Hostname: page-component-7bb8b95d7b-pwrkn Total loading time: 0 Render date: 2024-09-26T06:32:28.111Z Has data issue: false hasContentIssue false

A tree-ring cellulose extraction device adapted to radiocarbon analysis

Published online by Cambridge University Press:  26 September 2024

Pengyu Lin
Affiliation:
School of Geography and Ocean Science, Nanjing University, Nanjing, Jiangsu, 210023, China
Yesi Zhao
Affiliation:
School of Geography and Ocean Science, Nanjing University, Nanjing, Jiangsu, 210023, China Frontiers Science Center for Critical Earth Material Cycling, Nanjing University, Nanjing, Jiangsu, 210023, China
Hongyan Zhang*
Affiliation:
School of Geography and Ocean Science, Nanjing University, Nanjing, Jiangsu, 210023, China Frontiers Science Center for Critical Earth Material Cycling, Nanjing University, Nanjing, Jiangsu, 210023, China
Thomas Wieloch
Affiliation:
Department of Forest Genetics and Plant Physiology, Swedish University of Agricultural Sciences, Umeå Plant Science Centre, Umeå, 90183, Sweden
Yao Gu
Affiliation:
School of Geography and Ocean Science, Nanjing University, Nanjing, Jiangsu, 210023, China
Chenghong Liang
Affiliation:
School of Geography and Ocean Science, Nanjing University, Nanjing, Jiangsu, 210023, China
Feng Chen
Affiliation:
Yunnan Key Laboratory of International Rivers and Transboundary Eco-Security, Institute of International Rivers and Eco-Security, Yunnan University, Kunming, Yunnan, 650091, China
Huayu Lu*
Affiliation:
School of Geography and Ocean Science, Nanjing University, Nanjing, Jiangsu, 210023, China Frontiers Science Center for Critical Earth Material Cycling, Nanjing University, Nanjing, Jiangsu, 210023, China
*
Corresponding authors: Hongyan Zhang; Email: hongyan@nju.edu.cn and Huayu Lu; Email: huayulu@nju.edu.cn
Corresponding authors: Hongyan Zhang; Email: hongyan@nju.edu.cn and Huayu Lu; Email: huayulu@nju.edu.cn
Rights & Permissions [Opens in a new window]

Abstract

Tree-ring cellulose is a commonly used material for radiocarbon analysis. Extracting cellulose is labor-consuming and several devices that enable batchwise extraction have been developed. However, these devices bear the risk of sample contamination. The present study describes a new device which improves upon two aspects of currently available devices. First, to prevent cross-sample-contamination, we redesigned the drainage module to enable independent removal of chemical waste from each individual sample funnel. Second, we added covers to the sample funnels to reduce the risk of external contamination. Cellulose purity (i.e., holocellulose) was confirmed by Fourier Transform Infrared (FTIR) Spectroscopy. Furthermore, accuracy of the radiocarbon analysis was confirmed by results of 14C-blank samples and samples of known age. In conclusion, while maintaining labor-saving, our modified device significantly reduces the risk of sample contamination during extraction of tree-ring cellulose.

Type
Technical Note
Copyright
© The Author(s), 2024. Published by Cambridge University Press on behalf of University of Arizona

Introduction

Radiocarbon (14C) analysis by high-precision accelerator mass spectrometry (AMS) is an increasingly active area of research in dendrochronology and the Earth sciences. Radiocarbon analysis of tree-ring time series is used to construct 14C calibration curves (Hogg et al. Reference Hogg, Heaton, Hua, Palmer, Turney, Southon, Bayliss, Blackwell, Boswijk and Bronk Ramsey2020; Hua et al. Reference Hua, Turnbull, Santos, Rakowski, Ancapichún, De Pol-Holz, Hammer, Lehman, Levin and Miller2021; Reimer et al. Reference Reimer, Austin, Bard, Bayliss, Blackwell, Bronk Ramsey, Butzin, Cheng, Edwards and Friedrich2020), which enable determination of the calendar age of a sample (e.g., a prehistory artifact) from its 14C age. Furthermore, tree-ring radiocarbon analysis can provide information about solar activity during the pre-telescope period, such as long solar cycles, and therefore contributes to a better understanding of the physical basis of the sun’s activity (Damon and Sonett Reference Damon, Sonett, Sonett, Giampapa and Mattheus1991; Suess Reference Suess1980; Stuiver and Braziunas Reference Stuiver and Braziunas1993; Usoskin et al. Reference Usoskin, Solanki, Krivova, Hofer, Kovaltsov, Wacker, Brehm and Kromer2021). In addition, exceptional increases in tree-ring 14C abundance, as observed for events in the years 774 AD and 993 AD, may indicate past occurrences of solar super-flares (Büntgen et al. Reference Büntgen, Wacker, Galván, Arnold, Arseneault, Baillie, Beer, Bernabei, Bleicher and Boswijk2018; Miyake et al. Reference Miyake, Nagaya, Masuda and Nakamura2012, Reference Miyake, Masuda and Nakamura2013; Usoskin et al. Reference Usoskin, Kromer, Ludlow, Beer, Friedrich, Kovaltsov, Solanki and Wacker2013; Uusitalo et al. Reference Uusitalo, Arppe, Hackman, Helama, Kovaltsov, Mielikäinen, Mäkinen, Nöjd, Palonen and Usoskin2018).

