Hostname: page-component-cd9895bd7-jn8rn Total loading time: 0 Render date: 2024-12-18T20:23:40.891Z Has data issue: false hasContentIssue false

Biogenic carbonate samples: Preliminary tests on chemical protocols for radiocarbon analysis

Published online by Cambridge University Press:  05 December 2024

M I Oliveira
Affiliation:
Laboratório de Radiocarbono (LAC-UFF), Instituto de Física, Universidade Federal Fluminense (UFF), Av. Gal. Milton Tavares de Souza, s/n, Niterói, 24210-346, Rio de Janeiro, Brazil Programa de Pós-graduação em Geoquímica, Instituto de Química, Universidade Federal Fluminense (UFF), Outeiro São João Batista, s/n, Niterói, 24210-141, Brazil
C Carvalho*
Affiliation:
Laboratório de Radiocarbono (LAC-UFF), Instituto de Física, Universidade Federal Fluminense (UFF), Av. Gal. Milton Tavares de Souza, s/n, Niterói, 24210-346, Rio de Janeiro, Brazil Programa de Pós-graduação em Geoquímica, Instituto de Química, Universidade Federal Fluminense (UFF), Outeiro São João Batista, s/n, Niterói, 24210-141, Brazil
D Tremmel
Affiliation:
Laboratório de Radiocarbono (LAC-UFF), Instituto de Física, Universidade Federal Fluminense (UFF), Av. Gal. Milton Tavares de Souza, s/n, Niterói, 24210-346, Rio de Janeiro, Brazil Programa de Pós-graduação em Geoquímica, Instituto de Química, Universidade Federal Fluminense (UFF), Outeiro São João Batista, s/n, Niterói, 24210-141, Brazil
C Silveira
Affiliation:
Programa de Pós-graduação em Geoquímica, Instituto de Química, Universidade Federal Fluminense (UFF), Outeiro São João Batista, s/n, Niterói, 24210-141, Brazil
A A Brito
Affiliation:
Laboratório de Radiocarbono (LAC-UFF), Instituto de Física, Universidade Federal Fluminense (UFF), Av. Gal. Milton Tavares de Souza, s/n, Niterói, 24210-346, Rio de Janeiro, Brazil
F M Oliveira
Affiliation:
Laboratório de Radiocarbono (LAC-UFF), Instituto de Física, Universidade Federal Fluminense (UFF), Av. Gal. Milton Tavares de Souza, s/n, Niterói, 24210-346, Rio de Janeiro, Brazil Programa de Pós-graduação em Química, Instituto de Química, Universidade Federal Fluminense (UFF), Outeiro São João Batista, s/n, Niterói, 24210-141, Brazil
V N Moreira
Affiliation:
Laboratório de Radiocarbono (LAC-UFF), Instituto de Física, Universidade Federal Fluminense (UFF), Av. Gal. Milton Tavares de Souza, s/n, Niterói, 24210-346, Rio de Janeiro, Brazil Programa de Pós-graduação em Geoquímica, Instituto de Química, Universidade Federal Fluminense (UFF), Outeiro São João Batista, s/n, Niterói, 24210-141, Brazil
K Macario
Affiliation:
Laboratório de Radiocarbono (LAC-UFF), Instituto de Física, Universidade Federal Fluminense (UFF), Av. Gal. Milton Tavares de Souza, s/n, Niterói, 24210-346, Rio de Janeiro, Brazil Programa de Pós-graduação em Geoquímica, Instituto de Química, Universidade Federal Fluminense (UFF), Outeiro São João Batista, s/n, Niterói, 24210-141, Brazil
L Bastos
Affiliation:
Laboratório de Instrumentação Nuclear (LIN-COPPE). COPPE, Universidade Federal do Rio de Janeiro (UFRJ), Av. Horácio Macedo, 2030, bloco I, sala I-133, Rio de Janeiro, 21941-914, Rio de Janeiro, Brazil
M Moreira
Affiliation:
Programa de Pós-graduação em Geoquímica, Instituto de Química, Universidade Federal Fluminense (UFF), Outeiro São João Batista, s/n, Niterói, 24210-141, Brazil
R T Lopes
Affiliation:
Laboratório de Instrumentação Nuclear (LIN-COPPE). COPPE, Universidade Federal do Rio de Janeiro (UFRJ), Av. Horácio Macedo, 2030, bloco I, sala I-133, Rio de Janeiro, 21941-914, Rio de Janeiro, Brazil
*
Corresponding author: C Carvalho; Email: [email protected]
Rights & Permissions [Opens in a new window]

Abstract

Most of the carbonate samples have a basic well-defined pretreatment protocol for 14C-AMS dating, but particularities of specific organisms have to be treated with care. This is the case of stromatolite samples, in which carbonate is formed by biogenesis and also has a porous structure that could contain recent organic material as a contaminant. In this work, we analyzed the differences in the radiocarbon content by using organic matter removals before chemical treatment with HCl: sodium hypochlorite (NaOCl) a 0.7M solution with pH ∼11, and hydrogen peroxide (H2O2) an 8.8M solution with pH ∼5. These treatments were chosen because they are the most used in stromatolite samples for geochemical analysis. To compare the impact of the organic matter removal treatments in stromatolite samples we also processed them as regular carbonate samples for radiocarbon analysis, with no organic matter removal (control samples). X-ray diffraction and X-ray fluorescence have been used to obtain mineral and elemental characterization, respectively. H2O2 could not influence the results of Mg-calcite concentrate samples. The use of NaOCl appears to have been effective in preserving more material than H2O2 independent of the mineralogical composition of the stromatolite layers. The F14C results after HCl etching for Mg-calcite concentrated samples were similar to those without etching suggesting that the HCl etching does not impact the results in this case. The organic matter removal is more important than the etching procedure for stromatolite samples. NaOCl is more indicated to be used as chemical pretreatment for radiocarbon analysis purposes independent of the mineral matrix of samples.

