Next Article in Journal
Effect of Different NaOH Solution Concentrations on Mechanical Properties and Microstructure of Alkali-Activated Blast Furnace Ferronickel Slag
Previous Article in Journal
Synthesis and Structural Analysis of Ternary Ca–Al–Fe Layered Double Hydroxides with Different Iron Contents
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Protocol Comparison for Organic Residue Analyses from Waterproofing Materials and Shards of Roman Archaeological Amphorae

1
IMBE UMR7263/IRD237, Avignon University/CNRS/IRD/AMU, Restoration Engineering of Natural and Cultural Heritage, 84000 Avignon, France
2
Dipartimento di Biologia Ambientale, Sapienza Università di Roma, Piazzale Aldo Moro, 5, 00185 Roma, Italy
3
French National Center for Scientific Research, Centre Camille Jullian UMR 7299, Maison Méditerranéenne des Sciences de l’Homme, CEDEX 2, 13097 Aix-en-Provence, France
4
Soprintendenza APAB National per le Province di Frosinone e Latina, Via Pompeo Magno 2, 00192 Rome, Italy
*
Authors to whom correspondence should be addressed.
Submission received: 20 September 2021 / Revised: 11 October 2021 / Accepted: 15 October 2021 / Published: 26 October 2021
(This article belongs to the Section Inorganic Crystalline Materials)

Abstract

:
With the aim of addressing the impact of extractive protocols in molecular characterization of ceramic content, sixteen archaeological shards and waterproofing coatings of Roman amphorae were studied to compare the extractive capacities of protocols prevalently mentioned in wine amphorae analysis. A microwave-assisted protocol is developed in order to esterify grape-derivative markers from archaeological pitch and shard. Gas chromatography-mass spectrometry is used to highlight the great capacities of a two-step protocol that combines organic extraction with BF3-etherate complex butylation applied on archaeological shards. Instead, simultaneous alkaline fusion and direct-resin acid-catalyzed butylation are favored for the characterization of waterproofing material. The identification of tartaric acid, together with succinic, fumaric pyruvic and syringic acids provide valuable insights on the archaeological grape-derivative content, possibly wine. Diterpenic markers highlighted Pinus pitch and wood tar, originally used to waterproof the amphorae. Since markers are reliable tools in organic residue analyses, protocols exhibiting high extractive capacities are favored to avoid false conclusions drawn through the absence of markers.

Graphical Abstract

1. Introduction

In the early 1990s, Evershed introduced the archaeological biomarker concept to trace back the original use of potteries. Focusing on either the carbon structures or the pattern distributions, molecules act as chemical fingerprints [1]. They can reveal information regarding the initial composition, natural ageing, anthropic degradation as well as contamination. This way, organic residue analyses became an established field of research; fundamental to address the archaeological content by rending microremains identifiable [2]. For instance, the earliest consumption of wine could be dated back to the Neolithic in the South Caucasus through molecular markers [3].
Although pioneering studies laid the foundation using Feigl spots, infrared spectroscopy or HPLC to identify tartaric acid [4,5,6], the emergence of chromatographic tools offered great substitutes to traditional methods to prevent from false-positives [7,8,9]. As rightly criticized in literature, neither molecular identification through sole retention or migration time, nor the precision of UV detection are reliable [10,11]. Enhancing the specificity and sensitivity required for wine markers identification as well as providing structural information, gas chromatography coupled with mass spectrometry (GC-MS) became frequently employed [12,13,14,15]. Although recent techniques on the cutting edge of technology showed great analytical advancements, equipment costs remain highly limiting. Even though tandem MS mode or selected ion monitoring mode in liquid or gas chromatography coupled with MS were published to grandly improve the limit of detection to specifically target the presence of archaeological wines [3,16,17,18,19], their usage remains rare since it implies ultra-advanced equipment. GC-MS consequently appears as a good compromise to turn routine analyses into efficient molecular markers searches, decisive for their identification. This way, fermented grape-beverage has been evidenced in archaeological ceramic jars and amphorae through the presence of tartaric acid [6,20]. Although the molecule is not exclusively produced by grapes, its concentration remains higher than in other exotic plant sources such as tamarind, yellow plums [18] or pomegranates [21]. Moreover, it better survives upon archaeological time compared to other grape acids [9]. For this reason, it is usually considered as a grape biomarker when supported by archaeological contexts and/or archaeobotanical grape evidence [3,11,12,22]. Although synthesized by numerous plants and fungi, pyruvic, fumaric and malic acids are additionally produced during the fermentation process [23]. Syringic acid highlights red wine since it arises from malvidin, the pigment responsible for dark grape coloration [16,18].
Beyond the presence of markers, two concepts need however to be separated: the extraction and the analytical detection. While analytical developments focused on the former, the latter remains of a first interest. Indeed, robust protocols must be employed to increase the chance of extracting tartaric acid and other grape beverage markers from the ceramic matrix. Since the biomarker concept relies on the presence/absence of characteristic features, it becomes necessary to avoid false-negative induced by extraction strength flaw [24]. Among the protocols most mentioned to promote the breaking of the intramolecular hydrogen bonding between tartrate salts and the silicate-rich ceramic, either acid or alkaline conditions are favored [9,12,25]. Alkaline treatment is followed by acidification to make tartaric acid more soluble in ethyl acetate before extraction [13,26]. Differently, the acido-catalyzed butylation of the unsolved material after polar extraction greatly evidenced grape derivatives in Neolithic, Etruscan or even Early Medieval Islamic jars [12,27,28]. The esterification of organic markers aimed at enhancing their solubility into the extractive solvent.
The primary aim of this paper is to describe an adapted microwave-assisted protocol for the acid-catalyzed extraction in order to ensure easy-going and accelerated approach for grape-derivative detection, applicable on both organic and inorganic artefacts. Using GC-MS, we additionally provide a comparative study of the extractive capacities offered by alkaline, acidic and polar extractions with the specific aim of proposing a robust and cost-effective methodology directly applicable on archaeological artefacts (waterproofing material and shard). In this study, three protocols were conducted depending on the nature of the samples. The first one consisted of a basic extraction applied on both pitch and shard. The second one encompassed a two-step protocol with consecutive lipid extraction, including microwave-assisted optimization for the butylation. This protocol was tested on pitch and shard. The last protocol corresponded to a variant of the second one, with the application of the two-step extraction handled separately (i.e., not consecutively but in two different ways) and both steps directly conducted on the waterproofing matter. This study was carried on sixteen Roman amphorae coming from two different maritime archaeological sites. A total of ten shards and eight resinous coatings were investigated in order to promote a great representativeness of the comparison and prevent from any bias that would arise from local or material specificities. A glossary of all of the identified compounds was presented at the end of the article.

