table S2).

CONCLUSIONS

A new high-resolution set of radiocarbon data from the sub-annual oak tree rings of Grabie, southern Poland, is presented in this paper. This radiocarbon concentration dataset from the EW and LW rings covers the SEP event around 660 BCE. This new data record shows a prolonged increase in Δ14C between 665–663 BCE. It started with an increase of 8.9 ± 2.8‰ in Δ14C between 665 and 664 BCE (Rakowski et al. Reference Rakowski, Krąpiec, Huels, Pawlyta, Hamann and Wiktorowski2019), and lasted until 661 BCE, which is similar (and in agreement) to previously reported observations. In comparison, the radiocarbon data collected for this period presented by Park et al. (Reference Park, Southon, Fahrni, Creasman and Mewaldt2017) and Sakurai et al. (Reference Sakurai, Tokanai, Miyake, Horiuchi, Masuda, Miyahara, Ohyama, Sakamoto, Mitsurani and Moriya2020), provide comparable data patterns with a prolonged increase until 662 BCE. The most probable time for a major event, which caused a significant change in the radiocarbon concentration in the tree rings, is the late spring (May–June, before beginning of LW formation) of 663 BCE.

SUPPLEMENTARY MATERIAL

For supplementary material accompanying this paper visit https://doi.org/10.1017/RDC.2023.79

ACKNOWLEDGMENTS

This work was partly supported by the National Science Centre, Poland, grant UMO-2017/25/B/ST10/02329, EUROPLANET24 grant 20-EPN2-018 and Nagoya University ISEE –International Joint Research Program-00027. Publication supported as part of the Excellence Initiative – Research University program implemented at the Silesian University of Technology, year 2022 (14/020/SDU/10-27-01).

Footnotes

Selected Papers from the 24th Radiocarbon and 10th Radiocarbon & Archaeology International Conferences, Zurich, Switzerland, 11–16 Sept. 2022.