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

Introduction

Stromatolites are biologically induced organomineral layered sedimentary formations and their metabolic activities induce conditions for precipitation (Dupraz et al. Reference Dupraz, Reid, Braissant, Decho, Norman and Visscher2009). Its structure can host diverse microbial communities (Baumgartner et al. Reference Baumgartner, Spear, Buckley, Pace, Reid, Dupraz and Visscher2009), including bacteria, algae, and archaea (Lepot et al. Reference Lepot, Benzerara and Brown2008). Some bacteria within stromatolites may perform processes that can affect radiocarbon analyses (Barker and Fritz Reference Barker and Fritz1981). For example, methane-producing bacteria can introduce carbon from different sources, while sulfate-reducing bacteria can influence carbonate precipitation (Andres et al. Reference Andres, Sumner, Reid and Swart2006; Visscher et al. Reference Visscher, Reid and Bebout2000). These microbial processes are intimately linked with mineral formation (Reid et al. Reference Reid, Dupraz, Visscher, Sumner, Krumbein, Paterson and Zavarzin2003) but are also influenced by interconnected local factors such as light, pH, and temperature (Bowlin et al. Reference Bowlin, Klaus, Foster, Andres, Custals and Reid2012). Therefore, understanding the microbial community and their metabolic activities provides insights into mineralogical properties and potential variations in 14C content.

The reservoir effect on stromatolites is particularly important in calcareous regions rich in ancient carbonates. This effect is well-documented such as in Carreira et al. (Reference Carreira, Marques, Graça and Aires-Barros2008), where regions high in limestone are largely devoid of radiocarbon due to prolonged isolation from the atmospheric carbon cycle, hence, adding carbon with minimal 14C into surrounding waters. Stromatolites may assimilate this carbon during their growth, causing an overestimated age during radiocarbon dating due to the inclusion of older carbon sources that lack radiocarbon (Brook et al. Reference Brook, Cherkinsky, Railsback, Marais and Hipondoka2013; Jull et al. Reference Jull, Burr and Hodgins2013). As seen, organisms residing in areas with low carbonate influence are not subject to this effect (Macario et al. Reference Macario, Alves and Carvalho2016).

The study of the formation of these structures is very important because it is believed that stromatolites have dominated 80% of all geological formation on planet Earth and are the evidence of the earliest life found, dating to around 3.5 billion years (Grotzinger and Knoll Reference Grotzinger and Knoll1999; Hofmann Reference Hofmann1969, Reference Hofmann1973; Riding and Awramik Reference Riding and Awramik2000; Vologdin Reference Vologdin1962; Walter Reference Walter1976). Several paleoclimate reconstruction studies have been performed using the geochemistry proxies stored within the carbonate matrix over the last years (Bahniuk Reference Bahniuk2013; Birgel et al. Reference Birgel, Meister, Lundberg, Horath, Bontognali, Bahniuk, De Rezende, Vasconcelos and Mckenzie2015; Carvalho et al. Reference Carvalho, Oliveira, Macario, Guimarães, Keim, Sabadini-Santos and Crapez2017; Iespa et al. Reference Iespa, Iespa and Borghi2012; Silva e Silva and Senra 2000; Vasconcelos and McKenzie Reference Vasconcelos and McKenzie1997; Vasconcelos et al. Reference Vasconcelos, Warthmann, McKenzie, Visscher, Bittermann and van Lith2006). Radiocarbon dating is an important tool for a better understanding of these records. However, stromatolite has no well-defined pretreatment for being dated by the 14C-AMS technique.

In general, the 14C-AMS dating protocol for carbonate samples is based on a simple removal of the outer layer with a sandblaster or a chemical etching pretreatment using hydrochloric acid (HCl), to remove possible external contamination. Then, hydrolysis is performed by using orthophosphoric acid (H3PO4), to convert the carbonate into carbon dioxide, by the following equation (Burman et al. Reference Burman, Gustafsson, Segl and Schmitz2005):

$${\rm {3CaCO_3} + \;{2H_3}{PO_4} \rightarrow {Ca_3}{(PO_4)_2} \;+ 3H_2O\,+\,{3CO_2}}$$

This acid does not convert organic matter to carbon dioxide, but some authors affirm that through this reaction it is possible to produce this gas and/or other molecular gaseous species (Bowen Reference Bowen1966, Reference Bowen1991; Epstein et al. Reference Epstein, Buchsbaum, Lowenstam and Urey1951, Reference Epstein, Buchsbaum, Lowenstam and Urey1953; Falster et al. Reference Falster, Delean and Tyler2018; Oehlerich et al. Reference Oehlerich, Baumer, Lücke and Mayr2013; Weber et al. Reference Weber, Deines, Weber and Baker1976). Under certain conditions, CO2 can be produced from organic matter in the acid environment if the organo-mineral interaction bonds are broken and the organic matter is more accessible to hydrolysis reactions which is not the case because H3PO4 is a weak acid that is widely used to obtain carbon dioxide from carbonate samples for radiocarbon purposes (Wacker et al. Reference Wacker, Fülöp, Hajdas, Molnár and Rethemeyer2013a, Reference Wacker, Lippold, Molnár and Schulz2013b). Concerning stromatolites structure and composition it is necessary to consider the question: what is the impact of performing or not performing a chemical pretreatment to remove organic matter from the samples? Chaduteau et al. (Reference Chaduteau, Ader, Lebeau, Landais and Busigny2021) realized a similar test in δ13C analysis of carbonate samples and proved that it makes a difference in the results. Key et al. (Reference Key, Smith, Phillips and Forrester2020) also assess the impact of the most used pretreatment methods of organic matter removal on δ13C and δ18O ratios for carbonate samples from taxonomic groups with complex mineralogies. The main purpose of this work is to attempt to answer this question by presenting F14C results for 14C-AMS analysis, testing the extraction of organic matter and acid etching on subfacies of a stromatolite specimen which can have different mineral compositions in its growth layers.

