Hostname: page-component-7bb8b95d7b-l4ctd Total loading time: 0 Render date: 2024-09-27T03:45:08.691Z Has data issue: false hasContentIssue false

RETREATMENT OF BONE MATERIAL IN THE GLIWICE RADIOCARBON LABORATORY USING ULTRAFILTRATION

Published online by Cambridge University Press:  19 March 2024

Fatima Pawełczyk*
Affiliation:
Institute of Physics – CSE, Silesian University of Technology, Konarskiego 22B, 44-100 Gliwice, Poland
Magdalena Niedziałkowska
Affiliation:
Mammal Research Institute Polish Academy of Sciences, Stoczek 1, 17-230 Białowieża, Poland
Sławomira Pawełczyk
Affiliation:
Institute of Physics – CSE, Silesian University of Technology, Konarskiego 22B, 44-100 Gliwice, Poland
Natalia Piotrowska
Affiliation:
Institute of Physics – CSE, Silesian University of Technology, Konarskiego 22B, 44-100 Gliwice, Poland
Maciej Sykut
Affiliation:
Mammal Research Institute Polish Academy of Sciences, Stoczek 1, 17-230 Białowieża, Poland
*
*Corresponding author. Email: fatima.pawelczyk@polsl.pl
Rights & Permissions [Opens in a new window]

Abstract

Preparation of bones for radiocarbon (14C) dating is still quite a challenge for researchers. The methods are being tested and improved, to increase reliability of dating results and to verify the previous ones. In this work, a set of gelatine samples, extracted from Cervus elaphus and Cervus canadensis bones from various sites in Europe and a set of human bones from archaeological sites in Poland were subjected to retreatment using ultrafiltration in Gliwice Radiocarbon Laboratory. The tested samples represent a wide range of ages, from older than 40,000 14C years BP to modern. The prepared material was subjected to the measurement of C/N atomic ratios and 14C dating using the accelerator mass spectrometry (AMS) technique. Also, the stable isotopes (δ13C and δ15N) values were determined. In a few cases ultrafiltration allows to improve gelatine quality for long-stored samples, by increasing the %C and %N as well as decreasing C/Nat ratios. Nevertheless, this effect was not observed for majority of the samples. Remeasurements of long-term stored samples give mostly the same 14C ages for ultrafiltered ones and for those without ultrafiltration.

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

INTRODUCTION

Radiocarbon (14C) dating of bone material has been a challenge for researchers since the first attempts, giving results contradicting archaeological context. The age offset can be caused by contamination with exogenous carbon. In a natural burial environment these are mostly humic acids reacting with the collagen triple helix (Van Klinken and Hedges Reference Van Klinken and Hedges1995). The geochemistry of the environment in which the bone was deposited is important for the collagen degradation and contamination. Another source of contamination with exogenous carbon, in case of archaeological collections, can be preservation agents, like wax, consolidants, adhesives, varnishes, glue or resin (Fedi et al. Reference Fedi, Caforio, Liccioli, Mandò, Salvini and Taccetti2014; Brock et al. Reference Brock, Dee, Hughes, Snoeck, Staff and Ramsey2018; Crann and Grant Reference Crann and Grant2019). These substances cause the effect of altering the 14C date, with the magnitude depending on the age of the bones, their preservation level, and the age of exogenous carbon. Because of that, the results obtained on bone samples in the past were often rejected (Taylor Reference Taylor, Taylor, Long and Kra1992).

Today, 14C dating is done on the extracted collagen or its fractions (not bulk bone) and even small samples, below 100 mg (Fewlass et al. Reference Fewlass, Tuna, Fagault, Hublin, Kromer, Bard and Talamo2019), or very old samples, near the limit of the 14C method (Hajdas et al. Reference Hajdas, Michczynski, Bonani, Wacker and Furrer2009) were dated successfully. Most of the methods are based on the classical Longin method of collagen isolation (Longin Reference Longin1971), supplemented by the step of ultrafiltration (Brown et al. Reference Brown, Nelson, Vogel and Southon1988), which removes low-molecular weight material. The ultrafiltration method has been a subject of wide research since it was first used. Different preparation improvements have been tested to solve the encountered problems, like different acid strength and decalcification time (Shammas 2009; Talamo and Richards Reference Talamo and Richards2011), gelatinization temperature (Beaumont et al. Reference Beaumont, Beverly, Southon and Taylor2010) or different ultrafilters (Talamo et al. Reference Talamo, Fewlass, Maria and Jaouen2021). It was also noticed that the type of vessels and time of heating (as well as heating apparatus used) may affect the gelatinization step (Brock et al. Reference Brock, Geoghegan, Thomas, Jurkschat and Higham2013a). Some of the researchers underline that there is a risk of sample contamination at the ultrafiltration step (e.g., Bronk Ramsey et al. Reference Bronk Ramsey, Higham, Bowles and Hedges2004; Hüls et al. Reference Hüls, Grootes and Nadeau2007, Reference Hüls, Grootes and Nadeau2009; Beaumont et al. Reference Beaumont, Beverly, Southon and Taylor2010; Brock et al. Reference Brock, Higham and Ramsey2013b), so the improvements of ultrafiltration protocols covered also increased intensity of the ultrafilters precleaning (Bronk Ramsey et al. Reference Bronk Ramsey, Higham, Bowles and Hedges2004; Brock et al. Reference Brock, Ramsey and Higham2007).

