Skip to main content

Advertisement

Log in

Charting a Future Course for Organic Residue Analysis in Archaeology

  • Published:
Journal of Archaeological Method and Theory Aims and scope Submit manuscript

Abstract

Working hypotheses, which draw upon as many relevant disciplines as possible to derive the maximum information from a very limited database, are key to the highly interdisciplinary field of organic residue analysis in archaeology, a branch of biomolecular archaeology. Archaeology and chemistry are most important for effectively developing and testing such hypotheses, but botany, zoology, geology, etc. also need to be taken into account. Archaeologically, the goal is to obtain as many relevant samples as possible from the best preserved and dated contexts, which have been subjected to the least degradation and disturbance by later natural processes and human handling, including washing and conservation treatment. Chemically, molecular biomarkers of natural products need to be defined and identified by the best and most appropriate techniques, together with bioinformatics searches and assessment of degradation. With ever-improving techniques and new data, previously analyzed samples need to be retested and hypotheses possibly reformulated. Consideration of three case studies illustrates this holistic approach to inductive hypothesis generation and deductive testing: (1) new chemical findings that attest to grape wine in amphoras on board the 14th c. B.C. Uluburun ship, the earliest recorded Mediterranean wreck; (2) recently published research on beeswax/mead in Chalcolithic Israel and Neolithic China and Poland; and (3) recent articles on milk products from 2nd millennium B.C. Central Asia and Neolithic Poland. Potential pitfalls leading to weak hypotheses and mistaken conclusions are described, and a more productive approach is proposed.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

References

  • Amber Research Laboratory of Vassar College, chemical database (maintained by the National Gallery, London. http://cima.ng-london.org.uk/arl/. Accessed 11 Feb 2015.

  • Badler, V. R., McGovern, P. E., & Michel, R. H. (1990). Drink and be merry!: Infrared spectroscopy and ancient Near Eastern wine. In W. R. Biers & P. E. McGovern (Eds.), Organic contents of ancient vessels: Materials analysis and archaeological investigation (pp. 25–36). MASCA Research Papers in Science and Archaeology, vol. 7. Philadelphia: MASCA, University of Pennsylvania Museum, University of Pennsylvania.

  • Barnard, H., Dooley, A. N., Areshian, G., Gasparyan, B., & Faull, K. F. (2011). Chemical evidence for wine production around 4000 BCE in the late Chalcolithic near eastern highlands. Journal of Archaeological Science, 38, 977–984.

    Article  Google Scholar 

  • Bass, G. F., Pulak, C., Collon, D., & Weinstein, J. M. (1989). The Bronze Age shipwreck at Uluburun: 1986 campaign. American Journal of Archaeology, 93, 1–29.

    Article  Google Scholar 

  • Beck, C. W., Stout, E. C., Wovkulich, K. M., & Phillips, A. J. J. (2007). Analysis of organic remains in the fabric of Minoan and Mycenaean pottery sherds by gas chromatograph-mass spectrometry. In Y. Tzedakis, H. Martlew, & M. K. Jones (Eds.), Archaeology meets science: Biomolecular investigations in Bronze Age Greece; the primary scientific evidence 1997–2003 (pp. 12–47). Oxford: Oxbow.

    Google Scholar 

  • Bogucki, P. I. (1984). Ceramic sieves of the Linear Pottery Culture and their economic implications. Oxford Journal of Archaeology, 3, 15–30.

    Article  Google Scholar 

  • Boulton, N., & Heron, C. (2000). Chemical detection of ancient wine. In P. T. Nicholson & I. Shaw (Eds.), Ancient Egyptian materials and technology (pp. 599–603). Cambridge: Cambridge University.

    Google Scholar 

  • Carlson, D. N. (2003). The classical Greek shipwreck at Tektaš, Burnu, Turkey. American Journal of Archaeology, 107, 581–600.

    Article  Google Scholar 

  • Cavalieri, D., McGovern, P. E., Hartl, D. L., Mortimer, R., & Polsinelli, M. (2003). Evidence for S. cerevisiae fermentation in ancient wine. Journal of Molecular Evolution, 57, S226–S232.