Cellulose is commonly utilized in tree-ring based isotope analysis. As a major structural component of xylem cell walls, cellulose does not migrate between tree rings and is therefore unambiguously linked to the year of tree-ring formation (Leavitt and Danzer Reference Leavitt and Danzer1993). Cellulose composition is comparatively simple, thus avoiding potential bias due to varying proportions of constituents between tree-rings (Rinne et al. Reference Rinne, Boettger, Loader, Robertson, Switsur and Waterhouse2005). Early studies recommended the use of α-cellulose for radiocarbon analysis since α-cellulose only contains D-glucose monomers (Head Reference Head1979; Hoper et al. Reference Hoper, McCormac, Hogg, Higham and Head1998; Stuiver and Quay Reference Stuiver and Quay1981). However, more recent studies have shown that 14C abundance of holocellulose is not systematically different from α-cellulose (Capano et al. Reference Capano, Miramont, Guibal, Kromer, Tuna, Fagault and Bard2018; Hajdas et al. Reference Hajdas, Hendriks, Fontana and Monegato2017; Lange et al. Reference Lange, Nordby, Murphy, Hodgins and Pearson2019; Michczyńska et al. Reference Michczyńska, Krąpiec, Michczyński, Pawlyta, Goslar, Nawrocka, Piotrowska, Szychowska-Krąpiec, Waliszewska and Zborowska2018; Southon and Magana Reference Southon and Magana2010; Staff et al. Reference Staff, Reynard, Brock and Bronk Ramsey2014). Since holocellulose extraction from wood is less labor-intensive than α-cellulose extraction, holocellulose has become the preferred analytical substrate for reconstructing interannual atmospheric 14C content (Büntgen et al. Reference Büntgen, Wacker, Galván, Arnold, Arseneault, Baillie, Beer, Bernabei, Bleicher and Boswijk2018; Brehm et al. Reference Brehm, Bayliss, Christl, Synal, Adolphi, Beer, Kromer, Muscheler, Solanki and Usoskin2021, Reference Brehm, Christl, Knowles, Casanova, Evershed, Adolphi, Muscheler, Synal, Mekhaldi and Paleari2022; Güttler et al. Reference Güttler, Wacker, Kromer, Friedrich and Synal2013; Wacker et al. Reference Wacker, Güttler, Goll, Hurni, Synal and Walti2014).

To extract cellulose (either holocellulose or α-cellulose) from wood materials is a central step in the preparation of tree-ring samples for radiocarbon analysis. Cellulose extraction involves the removal of resins and lignin via organic solvent extraction and chlorination (Green Reference Green and Whistler1963; Leavitt and Danzer Reference Leavitt and Danzer1993), or alkaline extraction and chlorination (Rinne et al. Reference Rinne, Boettger, Loader, Robertson, Switsur and Waterhouse2005; Wieloch et al. Reference Wieloch, Helle, Heinrich, Voigt and Schyma2011). The cellulose extraction process requires repeated addition and removal of reagents as well as washing and is, therefore, a highly labor-intensive process. Consequently, dendro-scientists have developed various devices to facilitate batchwise extraction of cellulose. These include filter bags (Cullen and Macfarlane Reference Cullen and Macfarlane2005; Leavitt and Danzer Reference Leavitt and Danzer1993), filter tubes or funnels with or without polytetrafluorethylene (PTFE) drainage blocks (Andreu-Hayles et al. Reference Andreu-Hayles, Levesque, Martin-Benito, Huang, Harris, Oelkers, Leland, Martin-Fernández, Anchukaitis and Helle2019; Harada et al. Reference Harada, Watanabe, Nakatsuka, Tazuru-mizuno, Horikawa, Sugiyama, Tsuda and Tagami2014; Loader et al. Reference Loader, Robertson, Barker, Switsur and Waterhouse1997; Fogtmann-Schulz et al. Reference Fogtmann-Schulz, Kudsk, Adolphi, Karoff, Knudsen, Loader, Muscheler, Trant, Østbø and Olsen2021; Wieloch et al. Reference Wieloch, Helle, Heinrich, Voigt and Schyma2011), and plate splines with PTFE punching sheets (Kagawa et al. Reference Kagawa, Sano, Nakatsuka, Ikeda and Kubo2015; Li et al. Reference Li, Labbé, Driese and Grissino-Mayer2011). With regard to radiocarbon analysis, however, all currently available devices carry the potential risk of external contamination and internal cross-sample-contamination due to either direct contact between tree-ring samples (plate splines) or their interconnected storage spaces (filter bags and funnels). For instance, Santos et al. (Reference Santos, Granato-Souza, Barbosa, Oelkers and Andreu-Hayles2020) report that the extraction procedure designed by Andreu-Hayles et al. (Reference Andreu-Hayles, Levesque, Martin-Benito, Huang, Harris, Oelkers, Leland, Martin-Fernández, Anchukaitis and Helle2019) can result in modern carbon contamination and suggest it is unsuitable for analyses of preglacial samples. This is a significant concern, especially because Sookdeo et al. (Reference Sookdeo, Kromer, Büntgen, Friedrich, Friedrich, Helle, Pauly, Nievergelt, Reinig and Treydte2020) recommend that, for high-precision radiocarbon analysis, cellulose extraction for tree-ring samples should be performed simultaneously with blank samples. Thus, a device for high-throughput extraction of tree-ring cellulose suitable for radiocarbon analysis (i.e., mitigating the risk of external and internal contamination) is currently unavailable.

To fill this gap, we adapted the batchwise cellulose extraction device developed by Wieloch et al. (Reference Wieloch, Helle, Heinrich, Voigt and Schyma2011), known as “multiple sample isolation system for solids” (MSISS) to radiocarbon analysis. First, to prevent cross-sample-contamination, we redesigned the drainage module to enable independent removal of chemical waste from each individual sample funnel. Second, to reduce the risk of external contamination, we added covers to the sample funnels. In the following sections, we start with a detailed illustration of the modified device, followed by an evaluation of its suitability for tree-ring cellulose extraction for radiocarbon analysis.

Description of the device

Two key modifications were implemented to adapt the original MSISS:

  1. 1. Independent drainage channels: In the original MSISS, the drainage channels were interconnected. This bears the risk of transferring chemicals and contaminants among tree-ring samples. However, in the modified version, each tree-ring sample has its own separate drainage channel, which prevents cross-sample-contamination (Figure 1a).

  2. 2. Funnel covers: In the original MSISS, the funnels are uncovered during entire experiment. This bears the risk of external contaminants entering the funnels. To address this issue, PTFE covers have been added to each funnel in the modified device (Figure 1a–d). These covers effectively prevent contaminants from falling into the funnel during the extraction process (Figure 1e). Additionally, the covers minimize volatilization of ClO2 and Cl2 gases produced during the acid NaClO2 treatment step. Each cover is equipped with a short handle for effortless lifting and replacing.

Figure 1. An illustration of the device and sliced samples used to extract tree-ring cellulose for radiocarbon analysis. (a) The three-dimensional structural diagram of the modified device. The left part of the diagram shows the internal drainage channels. (b-c) An image and side view of the device: 1. silicone tubing outlet used to remove chemical waste; 2. PTFE cover used to prevent external contamination with its side view shown at the top left; 3. glass funnel used for loading tree-ring sample; 4. silicone tubing used to connect funnels to the drainage module; 5. PTFE screw; 6. sintered glass filter disc inside the funnel (similarly, hereinafter); 7. sliced wood samples inside the funnel (similarly, hereinafter). The red curve with arrows indicates the flow path of chemical waste during the pumping process. (d) Experimental setup used here for batchwise extraction of tree-ring cellulose. For chemical treatments, six devices containing 108 tree-ring samples were heated in a water bath (600 mm × 300 mm). With a vacuum pump, chemical waste can be pumped through the outlets at the top of PTFE block. (e) Contaminants (black arrow) that fell on the cover during the experiment. (f) The funnel. The filter disc at the base of the funnel has an aperture size of ca. 30–50 μm. (g) The wood samples that were sliced into pieces around 3×3×0.5 mm3 in size.