Besides ultrafiltration, other methods have also been used. It was proved that chromatographic isolation of collagen-specific peptides by HPLC improves 14C dating of contaminated bones (Van Klinken and Hedges Reference Van Klinken and Hedges1992; Van Klinken et al. Reference Van Klinken, Bowles and Hedges1994). Hydroxyproline is supposed to be a bone-specific biomarker, since it constitutes about 10% of bone collagen (Marom et al. Reference Marom, McCullagh, Higham, Sinitsyn and Hedges2012, Reference Marom, McCullagh, Higham and Hedges2013; Marom-Rotem Reference Marom-Rotem2012). According to this, a more specific approach is, for example, the chromatographic separation of hydroxyproline or purification of amino acids with XAD-2 resin (e.g., Stafford et al. Reference Stafford, Duhamel, Haynes and Brendel1982; Gillespie et al. Reference Gillespie, Hedges and Wand1984; Stafford et al. Reference Stafford, Brendel and Duhamel1988; McCullagh et al. Reference McCullagh, Marom and Hedges2010; Marom et al. Reference Marom, McCullagh, Higham and Hedges2013; Deviese et al. Reference Deviese, Comeskey, McCullagh, Bronk Ramsey and Higham2018; Minami and Nakamura Reference Minami and Nakamura2000; Yuan et al. Reference Yuan, Wu, Gao, Wang, Cai, Liu, Li and Ma2000). Another preparation method is the reaction of ninhydrin with amino acids, performed by Nelson (Reference Nelson1991), who proved the applicability of this method to any protein, as it allowed to obtain a well-purified sample. Additionally, some experiments to date the bioapatite fraction have been proved successful (e.g., Cherkinsky Reference Cherkinsky2009). For an overview of methods used for preparation bones for 14C dating see Herrando-Pérez (Reference Herrando-Pérez2021).

When 14C dating bones, an assessment of gelatine quality should be performed. A few criteria for that assessment have been adopted. In this study we used the values for %C, %N and C/Nat ratios as quality indicator, with range of the values published by Ambrose (Reference Ambrose1990): 15.3–47% for carbon and 5.5–17.3% for nitrogen, by mass. The C/N ratio for collagen is 3.2 and samples with C/N value between 2.9 and 3.6 are considered acceptable, while higher values suggest contamination with exogenous carbon (DeNiro Reference DeNiro1985; Ambrose Reference Ambrose1990). Another marker is %N in raw bone, which is an estimation for the amount of the surviving collagen and should be at least 5% for a bone to be considered as useful (DeNiro and Weiner Reference DeNiro and Weiner1988; Van Klinken and Hedges Reference Van Klinken and Hedges1992).

In this study, we analyze human and animal bones of different ages that come from many different places in Europe and Western Asia. Among them, there are samples from bones of red deer (Cervus elaphus) and wapiti (Cervus canadensis). The red deer is a savanna-type deer with a mixed feeding strategy that inhabits mainly Europe and southwestern Asia. Wapiti is a more cold-adapted open-country grazer that inhabits dry, cold, continental regions of East Asia and North America (Geist Reference Geist1998).

A large set of red deer and wapiti bones was dated between 2015 and 2019 to investigate the spatial distribution of this species in Europe and western Asia since the Late Pleistocene (Doan et al. Reference Doan, Zachos, Wilkens, Vigne, Piotrowska, Stanković, Jędrzejewska, Stefaniak and Niedziałkowska2017; Niedziałkowska et al. Reference Niedziałkowska, Doan, Górny, Sykut, Stefaniak, Piotrowska, Jędrzejewska, Ridush, Pawełczyk and Mackiewicz2021; Doan et al. Reference Doan, Niedziałkowska, Stefaniak, Sykut, Jędrzejewska, Ratajczak-Skrzatek, Piotrowska, Ridush, Zachos and Popović2022). These samples came from various types of sites—caves, river valleys, dunes and sites located in the vicinity of human settlements (for detailed information about the localities, species and type of material used see the supplementary material). Some of the bone samples were very poorly preserved and gave unsatisfactory results of carbon and nitrogen concentrations and carbon to nitrogen atomic ratios. Therefore, we decided to test them again, using a pretreatment extended with ultrafiltration.

Furthermore, we used a set of seven archaeological samples from Poland that came from human bones and deer antler. These samples are numbered 1–7 and have to remain anonymized due to copyright. Finally, 21 samples were chosen for 14C analysis by accelerator mass spectrometry (AMS). Additionally, three reference samples of known age were tested. Our goal was to conduct ultrafiltration tests in the Gliwice Radiocarbon Laboratory to implement this method as a standard method of bone preparation for the AMS 14C dating. We also wanted to investigate the effect of ultrafiltration on long-stored collagen samples.