References

REFERENCES

Becker, B. 1993. An 11,000 year German oak and pine chronology for radiocarbon calibration. Radiocarbon 35, 201213.CrossRefGoogle Scholar
Brehm, N, Christl, M, Adolphi, F, Muscheler, R, Synal, H-A, Mekhaldi, F, Paleari, C, Leuschner, HH, Bayliss, A, Nicolussi, K, et al. 2022. Tree rings reveal two strong solar proton events in 7176 and 5259 BCE. Nature Communications. doi: 10.21203/rs.3.rs-753272/v1.CrossRefGoogle Scholar
Büntgen, U, Wacker, L, Galván, JD, Arnold, S, Arseneault, D, Baillie, M, Beer, J, Bernabei, M, Bleicher, N, Boswijk, G, et al. 2018. Tree rings reveal globally coherent signature of cosmogenic radiocarbon events in 774 and 993 CE. Nature Communications 9(1):3605. doi: 10.1038/s4167-018-06036-0 CrossRefGoogle ScholarPubMed
Ermich, K. 1959. The investigation of the seasonal course of the diameter growth of Pinus silvestris L. and Quercus robur L. Acta Societatis Botanicorum Poloiae 28(1):1563. In Polish.CrossRefGoogle Scholar
Fogtmann-Schulz, A, Ostbo, SM, Nielsen, SGB, Olsen, J, Karoff, C, Knudsen, MF. 2017. Cosmic ray event in 994 CE recorded in radiocarbon from Danish oak. Geophysical Research Letters 44(16):86218628.CrossRefGoogle Scholar
Güttler, D, Adolphi, F, Beer, J, Bleicher, N, Boswijk, G, Christl, M, Hogg, A, Palmer, j, Vockenhuber, C, Wacker, L, Wunder, J. 2015. Rapid increase in cosmogenic 14C in AD 775 measured in New Zealand kauri trees indicates short-lived increase in 14C production spanning both hemispheres. Earth and Planetary Science Letters 411:290297.CrossRefGoogle Scholar
Güttler, D, Beer, J, Bleicher, N, Boswijk, G, Hogg, AG, Palmer, JG, Vockenhuber, C, Wacker, L, Wunder, J. 2015. Worldwide detection of a rapid increase of cosmogenic 14C in AD 775. Poster presented at the Nuclear Physics in Astrophysics.Google Scholar
Hakozaki, M, Miyake, F, Nakamura, T, Kimura, K, Masuda, K, Okuno, M. 2018. Verification of the annual dating of the 10th century Baitoushan volcano eruption based on an AD 774–775 radiocarbon spike. Radiocarbon 60(1):261268.CrossRefGoogle Scholar
Hogg, AG, Heaton, TJ, Hua, Q, Palmer, JG, Turney, CSM, Wacker, L. 2020. SHCal20 Southern Hemisphere calibration, 0–55,000 years cal BP. Radiocarbon 62(4):759778. doi: 10.1017/RDC.2020.59 CrossRefGoogle Scholar
Holmes, RL. 1999. User’s Manual for Program COFECHA. Tucson (AZ): University of Arizona.Google Scholar
Janovics, R, Futó, I, Molnár, M. 2018. Sealed tube combustion method with MnO2 for AMS 14C measurement. Radiocarbon 60(5):13471355.CrossRefGoogle Scholar
Jull, AJT, Panyushkina, IP, Lange, TE, Kukarskih, VV, Myglan, VS, Clark, KJ, Salzer, MW, Burr, GS, Leavitt, SW. 2014. Excursions in the 14C record at A.D. 774–775 in tree rings from Russia and America. Geophysical Research Letters 41(8):30043010.CrossRefGoogle Scholar
Jull, AJT, Panyushkina, I, Miyake, F, Masuda, K, Nakamura, T, Mitsutani, T, Lange, TE, Cruz, R, Baisan, C, Janovics, R, et al. 2018. More rapid 14C excursions in the tree-ring record: a record of different kind of solar activity at about 800 BC? Radiocarbon 60(4):12371248.CrossRefGoogle Scholar
Kanzawa, K, Miyake, F, Horiuchi, K, Sasa, K, Takano, K, Matsumura, M, et al. 2021. High-resolution 10Be and 36Cl data from the Antarctic Dome Fuji ice core (∼100 years around 5480 BCE): an unusual grand solar minimum occurrence? Journal of Geophysical Research: Space Physics 126:e2021JA029378. doi: 10.1029/2021JA029378 CrossRefGoogle Scholar