    Article  Google Scholar 

  • Colombini, M. P., Giachi, G., Modugno, F., & Ribechini, E. (2005). Characterisation of organic residues in pottery vessels of the Roman age from Antinoe (Egypt). Microchemical Journal, 79, 83–90.

    Article  Google Scholar 

  • Crane, E. (1999). The world history of beekeeping and honey hunting. New York: Routledge.

    Google Scholar 

  • Dalby, A. (2011). Geoponika: Farm work. Devon, Eng.: Prospect Books.

    Google Scholar 

  • DeNiro, M. J., & Epstein, S. (1977). Mechanism of carbon isotope fractionation associated with lipid synthesis. Science, 197, 261–263.

    Article  Google Scholar 

  • Dickson, J. H. (1978). Bronze Age mead. Antiquity, 52, 108–113.

    Article  Google Scholar 

  • Dr. Duke’s Phytochemical and Ethnobotanical Databases. Beltsville, MD: US Department of Agriculture, Agricultural Research Service. http://www.ars-grin.gov/duke/. Accessed 11 Feb 2015.

  • Evershed, R. P. (2008). Organic residue analysis in archaeology: the archaeological biomarker revolution. Archaeometry, 50, 895–924.

    Article  Google Scholar 

  • Gougeon, R. D., Lucio, M., Frommberger, M., Peyron, D., Chassagne, D., Alexandre, H., et al. (2009). The chemodiversity of wines can reveal a metabologeography expression of cooperage oak wood. Proceedings of the National Academy of Sciences of the United States of America, 106, 9174–9179.

    Article  Google Scholar 

  • Guasch-Jané, M. R., Ibern-Gómez, M., Andrés-Lacueva, C., Jáuregui, O., & Lamuela-Raventós, R. M. (2004). Liquid chromatography with mass spectrometry in tandem mode applied for the identification of wine markers in residues from Ancient Egyptian vessels. Analytical Chemistry, 76, 1672–1677.

    Article  Google Scholar 

  • Haldane, C. (1993). Direct evidence for organic cargoes in the Late Bronze Age. World Archaeology, 24, 348–360.

    Article  Google Scholar 

  • Hayes, J. M. (2001). Fractionation of the isotopes of carbon and hydrogen in biosynthetic processes. In J. W. Valley & D. R. Cole (Eds.), Stable isotope geochemistry (pp. 225–278). Reviews in Mineralogy and Geochemistry, vol. 43. Washington, D. C.: Mineralogical Society of America.

  • Hope, C. (1978). Excavations at Malkata and the Birket Habu 1971–1974; V. Jar Sealings and Amphorae of the 18th dynasty: A technological study. Warminster: Aris & Phillips.

    Google Scholar 

  • Hornsey, I. S. (2003). A history of beer and brewing. Cambridge: Royal Society of Chemistry.

    Google Scholar 

  • Hostetter, E., Beck, C. W., & Stewart, D. R. (1994). A bronze situla from Tomb 128, Valle Trebba: chemical evidence for resinated wine at Spina. Studi Etruschi, 59, 211–225.

    Google Scholar 

  • Itan, Y., Powell, A., Beaumont, M. A., Burger, J., & Thomas, M. G. (2009). The origins of lactase persistence in Europe. PLoS Computational Biology, 5, e1000491.

    Article  Google Scholar 

  • Jaillon, O., Aury, J.-M., Noe, B., Policriti, A., Clepet, C., Casagrande, A., et al. (2007). The grapevine genome sequence suggests ancestral hexaploidization in major angiosperm phyla. Nature, 449, 463–467.

    Article  Google Scholar 

  • Katz, S. E. (2012). The art of fermentation: An in-depth exploration of essential concepts and processes from around the world. White River Junction, VT: Chelsea Green.

    Google Scholar 

  • Kirk-Othmer Encyclopedia of Chemical Technology (2004). Hoboken: Wiley-Interscience. http://mrw.interscience.wiley.com/emrw/9780471238966/home/. Accessed 11 Feb 2015.