MATERIALS AND METHODS

In this work, we prepared two batches of samples, previously dated in our laboratory. Batch 1 comprised seven samples from human bones and deer antler from four archaeological sites located in Poland (see Table 2). The gelatine was extracted previously in our laboratory, following a standard preparation. It included mechanical abrasion of the bones, cleaning in demineralized water in an ultrasonic bath, then drying and grinding in a hand mortar to ∼1 mm particles. The gelatine extraction was performed according to the Longin method (Longin Reference Longin1971), modified by Piotrowska and Goslar (Reference Piotrowska and Goslar2002), with the use of hydrochloric acid and alkali solution at room temperature. The 0.5 M hydrochloric acid was replaced several times, and the reaction was considered complete when the pH stabilized at < 1 and no bubbles were observed. The insoluble residues were rinsed with demineralized water and treated with 0.1 M sodium hydroxide for 30 min. Then, after rinsing with demineralized water and acid, the residues were subjected to gelatinisation in acid (pH=3) for at least 12 hours in 80°C and dried.

In the present study, these gelatine samples were redissolved and filtered through precleaned 9 mL Ezee FilterTM separators (Elkay). These separators are polypropylene tubes with sintered polyethylene filters. Then, every sample was divided volumetrically into two equal subsamples: A and B. Set “A” was freeze dried (Alpha 1-2 LD plus Martin Christ) after filtration, then graphitized and measured. Set “B” was additionally subjected to the ultrafiltration. This step was carried out using Millipore Amicon® Ultra-15 ultrafiltration tubes (Merck), carefully precleaned following the protocol used at the Ion Beam Physics Laboratory at the ETH Zurich, which is based on Bronk Ramsey et al. (Reference Bronk Ramsey, Higham, Bowles and Hedges2004) and Brock et al. (Reference Brock, Ramsey and Higham2007). The ultrafilters are produced from high recovery Ultracel regenerated cellulose membrane of 30 kD molecular weight cut-off. The cleaning protocol included rinsing in demineralized water, then ultrasonic bath for 15 min and triple centrifugation to remove any residual humectant added by the manufacturer. The gelatine samples with an initial volume of 15 mL were then centrifuged at 4400 rpm for 5–20 min to remove the fraction <30 kD and collect the heavy-molecule fraction >30 kD. The obtained supernatant having a volume of at least 1500 µL was then freeze dried, graphitized, and measured in the same way as in the set “A”. Additionally, batch 1 included three gelatine samples: SIRI C, VIRI E and VIRI H from international 14C intercomparison programmes (Scott et al. Reference Scott, Cook and Naysmith2010, Reference Scott, Naysmith and Cook2017; see Table 1). They were also divided into two subsets—A and B, as above—and subjected to the same 14C preparation and dating procedures.

Table 1 Results of weight% of C and N, C/N atomic ratios (C/Nat) determinations, and 14C measurements for international 14C intercomparison samples (from batch 1); Treatment methods: A = without ultrafiltration, B = with ultrafiltration; Previously obtained 14C age values after Piotrowska et al. (Reference Piotrowska, Tomczyk, Pawełczyk and Stanaszek2019) are listed in italics as set C (without the ultrafiltration step); Consensus values after Scott et al. (Reference Scott, Cook and Naysmith2010) for VIRI and Scott et al. (Reference Scott, Naysmith and Cook2017) for SIRI.

Batch 2 consisted of 14 gelatine samples of red deer and wapiti bones from Russia, Italy, Belarus, Ukraine, Armenia, and Austria, prepared in our laboratory and AMS measured in 2015–2019 (see also the supplementary material). The preparation was slightly different—with the exclusion of alkali treatment, to minimize sample loss. The bone samples, after being abraded and cleaned in an ultrasonic bath in demineralized water, were dried and ground in a ball mill (Pulverisette 6, Fritsch). Next, the bone powder was demineralized in hydrochloric acid and the insoluble residues were rinsed with demineralized water and subjected to gelatinisation, same as batch 1. In this study, the dried gelatine samples stored at room temperature were redissolved and then filtered with precleaned Ezee Filters. Next, every sample was subjected to ultrafiltration and then freeze dried, graphitized, and measured in the same way as batch 1, serie B.

The graphitization step was performed using an AGE-3 system with a VarioMicroCube by Elementar elemental analyzer and an automated graphitization unit (Nemec et al. Reference Nemec, Wacker and Gäggeler2010; Wacker et al. Reference Wacker, Nemec and Bourquin2010). The analyzer was calibrated with the use of aspartic acid and glutamic acid as reference materials to obtain the weight percent of C and N. C/Nat were calculated taking into account the atomic masses of C and N, according to the formula:

$$C/{N_{at}} = {{{\% C}}\over{{\% N}}}*{{{14}}\over{{12}}}$$

The 14C concentrations in all samples, Oxalic Acid II standards, coal and blank samples were measured by the Poznan Radiocarbon Laboratory, Poland (Goslar et al. Reference Goslar, Czernik and Goslar2004).