Kovaltsov, GA, Mishev, A, Usoskin, IG. 2012. A new model of cosmic production of radiocarbon 14C in the atmosphere. Earth and Planetary Sciences Letters 337–338:114120.CrossRefGoogle Scholar
Krąpiec, M. 2001. Holocene dendrochronological standards for subfossil oaks from the area of Southern Poland. Studia Quaternaria 18:4763.Google Scholar
Krąpiec, M, Rakowski, AZ, Pawlyta, J, Wiktorowski, D, Bolka, M. 2020. Absolute dendrochronological scale for pine (Pinus sylvestris L.) from Ujscie (NW Poland), dated using rapid atmospheric 14C changes. Radiocarbon. doi: 10.1017/RDC.2020.116 CrossRefGoogle Scholar
Krawczyk, A, Krąpiec, M. 1995. Dendrochronologiczna baza danych. Materiały II Krajowej Konferencji: Komputerowe wspomaganie badań naukowych (Dendrochronological database. Proceedings of the Second Polish Conference on Computer Assistance to Scientific Research). Wrocław: 247–252. In Polish.Google Scholar
Kuitems, M, Wallace, BL, Lindsay, C. et al. 2022. Evidence for European presence in the Americas in ad 1021. Nature 601:388391. doi: 10.1038/s41586-021-03972-8 CrossRefGoogle ScholarPubMed
Leuschner, H-H, Delorme, A. 1988. Tree-ring work in Göttingen: absolute oak chronologies back to 6255 BC. PACT 22:123132.Google Scholar
Mekhaldi, F, Muscheler, R, Adolphi, F, Aldahan, A, Beer, J, McConnell, JR, Possnert, G, Sigl, M, Svensson, A, Synal, H-A, et al. 2015. Multiradionuclide evidence for the solar origin of the cosmic-ray events of AD 774/5 and 993/4. Nature Communications 6:8611.CrossRefGoogle ScholarPubMed
Michczyńska, DJ, Krąpiec, M, Michczyński, A, Pawlyta, J, Goslar, T, Nawrocka, N, Piotrowska, N, Szychowska-Krąpiec, E, Waliszewska, B, Zborowska, M. 2018. Different pretreatment methods for 14C dating of Younger Dryas and Allerød pine wood (Pinus sylvestris L.). Quaternary Geochronology 48:3844.CrossRefGoogle Scholar
Miyahara, H, Tokanai, F, Moriya, T, Takeyama, M, Sakurai, H, Ohyama, M, Horuchi, K, Hotta, H. 2022. Recurrent large-scale solar proton events before the onset of the Wolf Grand Solar Minimum. Geophysical Research Letters 49. doi: 10.1029/2021GL097201 CrossRefGoogle Scholar
Miyake, F, Jull, AJT, Panyushkina, IP, Wacker, L, Salzer, M, Baisan, CH, Lange, T, Cruz, R, Masuda, K, Nakamura, T. 2017. Large 14C excursion in 5480 BC indicates an abnormal sun in the mid-Holocene. Proceedings of the National Academy of Sciences of the United States of America 114(5):881884. doi: 10.1073/pnas.1613144114.CrossRefGoogle Scholar
Miyake, F, Masuda, K, Nakamura, T. 2013. Another rapid event in the carbon-14 content of tree rings. Nature Communications 4:1748. doi: 10.1038/ncomms2873.CrossRefGoogle ScholarPubMed
Miyake, F, Nagaya, K, Masuda, K, Nakamura, T. 2012. A signature of cosmic-ray increases in AD 774–775 from tree rings in Japan. Nature 486(7402):240242.CrossRefGoogle ScholarPubMed
Miyake, F, Masuda, K, Hakozaki, M, Nakamura, T, Tokanai, F, Kato, K, Kimura, K, Mitsutani, T. 2014. Verification of the cosmic-ray event in AD 993–994 by using a Japanese Hinoki tree. Radiocarbon 56(3):11841194.CrossRefGoogle Scholar
Molnar, M, Rinyu, L, Veres, M, Seiler, M, Wacker, L, Synal, H-A. 2016. EnvironMICADAS: A Mini 14C AMS with enhanced gas ion source interface in the Hertelendi Laboratory of Environmental Studies (HEKAL), Hungary. Radiocarbon 55(2):338344.CrossRefGoogle Scholar