2. Materials

2.1. Archaeological Samples

This comparative study integrated 16 archaeological amphorae coming from two different contexts. A total of 11 amphorae were excavated from the shipwreck of Planier 3 (France) and 5 came from the ancient anchorage of San Felice Circeo (Italy). The amphorae were selected to increase the variety between objects in order to make the protocol comparison the more likely to be applied on any further artefacts. Thus, the objects present a strong variability in terms of ceramic pastes, provenance and marine archaeological context, visual presence of coating and conservative conditions (Table 1). Although they all belong to the Roman period, they came from maritime context that can be interpreted differently. The amphorae from Planier 3 were all excavated from the same shipwreck, sunk in 49 BC near the Planier’s island (43°11′54″ N; 5°13′48″ E) close to the Marseille coasts (France). Coming from the region of Brindes, the cargo might have stopped first in the Sinus Tarentinus region to load Lamboglia 2 amphorae and then at Pozzuoli before heading towards the Narbonnaise. Excavation campaigns conducted from 1968 to 1975 revealed the important distribution of Dressel 1B, Lamboglia 2, ovoid and Brindisium amphorae in the cargo [29]. Along with the tremendous quantity of amphorae, Planier 3 exhibits the more important evidence of archaeological pouzzolane stoppers. Different mineral pigments were also reported in the cargo, such as realgar, white lead and the precious Egyptian blue of which the production was settled in Pozzuoli [30]. Such discoveries outlined the luxurious assumptions of traded products transported on the ship.
Five amphorae were excavated in the ancient anchorage of San Felice Circeo (41°13′49.0″ N; 13°06′30.1″ E) located 90 km SE of Roma. They were uncovered due to the winter storm of 2018, along with an important scattering of archaeological records located few hundred meters from the coast at a depth of 5 to 7 m under the sea level. The Soprintendenza (the local Office of the Italian Ministry of Culture) supervises ongoing excavation campaigns. The time scale established by the archaeological finds ranges from the Republican period through the Late Roman period to the post-medieval period. The diversity of the excavated ceramics, in terms of morphology and the time period they belong to provided valuable insights of ancient anchorage (Delpino and Melandri, unpublished paper) although the hypothesis of a shipwreck cannot be excluded.
The Late Greco-Italian/Dressel 1 typologies refer to 150 to 10 BC. They originate from south-central Italy, from Campania to Etruria, as attested by important kiln finds along the coastal area [31]. These amphorae were widely used to trade wine in the Mediterranean Basin, from Central Europe to Spain via Gaul. Lamboglia 2 amphora originate from the Adriatic coast [32,33]. They have largely contributed to the trade of wine in the western Mediterranean [34]. Chios amphora were surely meant for wine trading since the island was prized for the great quality of wine they produced [35]. Maña C2 amphora arose from the Punic tradition amphorae, originating from North African coasts. Although oil and fish preserves are frequently mentioned in such typology, the amphora was included in the study [36]. Dressel 5, also referenced as Rhodian amphorae, usually contained wine from the Aegean coast [37]. The present study focuses on 8 samples from waterproofing coatings and 10 samples of ceramic shards (see Table 1). Sample Nos. 749 and 1014 were analyzed with both the coating and the ceramic clay.

2.2. Solvents and Reagents

All of the organic solvents were of analytical grade. Methanol (MeOH), dichloromethane (DCM), diethyl ether (DEE), ethyl acetate and KOH were purchased by Merck (Darmstadt, Germany). Hexane and N,O-Bis(trimethylsilyl)trifluoroacetamide/trimethylchlorosilane (BSTFA/TMCS) and commercial standard molecules such as maleic, succinic, fumaric, malic, pyruvic, tartaric, syringic and dehydroabietic (DHA) acids were supplied from Sigma-Aldrich (Darmstadt, Germany). Anhydrous butanol and cyclohexane were purchased by Acros Organics (Illkirch France) and BF3 diethyl etherate from Alfa Aesar (Kandel, Germany). The fresh colophony resin standard was purchased by Kremer Pigmente GmbH & Co. KG (Aichstetten, Germany).

3. Methods

3.1. Optimization of the Acid-Catalyzed Esterification: Microwave-Assisted-Butylation

The protocol aiming at butylating wine acid markers was adapted from Garnier and Valamoti [12]. It was firstly developed for tartaric acid before being extended to other standard molecules of maleic, succinic, fumaric, pyruvic, malic, syringic acids. A total of 5 mg of commercial standard were treated with a mixture of boron trifluoride, butan-1-ol and cyclohexane (1:2:4 v/v/v) in a sealed vial placed in CEM Discover® LabMate microwave synthesizer (MW) (CEM Corp., Orsay, France). The instrument was used in single-mode (50 Hz; 300 W maximum output power). An infrared sensor positioned below the circular vessel continuously measure the routine temperature. The self-adaptive circular waveguide technique allows the circular cavity to automatically be adjusted in order to optimize the energy provided for the reaction [38]. Continuous pressure measurements permit on-the-fly changes for power control to maintain a maximal temperature, set at 80 °C with the CEM SynergyTM software (software version 0.9, Kamp-Lintfort, Germany). Different reaction times were tested (single dynamic cycle of 5 min, 10 min and three successive dynamic cycles of 5 min each) to evaluate the heating run necessary for the butylation with high stirring speed. The butylation advancement was followed by thin-layer chromatography with cyclohexane:ethyl acetate (1:1 v/v). After the butylation was completed, the solution was neutralized with a saturated solution of sodium carbonate. The esterified compounds were extracted two times with DEE. The combined organic fractions were washed twice with Milli-Q water and dried with anhydrous sodium sulphate before filtration on a PTFE cartridge (0.45 µm).

3.2. Analytical Procedures for Inorganic Shards

For the comparative analyses of the archaeological shards, two protocols were applied: a basic extraction adapted from Pecci et al. [9] and a two-step lipid extraction including the MW-assisted optimization for the butylation. To reduce the risk of external contamination, a thin layer was initially removed from the inner surface of the shard before sampling the drilled ceramic over 1–2 mm in depth.
On the one hand, 100 mg of crushed shard were extracted three times with KOH (1 M; 2 mL) using an ultrasound probe (VCX 130 Vibra-Cell Sonics, Sonics and materials, Newtown, CT, USA) for 3 min. After centrifugation, the successive extracts were combined and acidified up to pH 2. The organic phase was extracted 3 times with ethyl acetate (3 mL), filtered on a PTFE cartridge (0.45 µm) and evaporated to dryness.
On the other hand, 100 mg of crushed shard were extracted 3 times with DCM:MeOH (1:1 v/v) with the ultrasound probe for 3 min. After centrifugation, the organic supernatant of was filtered with PTFE (0.45 µm) and evaporated to dryness. This first step of the lipid extraction was recorded as 1LE. The remaining powder after the organic extraction was treated for the MW-assisted butylation with a mixture of BF3 etherate complex, butan-1-ol and cyclohexane (1:2:4 v/v/v) for 3 times 5 min. After neutralization with a saturated sodium carbonate solution, the organic fractions were extracted with DEE and washed twice with H2O before drying over sodium sulphate. The corresponding extract was labelled 2LE-MW. After filtration on a PTFE cartridge (0.45 µm), both extracts (1LE and 2LE-MW) were individually evaporated under a gentle N2 stream.

3.3. Analytical Procedures for Organic Coatings

For the comparative analyses of the archaeological coatings, three protocols were tested. The first one was the basic extraction already reported (see Section 3.2). The second was the adapted two-step lipid extraction with MW optimization for the butylation as previously described (see Section 3.2). The only difference was the starting amount of 50 mg of organic coating versus 100 mg of shard. The third protocol consisted of the same two-step lipid extraction, with each step independently conducted on the coating: the organic extraction with DCM:MeOH (1LE) and the extraction after butylation (R-2LE-MW) were hence handled separately. A total of 10 mg of coating material were necessary for the direct-pitch esterification. Butylation was performed as previously described for the MW parameter and heating runs.