Materials and methods

The stromatolite specimen used in this work was collected in the Salgada Lagoon (21°54'10''S e 41°00'30''W) in January 2016. This lagoon is located in the coastal area of Rio de Janeiro State, Brazil, being part of the Paraíba do Sul deltaic river complex (Lamego Reference Lamego1955). This is one of the few hypersaline lagoons that still have stromatolite formation nowadays. Initially, X-ray microtomography (micro-CT) was performed on the raw material to identify five stromatolite subfacies the analysis was performed at the Nuclear Instrumentation Laboratory at Rio de Janeiro State University (LIN-COPPE/UFRJ). This is a non-destructive technique that allows the inspection of the internal structure of the sample and the Phoenix Vtomex|m GE micro-CT system was used. Reconstruction of the sample’s volume was possible through software datos x (Version 2.5.0) and CTAn software (Version 1.18.4.0) was used to perform the image segmentation and quantitative and qualitative analyses. The segmentation allowed the division of the image, which was in different gray levels, into just black and white. The micro-CT examination of the raw material allowed the detailed identification of five subfacies as shown in Figure 1. SB01 layer is the oldest one, which corresponds to the lagoon floor, and SB05 is the most recent, on the top of the stromatolite. The subfacies structure showed different degrees of porosity, as follows: SB01=18.98%, SB02=40.73%, SB03=35.74%, SB04=11.85%, and SB05= 8.99%.

Figure 1. Stromatolite Subfacies limits from the bottom (SB01) to the top (SB05) obtained through micro-CT analysis.

The elemental and mineral composition were analyzed by X-ray fluorescence (XRF) and X-ray diffraction (XRD), respectively, before starting treatments to verify whether any differences could impact the radiocarbon results. For 14C pretreatment, the three intermediary samples were submitted to different organic matter treatments: hydrogen peroxide (H2O2), sodium hypochlorite (NaOCl), and control, without organic matter removal. The latter is a common carbonate protocol for radiocarbon dating. Then, the samples were subdivided into two groups: with and without etching using hydrochloric acid (HCl) and proceeded to the acid hydrolysis and graphitization protocol for 14C-AMS dating at the Radiocarbon Laboratory of the Fluminense Federal University (LAC-UFF). From the results obtained in the present work, we expect to better understand the stromatolite sample handling to improve the protocols for radiocarbon dating of biogenic carbonate samples by eliminating possible organic contamination. The stromatolite analyzed in this work was 10 cm high and 30 cm in diameter (Figure 2A).

Figure 2. (A) stromatolite sample of Salgada Lagoon; (B) sample extraction.

Five subsamples were extracted using a 7 mm wall drill (Figure 2B) and crushed in an agate mortar to be homogenized at LAC-UFF to perform XRF and XRD composition analysis.

XRF analysis was used to determine the elemental composition of subsamples. The sample preparation and analysis were made at Laboratório de Oceanografia Operacional e Paleoceonografia at Fluminense Federal University (LOOP-UFF). A benchtop Epsilon 3X energy-dispersive X-ray fluorescence spectrophotometer (EDXRF) from PANalytical was used. It has an X-ray tube with a silver anode and a 50 µm beryllium window, a power of 9 W, a current of 1 mA, and a voltage of 50 kV, using air and helium as carrier gas. The detector is a high-resolution Silicon Drift detector, typical of 135 eV, and a thin window of 8 µm (Be).

XRD analysis was used to investigate the crystalline structure of the powdered samples. The sample preparation and analysis were undertaken at the Geochemistry Department at Fluminense Federal University. A BrukerD2 Phaser (Cu Kα radiation) model was operated in a Bragg–Brentano θ/θ configuration, with the diffraction patterns being collected in a flat geometry with a range between 3 and 100 degrees, steps of 0.02 degrees, and accumulation time of 3.0 s per step. Phase identification was done with EVA® software. The XRD data were refined following the Rietveld method using the DIFFRAC.SUITE TOPAS® software (McCusker et al. Reference Mccusker, Von Dreele, Cox, Louer and Scardi1999; Toby Reference Toby2006).

Afterwards, the three intermediary subfacies were selected for the organic matter removal pretreatment test for 14C dating. For the test, all samples were prepared and analyzed at LAC-UFF. Each subfacie was subdivided to perform the organic removal pretreatment test as follows: [T-A]: Samples without organic matter removal; [T-B]: organic matter removal using a solution of 0.7 M of sodium hypochlorite that has pH ∼11 (NaOCl, Isofar, Brazil, 4–6% purum p.a., CAS 7722-84-1, Lot. n. 245/2015), and [T-C]: a solution of 8.8 M of hydrogen peroxide that has pH ∼5 (H2O2, Isofar, Brazil, 30% purum p.a., CAS 7681-52-9). NaOCl and H2O2 were selected because they are powerful oxidation agents and are the most used in organic matter treatments for stromatolite samples in geochemical analysis. A treatment diagram is presented in Figure 3. The treatment T-A-1 can be referred to as a “control” treatment not in the classical sense in terms of expected values. It can be referred to as a control treatment compared to the others since the sample was untreated. The T-A-2 treatment consists of etching the sample without organic matter removal. For both organic matter removal treatments, 3 ml of NaOCl or H2O2 were added to approximately 100 mg of dry sample and left to react for 1h at 60 °C in the dry bath. Each process was repeated 5 times and at the end, samples were washed 3 times with ultrapure water and dried at 90ºC to determine the mass loss percentage.

Figure 3. Steps of the treatment tests performed.

After the organic matter removal samples were split in two aliquots to proceed with the carbonate etching, as shown in Figure 3. An aliquot of the samples went straight to CO2 conversion without etching addition (T-B-1 and T-C-1), and another aliquot, with approximately 40 mg, was chemically treated with 1.7 mL of 0.1M HCl and stayed overnight at 90ºC in the dry bath (T-A-2, T-B-2, and T-C-2) for a 25% etching, following the LAC-UFF etching standard protocol. At the end of the process, samples were dried at 90ºC and proceeded to CO2 conversion.