  • Łuczaj, Ł., & Szymański, W. M. (2007). Wild vascular plants gathered for consumption in the Polish countryside: a review. Journal of Ethnobiology and Ethnomedicine, 3, 17–39.

    Article  Google Scholar 

  • Madej, T., Pirożnikow, E., Dumanowski, J., & Łukasz, Ł. (2014). Juniper beer in Poland: the story of the revival of a traditional beverage. Journal of Ethnobiology, 34, 84–103.

    Article  Google Scholar 

  • McGovern, P. E. (1997). Wine of Egypt’s golden age: an archaeochemical perspective. Journal of Egyptian Archaeology, 83, 69–108.

    Article  Google Scholar 

  • McGovern, P. E. (2003/2007). Ancient wine: The search for the origins of viniculture. Princeton: Princeton University.

  • McGovern, P. E. (2009/2010). Uncorking the past: The quest for wine, beer, and other alcoholic beverages. Berkeley: University of California.

  • McGovern, P. E., Sever, T., Myers, L. J. W., Myers, E. E., Bevan, B., Miller, N. F., et al. (1995). Science in archaeology: a review. American Journal of Archaeology, 99, 79–142.

    Article  Google Scholar 

  • McGovern, P. E., Glusker, D. L., Exner, L. J., & Voigt, M. M. (1996). Neolithic resonated wine. Nature, 381, 480–481.

    Article  Google Scholar 

  • McGovern, P. E., Glusker, D. L., Moreau, R. A., Nuñez, A., Simpson, E., Butrym, E. D., et al. (1999). A funerary feast fit for King Midas. Nature, 402, 863–864.

    Article  Google Scholar 

  • McGovern, P. E., Glusker, D. L., & Exner, L. J. (2001). The organic contents of the Tomb U-j Syro-Palestinian Type Jars: resinated wine flavored with fig. In U. Hartung (Ed.), Umm el-Qaab II: Importkeramik aus dem Friedhof U in Abydos (Umm el-Qaab) und die Beziehungen Ägyptens zu Vorderasien im 4. Jahrtausend v. Chr. (pp. 399–403). Archäologische Veröffentlichungen, 92. Mainz: P. von Zabern.

  • McGovern, P. E., Zhang, J., Tang, J., Zhang, Z., Hall, G. R., Moreau, R. A., et al. (2004). Fermented beverages of pre- and proto-historic China. Proceedings of the National Academy of Sciences of the United States of America, 101, 17593–17598.

    Article  Google Scholar 

  • McGovern, P. E., Glusker, D. L., Exner, L. J., & Hall, G. R. (2007). The chemical identification of resinated wine and a mixed fermented beverage in Bronze Age pottery vessels of Greece. In Y. Tzedakis, H. Martlew, & M. K. Jones (Eds.), Archaeology meets science: Biomolecular investigations in Bronze Age Greece; the primary scientific evidence 1997–2003 (pp. 169–218). Oxford: Oxbow.

    Google Scholar 

  • McGovern, P. E., Mirzoian, A., & Hall, G. R. (2009). Ancient Egyptian herbal wines. Proceedings of the National Academy of Sciences of the United States of America, 106, 7361–7366.

    Article  Google Scholar 

  • McGovern, P. E., Christofidou-Solomidou, M., Wang, W., Dukes, W. F., Davidson, T., & El-Deiry, W. S. (2010). Anticancer activity of botanical compounds in ancient fermented beverages. International Journal of Oncology, 37, 5–21.

    Article  Google Scholar 

  • McGovern, P. E., Hall, G. R., & Mirzoian, A. (2013a). A biomolecular archaeological approach to “Nordic grog.” (pp. 112–131). Danish Journal of Archaeology, Journal: Danish Journal of Archaeology.

    Google Scholar 

  • McGovern, P. E., Luley, B. P., Rovira, N., Mirzoian, A., Callahan, M. P., Smith, K. E., et al. (2013b). The beginning of viniculture in France. Proceedings of the National Academy of Sciences of the United States of America, 110(25), 10147–10152.