3.4. GC-MS Analyses

GC-MS analyses were carried out on a Thermo Scientific™ Focus system equipped with an Al 3000 autosampler and an ITQ™ 700 Series Ion Trap Mass Spectrometer (ThermoFisher Scientific, Illkirch-Graffenstaden, France). A ThermoGOLD™ TG-5MS fused silica capillary column (5% diphenyl; 95% dimethyl polysiloxane) of 30 m length × 0.25 mm i.d. × 0.25 μm thickness ensured the separation of the mixture carried with helium at a constant flow rate of 1 mL min−1. A total of 1 µL solution was injected in splitless mode at 250 °C. Transfer line, ion trap and manifold temperatures were respectively 300 °C, 200 °C and 50 °C. Mass spectra were recorded in electron impact mode with an electron ionization energy of 70 eV, with ionization time of 25,000 µs. Scan are recorded in the range of 40–650 m/z. The oven temperature stayed isothermal for 2 min at 50 °C, increased at 8 °C/min to 140 °C held for 2 min before heating to 160 °C at 2.5 °C/min and finally 330 °C at 15 °C/min and held for 3 min.
After evaporation, all of the extracts were derivatized with BSTFA (200 µL, 70 °C, 30 min) before injection in splitless mode in GC-MS in 200 µL of hexane:DCM (1:1 v/v).
Data treatment was performed on XcaliburTM software (software version 4.3, Thermo Scientific, Waltham, MA, USA). Peak identification was achieved by comparison of retention time and mass spectra with molecular commercial standards and from the NIST MS Search 2.0 database (last access in May 2021).

3.5. Radar Plot Construction

The building of the radar plots (Figure 1 and Figure 2) relies on the presence/absence of 22 targeted molecules, that account for the characterization of the grape derivatives content and the resinous coating. Among them, 7 organic acids referred to grape composition (tartaric and syringic acids), fermentation (maleic, succinic, pyruvic, fumaric, malic acids) and 15 diterpenic derivatives indicative of the nature of the coating (dehydroabietic acid DHA, dehydroabietic methyl ester DHAM and retene) and its ageing (3-hydroxy-, 7-hyrdoxy-, 15-hydroxy-; 7,15-dihydroxy-, 7-oxo- and 15-hydroxy-7-oxo-DHA and DHAM-derivatives). We notably emphasized on highly significant markers, indispensable for grape derivative and coating identification (i.e., tartaric acid, retene and DHAM compounds).

4. Results and Discussion

4.1. Optimization of the Acid-Catalyzed Butylation

The esterification advancement, followed by thin-layer chromatography, ensured the effective butylation of standard molecules after three dynamic cycles of 5 min each. Butylated compounds were additionally controlled with infrared spectroscopy (FT-IR) and NMR (1H, 13C). GC-MS analyses allowing dibutyl tartrate (DBT) to be characterized with fragment ions at m/z 276, 305 and 391. The same butylation protocol was then applied on commercial standards acids considered as dark grape (i.e., syringic acid) and fermentation markers (i.e., maleic, succinic, pyruvic, fumaric and malic acids) to obtain their retention time and fragmentation patterns.

4.2. Extracting Capacities Comparison on Archaeological Shards

Archaeological artefacts being different from each other, the studied shards cannot be generalized to a set of amphorae. Protocol comparisons and analyses interpretations must respect individual molecular specificities, turning conclusions on the extraction capacities to be singular and object dependent. For this reason, the use of a radar plot was favored to independently outline the number of molecules extracted by each protocol and for each of the ten archaeological shards (Figure 1). The protocol comparison applied on shards encompassed: (i) an alkaline fusion with KOH extraction, (ii) a DCM-MeOH organic extraction and (iii) its coupling with BF3-catalyzed MW-butylation applied on the dried remaining powder after the organic extraction (2LE-MW extract).
For all the 10 shards, only the BF3-catalyzed butylation allowed DBT, highlighting the presence of tartaric acid (Figure 1). Partially dissolving the ceramic clay, BF3, optimized the release of the organic compounds strongly bonded, or even polymerized [25,39]. Increasing the apolar character of the esterified acids, butylation favored their rapid extraction in cyclohexane, hence favoring the butylation of remaining acids by shifting the equilibrium [12]. From there, the extraction from the co-solvent is enhanced with DEE that has a low dielectric constant solvent.
Even though hydroxyl anions arising from the alkaline fusion are supposed to interact with the ceramic matrix to enhance the release of bonded acids [9,13,26], no tartaric acid could be identified with this protocol (Figure 1). By shifting the solubility equilibrium of tartrate salts, KOH should favor the bonding cleavage with the ceramic and leave the marker soluble in the aqueous phase [9,16]. In fact, tartaric acid resists archaeological time thanks to the formation of salts that strongly interact with the clay matrix [6]. Conversion into free tartaric acid would ensure the recovery in the aqueous phase before extraction with ethyl acetate. Extraction issues could originate from the insufficient alkaline robustness or from the poor solubility of tartaric acid in ethyl acetate [11,16]. The hypothesis of tartrate salt that would be formed once the alkaline fusion released tartaric acid from the ceramic cannot be ruled out [11]. It is worth noting the limit of detection involved in every protocols. Starting from pure standard, quantitative analysis comparing the amount of tartaric acid recovered after extraction reported to identify 77% of the acid with butylation while it did not reach 0.1% with alkaline fusion [11]. Additionally, Garnier and Valamoti reported the detection of tartaric acid up to 10 ng/g shard with the acido-catalyzed protocol [12]. In conclusion, neither KOH fusion, nor the organic extraction with DCM-MeOH were suitable for the characterization of grape derivatives.
Aside from the considerable extraction of DBT, esterification also accounted for the extraction of grape acids (Table 2). Maleic acid was only characterized with butylation (m/z 99; 117). Malic acid, although hardly characterized with KOH fusion, was always identified as dibutyl malate (m/z 101; 145; 161; 303). The most important increase of molecules extracted was observed for the amphora No. 6904, where three fermentation acids (over the five considered) could be identified with butylation, hitherto absent with alkaline fusion and DCM:MeOH extraction. Surprisingly, pyruvic acid was directly extracted with traditional solvents in nine shards. Neither KOH nor butylation reached such great extent (Table 2) and the molecule that originates from malolactic fermentation [23] was only identified in two shards with alkaline fusion and never recovered as dibutylcetal. To the contrary, fumaric acid which is considered as marker of alcoholic fermentation [39], was preferentially extracted with alkaline fusion and never identified with butylation (Table 2). Since maleic, succinic, pyruvic, fumaric and malic acids can originate from the fermentation of grapes, they are considered as fermentation markers of wine. However, to compensate for their lack of exclusivity towards grape fermentation only, the greater the number of fermentation markers extracted, the more reliable the fermentation assumption. For this reason, the combination of extractive protocols would allow the number of molecules extracted to be increased. Since only the butylation proved to surely trace tartaric acid, the most fruitful coupling would include it to widen its extractive capacities.
Although syringic acid is naturally present in many plants and in wood lignin [13,18], malvidin origin has been clearly evidenced [16]. After the reaction of crushed grapes with yeasts present at the fruit surface, the alcohol produced by the spontaneous fermentation starts hydrolyzing malvidin contained in the dark grape skins. To surely state on red grapes, the benefits of a two-step protocol with the first part targeting the free syringic acid are not to be demonstrated anymore [12,26]. Indeed, while the first extraction can include free syringic acid, the second one rather focuses on the acids that are deeply impregnated in the clay matrix. The acid-catalyzed extraction hence promotes the tracing of malvidin origin. Following Table 2, none of the samples revealed the presence of syringic acid as free acid and butyl syringate was identified in almost all the samples (m/z 240, 296, 311, 326), shedding light on the dark-color beverage contained in amphorae. The Chios amphora was the only one that did not show the dark marker, highlighting the white nature of the content.
Coupling organic extraction with esterification extended the overall number of identified compounds from two to ten molecules per sample over the twenty-two molecules targeted (Figure 1). No DHA was extracted for the amphorae Nos. 6904, 6566 and 6828c using alkaline fusion whereas this compound was extracted with DCM:MeOH for all the items (Table 2). Exceptions were nonetheless noticed for amphorae Nos. 6828b and 6793 where organic solvents barely supported the extraction of diterpene derivatives, hence grandly diminishing the number of characterized molecules (Figure 1). Oxidized diterpenes, identified in all the amphorae, provided insights on the ageing of the pitch that occurred through oxygen incorporation into the DHA skeleton to form peroxide intermediates. Hydroxyl and ketone derivatives are produced by peroxide reduction or dehydration, respectively [40].
Almost all the protocols extracted DHAM markers, traducing an archaeological use of Pinaceae wood tar [41] (Figure 1). Indeed, DHAM characterizes wood distillation: DHA is esterified by the methanol contained in the wood during pyrolysis at high temperature. Interestingly, the amphora from Chios (No. 6904) is the unique potsherd only containing Pinus pitch. Neither DHAM derivatives, nor triterpenoids were identified yet the island was famous for its mastic resin of Pistacia lentiscus L. [42]. The presence of resinated wine such as retsina can be assumed since the practice was common in Greece. Aside from antibacterial properties, small amounts of Pinus pitch could be added during the alcoholic fermentation to flavor and color wines [43]. No retene was observed for the amphorae Nos. 6566 and 6828c with alkaline fusion whereas it did with DCM:MeOH (Figure 1). Even though no chemical explanation can afford it, the archaeological meaning refers to the heating treatment of the resin [40]. Thermal degradation above 300 °C induces abietane aromatization into retene. Only the amphora No. 6545 (Dressel 5), where retene has never been identified, was coated with a resinous material that had not been produced under a high temperature.