The modified device has an inverted T-shaped drainage module with staggered L-shaped drainage channels inside. The upper part of the drainage channel functions as an outlet (Figure 1a-d). On both sides of the drainage module, 18 funnels are arranged with PTFE covers. Numbers 1 to 18 are engraved next to the corresponding funnel for easy identification. Prior to chemical treatment, wood samples are placed into the funnels. Subsequently, chemical reagents are manually added to the funnels using a pipette gun. Once the chemicals are added, the funnels must be covered immediately to prevent contaminants from entering the system. Removing chemical waste from funnels can be achieved using a vacuum pump to create a negative pressure environment inside the drainage module. The silicone tube is attached to the outlet, allowing chemical waste to flow automatically into a collecting bottle. The covers do not need to be adjusted while chemical waste is pumping out because they are not air-tight seals. A filter disc (ca. 30–50 μm aperture size) is fixed at the base of each funnel, keeping the wood samples inside funnels during pumping (Figure 1c, f). The process of pumping chemical waste from each funnel just takes 3 to 5 seconds when using a vacuum pump with a suction rate of 10 L/min. With six drainage modules in a water bath (length: 600 mm; width: 300 mm; height: 130 mm), 108 tree-ring samples can be processed simultaneously (Figure 1d). The entire procedure must be conducted under a fume hood, since harmful gases are released during cellulose extraction.

Overall, these modifications are shown to significantly reduce the risk of contamination during cellulose extraction while maintaining the high efficiency of the existing device. For more information on the device design and availability, please refer to Text S1.

Materials and methods

Wood samples

Five wood samples of different ages were used to assess the suitability of the modified device for radiocarbon analysis (Table 1). The samples are IAEA-C9, a subfossil kauri wood from peat bogs in New Zealand (Hogg et al. Reference Hogg, Higham, Robertson, Beukens, Kankainen, McCormac, van der Plicht and Stuiver1995); NF142, a fossil pine wood from the late Miocene layer in Xianfeng Basin, Yunnan, China (Wang et al. Reference Wang, Oskolski, Jacques, Wang and Zhou2017); KB, a kauri wood from Marine Isotope Stage 7 in New Zealand (Sookdeo et al. Reference Sookdeo, Kromer, Büntgen, Friedrich, Friedrich, Helle, Pauly, Nievergelt, Reinig and Treydte2020); LWQ1796, a tree-ring sample of known age (1796 AD) from an absolutely dated spruce tree-ring-width chronology (LWQ; Figure S1; Chen et al. Reference Chen, Shang, Panyushkina, Meko, Li, Yuan, Yu, Chen, He and Luo2019) from Riwoche County in Xizang, China; AD1515, an architecture oak wood with a known age (1515 AD) from A. Bayliss of Historic England (Bayliss et al. Reference Bayliss, Howard and Tyers2023).

Table 1. General information about the wood samples used for assessing the suitability of our modified tree-ring cellulose extraction device for radiocarbon analysis

* Infinite ages denote ages that lie beyond the 14C dating limit (Taylor and Bar-Yosef Reference Taylor and Bar-Yosef2016). The 14C abundance of an infinite-aged sample is indistinguishable from background values.

Holocellulose extraction

The protocol for holocellulose extraction was conducted according to the revised Base-Acid-Base-Acid-Bleaching procedure as shown in Figure 2 (Němec et al. Reference Němec, Wacker, Hajdas and Gäggeler2010; Sookdeo et al. Reference Sookdeo, Kromer, Büntgen, Friedrich, Friedrich, Helle, Pauly, Nievergelt, Reinig and Treydte2020). Prior to chemical processing, the glass funnels and 10 mL glass vials were heated at 500°C for 4 hr in a Muffle oven (Lindberg/Blue M™, Thermo Scientific™, USA) to remove carbon contaminants. To attain the desired level of cleanliness, the glass funnels, PTFE drainage modules, silicone tubes, and covers were immersed for 24 hr in 1M HNO3 and ultrapure water respectively. To ensure an efficient chemical extraction of cellulose, all wood samples were sliced into grains of ca. 3×3×0.5 mm3 (Figure 1g). To reduce the corrosive effects of alkaline solution on the sintered glass filter discs inside the funnels, the first step was executed in the glass vials inside an oven (DHG-9123A, SANFA®, China). Subsequent steps were performed in a water bath environment using our modified extraction devices. The procedure is carried out under a fume hood, due to the release of harmful gases during cellulose extraction. In general, the second base step and the bleaching step should be repeated 3 times. However, fossil wood samples typically require more repetitions than other samples. The samples need to be rinsed in ultrapure water 3 times after the base steps and the first acid step, and 6 times after the bleaching step.

Figure 2. Protocol for holocellulose extraction used in this study. This protocol is a modified version of the Base-Acid-Base-Acid-Bleaching protocol (Němec et al. Reference Němec, Wacker, Hajdas and Gäggeler2010; Sookdeo et al. Reference Sookdeo, Kromer, Büntgen, Friedrich, Friedrich, Helle, Pauly, Nievergelt, Reinig and Treydte2020). Day 1 (blue): The glass funnels and vials are heated at 500°C in a Muffle oven for 4 hr. After cooling down, the vials are labelled and loaded with wood samples. Day 2 (red): Wood samples are treated with 1M NaOH in the vials at 50°C inside an oven for one day. Meanwhile, all parts of the device including glass funnels, PTFE drainage modules, silicone tubes, and covers are immersed in 1M HNO3 for 24 hrs. Day 3 (gray): Wood samples continue to be treated with 1M NaOH for one day. Meanwhile, all parts of the device are rinsed and immersed in ultrapure water for 24 hr. Day 4 (green): The device is assembled. Wood samples are transferred to funnels followed by ultrapure water rinsing for 3 times. Wood samples are treated with 1M HCl at 75°C inside a water bath for 1 hr then rinsed for 3 times. Subsequently, the samples are treated with 1M NaOH for 1 hr. This step should be repeated until the supernatant remains colorless. After being rinsed 3 times, wood samples are treated with 1M HCl again. Day 5 (yellow): The wood samples are bleached for 1 hr with HCl-acidified 5% NaClO2 (pH = 2). This step should be repeated until the wood transforms into holocellulose and the color turns into pure white. After being rinsed 6 times, the holocellulose samples are transferred to vials and then homogenized using an ultrasonic cell crusher. Day 6 & 7 (brown): The holocellulose samples are frozen and lyophilized.