Furthermore, for eight samples, stable carbon, and nitrogen isotopes (δ13C, δ15N) were determined. Freeze-dried gelatine samples were weighed in tin capsules. If the volume of the sample allowed, three subsamples of the gelatine sample were prepared for measurements. The analysis was performed by the Gliwice Mass Spectrometry Laboratory using an IsoPrime EA-CF-IRMS continuous-flow isotope ratio mass spectrometer connected to the EuroVector elemental analyzer. The instrument precision is 0.1‰ for δ13C and 0.3‰ for δ15N for one subsample. The obtained carbon and nitrogen isotope measurements were calibrated to the VPDB and AIR standards, respectively (Gonfiantini et al. Reference Gonfiantini, Rozanski and Stichler1990; Misarti et al. Reference Misarti, Finney, Maschner and Wooller2009).

RESULTS AND DISCUSSION

The results of the batch 1 measurements are shown in Table 1 for the reference samples (SIRI C, VIRI H, and E) and in Table 2 for the Polish archaeological samples. Table 3 shows the measurement results for the batch 2—a set of red deer and wapiti samples. All results are presented for both treatment ways: without (A) and with (B) the ultrafiltration step, in comparison to the ones obtained previously without ultrafiltration (C). Figure 1 shows the comparison of the C and N concentration measurements of samples treated with and without ultrafiltration. C/Nat and the 14C ages of the investigated samples are shown in Figure 2. Figure 3 shows the results of the stable isotope measurements of the investigated samples.

Table 2 Results of the% weight of C and N, C/Nat and 14C measurements of anonymized archaeological samples—human bones but GdA-6163 (Sample 7)—deer antler (from batch 1). Samples were divided into two sets of subsamples: A, treated without ultrafiltration, and B, treated with ultrafiltration. In addition, the values obtained previously were listed in italics (= Set C, without the ultrafiltration step). Small samples (below 200 mg of raw sample) were marked with an asterisk(*).

Table 3 Results of weight % of C and N, C/Nat, stable isotope ratios (δ13C, δ15N) and 14C measurements of red deer and wapiti bone samples from Europe and Western Asia (Doan et al. 2017; Reference Doan, Niedziałkowska, Stefaniak, Sykut, Jędrzejewska, Ratajczak-Skrzatek, Piotrowska, Ridush, Zachos and Popović2022; Niedziałkowska et al. Reference Niedziałkowska, Doan, Górny, Sykut, Stefaniak, Piotrowska, Jędrzejewska, Ridush, Pawełczyk and Mackiewicz2021); Treatment methods: B = with ultrafiltration (this study), C = without ultrafiltration (previous studies; in italics). When only one sample was available for stable isotope ratio measurement, it was marked with an asterisk (*).

Figure 1 Changes in C and N concentration measured in all investigated samples; preparation without ultrafiltration step—open symbols; preparation with ultrafiltration step—filled symbols. The majority of anonymized Polish archaeological samples marked with an ellipse. Numbers indicate GdA laboratory codes.

Figure 2 C/Nat ratios and 14C ages. Preparation without ultrafiltration step: open symbols; preparation with ultrafiltration step: filled symbols. Numbers indicate GdA laboratory codes.

Figure 3 The stable isotope (C and N) composition of the investigated samples. Preparation without ultrafiltration step: open symbols; preparation with ultrafiltration step: filled symbols. Numbers indicate GdA laboratory codes.

Remeasurements of the reference samples show a good quality of gelatine preservation, with C/Nat values 3.3–3.4 (Table 1). The C/Nat values do not differ by more than 0.2 in comparison to the previous ones, which is consistent with the values obtained in intercomparison studies (Sealy et al. Reference Sealy, Johnson, Richards and Nehlich2014) and the usual scatter of C/N measurements as shown by Scirè Calabrisotto et al. (Reference Scirè Calabrisotto, Fedi, Caforio, Bombardieri and Mando2013) and Svyatko et al. (Reference Svyatko, Mertz and Reimer2015).

The results of the 14C remeasurements (A and B) of the VIRI H show perfect consistency. In case of the sample SIRI C, subsamples A and B are consistent with each other within 1 σ and with the C subsample within 2 σ. It has to be underlined that SIRI C is a very old sample and the uncertainties of measurements are relatively high, over 1000 years. Subsamples A and C of the sample VIRI E show consistency within 1 σ with each other and 2 σ with the ultrafiltered one (B). Each of the VIRI E subsamples gave a result younger than the consensus value but the result of the ultrafiltered sample (B) was the closest one.

It should be underlined that the time interval between previous and current graphitizations and measurements (i.e., the storage time of gelatine samples) varied between 2 and 6 years (see Tables 1 and 2). It may be possible that during this time there was some absorption of modern atmospheric CO2 as well as microbial activity.

In the case of six samples prepared from raw bone material (Samples 1–6), we observed that ultrafiltration resulted in a significantly lower gelatine yield compared to simple filtration with Ezee Filters (see Table 2). The gelatine yield for samples treated without ultrafiltration reached between 3.6 and 15.5%, while after ultrafiltration the values decreased and ranged between 0.5 and 9.5%. This fact is not surprising taking into account that ultrafiltration is aimed at discarding the filtrate which contains degraded collagen chains and other unwanted molecules.