    Article  Google Scholar 

  • Murray, M. A. (2000). Viticulture and wine production. In P. T. Nicholson & I. Shaw (Eds.), Ancient Egyptian materials and technology (pp. 577–608). Cambridge: Cambridge University.

    Google Scholar 

  • Myles, S., Boyko, A. R., Owens, C. L., Brown, P. J., Grassi, F., Aradhya, M. K., et al. (2011). Genetic structure and domestication history of the grape. Proceedings of the National Academy of Sciences of the United States of America, 108, 3530–3535.

    Article  Google Scholar 

  • Namdar, D., Neumann, R., Goren, Y., & Weiner, S. (2009). The contents of unusual cone-shaped vessels (cornets) from the Chalcolithic of the southern Levant. Journal of Archaeological Science, 36, 629–636.

    Article  Google Scholar 

  • Nelson, M. (2005). The Barbarian’s beverage: A history of beer in ancient Europe. London: Routledge.

    Google Scholar 

  • Nigra, B. T., Faull, K. F., & Barnard, H. (2015). Analytical chemistry in archaeological research. Analytical Chemistry, 87, 3–18.

    Article  Google Scholar 

  • Pulak, C. (1988). The Bronze Age shipwreck at Ulu Burun: 1985 campaign. American Journal of Archaeology, 92, 1–37.

    Article  Google Scholar 

  • Pulak, C. (1998). The Uluburun shipwreck: an overview. International Journal of Nautical Archaeology, 27, 188–224.

    Article  Google Scholar 

  • Pulak, C. (2005). Discovering a royal ship from the age of King Tut: Uluburun, Turkey. In G. F. Bass (Ed.), Beneath the seven seas: Adventures with the Institute of Nautical Archaeology (pp. 34–47). London: Thames & Hudson.

    Google Scholar 

  • Pulak, C. (2008). The Uluburun shipwreck and Late Bronze Age trade. In J. Aruz, K. Benzel, & J. M. Evans (Eds.), Beyond Babylon: Art, trade, and diplomacy in the second millennium B.C (pp. 289–385). New York: Metropolitan Museum of Art.

    Google Scholar 

  • Regert, M. (2007). Elucidating pottery function using a multi-step analytical methodology combining infrared spectroscopy, mass spectrometry and chromatographic procedures. In J. W. Eerkens & H. Barnard (Eds.), Theory and practice of archaeological residue analysis (pp. 61–76). British Archaeological Reports International Series 1650. Oxford: Archaeopress.

  • Ribechini, E., Colombini, M. P., Giachi, G., Modugno, F., & Pallecchi, P. (2009a). A multi-analytical approach for the characterization of commodities in a ceramic jar from Antinoe (Egypt). Archaeometry, 51, 480–494.

    Article  Google Scholar 

  • Ribechini, E., Orsini, S., Silvano, F., & Colombini, M. P. (2009b). Py-GC/MS, GC/MS and FTIR investigations on Late Roman-Egyptian adhesives from opus sectile: new insights into ancient recipes and technologies. Analytica Chimica Acta, 638, 79–87.

    Article  Google Scholar 

  • Ridgway, D. (1997). Nestor’s cup and the Etruscans. Oxford Journal of Archaeology, 16, 325–344.

    Article  Google Scholar 

  • Salque, M., Bogucki, P. I., Pyzel, J., Sobkowiak-Tabaka, I., Grygiel, R., Szmyt, M., et al. (2013). Earliest evidence for cheese making in the sixth millennium B.C. in northern Europe. Nature, 493, 522–525.

    Article  Google Scholar 

  • Singleton, V. L. (1995). An enologist’s commentary on ancient wines. In P. E. McGovern, S. J. Fleming, & S. H. Katz (Eds.), The origins and ancient history of wine (pp. 67–77). New York: Gordon and Breach.