After chemical processing, the treated samples were transferred to clean glass vials, homogenized using an ultrasonic cell crusher (JY96-IIN, LICHEN®, China) and then subjected to lyophilization for 48 hr using a freeze dryer (Alpha 1-4 LSCbasic, Christ™, Germany). Subsequently, FTIR (Nicolet™ iS50, Thermo Scientific™, USA) was used to test the holocellulose purity (Kagawa et al. Reference Kagawa, Sano, Nakatsuka, Ikeda and Kubo2015; Schollaen et al. Reference Schollaen, Baschek, Heinrich, Slotta, Pauly and Helle2017). Extracting holocellulose from 108 samples was completed in less than a week. Since the system is scalable, higher sample throughputs are possible.

Radiocarbon analysis

Each treated sample (2.4 ± 0.1 mg) was wrapped in a tin foil vessel. The sealed samples were combusted at 920 °C in an organic elemental analyzer (vario ISOTOPE SELECT, ELEMENTAR™, Germany) and graphitized at 580 °C with hydrogen gas and iron powder in an automated graphitization system (AGE3, IONPLUS™, Switzerland). F14C (Reimer et al. Reference Reimer, Brown and Reimer2004) measurements were performed on graphitized samples at the Laboratory of AMS Dating and the Environment at Nanjing University on a compact radiocarbon AMS system (MICADAS, IONPLUS™, Switzerland; Wacker et al. Reference Wacker, Bonani, Friedrich, Hajdas, Kromer, Němec, Ruff, Suter, Synal and Vockenhuber2010a). Fractionation and background correction was conducted automatically using the BATS program (Wacker et al. Reference Wacker, Christl and Synal2010b) based on results of standard (OXII, NIST SRM 4990C) and 14C-blank (Phthalic Acid, PhA, Sigma Aldrich®, PN-320064-500g) samples. ΔF14C, which is defined as the deviation between the measured F14C of each replicate and their weighted mean (Santos et al. Reference Santos, Komatsu, Renteria, Brandes, Leong, Collado-Fabbri and De Pol-Holz2023) was calculated to demonstrate scatter of the replicates. Reduced χ22 red, i.e., MSWD; Wendt and Carl Reference Wendt and Carl1991) was used to check the consistency of replicates.

Results

The FTIR results of both untreated and treated wood sample (LWQ1796) along with the international standard cellulose reference sample IAEA-CH3 are shown in Figure 3. In contrast to the wood spectrum, absorption peaks caused by resin lignin (dotted lines, 1596 cm–1, 1510 cm–1, 1455 cm–1, 1265 cm–1 and 1227 cm–1; Pandey and Pitman Reference Pandey and Pitman2003; Schollaen et al. Reference Schollaen, Baschek, Heinrich, Slotta, Pauly and Helle2017) are absent in the spectrum of treated sample. The spectrum of treated sample agrees with that of IAEA-CH3, confirming that lignin had been removed using our modified device. Notably, the absorption peak related to hemi-cellulose (dashed line, 1735 cm–1; Schollaen et al. Reference Schollaen, Baschek, Heinrich, Slotta, Pauly and Helle2017) is not obvious in the spectrum of treated sample, indicating a very low content of hemicellulose in our chemically treated sample. Similar result was produced by Khumalo et al. (Reference Khumalo, Svarva, Zurbach and Nadeau2024). However, we referred to the chemically treated sample as holocellulose because it did not undergo the standard α-cellulose extraction procedure involving alkaline treatment with 17.5% (w/v) NaOH solution (Anchukaitis et al. Reference Anchukaitis, Evans, Lange, Smith, Leavitt and Schrag2008; Gaudinski et al. Reference Gaudinski, Dawson, Quideau, Schuur, Roden, Trumbore, Sandquist, Oh and Wasylishen2005; Němec et al. Reference Němec, Wacker, Hajdas and Gäggeler2010).

Figure 3. The normalized FTIR absorbance spectra of untreated LWQ1796 sample (wood, black line), chemically treated LWQ1796 sample using our modified cellulose-extraction device (holocellulose, red line), and the reference sample IAEA-CH3 (cellulose, blue line). The dotted and dashed vertical lines denote the absorption bands related to lignin and hemicellulose, respectively (Pandey and Pitman Reference Pandey and Pitman2003; Schollaen et al. Reference Schollaen, Baschek, Heinrich, Slotta, Pauly and Helle2017).

The results of the radiocarbon analysis on treated 14C-blank wood samples are shown in Figure 4a–b (the data are listed in Table S1). NF142 and KB, which currently lack reference values, are compared directly with PhA without background correction (Figure 4a). The F14Cwm values (weighted mean of F14C values) for NF142 and KB are 0.0023 ± 0.00003 (n = 10) and 0.0019 ± 0.00005 (n = 4), respectively, which are comparable to PhA (0.0021 ± 0.00004, n = 6). The F14Cwm value of IAEA-C9 is 0.0021 ± 0.0002 (n = 8; i.e., 0.21 ± 0.02 pMC; Figure 4b), which concurs with previously reported reference values (0.12 to 0.21 pMC, Hogg et al. Reference Hogg, Higham, Robertson, Beukens, Kankainen, McCormac, van der Plicht and Stuiver1995; 0.2 ± 0.05 pMC, Scott Reference Scott2003). The ΔF14C of IAEA-C9 varies in the range −0.0006 to 0.0004. The χ2 red value of the IAEA-C9 replicates is 0.5, indicating that the replicated radiocarbon measurements are consistent. These results demonstrate that the 14C-blank wood samples were not influenced by other samples and confirm that no modern contamination was introduced during cellulose extraction.