As one of results of the ultrafiltration process, we observed changes in the C and N concentration and C/N atomic ratios. The concentration of C and N increased in most of the samples examined (see Figure 1). In addition, in most cases, the C/Nat improved to the acceptable value for well-preserved collagen, changing even from 4.6 to 3.2 in the case of the sample GdA-5883 (see Figure 2). However, for two of three small samples GdA-5958 and GdA-6099, such improvement was not observed (see Table 2).

The improvement of carbon and nitrogen parameters was expected as a result of ultrafiltration—apparently there were impurities, some of which added weight, and which were removed during preparation. In general, they caused an increase in the carbon and nitrogen contents. The C/Nat indicate these impurities contained excessive carbon atoms, which is characteristic of humic acid contamination. After applied pretreatment the C/N atomic ratio decreased, which suggests that this contamination was reduced.

Humic acids suspected to contaminate bone samples can also be expected to have ages different than a bone sample. In fact, after removal of carbon-bearing contamination with ultrafiltration, older 14C ages were obtained for some of the examined cases (i.e. GdA-4225, GdA-4593, GdA-4617). However, the altering effect is more visible for the set of wapiti and red deer bones than for archaeological samples from Poland. It can be explained by the different conditions in which bones have been deposited over time. Some of the samples of red deer and wapiti bones were found in caves, where conditions are unfavourable in terms of bone preservation (Collins et al. Reference Collins, Nielsen-Marsh, Hiller, Smith, Roberts, Prigodich, Wess, Csapò, Millars and Turner-Walker2002).

There is a visible shift in δ15N values after ultrafiltration in almost every case. Values of δ15N increased on average by 1.45‰. As for δ13C, no systematic shift in the values was observed (Figure 3).

This situation can be explained by the presence of another component, whose δ15N was low. As the samples are very different and come from different places, with different environmental conditions, one of the possible explanations for such a systematic shift in δ15N is the effect of preparation with ultrafilters that screen out short and degraded chains of protein molecules.

Long-term storage of the gelatine samples results in younger 14C ages of the samples. It is visible in the majority of cases samples labeled “C” gave 14C ages older than samples “A” (see e.g., sample GdA-5339 /SIRI C/ in the Table 1 or sample GdA-6097 in the Table 2). It is probably caused by some microbial activity, since samples “C” were measured immediately after preparation and samples “A” after some time, from a few months to a few years. However, this effect can be reduced by applying ultrafiltration (samples labeled “B”). It is visible in both batches of samples, with some exceptions. The results of batch 1 (Tables 1 and 2) show that remeasurements of long-term stored samples give slightly older 14C ages for ultrafiltered ones (set “B”) than for those without ultrafiltration (set “A”), however taking into account the uncertainties, the results mostly stay equal. The remeasurements of the red deer and wapiti bones, with the use of ultrafiltration (set “B”) gave, in most cases, older ages than set “C” (measured without ultrafiltration). One of the most spectacular cases was sample GdA-5883, for which after ultrafiltration the %C and %N increased, reaching acceptable levels. The C/Nat value was lowered from 4.6 to 3.2 and the 14C age increased from 8515 ± 35 to 13210 ± 80 BP. On the other hand, for poorly preserved bone GdA-4227 the ultrafiltration step did not improve %C and %N to an acceptable level, the C/Nat was also too low. The age obtained after ultrafiltration was significantly younger than before. Altogether it makes this result unreliable.

CONCLUSIONS

We proved the potential utility of ultrafiltration in contamination removal. Although the 14C ages of bone samples must be evaluated carefully with respect to potential contamination and treatment efficiency, especially for long-stored samples, this method is still highly popular and useful. Ultrafiltration helps to improve the quality of gelatine as evidenced by %C and %N as well as C/Nat, by removal of contaminants with the membrane. The cases of poorly preserved bones are still challenging for 14C dating (see the example of GdA-4227), which raises doubts whether the reliable results are possible to achieve. Similarly, the results obtained from small samples of bone, <200 mg, are problematic to interpret in our case. These results might not be reliable due to probable non-linear effects in elemental analysis. The VarioMicroCube Elemental Analyzer, used in our laboratory, gives a greater scatter of results for small samples. This effect cannot be avoided, despite careful calibration.

One of the effects of long-term storage of gelatine samples at room temperature might be some disturbance in the 14C age of the samples obtained. The use of ultrafiltration allows to improve the results for samples stored in such unfavorable conditions, as demonstrated for both very old (>30000 BP) and young samples. Even though, to avoid potential contamination, the time of storage between preparation and 14C measurements should be minimized.

ACKNOWLEDGMENTS

The research was made possible through the subsidies of the Silesian University of Technology (14/020/BKM20/0004 and 14/020/BKM21/0013). The research was also supported by the SUT grant for the maintenance and development of the research potential (14/020/BK_21/0007), the Ministry of Science and Education through a special grant for the 14C and Mass Spectrometry Laboratory (209848/E-367/SPUB/2018/1), and by the National Science Centre (decision no. DEC-2013/11/B/NZ8/00888).