4.3. Extracting Capacities Comparison on Archaeological Coatings

The protocol comparison applied on coatings encompassed: (i) an alkaline fusion with KOH extraction, (ii) a DCM:MeOH organic extraction, (iii) its coupling with BF3-catalyzed MW-butylation applied on the dried remaining powder after the organic extraction (2LE-MW) and (iv) the same BF3-catalyzed MW-butylation but directly applied on the coating matter (R-2LE-MW).
The radar plot (Figure 2) overviews the extractive molecular capacities of the different protocols for each of the eight coatings. Samples Nos. 1014 and 749 have already been discussed through the analyses of their associated shards. Coatings and shards analyses provided identical insights on the nature of the coating and the content even though poor, insignificant molecular differences regarding DHA and DHAM derivatives were observed (Table 2 and Table 3).
Tartaric acid was successfully identified in all of the eight coatings when the butylation was directly applied on the pitch (R-2LE-MW) (Figure 2). In comparison, when it was applied after organic extraction (2LE-MW), the esterification allowed DBT to be identified from only seven pitch samples. The amphora No. 1014 (Dressel 1) did not exhibit DBT in the butylated fraction, which can be related to the insufficient remaining sample quantity. Indeed, the remaining powders were usually less than 1 mg, a major part of the initial mass being dissolved during DCM:MeOH extraction. Moreover, DBT was identified as traces for all the samples in 2LE-MW, which clearly argued for the limited quantity effect. The protocol was therefore hardly suitable for content identification. Except for the amphorae Nos. SFC1 and SFC3, the alkaline fusion did not allow tartaric acid to be extracted (Figure 2), which mirrors the unsuitability already reported for shards.
Considering fermented markers, pyruvic acid was only identified with butylation, as dibutylcetal (m/z 61, 117, 173) [44]. Maleic acid was never characterized. Succinic and fumaric acids were respectively rarely and never identified with butylation, although they were successfully extracted with alkaline fusion or organic solvent extraction (Table 3). Moreover, KOH fusion allowed eight additional grape-acids (i.e., succinic, fumaric and malic acids) to be extracted from five amphorae (Nos. 6570a, SFC1, SFC3, SFC4 and SFC5) that traditional solvents did not provide.
Syringic acid was not detected in any of the amphorae SFC2, SFC4 and SFC5 (Table 3), hence suggesting a white winemaking process. On the contrary, red beverages were conjectured for all the other samples (Nos. 1014, 749, 6570a, SFC1 and SFC3). Although a two-step protocol with successive alkaline fusion and butylation would favor the identification of free syringic acid, it is technically hardly feasible because the alkaline fusion left an aqueous matrix difficult to dry completely. Remaining water strongly reacts with BF3, thus inhibiting butylation (data not shown).
Interestingly, diterpenic markers could not be addressed after butylation. Although diterpenic acids should have been recovered as butylated-derivatives, pimarane and abietane seemed to have undergone transformation. Applied on a standard of pimaric acid (Figure 3C), butylation gave rise to a wide distribution of unidentified diterpenic compounds, of which patterns belong to pimarane (m/z 241; 359). The butylation of standard colophony similarly produced unidentified compounds of abietane skeletons (m/z 239; 372) together with the same unidentified molecules already observed with pimaric acid butylation (Figure 3B,C). Induced by the harsh Lewis-acid conditions, diterpenic skeletons were reported to undergo isomerization, skeletal transposition, isomerization, rearrangement and proton migration [45,46,47]. Moreover, the butylation of standard colophony selectively esterified the diterpenoids present in the resin (Figure 3A,B). Only butyl dehydroabietate and pimarate could be identified (m/z 239; 356 and m/z 241; 343, respectively). The remaining presence of DHA in the butylated fraction outlined the incomplete esterification of diterpenoids.
Aside from showing unequivocal extractive capacities to target tartaric acid when directly applied on the pitch, the butylation is consequently not self-sufficient to describe the resinous material. The protocol must be coupled to broaden the scope of characterization and alkaline fusion coupled to direct-pitch butylation managed to give rise to more quantitative extractions (Figure 2). Emphasizing on coating markers, the numerical difference reverted to the extractive capacities of alkaline fusion versus traditional solvents and homogeneously concerned DHA derivatives (hydroxy-DHA for all the coatings, oxo-DHA for six of them, dihydroxy-DHA for four of them and hydroxy-oxo-DHA for three pitch over eight). Again, the important presence of oxidized Pinaceae diterpenoids highlighted the significant ageing of Pinus pitch. Retene and DHAM markers, characterized in all the coatings independently from the protocol employed (Figure 2), attested of Pinaceae wood tar produced under high temperature pyrolytic treatment.