The CO2 was obtained by acid hydrolysis using the orthophosphoric acid (H3PO4, 85%) protocol. After CO2 purification using temperature traps in the vacuum line, graphitization took place at 550°C (7h) with TiH2, Zn and iron used as catalysts in an inner tube (Macario et al. Reference Macario, Alves, Moreira, Oliveira and Chanca2017a, Reference Macario, Oliveira, Moreira, Alves and Carvalho2017b). Graphitized samples were measured in a NEC 250 kV single stage (SSAMS) system (Macario et al. Reference Macario, Gomes, Anjos, Carvalho, Linares, Alves, Oliveira, Castro, Chanca, Silveira, Pessenda, Moraes, Campos and Cherinsky2013, Reference Macario, Oliveira, Carvalho, Santos, Xu, Chanca, Alves, Jou, Oliveira, Brandão, Moreira, Muniz, Linares, Gomes, Anjos, Castro, Anjos, Marques and Rodrigues2015). Typical currents were 50 μA12C–1 (measured at the low energy Faraday cup). Graphite standard and calcite blanks yielded average 14C/13C ratios of 6×10–13 and 7×10–13, respectively. The average machine background was around 50 kyr for the unprocessed graphite, while the average precision ranged from 0.3 to 0.5%. Data analyses were carried out on LACAMS software (Castro et al. Reference Castro, Macario and Gomes2015).

Results and discussion

XRF analysis

From XRF analysis the elemental composition of subsamples was determined showing a low concentration of Fe and Mn in the stromatolite chemical composition when compared to Mg concentration was observed in XRF analysis results (Table 1). For more detailed information and to obtain the mineralogical composition we conducted XRD analysis.

Table 1. XRF analysis results from stromatolite samples

XRD analysis

As shown by Moreira et al. (Reference Moreira, Macario, Guimarães, Dias, Araujo, Jesus and Douka2020), fossils of marine organisms are likely to undergo dissolution and recrystallization. During this process, carbonate exchanges may occur, which leads to the entry of exogenous carbon, thus altering the radiocarbon concentration. The same could occur to stromatolite material during the formation of layers. Therefore, we carried out an XRD analysis to verify the crystalline composition of our samples. The results have shown that the subfacies presented are mainly calcite composition but with three types according to the d104 typical peak. There is a mixture of calcite and Mg-calcite with different proportions. Table 2 presents d104 (A) results and MgCO3 (%) concentrations, where SB05 and SB04 presented low MgCO3 concentrations, 12% and 16%, respectively. SB02 presented 24% of MgCO3 and SB03 and SB01 presented high and close concentrations, 42% and 41%, respectively. It is important to mention that the balance between precipitation and dissolution of MgCO3 is a process that is influenced by the metabolic activity of the microbial community of the stromatolite during its growth, as well as by the physicochemical properties of the surrounding water, such as pH, temperature, saturation state, and the concentration of Mg2+ and carbonate ions in the environment. The concentration of MgCO3 in the growth layers can give information on the sedimentation environment in the past, but that is not the main goal of this work and will not be discussed in the present paper. Figure 4 shows the detailed d104 peaks from XRD analysis. Samples SB04 and SB05 have more calcite (the two blue peaks on the left of Figure 4). Otherwise, samples SB01, SB02, and SB03 are more Mg-calcite enriched. Samples SB01 and SB03 stand out as very Mg-enriched calcite samples (the black and green peaks on the right of Figure 4). The d104 values below 2.99 are typical Mg-enriched calcite samples. This appears to be related to a change in the crystallographic unit cell from the Mg-calcite to be like the one from the mineral kutnohorite (Ca (Mn, Fe, Mn)(CO3)2) a rare mineral from the dolomite group and not present here (Graff Reference Graff1961). Some authors also had a large shift in diffractogram peak (from 2.99A to 2.94A related to Mg-calcite) as a result of a very high-magnesian calcite (VHMC) or a disordered dolomite, as pointed out by Zhang et al. (Reference Zhang, Xu, Konishi and Roden2010).

Table 2. Stromatolite subfacies d104 (A) results and MgCO3 (%) concentrations, where d104 (A) higher than 2.99 are related to low MgCO3 (%) concentrations (SB04 and SB05) and d104 (A) lower than 2.99 have high MgCO3 (%) concentrations (SB03, SB02 and SB01)

Figure 4. Comparison of diffractograms presenting the d104 calcite characteristic peak in Angstrom (A).

As can be observed in Table 3, subfacie SB04 had the lowest concentration of Mg-calcite (36%) and the highest concentration of calcite (57%). The highest concentration of Mg-calcite and lowest concentration of calcite were found in SB01 85% and 5% respectively. SB02 and SB03 have similar mineralogical values, although SB03 has a higher Mg-calcite concentration. Carbonate precipitation and recrystallization in stromatolites can manifest in various mineral forms, such as calcite, aragonite, or dolomite, and are influenced by physicochemical, biological, and environmental conditions. The subfacie SB02, which presented 12% of aragonite in its composition, was probably subjected to different conditions during the layer growth. If occurs, the recrystallization in the same layer while studying stromatolite well-defined sequential growth layers seems to not impact the radiocarbon dating discussion.