SUPPLEMENTARY MATERIAL

To view supplementary material for this article, please visit https://doi.org/10.1017/RDC.2024.32

References

REFERENCES

Ambrose, SH. 1990. Preparation and characterisation of bone and tooth collagen for isotopic analysis. Journal of Archaeological Science 17:431451.CrossRefGoogle Scholar
Beaumont, W, Beverly, R, Southon, J, Taylor, RE. 2010. Bone preparation at the KCCAMS laboratory. Nuclear Instruments and Methods in Physics Research B 268(7–8):906909.CrossRefGoogle Scholar
Brock, F, Dee, M, Hughes, A, Snoeck, C, Staff, R, Ramsey, CB. 2018. Testing the effectiveness of protocols for removal of common conservation treatments for radiocarbon dating. Radiocarbon 60(1):3550.CrossRefGoogle Scholar
Brock, F, Geoghegan, V, Thomas, B, Jurkschat, K, Higham, TFG. 2013a. Analysis of bone “collagen” extraction products for radiocarbon dating. Radiocarbon 55(2):445463.CrossRefGoogle Scholar
Brock, F, Higham, T, Ramsey, CB. 2013b. Comments on the use of Ezee-filters™ and ultrafilters at ORAU. Radiocarbon 55(1):211212.CrossRefGoogle Scholar
Brock, F, Ramsey, CB, Higham, T. 2007. Quality assurance of ultrafiltered bone dating. Radiocarbon 49(2):187192.CrossRefGoogle Scholar
Bronk Ramsey, C, Higham, T, Bowles, A, Hedges, R. 2004. Improvements to the pretreatment of bone at Oxford. Radiocarbon 46(1):155–63.CrossRefGoogle Scholar
Brown, TA, Nelson, DE, Vogel, JS, Southon, JR. 1988. Improved Collagen extraction by modified Longin method. Radiocarbon 30:171177.CrossRefGoogle Scholar
Cherkinsky, A. 2009. Can we get a good radiocarbon age from “bad bone”? Determining the reliability of radiocarbon age from bioapatite. Radiocarbon 51(2):647655.CrossRefGoogle Scholar
Collins, MJ, Nielsen-Marsh, CM, Hiller, J, Smith, C I, Roberts, JP, Prigodich, RV, Wess, TJ, Csapò, J, Millars, AR, Turner-Walker, G. 2002. The survival of organic matter in bone: a review. Archaeometry 44(3):383394.CrossRefGoogle Scholar
Crann, CA, Grant, T. 2019. Radiocarbon age of consolidants and adhesives used in archaeological conservation. Journal of Archaeological Science: Reports 24:10591063.Google Scholar
DeNiro, MJ. 1985. Postmortem preservation and alteration of in vivo bone collagen isotope ratios in relation to palaeodietary reconstruction. Nature 317(6040):806809.CrossRefGoogle Scholar
DeNiro, MJ, Weiner, S. 1988. Use of collagenase to purify collagen from prehistoric bones for stable isotopic analysis. Geochimica et Cosmochimica Acta 52(10):24252431.CrossRefGoogle Scholar
Deviese, T, Comeskey, D, McCullagh, J, Bronk Ramsey, C, Higham, T. 2018. New protocol for compound-specific radiocarbon analysis of archaeological bones. Rapid Communications in Mass Spectrometry 32:373379.CrossRefGoogle ScholarPubMed
Doan, K, Niedziałkowska, M, Stefaniak, K, Sykut, M, Jędrzejewska, B, Ratajczak-Skrzatek, U, Piotrowska, N, Ridush, B, Zachos, FE, Popović, D, et al. 2022. Phylogenetics and phylogeography of red deer mtDNA lineages during the last 50 000 years in Eurasia. Zoological Journal of the Linnean Society 194(2):431456.CrossRefGoogle Scholar
Doan, K, Zachos, FE, Wilkens, B, Vigne, J-D, Piotrowska, N, Stanković, A, Jędrzejewska, B, Stefaniak, K, Niedziałkowska, M. 2017. Phylogeography of the Tyrrhenian red deer (Cervus elaphus corsicanus) resolved using ancient DNA of radiocarbon-dated subfossils. Scientific Reports 7(1):19.Google Scholar
Fedi, ME, Caforio, L, Liccioli, L, Mandò, PA, Salvini, A, Taccetti, F. 2014. A simple and effective removal procedure of synthetic resins to obtain accurate radiocarbon dates of restored artworks. Radiocarbon 56(3):969979.CrossRefGoogle Scholar
Fewlass, H, Tuna, T, Fagault, Y, Hublin, JJ, Kromer, B, Bard, E, Talamo, S. 2019. Pretreatment and gaseous radiocarbon dating of 40–100 mg archaeological bone. Scientific Reports 9(1):111.CrossRefGoogle ScholarPubMed
Geist, V. 1998. Deer of the world: their evolution, behaviour, and ecology. Stackpole Books.Google Scholar
Gillespie, R, Hedges, RE, Wand, JO. 1984. Radiocarbon dating of bone by accelerator mass spectrometry. Journal of Archaeological Science 11(2):165170.CrossRefGoogle Scholar
Gonfiantini, R, Rozanski, K, Stichler, W. 1990. Intercalibration of environmental isotope measurements: the program of the International Atomic Energy Agency. Radiocarbon 32(3):369374.CrossRefGoogle Scholar
Goslar, T, Czernik, J, Goslar, E. 2004. Low-energy 14C AMS in Poznań radiocarbon laboratory, Poland. Nuclear Instruments and Methods in Physics Research B 223:511.CrossRefGoogle Scholar
Hajdas, I, Michczynski, A, Bonani, G, Wacker, L, Furrer, H. 2009. Dating bones near the limit of the radiocarbon dating method: study case mammoth from Niederweningen, ZH Switzerland. Radiocarbon 51:675680.CrossRefGoogle Scholar
Herrando-Pérez, S. 2021. Bone need not remain an elephant in the room for radiocarbon dating. Royal Society Open Science 8(1):201351.CrossRefGoogle Scholar
Hüls, CM, Grootes, PM, Nadeau, MJ. 2009. Ultrafiltration: boon or bane? Radiocarbon 51(2):613625.CrossRefGoogle Scholar
Hüls, MC, Grootes, PM, Nadeau, MJ. 2007. How clean is ultrafiltration cleaning of bone collagen? Radiocarbon 49(2):193200.CrossRefGoogle Scholar
Longin, R. 1971. New methods of collagen extraction for radiocarbon dating. Nature 230(5291):241242.CrossRefGoogle ScholarPubMed
Marom, A, McCullagh, JS, Higham, TF, Hedges, RE. 2013. Hydroxyproline dating: Experiments on the 14C analysis of contaminated and low-collagen bones. Radiocarbon 55(2):698708.CrossRefGoogle Scholar