    Google Scholar 

  • Stern, B., Heron, C., Tellefsen, T., & Serpico, M. (2008). New investigations into the Uluburun resin cargo. Journal of Archaeological Science, 35, 2188–2203.

    Article  Google Scholar 

  • Unger, R. W. (2007). Beer in the Middle Ages and the Renaissance. Philadelphia: University of Pennsylvania.

    Google Scholar 

  • Vouillamoz, J. F., McGovern, P. E., Ergul, A., Söylemezoğlu, G., Tevzadze, G., & Grando, M. S. (2006). Genetic characterization and relationships of traditional grape cultivars from Transcaucasia and Anatolia. Plant Genetic Resources: Characterization & Utilization, 4, 144–518.

    Article  Google Scholar 

  • Ward, C. (2004). Plant remains. In G. F. Bass, S. Matthews, J. R. Steffy, & F. H. van Doorninck Jr. (Eds.), Serçe Limani: An eleventh-century shipwreck; the ship and its anchorage, crew, and passengers (pp. 495–511). College Station, TX: Texas A&M University.

    Google Scholar 

  • Warinner, C., Hendy J., Speller C., Cappellini E., Fischer R., Trachsel C., et al. (2014). Direct evidence of milk consumption from ancient human dental calculus. Scientific Reports, 4, article 7104. http://www.nature.com/srep/2014/141127/srep07104/full/srep07104.html. Accessed 11 Feb 2015.

  • Wild, H. (1966). Brasserie et panification au tombeau de Ti. Bulletin de l’Institut français d’archéologie orientale, 64, 95–120.

    Google Scholar 

  • Yang, Y., Shevchenkoc, A., Knaustc, A., Abuduresuled, I., Lid, W., Hud, X., & Wang, C. (2014). Proteomics evidence for Kefir dairy in early Bronze Age China. Journal of Archaeological Science, 45, 178–186.

    Article  Google Scholar 

Download references

Acknowledgments

We are especially grateful to Armen Mirzoian of the Beverage Alcohol Laboratory in the Scientific Services Division of the Alcohol and Tobacco Tax and Trade Bureau (TTB) for his LC-MS-MS analyses of the Uluburun samples and for method and extraction development targeting tartaric acid/tartrate. Significant support for the study came from Abdul Mabud of TTB’s Scientific Services Division.

W. Christian Petersen assisted in the GC-MS analyses at the Scientific Research and Analysis Laboratory of Winterthur Museum and Country Estate. He, together with Theodore Davidson and Joshua M. Henkin of the Penn Museum’s Biomolecular Archaeology Laboratory and Michael Gregg of Mount Allison University in New Brunswick, consulted on chemical and natural product matters. Rosa M. Lamuela-Raventós and Olga Jáuregui advised on LC-MS-MS analysis of grape/wine. Michael P. Callahan of the Solar System Exploration Division of Goddard Space Flight Center (NASA) reviewed the edited version of the paper for its scientific rationale. Łuczaj Łkasz of the High School of Humanities and Economics in Łódz and Peter Bogucki of Princeton University provided information on the natural products and archaeology of Poland. Rose Marie Belforti of the Finger Lakes Dexter Creamery conferred with us on kefir and other fermented milk products. We are grateful to the excavators and researchers of the Uluburun ship, especially Cemal Pulak and Cheryl Ward.

Curt W. Beck, former director of the Amber Research Laboratory at Vassar College and now sadly deceased, kindly provided the Uluburun terebinth resin samples via Cemal Pulak. The laboratory’s current director, Edith C. Stout, also made available a reference sample of modern Pistacia sp. (collected by Sabine Beckmann on Chios).

Figures 1, 4, 5, and 6 were prepared by Shih-han Samuel Lin, Jeffrey Hoyt, Anne Bomalaski, and Kimberly Leaman.

Conflict of Interest

The authors declare no conflicts of interest.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Patrick E. McGovern.

Appendices

Appendix 1: LC-MS-MS Analyses

In reanalyzing the ancient Egyptian sample from Scorpion I’s tomb at Abydos, we followed the general methodology of Guasch-Jané et al. (2004), with modification of the extraction procedure to detect tartaric acid/tartrate in the negative mode (McGovern et al. 2009).