Figure 4. Radiocarbon analytical results of the wood samples. (a) Comparison of measured F14C values of the 14C-blank (PhA, green diamond), NF142 (blue triangle) and KB samples (pink pentagon). Weighted means of individual F14C values (F14Cwm) are given in the legend (similarly, hereinafter; Table S1). Dotted lines represent F14Cwm, and colored bands represent the ±1σ interval (similarly hereinafter). (b) ΔF14C of the IAEA-C9 samples and the associated χ2 red value. The thinner dotted line with the gray band represents the reference value (F14Cref; Scott Reference Scott2003; similarly, hereinafter). (c–d) ΔF14C of the LWQ1796 and AD1515 samples, respectively. F14Cref is from Brehm et al. Reference Brehm, Bayliss, Christl, Synal, Adolphi, Beer, Kromer, Muscheler, Solanki and Usoskin2021. (e–f) Calibrated F14Cwm results of the LWQ1796 and AD1515 samples derived from the program Oxcal 4.4 (Bronk Ramsey Reference Bronk Ramsey2009) with IntCal20 used as reference (Reimer et al. Reference Reimer, Austin, Bard, Bayliss, Blackwell, Bronk Ramsey, Butzin, Cheng, Edwards and Friedrich2020).

Figures 4c–f show the results of our radiocarbon analysis on wood samples of known age. The ΔF14C of LWQ1796 and AD1515 varies within the range −0.0022 to 0.0032 (n = 9; Figure 4c) and −0.0012 to 0.0014 (n = 4; Figure 4d), respectively. The χ2 red values of LWQ1796 and AD1515 replicates is 0.7 and 0.2, respectively. According to ΔF14C and χ2 red, the replicates of both LWQ1796 and AD1515 are not significantly different. Although all AD1515 replicates have higher F14C values than the reference (F14Cref = 0.95805 ± 0.00014; Brehm et al. Reference Brehm, Bayliss, Christl, Synal, Adolphi, Beer, Kromer, Muscheler, Solanki and Usoskin2021), they are still in statistical agreement. The calibrated ages of LWQ1796 and AD1515 correspond to their known calendar ages (Figure 4e–f). These results indicate that our adapted device and methods are suitable for high-precision radiocarbon analysis.

Discussion

Exogenous contamination, possibly arising from inadequate cleaning of devices or use of inappropriate reagents, can impact high precision radiocarbon dating (Gaudinski et al. Reference Gaudinski, Dawson, Quideau, Schuur, Roden, Trumbore, Sandquist, Oh and Wasylishen2005; Anchukaitis et al. Reference Anchukaitis, Evans, Lange, Smith, Leavitt and Schrag2008). According to Santos et al. (Reference Santos, Granato-Souza, Barbosa, Oelkers and Andreu-Hayles2020), the use of current batchwise devices for tree-ring cellulose extraction yielded sub-optimal results for blank samples. This implies that substances originating from introduced contaminants and/or other wood samples inside the device may taint the blank samples. In this study, the original MSISS is adapted to extract holocellulose for radiocarbon analysis by converting the drainage channel from interconnected to independent, and adding funnel covers. The FTIR results show that our modified device can produce pure holocellulose, while the radiocarbon analysis revealed that there are no effects of contaminates on the blank or reference samples.

Pumping chemical waste out for each sample individually with our modified device requires more operations compared to the original MSISS, which has the capability of pumping dozens of samples simultaneously (Wieloch et al. Reference Wieloch, Helle, Heinrich, Voigt and Schyma2011). However, pumping out chemical waste for individual samples is still labor-saving compared to pipetting chemical waste from test tubes (Santos et al. Reference Santos, Komatsu, Renteria, Brandes, Leong, Collado-Fabbri and De Pol-Holz2023; Southon and Magana Reference Southon and Magana2010). Fogtmann-Schulz et al. (Reference Fogtmann-Schulz, Kudsk, Adolphi, Karoff, Knudsen, Loader, Muscheler, Trant, Østbø and Olsen2021) designed a system that is comparable to ours but contains more intricate and bigger funnels with valves. The smaller funnels of our device enable the simultaneous extraction of more samples. Meanwhile, our modification enables a laboratory currently equipped with MSISS to manufacture the new drainage modules and covers, while the glass funnels, silicone tubes, and vacuum pump remain unchanged. This is more convenient and economical than ordering and designing new devices.

In summary, the adapted tree-ring cellulose extraction device has considerable benefits in reducing contamination and saving labor in the removal of chemical waste. Although the adapted device was initially designed for radiocarbon analysis, considerable benefits can be conducted to ascertain its applicability to cellulose extraction for other plant tissues and stable isotope analysis.

Conclusions

An adapted device for cellulose extraction was developed at the Laboratory of AMS Dating and the Environment at Nanjing University. It is particularly suitable for radiocarbon analysis of large numbers of tree-ring samples based on the following innovations:

  1. (a) Independent drainage channels prevent cross-contamination among samples.

  2. (b) Funnel covers prevent sample contamination from the outside and enhance the efficiency of the NaClO2 treatment step.

The suitability of the adapted device for radiocarbon analysis is confirmed by FTIR and a number of radiocarbon test analyses. Researchers in related disciplines, such as stable isotope analysis, should find the device useful, particularly when there is an increased risk of error due to sample contamination.

Supplementary material

Information about the device design and availability is referred to Text S1. The LWQ tree-ring chronology and the radiocarbon data in Figure 4 are shown in Figure S1 and Table S1.

For supplementary material accompanying this paper visit https://doi.org/10.1017/RDC.2024.83

Acknowledgments

We thank Dr. Irka Hajdas (ETH Zurich, Switzerland) for her valuable comments as well as supplying the wood samples KB and AD1515. We are also grateful to Prof. Tao Su (Chengdu University of Technology, China) for supplying the wood sample NF142. We feel thankful to Prof. Michael Meadows (Nanjing University, China) for his contribution to improve the language. We would like to thank Editor-in-Chief, Prof. Timothy Jull, Associate Editor Prof. Steven Leavitt and two anonymous reviewers for their insightful comments. This research is supported by the National Natural Science Foundation of China (42171155; 42011530119; 42201014), the “GeoX” Interdisciplinary Research Funds for the Frontiers Science Center for Critical Earth Material Cycling, Nanjing University (2023300297) and “Formas - a Swedish Research Council for Sustainable Development” (2022-02833).