Marom, A, McCullagh, JS, Higham, TF, Sinitsyn, AA, Hedges, RE. 2012. Single amino acid radiocarbon dating of Upper Paleolithic modern humans. Proceedings of the National Academy of Sciences 109:68786881.CrossRefGoogle ScholarPubMed
Marom-Rotem, A. 2012. Development and application of an analytical method for radiocarbon dating bones using the amino acid hydroxyproline [doctoral dissertation, University of Oxford].Google Scholar
McCullagh, JS, Marom, A, Hedges, RE. 2010. Radiocarbon dating of individual amino acids from archaeological bone collagen. Radiocarbon 52:620634.CrossRefGoogle Scholar
Minami, M, Nakamura, T. 2000. AMS radiocarbon age for fossil bone by XAD-2 chromatography method. Nuclear Instruments and Methods in Physics Research B 172(1-4):462468.CrossRefGoogle Scholar
Misarti, N, Finney, B, Maschner, H, Wooller, MJ. 2009. Changes in northeast Pacific marine ecosystems over the last 4500 years: evidence from stable isotope analysis of bone collagen from archeological middens. The Holocene 19(8):11391151.CrossRefGoogle Scholar
Nelson, DE. 1991. A new method for carbon isotopic analysis of protein. Science 251:552554.CrossRefGoogle ScholarPubMed
Nemec, M, Wacker, L, Gäggeler, HW. 2010. Optimisation of the graphitisation process at AGE-1. Radiocarbon 52(3):13801393.CrossRefGoogle Scholar
Niedziałkowska, M, Doan, K, Górny, M, Sykut, M, Stefaniak, K, Piotrowska, N, Jędrzejewska, B, Ridush, B, Pawełczyk, S, Mackiewicz, P, et al. 2021. Winter temperature and forest cover have shaped red deer distribution in Europe and the Ural Mountains since the Late Pleistocene. Journal of Biogeography 48(1):147159.CrossRefGoogle Scholar
Piotrowska, N, Goslar, T. 2002. Preparation of bone samples in the Gliwice radiocarbon laboratory for AMS radiocarbon dating. Isotopes in Environmental and Health Studies 38:267275.CrossRefGoogle ScholarPubMed
Piotrowska, N, Tomczyk, J, Pawełczyk, S, Stanaszek, ŁM. 2019. Radiocarbon AMS dating of Mesolithic human remains from Poland. Radiocarbon 61(4):9911007.CrossRefGoogle Scholar
Scirè Calabrisotto, C, Fedi, ME, Caforio, L, Bombardieri, L, Mando, PA. 2013. Collagen quality indicators for radiocarbon dating of bones: new data on Bronze Age Cyprus. Radiocarbon 55(2–3):472480.CrossRefGoogle Scholar
Scott, E, Naysmith, P, Cook, G. 2017. Should archaeologists care about 14C intercomparisons? Why? A summary report on SIRI. Radiocarbon 59(5):15891596.Google Scholar
Scott, EM, Cook, GT, Naysmith, P. 2010. A report on phase 2 of the Fifth International Radiocarbon Intercomparison (VIRI). Radiocarbon 52:846858.CrossRefGoogle Scholar
Sealy, J, Johnson, M, Richards, M, Nehlich, O. 2014. Comparison of two methods of extracting bone collagen for stable carbon and nitrogen isotope analysis: Comparing whole bone demineralization with gelatinization and ultrafiltration. Journal of Archaeological Science 47(1):6469.CrossRefGoogle Scholar
Shammas, N, Walker, B, De La Torre, HM, Bertrand, C, Southon, J. 2019. Effect of ultrafilter pretreatment, acid strength and decalcification duration on archaeological bone collagen yield. Nuclear Instruments and Methods in Physics Research B 456:283286.CrossRefGoogle Scholar
Stafford, TW Jr, Brendel, K, Duhamel, RC. 1988. Radiocarbon, 13C and 15N analysis of fossil bone: removal of humates with XAD-2 resin. Geochimica et Cosmochimica Acta. 52(9):22572267.CrossRefGoogle Scholar
Stafford, TW Jr, Duhamel, RC, Haynes, CV Jr, Brendel, K. 1982. Isolation of proline and hydroxyproline from fossil bone. Life Sciences 31:931938.CrossRefGoogle ScholarPubMed
Svyatko, SV, Mertz, IV, Reimer, PJ. 2015. Freshwater reservoir effect on redating of Eurasian Steppe cultures: first results for Eneolithic and Early Bronze Age northeast Kazakhstan. Radiocarbon 57(4):625644.CrossRefGoogle Scholar
Talamo, S, Fewlass, H, Maria, R, Jaouen, K. 2021. “Here we go again”: the inspection of collagen extraction protocols for 14C dating and palaeodietary analysis. STAR: Science & Technology of Archaeological Research 7(1):6277.CrossRefGoogle ScholarPubMed
Talamo, S, Richards, M. 2011. A comparison of bone pretreatment methods for AMS dating of samples> 30,000 BP. Radiocarbon 53(3):443449.CrossRefGoogle Scholar
Taylor, RE. 1992. Radiocarbon dating of bone: To collagen and beyond. In: Taylor, RE, Long, A, Kra, RS, editors. Radiocarbon after four decades: an interdisciplinary perspective. New York: Springer-Verlag. p. 375402.CrossRefGoogle Scholar
Van Klinken, GJ, Bowles, AD, Hedges, REM. 1994. Radiocarbon dating of peptides isolated from contaminated fossil bone collagen by collagenase digestion and reversed-phase chromatography. Geochimica et Cosmochimica Acta 58(11):25432551.CrossRefGoogle Scholar
Van Klinken, GJ, Hedges, RE. 1992. Experiments on 14C dating of contaminated bone using peptides resulting from enzymatic cleavage of collagen. Radiocarbon 34(3):292295.CrossRefGoogle Scholar
Van Klinken, GJ, Hedges, RE. 1995. Experiments on collagen-humic interactions: speed of humic uptake, and effects of diverse chemical treatments. Journal of Archaeological Science 22(2):263270.CrossRefGoogle Scholar
Wacker, L, Nemec, M, Bourquin, J. 2010. A revolutionary graphitisation system: fully automated, compact and simple. Nuclear Instruments and Methods in Physics Research B 268(7–8):931934.CrossRefGoogle Scholar
Yuan, S, Wu, X, Gao, S, Wang, J, Cai, L, Liu, K, Li, K, Ma, H. 2000. Comparison of different bone pretreatment methods for AMS 14C dating. Nuclear Instruments and Methods in Physics Research B 172(1–4):424427.CrossRefGoogle Scholar
Figure 0