An even more refined extraction procedure was found to enhance the detection of tartaric acid/tartrate by LC-MS-MS for the Uluburun terebinth resin samples. Whole pieces of resin were stirred overnight in 1.5 mL of ammonium hydroxide (2.8 % by vol) and 1.5 mL methylene chloride. The mixture was centrifuged for at least 10 min at 4400 rpm to clarify the layers and cause any remaining materials and emulsions to precipitate. The upper basic aqueous layer was then removed, reduced in volume, filtered through a 0.45-μm membrane, placed into 200 μL vial inserts, and analyzed.

Two MRM transitions (149→87.1 and 149→73), not just one as we had previously done, at a specific retention time (0.80 min) were monitored for tartaric acid, as can be seen on Fig. 5. Together, these transitions and the resulting daughter ions are definitive for the acid’s presence/absence. Monitoring both transitions has not been applied to ancient samples before and provides very strong evidence for tartaric acid/tartrate in KW 181, KW 215, and probably KW 102.

The LC mobile phase conditions were also modified to achieve better peak shapes, as follows: A (water, 0.1 % v/v formic acid) and B (acetonitrile, 0.1 % v/v formic acid) had an initial composition of 98 % phase A and 2 % B, which was changed to 50 % A/50 % B over a 2-min period, and then equilibrated back to 98 % A/2 % B.

Accentuating the detection of a tartrate salt by this new method takes advantage of the much lower solubilities of potassium bitartrate and calcium tartrate, about 4 and 0.3 g/L in cold water, respectively, compared to the acid’s solubility of about 1400 g/L at 20 °C (Singleton 1995: 68). The salts readily form in wine and precipitate out in the lees. Calcium can interchange with potassium in highly calcareous pottery fabrics, which are characteristic of Eastern Mediterranean amphoras.

GC-MS was employed to further test KW 215, which had given a positive tartaric acid/tartrate by LC-MS-MS. The resin sample was completely dissolved in methylene chloride and treated with a small amount of formic acid to acidify any tartrate present to tartaric acid. Silylation was carried out with BSTFA (N,O-bis(trimethyl-silyl)trifluoro acetamide). The sample was injected splitless onto an HP-5MS column (5 % phenyl methyl siloxane) of an Agilent HP-6890 gas chromatograph equipped with a Hewlett-Packard 5973 mass selective detector. The key silylated tartaric acid ion at 219+ was tentatively identified by selected ion monitoring which enhances sensitivity.

Both LC-MS-MS and GC-MS yielded an approximately 0.25-ppm concentration of the acid/salt for sample KW 215, which was close to the GC-MS detection limit and only achievable in the selected ion mode.

Appendix 2: FT-IR Analyses

Although potentially many absorptions in the standard FT-IR region of 4000–400 cm−1 can be equivocal, careful observation, combined with rigorous methodology, enables the critical peaks for individual compounds in mixtures to be sorted out. Moreover, fine spectral details can be distinguished by extracting an ancient sample separately with methanol and other organic solvents (e.g., chloroform). While chloroform selectively extracts and accentuates the FT-IR spectra of low-polarity compounds, methanol highlights those of high-polarity compounds.

Figure 6 and Table 2 illustrate how powerful a combined solvent approach can be in distinguishing the carbonyl and acid hydroxyl absorptions of synthetic L-tartaric acid from those of ancient and modern Pistacia sp. resin. The higher-polarity tartaric acid, which was extracted by methanol, has a distinctive doublet in the 1740–1720 cm−1 carbonyl region, with a less intense shoulder at the lower wave number (frequency). Its hydroxyl absorption occurs in the 1450–1430 cm−1 region. By contrast, the carbonyl of the lower polarity terebinth resinous acids, which were extracted by chloroform, have a single intense absorption at 1710–1700 cm−1, and their hydroxyl absorption is in the 1470–1445 cm−1 region.