Table 3. Mineralogical pattern of Salgada lagoon stromatolite layers. ND: not detected; trace: value below 1%

Chemical treatments

After the previous characterization, three inner layers were selected to move forward to the organic matter removal test: SB02 and SB03 which presented similar composition, and SB04 which presented lower Mg-calcite and higher calcite concentrations. The percentage of mass loss after the process is indicated in Table 4. Concerning weight loss during the organic matter removal, NaOCl bleaching presented similar results for all subfacies and a smaller weight loss than H2O2. This may be caused by the oxidizing strength of H2O2 compared to NaOCl, which changes the inorganic matrix of the samples beyond the consumption of the organic matter. The same results were found by Chaduteau et al. (Reference Chaduteau, Ader, Lebeau, Landais and Busigny2021), in which H2O2 altered the preservation of the carbonate, showing inorganic matrix dissolution, which did not occur while using NaOCl. The weight loss during the H2O2 use suggests dissolution in all samples. As pointed out by Key et al. (Reference Key, Smith, Phillips and Forrester2020), many researchers reported that H2O2 is strongly corrosive to CaCO3 and that H2O2 at ∼ 30% has a pH of ∼5 which can promote the dissolution of carbonates. The highest weight loss (57%) was observed after the H2O2 treatment on SB04 which presented the highest concentration of calcite (57%). Some other factors can increase the susceptibility of dissolution as the Mg content and the duration of the treatment but our study did not evidentiate this.

Table 4. Mass loss during organic matter chemical pretreatment, where [T-B] is the organic matter removal treatment using sodium hypochlorite (NaOCl) and [T-C] using hydrogen peroxide (H2O2)

Following Reimer et al. (Reference Reimer, Brown and Reimer2004), radiocarbon results are reported by F14C values determined after the treatments and are shown in Table 5. It was not possible to determine the results of SB02 after T-C-1 due to probable loss during CO2 purification in the vacuum line. As a result, we would expect a reduction in F14C values for all samples after the organic matter removal step since it should remove any recent carbon that could have been incorporated into the original sample matrix. Comparing the results observed for group 1 (organic matter removal without etching) it is possible to observe this behavior when comparing T-A-1 (Control) and T-B-1 (NaOCl) results for all layers. When comparing T-C-1 (H2O2) with the control T-A-1 group the layer SB03, which has more Mg-calcite, presented a reduction in F14C values. On the other hand, layer SB04 presented a higher F14C value.

Table 5. Radiocarbon results (F14C) and its uncertainty after pretreatment tests. [T-A]: Control samples without organic matter removal; [T-B]: organic matter removal using sodium hypochlorite (NaOCl) and [T-C]: hydrogen peroxide (H2O2). The numbers 1 and 2 represent results for samples without and with HCl etching, respectively

The F14C results after HCl etching (group 2) for Mg-calcite concentrated samples (SB02 and SB03) were similar to those without etching. The CaCO3 concentrated (SB04) presented F14C results higher than the control (T-A-2) in both combined tests: NaOCl + etching (T-B-2) and H2O2 + etching (T-C-2).

In the graph (Figure 5) the results represent the F14C levels for each treatment [T-A, T-B, and T-C], grouped by subfacie. The black triangles represent results after treatments with no etching and the gray circles represent results after etching. The F14C results for SB02 and SB03 are quite similar with low discrepancy. SB04 had a higher discrepancy, which seems to be an effect of its mineral composition containing higher calcite and low Mg-calcite. The results with etching (gray) for SB02 and SB04 are higher after all treatments, the same behavior is not observed for SB03 which has the highest Mg-calcite concentration.

Figure 5. F14C results for each treatment [T-A, T-B, and T-C], grouped by subfacie. The black triangles represent results with no etching treatment and the gray circles represent results after etching.

Many works discussed the effects of previous treatments on stable isotope ratios for biogenic and inorganic carbonate samples to remove organic matter showing its dependence on the mineralogical matrices, duration of the treatment, and the importance of avoiding isotopic fractionation or isotopic exchange during the treatments (Key et al. Reference Key, Smith, Phillips and Forrester2020; Zhang et al. Reference Zhang, Lin and Yamada2020). Since the selection of chemical pretreatment to be used depends on the mineral composition, wrong choices can also affect radiocarbon results, especially for stromatolite samples that could have a complex mineralogy. Carvalho et al. (Reference Carvalho, Oliveira, Macario, Guimarães, Keim, Sabadini-Santos and Crapez2017) dated Mg-calcite-rich stromatolite samples by applying H2O2, which was the most used treatment for this kind of sample, and afterward followed the radiocarbon carbonate etching protocol. From the present study, it is possible to say that it was a good choice to apply the H2O2 protocol to Mg-calcite-rich samples as it doesn’t seem to impact results in the case of high Mg-calcite composition. Additionally, comparing the present work with Chaduteau et al. (Reference Chaduteau, Ader, Lebeau, Landais and Busigny2021) who also observed shifted δ13C values for calcite-rich samples after being treated with H2O2, showing it is less efficient for removing organic matter than NaOCl, it’s possible to say that we have observed the same for SB04 subfacie in the present work. Furthermore, H2O2 pretreatment can make the remaining organic matter more reactive to H3PO4, which is used to convert carbonate into CO2 for radiocarbon analysis and could influence F14C results. On the other hand, NaOCl pretreatment presented good results for any stromatolite mineral matrix. The F14C results after etching seem to be dependent on the mineral matrix but were not significantly different for the Mg-calcite-rich layer.

Conclusions

Before radiocarbon dating stromatolite samples it is important to identify the mineralogical composition before organic matter removal. Even though H2O2 could not influence the results of Mg-calcite concentrate samples, it can promote the dissolution of the carbonate matrix showing an expressive weight loss for the calcite concentrated layer. The use of NaOCl appears to have been effective in preserving more material than H2O2 independent of the mineralogical composition of the stromatolite layers. The F14C results after HCl etching for Mg-calcite concentrated samples were similar to those without etching suggesting that the HCl etching does not impact the results in this case. Since stromatolites are biogenic carbonate samples with some porosity, the organic matter removal has shown to be more important than the etching procedure. The CaCO3 concentrated sample presented F14C results higher than the control in H2O2 pretreatment and both combined tests: NaOCl + etching and H2O2 + etching. In this case, NaOCl is more indicated to be used as chemical pretreatment for radiocarbon analysis purposes independent of the mineral matrix of samples.

Acknowledgments

The authors would like to thank Brazilian financial agencies CNPq (Conselho Nacional de Desenvolvimento Científico e Tecnológico, Grant 315514/2020-5 and Grant 426338/2018-9 to Carvalho, C.; Grant 317397/2021-4 to Macario, K., Grant 309412/2019-6 to Oliveira, F.) and INCT-FNA 464898/2014-5), FAPERJ (Grant E-26/202.714/2018 to Carvalho, C.; Grant E-26/202615/2019 and E-26/200.540/2023 to Macario, K.; Grant E-26/201.320/2022 to Oliveira, F.), Project CLIMATE-PRINT-UFF (Grant 88887.310301/2018-00) for their support. This study was financed in part by the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior - Brasil (CAPES) - Finance Code 001. We also thank CNPq for the PhD’s scholarship of Oliveira, M.I. We have a special acknowledgement to Professor Cleverson Guizan who collected and donated the stromatolite specimen making this research possible and the anonymous reviewers who identified some aspects that could be improved to clarify the work done.

References

Andres, M, Sumner, D, Reid, R and Swart, P (2006) Isotopic fingerprints of microbial respiration in aragonite from Bahamian stromatolites. Geology 34(11), 973976. https://doi.org/10.1130/G22859A.1.CrossRefGoogle Scholar
Bahniuk, A (2013) Coupling Organic and Inorganic Methods to Study Growth and Diagenesis of Modern Microbial Carbonates, Rio de Janeiro State, Brazil: Implications for Interpreting Ancient Microbialite Facies Development. PhD dissertation. ETH Zurich, no. 20984.Google Scholar
Baumgartner, LK, Spear, JR, Buckley, DH, Pace, NR, Reid, RP, Dupraz, C and Visscher, PT (2009) Microbial diversity in modern marine stromatolites, Highborne Cay, Bahamas. Environmental Microbiology 11, 27102719. https://doi.org/10.1111/j.1462-2920.2009.01998.x.CrossRefGoogle ScholarPubMed
Barker, J and Fritz, P (1981) Carbon isotope fractionation during microbial methane oxidation. Nature 293, 289291. https://doi.org/10.1038/293289a0.CrossRefGoogle Scholar
Birgel, D, Meister, P, Lundberg, R, Horath, TD, Bontognali, TRR, Bahniuk, AM, De Rezende, CE, Vasconcelos, C and Mckenzie, JA (2015) Methanogenesis produces strong 13C enrichment in stromatolites of Lagoa Salgada, Brazil: a modern analogue for paleo-neoproterozoic stromatolites? Geobiology 13, 245266. https://doi.org/10.1111/gbi.12130 CrossRefGoogle Scholar
Bowen, R (1966) Paleotemperature Analysis. Amsterdam: Elsevier.Google Scholar
Bowen, R (1991) Isotopes and Climates. Barking: Elsevier Science.Google Scholar
Bowlin, E, Klaus, J, Foster, J, Andres, M, Custals, L and Reid, R (2012) Environmental controls on microbial community cycling in modern marine stromatolites. Sedimentary Geology 263–264, 4555. https://doi.org/10.1016/j.sedgeo.2011.08.002.CrossRefGoogle Scholar
Brook, GA, Cherkinsky, A, Railsback, LB, Marais, E and Hipondoka, MHT (2013) 14C dating of organic residue and carbonate from stromatolites in Etosha Pan, Namibia: 14C Reservoir effect, correction of published ages, and evidence of >8m deep lake during the Late Pleistocene. Radiocarbon 55(3), 11561163. https://doi.org/10.1017/S0033822200048062.CrossRefGoogle Scholar
Burman, J, Gustafsson, O, Segl, M and Schmitz, B (2005) A simplified method of preparing phosphoric acid for stable isotope analyses of carbonates. Rapid Commun. Mass Spectrom. 19(21), 30863088. https://doi.org/10.1002/rcm.2159 CrossRefGoogle Scholar
Carreira, P, Marques, J, Graça, R and Aires-Barros, L (2008) Radiocarbon application in dating “complex” hot and cold CO2-rich mineral water systems: A review of case studies ascribed to the northern Portugal. Applied Geochemistry 23(10), 28172828. https://doi.org/10.1016/j.apgeochem.2008.04.004.CrossRefGoogle Scholar
Carvalho, C, Oliveira, MI, Macario, K, Guimarães, RB, Keim, CN, Sabadini-Santos, E and Crapez, MAC (2017) Stromatolite growth in Lagoa Vermelha, southeastern coast of Brazil: evidence of environmental changes. Radiocarbon 60(2), 383393. https://doi.org/10.1017/RDC.2017.126 CrossRefGoogle Scholar
Castro, MD, Macario, KD and Gomes, PRS (2015) New software for AMS data analysis developed at IF-UFF Brazil. Nuclear Instruments and Methods in Physics Research B 361, 526530. https://doi.org/10.1016/j.nimb.2015.01.070 CrossRefGoogle Scholar
Chaduteau, C, Ader, M, Lebeau, O, Landais, G and Busigny, V (2021) Organic matter removal for continuous flow isotope ratio mass spectrometry analysis of carbon and oxygen isotope compositions of calcite or dolomite in organic-rich samples. Limnology and Oceanography: Methods 19, 523539. https://doi.org/10.1002/lom3.10442 Google Scholar
Dupraz, C, Reid, PR, Braissant, O, Decho, AW, Norman, RS and Visscher, PT (2009) Processes of carbonate precipitation in modern microbial mats. Earth-Science Reviews 96, 141162. https://doi.org/10.1016/j.earscirev.2008.10.005 CrossRefGoogle Scholar
Epstein, S, Buchsbaum, R, Lowenstam, HA and Urey, HC (1951) Carbonate-water isotopic temperature scale. Geological Society American Bulletin 62, 417. https://doi.org/10.1130/0016-7606(1951)62[417:CITS]2.0.CO;2 CrossRefGoogle Scholar
Epstein, S, Buchsbaum, R, Lowenstam, HA and Urey, HC (1953) Revised carbonate-water isotopic temperature scale. Geological Society American Bulletin 64, 13161326. https://doi.org/10.1130/0016-7606(1953)64[1315:RCITS]2.0.CO;2 CrossRefGoogle Scholar
Falster, G, Delean, S and Tyler, J (2018) Hydrogen peroxide treatment of natural lake sediment prior to carbon and oxygen stable isotope analysis of calcium carbonate. Geochemistry, Geophysics, Geosystem 19, 35833595. https://doi.org/10.1029/2018GC007575 CrossRefGoogle Scholar
Graff, DL (1961) Crystallographic tables for the rhombohedral carbonates. American Mineralogist 46, 12831316.Google Scholar
Grotzinger, JP and Knoll, AH (1999) Stromatolites in Precambrian carbonates: Evolutionary mileposts or environmental dipsticks. Annual Reviews of Earth and Planetary Sciences 27, 313358. https://doi.org/10.1146/annurev.earth.27.1.313 CrossRefGoogle ScholarPubMed
Hofmann, HJ (1969) Attributes of stromatolites. Geological Survey of Canada Paper 58, 69–39.Google Scholar
Hofmann, HJ (1973) Stromatolites: characteristics and utility. Earth-Science Reviews 9, 339373.CrossRefGoogle Scholar
Iespa, AAC, Iespa, CMD and Borghi, L (2012) Evolução paleoambiental da Lagoa Salgada utilizando microbialitos, com ênfase em microfácies carbonáticas. São Paulo, UNESP, Geociências 31(3), 71380.Google Scholar
Jull, A, Burr, G and Hodgins, G (2013) Radiocarbon dating, reservoir effects, and calibration. Quaternary International 299, 6471. https://doi.org/10.1016/j.quaint.2012.10.028.CrossRefGoogle Scholar
Key, MM Jr, Smith, AM, Phillips, NJ and Forrester, JS (2020) Effect of removal of organic material on stable isotope ratios in skeletal carbonate from taxonomic groups with complex mineralogies. Rapid Commun. Mass Spectrometry 34, e8901. https://doi.org/10.1002/rcm.8901 CrossRefGoogle ScholarPubMed
Lamego, AR (1955) Geologia das quadrículas de Campos, São tomé, Lagoa Feia e Xexé. Rio de Janeiro, DNPM/ DGM, Boletim 154, 160.Google Scholar
Lepot, K, Benzerara, K, Brown, G et al. (2008) Microbially influenced formation of 2,724-million-year-old stromatolites. Nature Geosci. 1, 118121. https://doi.org/10.1038/ngeo107.CrossRefGoogle Scholar
Macario, KD, Gomes, PRS, Anjos, RM, Carvalho, C, Linares, R, Alves, EQ, Oliveira, FM, Castro, M, Chanca, IS, Silveira, MFM, Pessenda, LCR, Moraes, LMB, Campos, TB and Cherinsky, A (2013) The Brazilian AMS Radiocarbon Laboratory (LAC-UFF) and the intercomparison of results with CENA and UGAMS. Radiocarbon 55(2), 325330. https://doi.org/10.1017/S003382220005743X CrossRefGoogle Scholar
Macario, KD, Oliveira, FM, Carvalho, C, Santos, GM, Xu, X, Chanca, IS, Alves, EQ, Jou, R, Oliveira, MI, Brandão, B, Moreira, VN, Muniz, M, Linares, R, Gomes, PRS, Anjos, RM, Castro, MD, Anjos, L, Marques, AN and Rodrigues, LF (2015) Advances in the graphitization protocol at the radiocarbon laboratory of the Universidade Federal Fluminense (LAC-UFF) in Brazil. Nuclear Instruments and Methods in Physics Research B 361, 402405. https://doi.org/10.1016/j.nimb.2015.03.081 CrossRefGoogle Scholar
Macario, K, Alves, E, Carvalho, C et al. (2016) The use of the terrestrial snails of the genera Megalobulimus and Thaumastus as representatives of the atmospheric carbon reservoir. Sci. Rep. 6, 27395. https://doi.org/10.1038/srep27395.CrossRefGoogle ScholarPubMed
Macario, KD, Alves, EQ, Moreira, VN, Oliveira, FM, Chanca, IS et al. (2017a) Fractionation in the graphitization reaction for 14C-AMS analysis: The role of Zn × the role of TiH2 . International Journal of Mass Spectrometry 423, 3945. https://doi.org/10.1016/j.ijms.2017.10.005 CrossRefGoogle Scholar
Macario, KD, Oliveira, FM, Moreira, VN, Alves, EQ, Carvalho, C et al. (2017b) Optimization of the amount of zinc in the graphitization reaction for radiocarbon AMS measurements at LAC-UFF. Radiocarbon 59(3), 885891. https://doi.org/10.1017/RDC.2016.42 CrossRefGoogle Scholar
Mccusker, LB, Von Dreele, RB, Cox, DE, Louer, D and Scardi, P (1999) Rietveld refinement guidelines. Journal of Applied Crystallography 32, 3650. https://doi.org/10.1107/S0021889898009856 CrossRefGoogle Scholar
Moreira, VN, Macario, KD, Guimarães, RB, Dias, FF, Araujo, JC, Jesus, P and Douka, K (2020) Aragonite fraction dating of vermetids in the context of paleo sea-level curves reconstruction. Radiocarbon 62(2), 335348. https://doi.org/10.1017/RDC.2020.7 CrossRefGoogle Scholar
Oehlerich, M, Baumer, M, Lücke, A and Mayr, C (2013) Effects of organic matter on carbonate stable isotope ratios (δ13C, δ18O values)—implications for analyses of bulk sediments. Rapid Communications in Mass Spectrometry 27, 707712. https://doi.org/10.1002/rcm.6492 CrossRefGoogle ScholarPubMed
Reid, P, Dupraz, CD, Visscher, PT and Sumner, DY (2003) Microbial processes forming marine stromatolites. In Krumbein, WE, Paterson, DM and Zavarzin, GA (eds), Fossil and Recent Biofilms. Dordrecht: Springer. https://doi.org/10.1007/978-94-017-0193-8_6.Google Scholar
Reimer, PJ, Brown, TA and Reimer, RW (2004) Discussion; reporting and calibration of post-bomb 14C data. Radiocarbon 46, 12991304. https://doi.org/10.1017/S0033822200033154 Google Scholar
Riding, R and Awramik, S (2000) Microbial Sediments. Berlin: Springer, 331.CrossRefGoogle Scholar
Silva e Silva LH and Senra MCE (2000) Estudo comparativo de esteiras microbianas recentes em duas lagoas hipersalinas. Revista Universidade Guarulhos, Geociências 5, 225227.Google Scholar
Toby, BH (2006) R factors in Rietveld analysis: How good is good enough? Powder Diffraction 21, 6770. https://doi.org/10.1154/1.2179804 CrossRefGoogle Scholar
Vasconcelos, C and McKenzie, JA (1997) Microbial mediation of modern dolomite precipitation and diagenesis under anoxic conditions (Lagoa Vermelha, Rio de Janeiro, Brazil). Journal of Sedimentary Research 67, 378390. https://doi.org/10.1306/D4268577-2B26-11D7-8648000102C1865D Google Scholar
Vasconcelos, C, Warthmann, R, McKenzie, J, Visscher, PT, Bittermann, AG and van Lith, Y (2006) Lithifying microbial mats in Lagoa Vermelha, Brazil: Modern Precambrian relics? Sedimentary Geology 185, 175183. https://doi.org/10.1016/j.sedgeo.2005.12.022 CrossRefGoogle Scholar
Visscher, PT, Reid, R and Bebout, B (2000) Microscale observations of sulfate reduction: Correlation of microbial activity with lithified micritic laminae in modern marine stromatolites. Geology 28(10), 919922. https://doi.org/10.1130/0091-7613(2000)28<919:MOOSRC>2.0.CO;2.2.0.CO;2>CrossRefGoogle Scholar
Vologdin, AG (1962) The oldest algae of the USSR. Academy of Sciences of the USSR: 657.Google Scholar
Wacker, L, Fülöp, R-H, Hajdas, I, Molnár, M and Rethemeyer, J (2013a) A novel approach to process carbonate samples for radiocarbon measurements with helium carrier gas. Nuclear Instruments and Methods in Physics Research B 294, 214217. https://doi.org/10.1016/j.nimb.2012.08.030.CrossRefGoogle Scholar
Wacker, L, Lippold, J, Molnár, M and Schulz, H (2013b) Towards radiocarbon dating of single foraminifera with a gas ion source. Nuclear Instruments and Methods in Physics Research B 294, 307310. https://doi.org/10.1016/j.nimb.2012.08.038.CrossRefGoogle Scholar
Walter, M (1976) Stromatolites: Developments in Sedimentology. Elsevier, 790.Google Scholar
Weber, J, Deines, N, Weber, PH and Baker, PA (1976) Depth related changes in 13C/12C ratio of skeletal carbonate deposited by Caribbean reef-frame building coral Montastrea annularis: Further implications of a model for stable isotope fractionation by scleractinian corals. Geochimica et Cosmochimica Acta 40, 3139. https://doi.org/10.1016/0016-7037(76)90191-5 CrossRefGoogle Scholar
Zhang, F, Xu, H, Konishi, H and Roden, EE (2010) A Relationship between d104 value and composition in the calcite-disordered dolomite solid-solution series. American Mineralogist 95, 16501656. https://doi.org/10.2138/am.2010.3414 CrossRefGoogle Scholar
Zhang, N, Lin, M, Yamada, K, et al. (2020) The effect of H2O2 treatment on stable isotope analysis (δ13C, δ18O, and Δ47) of various carbonate minerals. Chem. Geol. 532, 119352. https://doi.org/10.1016/j.chemgeo.2019.119352.CrossRefGoogle Scholar
Figure 0

Figure 1. Stromatolite Subfacies limits from the bottom (SB01) to the top (SB05) obtained through micro-CT analysis.

Figure 1

Figure 2. (A) stromatolite sample of Salgada Lagoon; (B) sample extraction.

Figure 2

Figure 3. Steps of the treatment tests performed.

Figure 3

Table 1. XRF analysis results from stromatolite samples

Figure 4

Table 2. Stromatolite subfacies d104 (A) results and MgCO3 (%) concentrations, where d104 (A) higher than 2.99 are related to low MgCO3 (%) concentrations (SB04 and SB05) and d104 (A) lower than 2.99 have high MgCO3 (%) concentrations (SB03, SB02 and SB01)

Figure 5

Figure 4. Comparison of diffractograms presenting the d104 calcite characteristic peak in Angstrom (A).

Figure 6

Table 3. Mineralogical pattern of Salgada lagoon stromatolite layers. ND: not detected; trace: value below 1%

Figure 7

Table 4. Mass loss during organic matter chemical pretreatment, where [T-B] is the organic matter removal treatment using sodium hypochlorite (NaOCl) and [T-C] using hydrogen peroxide (H2O2)

Figure 8

Table 5. Radiocarbon results (F14C) and its uncertainty after pretreatment tests. [T-A]: Control samples without organic matter removal; [T-B]: organic matter removal using sodium hypochlorite (NaOCl) and [T-C]: hydrogen peroxide (H2O2). The numbers 1 and 2 represent results for samples without and with HCl etching, respectively

Figure 9

Figure 5. F14C results for each treatment [T-A, T-B, and T-C], grouped by subfacie. The black triangles represent results with no etching treatment and the gray circles represent results after etching.