Table 1 Results of weight% of C and N, C/N atomic ratios (C/Nat) determinations, and 14C measurements for international 14C intercomparison samples (from batch 1); Treatment methods: A = without ultrafiltration, B = with ultrafiltration; Previously obtained 14C age values after Piotrowska et al. (2019) are listed in italics as set C (without the ultrafiltration step); Consensus values after Scott et al. (2010) for VIRI and Scott et al. (2017) for SIRI.

Figure 1

Table 2 Results of the% weight of C and N, C/Nat and 14C measurements of anonymized archaeological samples—human bones but GdA-6163 (Sample 7)—deer antler (from batch 1). Samples were divided into two sets of subsamples: A, treated without ultrafiltration, and B, treated with ultrafiltration. In addition, the values obtained previously were listed in italics (= Set C, without the ultrafiltration step). Small samples (below 200 mg of raw sample) were marked with an asterisk(*).

Figure 2

Table 3 Results of weight % of C and N, C/Nat, stable isotope ratios (δ13C, δ15N) and 14C measurements of red deer and wapiti bone samples from Europe and Western Asia (Doan et al. 2017; 2022; Niedziałkowska et al. 2021); Treatment methods: B = with ultrafiltration (this study), C = without ultrafiltration (previous studies; in italics). When only one sample was available for stable isotope ratio measurement, it was marked with an asterisk (*).

Figure 3

Figure 1 Changes in C and N concentration measured in all investigated samples; preparation without ultrafiltration step—open symbols; preparation with ultrafiltration step—filled symbols. The majority of anonymized Polish archaeological samples marked with an ellipse. Numbers indicate GdA laboratory codes.

Figure 4

Figure 2 C/Nat ratios and 14C ages. Preparation without ultrafiltration step: open symbols; preparation with ultrafiltration step: filled symbols. Numbers indicate GdA laboratory codes.

Figure 5

Figure 3 The stable isotope (C and N) composition of the investigated samples. Preparation without ultrafiltration step: open symbols; preparation with ultrafiltration step: filled symbols. Numbers indicate GdA laboratory codes.

Supplementary material: File

Pawełczyk et al. supplementary material

Pawełczyk et al. supplementary material
Download Pawełczyk et al. supplementary material(File)
File 20.6 KB