5. Conclusions

Beyond the successful MW parameters set to esterify molecules, the analyses show different trends. On the one hand, the butylation of standard acids familiar of wine composition is successfully achieved. Grape-acids were recovered as butylated products using MW. Indeed, the energy conduction of MW-assisted reactions being different from reflux [38], it is essential to verify the molecular integrity. On the other hand, even though butylation undoubtedly promoted tartrate extraction from ceramic and coating materials, the protocol, when solely applied, did not provide a representative picture of archaeological artefacts, specifically for coating material characterization. Although MW are widely encouraged to reduce reaction times and costs, butylation has to be strengthen up using another extractive protocol. For the analyses of shards, organic extraction with traditional solvents followed by esterification were remarkably complementary, offering more reliability for the identification of grape-derivatives in terms of number of extracted molecules and retene extraction.
In parallel, the butylation protocol, initially developed for inorganic potsherds, was successfully adapted to organic coatings, hence ensuring a reliable tracing of archaeological grape-beverage directly from the coating. Alkaline fusion allowed diterpenic markers to be extracted, concluding on the nature; the formulation and the ageing of the coating matter, as well as it suggested the presence of fermentation markers. Although a two-step protocol would prevent from free-acid misinterpretation, further investigation is needed to overcome BF3 reaction with water.
Except for the amphora from Chios coated with pine pitch, all the other artefacts were waterproofed with Pinus wood tar. All the samples validate the presence of archaeological grape contents, which could have been wine, vinegar or any other fermented derivatives [3]. Even though Mañà C2 (No. SFC3) are mostly considered for oil or fish preserves, the identification of wine markers may be interpreted in the light of a reutilization, that needs to be further investigated. Among the frequent typologies mentioned for wine trading, the presence of red beverage was confirmed in Dressel 5 (No. 6545), hitherto barely studied.

Author Contributions

Conceptualization, L.C., C.V., C.M.; methodology, L.C.; formal analysis, L.C.; investigation, L.C., C.M.; resources, F.O., C.D., A.C.; writing—original draft preparation, L.C., C.M.; writing—review and editing, L.C., F.O., C.D., A.C., C.V., D.M., C.M.; visualization, L.C.; supervision, C.M.; project administration, C.V., D.M., C.M.; funding acquisition, A.C., C.V., D.M., C.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research and the APC were funded by Marie Skłodowska-Curie Innovative Training Network H2020-MSCA-ITN-EJD ED-ARCHMAT Joint Doctorate, ESR9, Grant Number 766311. Work in Rome was funded by Sapienza University project DISCOVER (RM120172AE4E11D6).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data available on request due to restrictions, e.g., privacy or ethical. The data presented in this study are available on request from the corresponding author.

Acknowledgments

This research was financially supported by Marie Skłodowska-Curie Innovative Training Network, the ED-ARCHMAT European Joint Doctorate, H2020-MSCA-ITN-EJD ED-ARCHMAT Joint Doctorate (Project ESR9, Grant Agreement No. 766311). The authors wish to thank Nathalie Huet, research engineer in charge of preventive conservation and furniture management at the DRASSM (France).

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

Abbreviations

DBTDibutyl tartrate
DCMDichloromethane
DHADehydroabietic acid
DHAMDehydroabietic methyl ester
DiOH-DHA7,15-dihydroxy-dehydroabietic acid
DiOH-DHAM7,15-dihydroxy-dehydroabietic methyl ester
KOHPotassium hydroxide
OH-DHAHydroxy-dehydroabietic acid (3-hydroxy-dehydroabietic acid; 7-hydroxy-dehydroabietic acid; 15-hydroxy-dehydroabietic acid)
OH-DHAMHydroxy-dehydroabietic methyl ester (3-hydroxy-dehydroabietic methyl ester; 7-hydroxy-dehydroabietic methyl ester; 15-hydroxy-dehydroabietic methyl ester)
Oxo-DHA7-oxo-dehydroabietic acid
Oxo-DHAM7-oxo-dehydroabietic methyl ester
Oxo-OH-DHA7-oxo-15-hydroxy-dehydroabietic acid
Oxo-OH-DHAM7-oxo-15-hydroxy-dehydroabietic methyl ester