References

Anchukaitis, KJ, Evans, MN, Lange, T, Smith, DR, Leavitt, SW and Schrag, DP (2008) Consequences of a rapid cellulose extraction technique for oxygen isotope and radiocarbon analyses. Analytical Chemistry 80(6), 20352041.CrossRefGoogle ScholarPubMed
Andreu-Hayles, L, Levesque, M, Martin-Benito, D, Huang, W, Harris, R, Oelkers, R, Leland, C, Martin-Fernández, J, Anchukaitis, KJ and Helle, G (2019) A high yield cellulose extraction system for small whole wood samples and dual measurement of carbon and oxygen stable isotopes. Chemical Geology 504, 5365.CrossRefGoogle Scholar
Bayliss, A, Howard, R and Tyers, C (2023) Dendrochronological dating of known-age tree-ring radiocarbon standards from Windsor Castle, Berkshire. Historic England Research Report Series 68/2023.Google Scholar
Brehm, N, Bayliss, A, Christl, M, Synal, H, Adolphi, F, Beer, J, Kromer, B, Muscheler, R, Solanki, SK, Usoskin, I et al. (2021) Eleven-year solar cycles over the last millennium revealed by radiocarbon in tree rings. Nature Geoscience 14(1), 1015.CrossRefGoogle Scholar
Brehm, N, Christl, M, Knowles, TDJ, Casanova, E, Evershed, RP, Adolphi, F, Muscheler, R, Synal, H, Mekhaldi, F, Paleari, CI et al. (2022) Tree-rings reveal two strong solar proton events in 7176 and 5259 BCE. Nature Communications 13(1), 1196.CrossRefGoogle ScholarPubMed
Bronk Ramsey, C (2009) Bayesian analysis of radiocarbon dates. Radiocarbon 51(1), 337360.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.CrossRefGoogle ScholarPubMed
Capano, M, Miramont, C, Guibal, F, Kromer, B, Tuna, T, Fagault, Y and Bard, E (2018) Wood 14C dating with AixMICADAS: Methods and application to tree-ring sequences from the Younger Dryas event in the southern French Alps. Radiocarbon 60(1), 5174.CrossRefGoogle Scholar
Chen, F, Shang, H, Panyushkina, I, Meko, D, Li, J, Yuan, Y, Yu, S, Chen, F, He, D and Luo, X (2019) 500-year tree-ring reconstruction of Salween River streamflow related to the history of water supply in Southeast Asia. Climate Dynamics 53(11), 65956607.CrossRefGoogle Scholar
Cullen, LE and Macfarlane, C (2005) Comparison of cellulose extraction methods for analysis of stable-isotope ratios of carbon and oxygen in plant material. Tree Physiology 25(5), 563569.CrossRefGoogle ScholarPubMed
Damon, PE and Sonett, CP (1991) Solar and terrestrial components of the atmospheric 14C variation spectrum. In Sonett, CP, Giampapa, MS and Mattheus, MS (eds), The Sun in Time. Tucson: University of Arizona Press, 360388.Google Scholar
Fogtmann-Schulz, A, Kudsk, SGK, Adolphi, F, Karoff, C, Knudsen, MF, Loader, NJ, Muscheler, R, Trant, PLK, Østbø, SM and Olsen, J (2021) Batch processing of tree ring samples for radiocarbon analysis. Radiocarbon 63(1), 7789.CrossRefGoogle Scholar
Gaudinski, JB, Dawson, TE, Quideau, S, Schuur, EAG, Roden, JS, Trumbore, SE, Sandquist, DR, Oh, S and Wasylishen, RE (2005) Comparative analysis of cellulose preparation techniques for use with 13C, 14C, and 18O isotopic measurements. Analytical Chemistry 77(22), 72127224.CrossRefGoogle ScholarPubMed
Green, JW (1963) Wood cellulose. In Whistler, RL (ed), Methods in Carbohydrate Chemistry 3. New York: Academic Press, 921.Google Scholar
Güttler, D, Wacker, L, Kromer, B, Friedrich, M and Synal, HA (2013) Evidence of 11-year solar cycles in tree rings from 1010 to 1110 AD—progress on high precision AMS measurements. Nuclear Instruments and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms 294, 459463.CrossRefGoogle Scholar
Hajdas, I, Hendriks, L, Fontana, A and Monegato, G (2017) Evaluation of preparation methods in radiocarbon dating of old wood. Radiocarbon 59(3), 727737.CrossRefGoogle Scholar
Harada, M, Watanabe, Y, Nakatsuka, T, Tazuru-mizuno, S, Horikawa, Y, Sugiyama, J, Tsuda, T and Tagami, T (2014) Alpha-cellulose extraction procedure for the tropical tree sungkai (Peronema canescens Jack) by using an improved vessel for reliable paleoclimate reconstruction. Geochemical Journal 48(3), 299307.CrossRefGoogle Scholar
Head, MJ (1979) Structure and chemical properties of fresh and degraded wood: Their effects on radiocarbon activity measurements. Unpublished MSc thesis, The Australian National University, Canberra.Google Scholar
Hogg, AG, Heaton, TJ, Hua, Q, Palmer, JG, Turney, CS, Southon, J, Bayliss, A, Blackwell, PG, Boswijk, G, Bronk Ramsey, C et al. (2020) SHCal20: Southern Hemisphere calibration, 0–55,000 years cal BP. Radiocarbon 62(4), 759778.CrossRefGoogle Scholar
Hogg, AG, Higham, T, Robertson, S, Beukens, R, Kankainen, T, McCormac, FG, van der Plicht, J and Stuiver, M (1995) Radiocarbon age assessment of a new, near background IAEA 14C quality assurance material. Radiocarbon 37(2), 797803.CrossRefGoogle Scholar
Hoper, ST, McCormac, FG, Hogg, AG, Higham, TFG and Head, MJ (1998) Evaluation of wood pretreatments on oak and cedar. Radiocarbon 40(1), 4550.CrossRefGoogle Scholar
Hua, Q, Turnbull, JC, Santos, GM, Rakowski, AZ, Ancapichún, S, De Pol-Holz, R, Hammer, S, Lehman, SJ, Levin, I, Miller, JB et al. (2021) Atmospheric radiocarbon for the period 1950–2019. Radiocarbon 64(4), 723745.CrossRefGoogle Scholar
Kagawa, A, Sano, M, Nakatsuka, T, Ikeda, T and Kubo, S (2015) An optimized method for stable isotope analysis of tree rings by extracting cellulose directly from cross-sectional laths. Chemical Geology 393–394, 1625.CrossRefGoogle Scholar
Khumalo, WH, Svarva, HL, Zurbach, D and Nadeau, M (2024) Squeaky clean cellulose: Comparing pretreatment effectiveness on single tree rings and wooden laths. Radiocarbon Published online, 1–13. doi:10.1017/RDC.2024.20 CrossRefGoogle Scholar
Lange, TE, Nordby, JA, Murphy, PLO, Hodgins, GWL and Pearson, CL (2019) A detailed investigation of pretreatment protocols for high precision radiocarbon measurements of annual tree-rings. Nuclear Instruments and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms 455, 230233.CrossRefGoogle Scholar
Leavitt, SW and Danzer, SR (1993) Method for batch processing small wood samples to holocellulose for stable-carbon isotope analysis. Analytical Chemistry 65(1), 8789.CrossRefGoogle Scholar