Fig. 6
figure 6

The FT-IR analysis of L-tartaric acid is clearly different at key carbonyl stretch and hydroxyl bend absorptions from those of ancient (KW 215) and modern Pistacia sp

Table 2 Comparison of key FT-IR absorption of methanol and chloroform extracts of L-tartaric acid, tartrate, and Pistacia sp. resin

In light of their marked spectral differences, the discussion of the FT-IR spectra of tartaric acid and Pistacia sp. resin by Stern et al. (2008) (pp. 2197–2198) is confusing. They state that their Uluburun resin samples all have absorptions in the 1740–1720 cm−1 region, similar to those of tartaric acid. They then dismiss the FT-IR evidence. Yet, their own Fig. 4 shows that the carbonyl peak (viz., 1710–1700 cm−1) of the Uluburun resin samples is significantly lower and does not overlap with that of tartaric acid.

The FT-IR spectrum for syringic acid also differs from that of tartaric acid and Pistacia sp. resin, as Stern et al. (2008) show in their Fig. 2. Its carbonyl stretch is a single intense absorption at 1690 cm−1, and its OH bend is at 1450 cm−1.

The blanket statement by Stern et al. (2008) (p. 2197) that “these absorption bands are shared by a huge variety of organic substances and are clearly not restricted [added emphasis] to tartaric acid/tartrate or terebinth resin” is far from being clearly established and undermines the value of infrared spectrometry. For the spectrometrist, including Moreira and Santos 2004 which Stern et al. cite in support of their contention, the hydroxyl and carbonyl absorption bands are important discriminators of organic compounds. Certainly, Moreira and Santos would not confuse the spectrum of syringic acid with that of tartaric acid, as implied by Stern et al. (2008) (p. 2195). Stern et al. also assert that interferences from ethanol and acetic acid pose a problem in interpreting the FT-IR spectra of wine, again citing Moreira and Santos 2004. Since such compounds have volatilized and disappeared in ancient wine samples, their point is not relevant.

In summary, it is very important to carefully scrutinize and statistically deconvolute an IR spectrum, and also to run searches of unknown archaeological samples against the large databases which are now available. Stern et al. do not appear to have exercised such care in their analyses, and thus they write off the FT-IR data too quickly.

Three subsidiary points should also be noted:

  1. 1.

    Even if it were possible to confuse the IR spectra of L-tartaric acid and syringic acid (Stern et al. 2008: 2195), this is irrelevant to the analysis of an ancient wine sample. Syringic acid is not a “derivative” of malvidin (p. 2189), one of the red pigments in wine, but can only be produced as a breakdown product of the latter by alkaline fusion, which was beyond the expertise of ancient peoples. In fact, it is malvidin, not syringic acid, which is present in modern red wine at 200 mg/L, according to Singleton (1995: 70) (see below).

  2. 2.

    It is very puzzling that the modern red wine, analyzed by Stern et al. (2008) (p. 2196, Fig. 3), is dominated by calcium tartrate. Tartaric acid, at a concentration of about 1400 g/L at 20 °C, is the principal acid in wine and should be detected. Moreover, any tartrate that has precipitated out should be in the form of potassium bitartrate, not calcium tartrate. Again, Stern et al. do not detail their extraction procedure, if any, which might explain these anomalies. Their explanation for the origin of calcium tartarte in ancient samples is equally suspect.

  3. 3.

    Stern et al. (2008) do not appear to have included “ancient reference samples” in their databases, comparable to the inscribed Malkata amphoras of our studies. If they had, they might have been able to detect differences in the FT-IR spectra of “aged” and modern wine samples.

    Together and apart from the lack of care in the interpretation of the FT-IR spectra, the subsidiary points highlight the confusing, ill-founded argumentation of Stern et al. (2008).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

McGovern, P.E., Hall, G.R. Charting a Future Course for Organic Residue Analysis in Archaeology. J Archaeol Method Theory 23, 592–622 (2016). https://doi.org/10.1007/s10816-015-9253-z

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10816-015-9253-z

Keywords

Navigation