References

  1. Heron, C.; Evershed, R.P. The analysis of organic residues and the study of pottery use. Archaeol. Method Theory 1993, 5, 247–284. [Google Scholar] [CrossRef]
  2. Evershed, R.P. Organic residue analysis in archaeology: The archaeological biomarker revolution. Archaeometry 2008, 50, 895–924. [Google Scholar] [CrossRef]
  3. McGovern, P.; Jalabadze, M.; Batiuk, S.; Callahan, M.P.; Smith, K.E.; Hall, G.R.; Kvavadze, E.; Maghradze, D.; Rusishvili, N.; Bouby, L.; et al. Early Neolithic wine of Georgia in the South Caucasus. Proc. Natl. Acad. Sci. USA 2017, 114, E10309–E10318. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. McGovern, P.E.; Michel, R.H. The analytical and archaeological challenge of detecting ancient wine: Two case studies from the Ancient Near East. In The Origins and Ancient History of Wine; Routledge: Langhorne, PA, USA, 1996; pp. 57–65. [Google Scholar]
  5. McGovern, P.E.; Zhang, J.; Tang, J.; Zhang, Z.; Hall, G.R.; Moreau, R.A.; Nunez, A.; Butrym, E.D.; Richards, M.P.; Wang, C.-s.; et al. Fermented beverages of pre- and proto-historic China. Proc. Natl. Acad. Sci. USA 2004, 101, 17593–17598. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Michel, R.H.; McGovern, P.E.; Badler, V.R. The first wine & beer. Anal. Chem. 1993, 65, 408A–413A. [Google Scholar] [CrossRef]
  7. Garnier, N.; Richardin, P.; Cheynier, V.; Regert, M. Characterization of thermally assisted hydrolysis and methylation products of polyphenols from modern and archaeological vine derivatives using gas chromatography–mass spectrometry. Anal. Chim. Acta 2003, 493, 137–157. [Google Scholar] [CrossRef]
  8. Jerković, I.; Marijanović, Z.; Gugić, M.; Roje, M. Chemical profile of the organic residue from ancient amphora found in the adriatic sea determined by direct GC and GC-MS analysis. Molecules 2011, 16, 7936–7948. [Google Scholar] [CrossRef]
  9. Pecci, A.; Giorgi, G.; Salvini, L.; Cau Ontiveros, M.Á. Identifying wine markers in ceramics and plasters using gas chromatography–mass spectrometry. Experimental and archaeological materials. J. Archaeol. Sci. 2013, 40, 109–115. [Google Scholar] [CrossRef]
  10. Stern, B.; Heron, C.; Tellefsen, T.; Serpico, M. New investigations into the Uluburun resin cargo. J. Archaeol. Sci. 2008, 35, 2188–2203. [Google Scholar] [CrossRef]
  11. Drieu, L.; Rageot, M.; Wales, N.; Stern, B.; Lundy, J.; Zerrer, M.; Gaffney, I.; Bondetti, M.; Spiteri, C.; Thomas-Oates, J.; et al. Is it possible to identify ancient wine production using biomolecular approaches? STAR Sci. Technol. Archaeol. Res. 2020, 6, 16–29. [Google Scholar] [CrossRef]
  12. Garnier, N.; Valamoti, S.M. Prehistoric wine-making at Dikili Tash (Northern Greece): Integrating residue analysis and archaeobotany. J. Archaeol. Sci. 2016, 74, 195–206. [Google Scholar] [CrossRef]
  13. Pecci, A.; Borgna, E.; Mileto, S.; Dalla Longa, E.; Bosi, G.; Florenzano, A.; Mercuri, A.M.; Corazza, S.; Marchesini, M.; Vidale, M. Wine consumption in Bronze Age Italy: Combining organic residue analysis, botanical data and ceramic variability. J. Archaeol. Sci. 2020, 123, 105256. [Google Scholar] [CrossRef]
  14. Fujii, H.; Mazzitelli, J.B.; Adilbekov, D.; Olmer, F.; Mathe, C.; Vieillescazes, C. FT-IR and GC–MS analyses of Dressel IA amphorae from the Grand Congloué 2 wreck. J. Archaeol. Sci. Rep. 2019, 28, 102007. [Google Scholar] [CrossRef]
  15. Fujii, H.; Krausz, S.; Olmer, F.; Mathe, C.; Vieillescazes, C. Analysis of organic residues from the Châteaumeillant oppidum (Cher, France) using GC–MS. J. Cult. Herit. 2021, 51, 50–58. [Google Scholar] [CrossRef]
  16. Guasch-Jané, M.R.; Ibern-Gómez, M.; Andrés-Lacueva, C.; Jáuregui, O.; Lamuela-Raventós, R.M. Liquid chromatography with mass spectrometry in tandem mode applied for the identification of wine markers in residues from ancient Egyptian vessels. Anal. Chem. 2004, 76, 1672–1677. [Google Scholar] [CrossRef] [PubMed]
  17. Guasch-Jané, M.R.; Andrés-Lacueva, C.; Járegui, O.; Lamuela-Raventós, R.M. The origin of the ancient Egyptian drink Shedeh revealed using LC/MS/MS. J. Archaeol. Sci. 2006, 33, 98–101. [Google Scholar] [CrossRef]
  18. Barnard, H.; Dooley, A.N.; Areshian, G.; Gasparyan, B.; Faull, K.F. Chemical evidence for wine production around 4000 BCE in the Late Chalcolithic Near Eastern highlands. J. Archaeol. Sci. 2011, 38, 977–984. [Google Scholar] [CrossRef]
  19. Manzano, E.; Cantarero, S.; García, A.; Adroher, A.; Vílchez, J.L. A multi-analytical approach applied to the archaeological residues in Iberian glasses. Earliest evidences on the consumption of fermented beverages in votive rituals. Microchem. J. 2016, 129, 286–292. [Google Scholar] [CrossRef]
  20. Rageot, M.; Mötsch, A.; Schorer, B.; Bardel, D.; Winkler, A.; Sacchetti, F.; Chaume, B.; Della Casa, P.; Buckley, S.; Cafisso, S.; et al. New insights into Early Celtic consumption practices: Organic residue analyses of local and imported pottery from Vix-Mont Lassois. PLoS ONE 2019, 14, e0218001. [Google Scholar] [CrossRef] [PubMed]
  21. Hasnaoui, N.; Jbir, R.; Mars, M.; Trifi, M.; Kamal-Eldin, A.; Melgarejo, P.; Hernandez, F. Organic acids, sugars, and anthocyanins contents in juices of Tunisian pomegranate fruits. Int. J. Food Prop. 2011, 14, 741–757. [Google Scholar] [CrossRef] [Green Version]
  22. Pecci, A.; Domínguez-Bella, S.; Buonincontri, M.P.; Miriello, D.; De Luca, R.; Di Pasquale, G.; Cottica, D.; Bernal Casasola, D. Combining residue analysis of floors and ceramics for the study of activity areas at the Garum Shop at Pompeii. Archaeol. Anthropol. Sci. 2018, 10, 485–502. [Google Scholar] [CrossRef]
  23. Ribéreau-Gayon, P.; Glories, Y.; Maujean, A.; Dubourdieu, D. The Chemistry of Wine Stabilization and Treatment, Handbook of Enology; Johan Wiley & Sons: Hoboken, NJ, USA, 2006; Volume 2. [Google Scholar]
  24. Blanco-Zubiaguirre, L.; Olivares, M.; Castro, K.; Carrero, J.A.; García-Benito, C.; García-Serrano, J.Á.; Pérez-Pérez, J.; Pérez-Arantegui, J. Wine markers in archeological potteries: Detection by GC-MS at ultratrace levels. Anal. Bioanal. Chem. 2019, 411, 6711–6722. [Google Scholar] [CrossRef]
  25. Correa-Ascencio, M.; Evershed, R.P. High throughput screening of organic residues in archaeological potsherds using direct acidified methanol extraction. Anal. Methods 2014, 6, 1330. [Google Scholar] [CrossRef]