Li, Z, Labbé, N, Driese, SG and Grissino-Mayer, HD (2011) Micro-scale analysis of tree-ring δ18O and δ13C on α-cellulose spline reveals high-resolution intra-annual climate variability and tropical cyclone activity. Chemical Geology 284, 138147.CrossRefGoogle Scholar
Loader, NJ, Robertson, I, Barker, AC, Switsur, VP and Waterhouse, JS (1997) An improved technique for the batch processing of small wholewood samples to alpha-cellulose. Chemical Geology 136(3–4), 313317.CrossRefGoogle Scholar
Michczyńska, DJ, Krąpiec, M, Michczyński, A, Pawlyta, J, Goslar, T, Nawrocka, N, Piotrowska, N, Szychowska-Krąpiec, E, Waliszewska, B and 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
Miyake, F, Masuda, K and Nakamura, T (2013) Another rapid event in the carbon-14 content of tree rings. Nature Communications 4(1), 1748.CrossRefGoogle ScholarPubMed
Miyake, F, Nagaya, K, Masuda, K and Nakamura, T (2012) A signature of cosmic-ray increase in AD 774–775 from tree rings in Japan. Nature 486, 240242.CrossRefGoogle ScholarPubMed
Němec, M, Wacker, L, Hajdas, I and Gäggeler, H (2010) Alternative methods for cellulose preparation for AMS measurement. Radiocarbon 52(3), 13581370.CrossRefGoogle Scholar
Pandey, KK and Pitman, AJ (2003) FTIR studies of the changes in wood chemistry following decay by brown-rot and white-rot fungi. International Biodeterioration & Biodegradation. 52(3), 151160.CrossRefGoogle Scholar
Reimer, PJ, Austin, WEN, Bard, E, Bayliss, A, Blackwell, PG, Bronk Ramsey, C, 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.CrossRefGoogle Scholar
Reimer, PJ, Brown, TA and Reimer, RW (2004) Discussion: Reporting and calibration of post-bomb 14C data. Radiocarbon 46(3), 12991304.Google Scholar
Rinne, KT, Boettger, T, Loader, NJ, Robertson, I, Switsur, VR and Waterhouse, JS (2005) On the purification of α-cellulose from resinous wood for stable isotope (H, C and O) analysis. Chemical Geology 222(1), 7582.CrossRefGoogle Scholar
Santos, GM, Granato-Souza, D, Barbosa, AC, Oelkers, R and Andreu-Hayles, L (2020) Radiocarbon analysis confirms annual periodicity in Cedrela odorata tree rings from the equatorial Amazon. Quaternary Geochronology 58, 101079.CrossRefGoogle Scholar
Santos, GM, Komatsu, ASY, Renteria, JM Jr., Brandes, AFN, Leong, CA, Collado-Fabbri, S and De Pol-Holz, R (2023) A universal approach to alpha-cellulose extraction for radiocarbon analysis of 14C-free to post-bomb ages. Quaternary Geochronology 74, 101414. CrossRefGoogle Scholar
Schollaen, K, Baschek, H, Heinrich, I, Slotta, F, Pauly, M and Helle, G (2017) A guideline for sample preparation in modern tree-ring stable isotope research. Dendrochronologia 44, 133145.CrossRefGoogle Scholar
Scott, EM (2003) Section 1: The Fourth International Radiocarbon Intercomparison (FIRI). Radiocarbon 45(2), 135150.CrossRefGoogle Scholar
Sookdeo, A, Kromer, B, Büntgen, U, Friedrich, M, Friedrich, R, Helle, G, Pauly, M, Nievergelt, D, Reinig, F, Treydte, K et al. (2020) Quality dating: A well-defined protocol implemented at ETH for high-precision 14C-dates tested on Late Glacial wood. Radiocarbon 62(4), 891899.CrossRefGoogle Scholar
Southon, JR and Magana, AL (2010) A comparison of cellulose extraction and ABA pretreatment methods for AMS 14C dating of ancient wood. Radiocarbon 52(3), 13711379.CrossRefGoogle Scholar
Staff, RA, Reynard, L, Brock, F and Bronk Ramsey, C (2014) Wood pretreatment protocols and measurement of tree-ring standards at the Oxford Radiocarbon Accelerator Unit (ORAU). Radiocarbon 56(2), 709715.CrossRefGoogle Scholar
Stuiver, M and Braziunas, TF (1993) Sun, ocean, climate and atmospheric 14CO2: an evaluation of causal and spectral relationships. The Holocene 3(4), 289305.CrossRefGoogle Scholar
Stuiver, M and Quay, PD (1981) Atmospheric 14C changes resulting from fossil fuel CO2 release and cosmic ray flux variability. Earth and Planetary Science Letters 53(3), 349362.CrossRefGoogle Scholar
Suess, HE (1980) The radiocarbon record in tree rings of the last 8000 years. Radiocarbon 22(2), 200209.CrossRefGoogle Scholar
Taylor, RE and Bar-Yosef, O (2016) Radiocarbon Dating. 2nd edition. Walnut Creek: Left Coast Press.Google Scholar
Usoskin, IG, Kromer, B, Ludlow, F, Beer, J, Friedrich, M, Kovaltsov, GA, Solanki, SK and Wacker, L (2013) The AD775 cosmic event revisited: The sun is to blame. Astronomy & Astrophysics 552, L3.CrossRefGoogle Scholar
Usoskin, IG, Solanki, SK, Krivova, NA, Hofer, B, Kovaltsov, GA, Wacker, L, Brehm, N and Kromer, B (2021) Solar cyclic activity over the last millennium reconstructed from annual 14C data. Astronomy & Astrophysics 649, A141.CrossRefGoogle Scholar
Uusitalo, J, Arppe, L, Hackman, T, Helama, S, Kovaltsov, G, Mielikäinen, K, Mäkinen, H, Nöjd, P, Palonen, V, Usoskin, I et al. (2018) Solar superstorm of AD 774 recorded subannually by Arctic tree rings. Nature Communications 9, 3495.CrossRefGoogle ScholarPubMed
Wacker, L, Bonani, G, Friedrich, M, Hajdas, I, Kromer, B, Němec, M, Ruff, M, Suter, M, Synal, H and Vockenhuber, C (2010a) MICADAS: Routine and high-precision radiocarbon dating. Radiocarbon 52(3), 252262.CrossRefGoogle Scholar
Wacker, L, Christl, M and Synal, H (2010b) Bats: A new tool for AMS data reduction. Nuclear Instruments and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms 268(7–8), 976979.CrossRefGoogle Scholar
Wacker, L, Güttler, D, Goll, J, Hurni, JP, Synal, H and Walti, N (2014) Radiocarbon dating to a single year by means of rapid atmospheric 14C changes. Radiocarbon 56(2), 573579.CrossRefGoogle Scholar
Wang, H, Oskolski, AA, Jacques, FMB, Wang, Y and Zhou, Z (2017) Lignified woods of Pinus (Pinaceae) from the late Miocene of central Yunnan, China, and their biogeographic and paleoclimatic implications. Palaeoworld 26(3), 553565.CrossRefGoogle Scholar
Wendt, I and Carl, C (1991) The statistical distribution of the mean squared weighted deviation. Chemical Geology: Isotope Geoscience Section 86(4), 275285.Google Scholar
Wieloch, T, Helle, G, Heinrich, I, Voigt, M and Schyma, P (2011) A novel device for batch-wise isolation of α-cellulose from small-amount wholewood samples. Dendrochronologia 29(2), 115117.CrossRefGoogle Scholar
Figure 0