Nemec, M, Wacker, L, Hajdas, I, Gaggeler, H. 2010. Alternative methods for cellulose preparation for AMS measurement. Radiocarbon 52(2):13581370.CrossRefGoogle Scholar
O’Hare, P, Mekhaldi, F, Adolphi, F, Reisbeck, G, Aldahan, A, Anderberg, E, Beer, J, Christl, M, Fahrni, S, Synal, H-A, et al. 2019. Multiradionuclide evidence for an extreme solar proton event around 2,610 B.P. (∼660 BC). Proceedings of the National Academy of Sciences of the United States of America PNAS 116;(13):59615966. doi: 10.1073/pnas.1815725116.CrossRefGoogle ScholarPubMed
Olsson, IU, Possnert, G. 1992. 14C activity in different sections and chemical fractions of oak tree rings, A.D. 1938–1981. Radiocarbon 34(3):757767.CrossRefGoogle Scholar
Oppenheimer, C, Wacker, L, Xu, J, Galván, JD, Stoffel, M, Guillet, S, Corona, C, Sigl, M, Cosmo, ND, Hajdas, I, et al. 2017. Multi-proxy dating the “Millennium Eruption” of Changbaishan to late 946 CE. Quaternary Science Reviews 158:164171.CrossRefGoogle Scholar
Paleari, CI, Mekhaldi, F, Adolphi, F, Christl, M, Vockenhuber, C, Gautschi, P, Beer, J, Brehm, N, Erhardt, T, Synal, HA, et al. 2022. Cosmogenic radionuclides reveal an extreme solar particle storm near a solar minimum 9125 years BP. Nat Commun 11; (13 1):214. doi: 10.1038/s41467-021-27891-4. PMID: 35017519; PMCID: PMC8752676.CrossRefGoogle Scholar
Park, J, Southon, J, Fahrni, S, Creasman, PP, Mewaldt, R. 2017. Relationship between solar activity and Δ14C peaks in AD 775, AD 994, and 660 BC. Radiocarbon 59(4):11471156.CrossRefGoogle Scholar
Pavlov, A, Blinov, AV, Konstantinov, AN, Ostryakov, VN, Vasilyev, GI, Vdovina, MA, Volkov, PA. 2013. AD 775 pulse of cosmogenic radionuclides production as imprint of a Galactic gamma-ray burst. Mon. Not. R. Astron. Soc. 435(4):28782884.CrossRefGoogle Scholar
Philipsen, B, Feveile, C, Olsen, J, Sindbaek, SM. 2022. Single-year radiocarbon dating anchors Viking Age trade cycles in time. Nature 601. doi: 10.1038/s41586-021-04240-5 CrossRefGoogle Scholar
Possnert, G, Southon, J, Bard, E, ASTER, Team, Muschler, R. 2019. Multinuclide evidence for an extreme solar proton event around 2,610 B.P. (∼660 BC). PNAS 116(13):59615966. doi: 10.1038/s41598-019-57273-2 Google Scholar
Rakowski, AZ, Krąpiec, M, Huels, M, Pawlyta, J, Boudin, M. 2018. Increase in radiocarbon concentration in tree rings from Kujawy village (SE Poland) around AD 993–994. Radiocarbon 60(4):12491258. doi: 10.101/rdc.2018.74 CrossRefGoogle Scholar
Rakowski, AZ, Krąpiec, M, Huels, M, Pawlyta, J, Dreves, A, Meadows, J. 2015. Increase of radiocarbon concentration in tree rings from Kujawy village (SE Poland) around AD 774–775. Nuclear Instruments and Methods in Physics Research B 351:564568.CrossRefGoogle Scholar
Rakowski, AZ, Krąpiec, M, Huels, M, Pawlyta, J, Hamann, Ch, Wiktorowski, D. 2019. Abdupt increase of radiocarbon concentration in 660 BC in the tree rings from Grabie near Karkow (SE Poland). Radiocarbon 61(5):13271335.CrossRefGoogle Scholar
Reimer, PJ, Austin, WEN, Bard, E, Bayliss, A, Blackwell, PG, Ramsey, CB, Butzin, M, Cheng, H, Edwards, RL, Friedrich, M, et al. 2020. The IntCal20 Northern Hemisphere radiocarbon age calibration curve (0–55 cal kBP). Radiocarbon 62(4):725757. doi: 10.1017/RDC.2020.41 CrossRefGoogle Scholar
Rinn, F. 2005. TSAP-Win. Time series analysis and presentation for dendrochronology and related applications. User reference. Heidelberg.Google Scholar
Rinyu, L, Molnár, M, Major, I, Nagy, T, Veres, M, Kimák, Á, Wacker, L, Synal, H-A. 2013. Optimization of sealed tube graphitization method for environmental 14C studies using MICADAS. Nuclear Instruments and Methods in Physics Research B 294:270275.CrossRefGoogle Scholar
Sakurai, H, Tokanai, F, Miyake, F, Horiuchi, K, Masuda, K, Miyahara, H, Ohyama, M, Sakamoto, M, Mitsurani, T, Moriya, T. 2020. Prolonged production of 14C during the ∼660 BCE solar proton event from Japanese tree rings. Scientific Reports Nature 10:660 doi: 10.1038/s41598-019-57273-2 CrossRefGoogle ScholarPubMed
Santos, GM, Bird, MI, Pillans, B, Fifield, LK, Alloway, BV, Chappell, J, Hausladen, PA, Arneth, A. 2001. Radiocarbon dating of wood using different pretreatment procedures: application to the chronology of Rotoehu Ash, New Zealand. Radiocarbon 43(2A):239248.CrossRefGoogle Scholar
Speer, JH. 2010. Fundamentals of tree-ring research. Tucson (AZ): Univ. of Arizona Press. 368 p.Google Scholar
Stuiver, M, Polach, HA. 1977. Discussion: reporting of 14C data. Radiocarbon 19:355363.CrossRefGoogle Scholar
Synal, H-A, Stocker, M, Suter, M. 2007. MICADAS: a new compact radiocarbon AMS system. Nuclear Instruments and Methods in Physics Research B 259(1):713.CrossRefGoogle Scholar
Tarrasi, F, Marzaioli, F, Bouompane, R, Passariello, I, Capano, M, Helama, S, Oinonen, M, Nojd, P, Uusitalo, J, Jull, AJT, et al. 2020 Can the 14C production in 1055 CE be affected by SN1054? Radiocarbon 62(5):14031418. doi: 10.1017/RDC.2020.58 CrossRefGoogle Scholar
Usoskin, IG, Kovaltsov, GA. 2012. Occurrence of extreme solar particle events: assessment from historical proxy data. The Astrophysical Journal 757:92.CrossRefGoogle Scholar
Usoskin, IG, Kromer, B, Ludlow, F, Beer, J, Friedrich, M, Kovaltsov, GA, Solanki, S, Wacker, L. 2013. The AD775 cosmic event revisited: the Sun is to blame. Astron. Astrophys. 55:L3. doi: 10.1051/0004-6361/201321080.CrossRefGoogle Scholar
van der Plicht, J. 2007. Radiocarbon dating/variations in atmospheric 14C. In: Holloway R, Elias SA, editors. Encyclopedia of Quaternary Science. Elsevier. p. 2923–2931.Google Scholar
Wacker, L, Christl, M, Synal, H-A. 2010. BATS: a new tool for AMS data reduction. Nuclear Instruments and Methods in Physics Research B 268(7–8):976979.CrossRefGoogle Scholar
Wacker, L, Guttler, D, Goll, J, Hurni, J, Synal, H-A, Walti, N. 2014. Radiocarbon dating to a single year by means of rapid atmospheric 14C changes. Radiocarbon 56(2):573579. doi: 10.2458/56.17634.CrossRefGoogle Scholar
Figure 0

Figure 1 Location of Grabie village (50.0391 N, 19.992 E).

Figure 1

Figure 2 Results of Δ14C determination in Grabie site tree. New results are shown as stars. Results from Rakowski et al. (2019) are shown as empty crosses.

Figure 2

Figure 3 Results of Δ14C determination in Grabie site tree. New data are shown as red stars. Our results from Rakowski et al. (2019) are shown as empty, red crosses. Sakurai et al. (2020) results are shown as black triangles. Park et al. (2017) results are shown as black bow ties. Another data (HD, UB, UCIAMS, and QL) were extracted from IntCal20 raw data sets (http://calib.org). IntCal20 with its uncertainty is shown as a gray-shaded curve. (Please see online version for color figures.)

Supplementary material: File

Rakowski et al. supplementary material

Rakowski et al. supplementary material

Download Rakowski et al. supplementary material(File)
File 268.8 KB