  26. Pecci, A.; Clarke, J.; Thomas, M.; Muslin, J.; van der Graaff, I.; Toniolo, L.; Miriello, D.; Crisci, G.; Buonincontri, M.; Di Pasquale, G. Use and reuse of amphorae. Wine residues in Dressel 2–4 amphorae from Oplontis Villa B (Torre Annunziata, Italy). J. Archaeol. Sci. Rep. 2017, 12, 515–521. [Google Scholar] [CrossRef]
  27. Drieu, L.; Orecchioni, P.; Capelli, C.; Meo, A.; Lundy, J.; Sacco, V.; Arcifa, L.; Molinari, A.; Carver, M.; Craig, O.E. Chemical evidence for the persistence of wine production and trade in Early Medieval Islamic Sicily. Proc. Natl. Acad. Sci. USA 2021, 118, e2017983118. [Google Scholar] [CrossRef] [PubMed]
  28. Frère, D.; Garnier, N. Dairy Product and Wine in Funerary Rituals: The Case of a Hellenistic Etruscan Tomb. J. Hist. Archaeol. Anthropol. Sci. 2017, 1, 222–227. [Google Scholar] [CrossRef] [Green Version]
  29. Tchernia, A. Premiers Résultats Des Fouilles de Juin 1968 Sur L’épave 3 du Planier; Etudes Classiques: Bouches-du-Rhône, France, 1968; Volume III. [Google Scholar]
  30. Tchernia, A. Les fouilles sous-marines de Planier (Bouches-du-Rhône). C. R. Séances L’académie Inscr. B-Lett. 1969, 113, 292–309. [Google Scholar] [CrossRef]
  31. Olmer, F. Les Amphores Sont-Elles Utiles à la Chronologie de la fin de L’âge du Fer Annexe. Philippe Barral. Regards sur la Chronologie de la fin de l’âge du Fer (IIIe-Ier siècle avant J.-C.) en Gaule non Méditerranéenne. Actes de la Table Ronde Tenue à Bibracte “Chronologie de la fin de l’âge du Fer (IIIe-Ier siècle avant J.-C.) Dans l’est de la France et les Régions Voisines ”, 22, Bibracte. 2012. ISBN 978-2-909668-74-1. Available online: https://halshs.archives-ouvertes.fr/halshs-01328970/document (accessed on 20 September 2021).
  32. Cipriano, M.T.; Carre, M.-B. Production et typologie des amphores sur la côte adriatique de l’Italie. Amphores romaines et histoire economique. Dix ans de recherche. Actes du colloque de Sienne (22–24 mai 1986). Publ. L’école Française Rome 1989, 114, 67–104. [Google Scholar]
  33. Panella, C. Anfore e archeologia subacquea. In Archeologia Subacquea. Come Opera L’archeologo Sott’acqua. Storie Dalle Acque; All’insegna del Giglio: Firenze, Italy, 1998; pp. 531–559. [Google Scholar]
  34. Lamboglia, N. La Nave Romana di Albenga. Rev. Degli Stud. Liguri 1952, 18, 131–236. [Google Scholar]
  35. Okan, E.; Atila, C.; Akyol, A.A. The production of chios-style amphorae at a ceramic workshop in Phocaea (Foça). Mediterr. Archaeol. Archaeom. 2015, 15, 259–276. [Google Scholar] [CrossRef]
  36. Bonifay, M. Amphores de l’Afrique romaine: Nouvelles avancées sur la production, la typo-chronologie et le contenu. In Amphorae ex Hispania. Paisajes de producción y consumo, III Congreso internacional de la SECAH-EX OFFICINA HISPANA (Tarragone, 10-13 décembre 2014); Monografías Ex Officina Hispana III: Tarragona, Spain, 2016; pp. 595–611. [Google Scholar]
  37. Peña, J.T. Roman Pottery in the Archaeological Record; Cambridge University Press: Cambridge, UK, 2007; ISBN 9780511499685. [Google Scholar]
  38. Kappe, C.O.; Dallinger, D.; Murphree, S.S. Practical Microwave Synthesis for Organic Chemists: Strategies, Instruments, and Protocols; John Wiley & Sons: Hoboken, NJ, USA, 2008; ISBN 9783527314522. [Google Scholar]
  39. Garnier, N.; Bernal Casasola, D.; Driard, C.; Pinto, I.V. Looking for ancient fish products through invisible biomolecular residues in the Roman production vats from the Atlantic coast. J. Marit. Archaeol. 2018, 13, 285–328. [Google Scholar] [CrossRef]
  40. Mezzatesta, E.; Perraud, A.; Vieillescazes, C.; Mathe, C. GC–MS and PCA analyses of diterpenoids degradation state in 21 human mummies of Ancient Egypt dating from New Kingdom to Graeco-Roman Period. J. Cult. Herit. 2021, 47, 43–49. [Google Scholar] [CrossRef]
  41. Colombini, M.P.; Modugno, F. Organic Mass Spectrometry in Art and Archaeology; Colombini, M.P., Modugno, F., Eds.; John Wiley & Sons: Chichester, UK, 2009; ISBN 9780470741917. [Google Scholar]
  42. Modugno, F.; Ribechini, E.; Colombini, M.P. Chemical study of triterpenoid resinous materials in archaeological findings by means of direct exposure electron ionisation mass spectrometry and gas chromatography/mass spectrometry. Rapid Commun. Mass Spectrom. 2006, 20, 1787–1800. [Google Scholar] [CrossRef]
  43. Proestos, C.; Bakogiannis, A.; Psarianos, C.; Koutinas, A.A.; Kanellaki, M.; Komaitis, M. High performance liquid chromatography analysis of phenolic substances in Greek wines. Food Control 2005, 16, 319–323. [Google Scholar] [CrossRef]
  44. Kawamura, K. Identification of C2-C10 ω-oxocarboxylic acids, pyruvic acid, and C2-C3 α-dicarbonyls in wet precipitation and aerosol samples by capillary GC and GC/MS. Anal. Chem. 1993, 65, 3505–3511. [Google Scholar] [CrossRef]
  45. Delmond, B.; Taran, M.; Valade, J. Réarrangements de squelette au cours de l’isomérisation catalysée par l’éthérate de trifluorure de bore d’epoxyde diterpénique. Tetrahedron Lett. 1978, 19, 4791–4794. [Google Scholar] [CrossRef]
  46. Taran, M.; Delmond, B. Part. IV—Obtention de nouveaux squelettes diterpeniques tetracycliques lors d’isomerisation d’epoxydes-8,9 diterpeniques. Tetrahedron 1986, 42, 4795–4806. [Google Scholar] [CrossRef]
  47. Fujita, E.; Fuji, K.; Nagao, Y.; Node, M. The Chemistry on Diterpenoids in 1978. Part-III. Bull. Inst. Chem. Res. Kyoto Univ. 1979, 57, 385–410. [Google Scholar]
Figure 1. Radar plot of the shards.
Figure 1. Radar plot of the shards.
Crystals 11 01300 g001
Figure 2. Radar plot of the coating materials.
Figure 2. Radar plot of the coating materials.
Crystals 11 01300 g002
Figure 3. TIC Chromatograms. (A) Standard colophony extracted with DCM:MeOH; (B) standard colophony extracted with MW-assisted butylation and (C) standard pimaric acid extracted with MW-assisted butylation with m/z 241 chromatogram. ac.: acid; but.: butylated; *: unidentified compound.
Figure 3. TIC Chromatograms. (A) Standard colophony extracted with DCM:MeOH; (B) standard colophony extracted with MW-assisted butylation and (C) standard pimaric acid extracted with MW-assisted butylation with m/z 241 chromatogram. ac.: acid; but.: butylated; *: unidentified compound.
Crystals 11 01300 g003
Table 1. Archaeological amphorae investigated.
Table 1. Archaeological amphorae investigated.