Figure 1. An illustration of the device and sliced samples used to extract tree-ring cellulose for radiocarbon analysis. (a) The three-dimensional structural diagram of the modified device. The left part of the diagram shows the internal drainage channels. (b-c) An image and side view of the device: 1. silicone tubing outlet used to remove chemical waste; 2. PTFE cover used to prevent external contamination with its side view shown at the top left; 3. glass funnel used for loading tree-ring sample; 4. silicone tubing used to connect funnels to the drainage module; 5. PTFE screw; 6. sintered glass filter disc inside the funnel (similarly, hereinafter); 7. sliced wood samples inside the funnel (similarly, hereinafter). The red curve with arrows indicates the flow path of chemical waste during the pumping process. (d) Experimental setup used here for batchwise extraction of tree-ring cellulose. For chemical treatments, six devices containing 108 tree-ring samples were heated in a water bath (600 mm × 300 mm). With a vacuum pump, chemical waste can be pumped through the outlets at the top of PTFE block. (e) Contaminants (black arrow) that fell on the cover during the experiment. (f) The funnel. The filter disc at the base of the funnel has an aperture size of ca. 30–50 μm. (g) The wood samples that were sliced into pieces around 3×3×0.5 mm3 in size.

Figure 1

Table 1. General information about the wood samples used for assessing the suitability of our modified tree-ring cellulose extraction device for radiocarbon analysis

Figure 2

Figure 2. Protocol for holocellulose extraction used in this study. This protocol is a modified version of the Base-Acid-Base-Acid-Bleaching protocol (Němec et al. 2010; Sookdeo et al. 2020). Day 1 (blue): The glass funnels and vials are heated at 500°C in a Muffle oven for 4 hr. After cooling down, the vials are labelled and loaded with wood samples. Day 2 (red): Wood samples are treated with 1M NaOH in the vials at 50°C inside an oven for one day. Meanwhile, all parts of the device including glass funnels, PTFE drainage modules, silicone tubes, and covers are immersed in 1M HNO3 for 24 hrs. Day 3 (gray): Wood samples continue to be treated with 1M NaOH for one day. Meanwhile, all parts of the device are rinsed and immersed in ultrapure water for 24 hr. Day 4 (green): The device is assembled. Wood samples are transferred to funnels followed by ultrapure water rinsing for 3 times. Wood samples are treated with 1M HCl at 75°C inside a water bath for 1 hr then rinsed for 3 times. Subsequently, the samples are treated with 1M NaOH for 1 hr. This step should be repeated until the supernatant remains colorless. After being rinsed 3 times, wood samples are treated with 1M HCl again. Day 5 (yellow): The wood samples are bleached for 1 hr with HCl-acidified 5% NaClO2 (pH = 2). This step should be repeated until the wood transforms into holocellulose and the color turns into pure white. After being rinsed 6 times, the holocellulose samples are transferred to vials and then homogenized using an ultrasonic cell crusher. Day 6 & 7 (brown): The holocellulose samples are frozen and lyophilized.

Figure 3

Figure 3. The normalized FTIR absorbance spectra of untreated LWQ1796 sample (wood, black line), chemically treated LWQ1796 sample using our modified cellulose-extraction device (holocellulose, red line), and the reference sample IAEA-CH3 (cellulose, blue line). The dotted and dashed vertical lines denote the absorption bands related to lignin and hemicellulose, respectively (Pandey and Pitman 2003; Schollaen et al. 2017).

Figure 4

Figure 4. Radiocarbon analytical results of the wood samples. (a) Comparison of measured F14C values of the 14C-blank (PhA, green diamond), NF142 (blue triangle) and KB samples (pink pentagon). Weighted means of individual F14C values (F14Cwm) are given in the legend (similarly, hereinafter; Table S1). Dotted lines represent F14Cwm, and colored bands represent the ±1σ interval (similarly hereinafter). (b) ΔF14C of the IAEA-C9 samples and the associated χ2red value. The thinner dotted line with the gray band represents the reference value (F14Cref; Scott 2003; similarly, hereinafter). (c–d) ΔF14C of the LWQ1796 and AD1515 samples, respectively. F14Cref is from Brehm et al. 2021. (e–f) Calibrated F14Cwm results of the LWQ1796 and AD1515 samples derived from the program Oxcal 4.4 (Bronk Ramsey 2009) with IntCal20 used as reference (Reimer et al. 2020).

Supplementary material: File

Lin et al. supplementary material

Lin et al. supplementary material
Download Lin et al. supplementary material(File)
File 1.5 MB