AmphoraArchaeological SiteTypologyCoatingShardGrape DerivativesPinaceae Products
1014Planier 3Dressel 1XXFermentedWood tar
749Planier 3Lamboglia 2XXFermentedWood tar
6570aPlanier 3Dressel 1X FermentedWood tar
SFC1San Felice CirceoDressel 1X FermentedWood tar
SFC2San Felice CirceoDressel 1X FermentedWood tar
SFC3San Felice CirceoMañà C2X FermentedWood tar
SFC4San Felice CirceoGreek-ItalianX FermentedWood tar
SFC5San Felice CirceoLamboglia 2X FermentedWood tar
6904Planier 3Chios amphora XFermentedPitch
6828aPlanier 3Lamboglia 2 XFermentedWood tar
6828aPlanier 3Lamboglia 2 XFermentedWood tar
6565Planier 3Lamboglia 2 XFermentedWood tar
6793Planier 3Lamboglia 2 XFermentedWood tar
6545Planier 3Dressel 5 XFermentedWood tar
6566Planier 3Lamboglia 2 XFermentedWood tar
6828cPlanier 3Lamboglia 2 XFermentedWood tar
Table 2. Molecules identified in shards. Presence (+) and absence (-) of molecular markers under alkaline fusion (KOH ext.), organic extraction with DCM:MeOH, butylation applied on the remaining fraction (2LE-MW). The number of ‘+’ refers to the number of molecules present. ac.: acid; OH-DHA: hydroxy-DHA (i.e., 3-hydroxy-DHA; 7-hydroxy-DHA and 15-hydroxy-DHA); Oxo-DHA: 7-oxo-DHA; DiOH-DHA: 7,15-dihydroxy-DHA; Oxo-OH-DHA: 7-oxo-15-hydroxy-DHA. DHAM and oxidized derivatives (OH-DHAM, Oxo-DHAM, DiOH-DHAM and Oxo-OH-DHAM) refer to the same skeleton with methyl ester derivatives instead of the carboxylic acid function.
Table 2. Molecules identified in shards. Presence (+) and absence (-) of molecular markers under alkaline fusion (KOH ext.), organic extraction with DCM:MeOH, butylation applied on the remaining fraction (2LE-MW). The number of ‘+’ refers to the number of molecules present. ac.: acid; OH-DHA: hydroxy-DHA (i.e., 3-hydroxy-DHA; 7-hydroxy-DHA and 15-hydroxy-DHA); Oxo-DHA: 7-oxo-DHA; DiOH-DHA: 7,15-dihydroxy-DHA; Oxo-OH-DHA: 7-oxo-15-hydroxy-DHA. DHAM and oxidized derivatives (OH-DHAM, Oxo-DHAM, DiOH-DHAM and Oxo-OH-DHAM) refer to the same skeleton with methyl ester derivatives instead of the carboxylic acid function.
AmphoraProtocolMaleic ac.Succinic ac.Pyruvic ac.Fumaric ac.Malic ac.Tartaric ac.Syringic ac.ReteneDHAOH-DHAOxo-DHADiOH-DHAOxo-OH-DHADHAMOH-DHAMOxo-DHAMDiOH-DHAMOxo-OH-DHAM
1014KOH ext-+-+--+++++++--++--+
DCM:MeOH-++----+++++++-+++++-
2LE-MW++-++++
749KOH ext-+-++-++++++++++++++++
DCM:MeOH-++----+++++++++++++++
2LE-MW++--+++
6904KOH ext------------------
DCM:MeOH-++----++---------
2LE-MW++-+++-
6828AKOH ext-+-++-++++++++++++-++
DCM:MeOH-+++---+++++-+++++++++
2LE-MW-+--+++
6828BKOH ext-++++-+++++++++++-++
DCM:MeOH-++---+++++---+++++--
2LE-MW++--+++
6565KOH ext-++-+-+++----+++-+
DCM:MeOH-++----++++---++++-+
2LE-MW-+--+++
6793KOH ext-+----++++++-+-++++++
DCM:MeOH-++---++++---+----
2LE-MW-+--+++
6545KOH ext-------------+----
DCM:MeOH-+---------+-+--+-
2LE-MW-+--+++
6566KOH ext-------------++---
DCM:MeOH--+----++--+-++-+-
2LE-MW-+--+++
6828CKOH ext---+---------+----
DCM:MeOH--+----++--+-++++-
2LE-MW-+--+++
Table 3. Molecules identified in pitch coatings. Presence (+) and absence (-) of molecular markers under alkaline fusion (KOH ext.), organic extraction with DCM:MeOH, butylation applied on the remaining fraction (2LE-MW) or applied directly on the pitch (R-2LE-MW). The number of ‘+’ refers to the number of molecules present. ac.: acid; OH-DHA: hydroxy-DHA (i.e., 3-hydroxy-DHA; 7-hydroxy-DHA and 15-hydroxy-DHA); Oxo-DHA: 7-oxo-DHA; DiOH-DHA: 7,15-dihydroxy-DHA; Oxo-OH-DHA: 7-oxo-15-hydroxy-DHA. DHAM and oxidized derivatives (OH-DHAM, Oxo-DHAM, DiOH-DHAM and Oxo-OH-DHAM) refer to the same skeleton with methyl ester derivatives instead of the acid function.
Table 3. Molecules identified in pitch coatings. Presence (+) and absence (-) of molecular markers under alkaline fusion (KOH ext.), organic extraction with DCM:MeOH, butylation applied on the remaining fraction (2LE-MW) or applied directly on the pitch (R-2LE-MW). The number of ‘+’ refers to the number of molecules present. ac.: acid; OH-DHA: hydroxy-DHA (i.e., 3-hydroxy-DHA; 7-hydroxy-DHA and 15-hydroxy-DHA); Oxo-DHA: 7-oxo-DHA; DiOH-DHA: 7,15-dihydroxy-DHA; Oxo-OH-DHA: 7-oxo-15-hydroxy-DHA. DHAM and oxidized derivatives (OH-DHAM, Oxo-DHAM, DiOH-DHAM and Oxo-OH-DHAM) refer to the same skeleton with methyl ester derivatives instead of the acid function.
AmphoraProtocolMaleic ac.Succinic ac.Pyruvic ac.Fumaric ac.Malic ac.Tartaric ac.Syringic ac.ReteneDHAOH-DHAOxo-DHADiOH-DHAOxo-OH-DHADHAMOH-DHAMOxo-DHAMDiOH-DHAMOxo-OH-DHAM
1014KOH ext-+----+++++++++++++-++
DCM:MeOH-+-+--+++++-+-+++++++
2LE-MW------+
R-2LE-MW-++-+++
749KOH ext-+-++-++++++++++++-++
DCM:MeOH-+-++-+++++---+++++-+
2LE-MW--+-++-
R-2LE-MW-++-+++
6570aKOH ext-+-++-++++++++++++-++
DCM:MeOH-------+---+-+++++
2LE-MW----++-
R-2LE-MW--+-+++
SFC1KOH ext-+-+++-+++++++-++++-
DCM:MeOH-+-----++----+++++--
2LE-MW--+-+++
R-2LE-MW-++-+++
SFC2KOH ext-+-----+++---+-+--
DCM:MeOH-+-----++----++++--
2LE-MW----++-
R-2LE-MW--+-++-
SFC3KOH ext-+-+++++++++++-++++++-
DCM:MeOH-+-+---++----+++++-+
2LE-MW--+--++
R-2LE-MW--+-+++
SFC4KOH ext-+-----+++---++++++
DCM:MeOH-------++----++++--
2LE-MW----++-
R-2LE-MW--+-++-
SFC5KOH ext-+-+---+++++++++++++++
DCM:MeOH-+-----++----++++-+
2LE-MW--+--+-
R-2LE-MW-++-++-
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chassouant, L.; Olmer, F.; Delpino, C.; Celant, A.; Vieillescazes, C.; Magri, D.; Mathe, C. Protocol Comparison for Organic Residue Analyses from Waterproofing Materials and Shards of Roman Archaeological Amphorae. Crystals 2021, 11, 1300. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst11111300

AMA Style

Chassouant L, Olmer F, Delpino C, Celant A, Vieillescazes C, Magri D, Mathe C. Protocol Comparison for Organic Residue Analyses from Waterproofing Materials and Shards of Roman Archaeological Amphorae. Crystals. 2021; 11(11):1300. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst11111300

Chicago/Turabian Style

Chassouant, Louise, Fabienne Olmer, Chiara Delpino, Alessandra Celant, Cathy Vieillescazes, Donatella Magri, and Carole Mathe. 2021. "Protocol Comparison for Organic Residue Analyses from Waterproofing Materials and Shards of Roman Archaeological Amphorae" Crystals 11, no. 11: 1300. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst11111300

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop