Вы находитесь на странице: 1из 15

Journal of Archaeological Science 37 (2010) 971985

Contents lists available at ScienceDirect

Journal of Archaeological Science


journal homepage: http://www.elsevier.com/locate/jas

Remains of the day-preservation of organic micro-residues on stone tools


Geeske H.J. Langejans
Institute for Human Evolution and School of Geography, Archaeology and Environmental Studies and Hazenberg Archeologie, University of the Witwatersrand,
Private Bag 3 PO Box, WITS 2050, South Africa

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 22 July 2009
Received in revised form
5 November 2009
Accepted 21 November 2009

Here I report on the decay processes of microscopic organic residues left on stone tool surfaces after their
use. Residue analysis on ancient stone tools facilitates reconstruction of past activities. This study enables
predictions about the circumstances under which ancient residues preserve. Experimental tool sets with
modern residues were buried for a year in separate deposits at Sterkfontein, Sibudu (South Africa) and
Zelhem (the Netherlands) whose pH and geomorphology varied, they were then analysed using light
microscopy. Biological weathering mainly causes residue decay. In unstable environments rich in
microbes and micro-organisms, residues decay quickly. From an archaeological perspective this means
that sites that are stable, desiccated, waterlogged, extremely acidic or alkaline and extremely cold or hot
sites. Different residue types have different preservation optima and this may lead to a preservation and
perhaps interpretation bias. The preliminary predictive models presented in this paper could aid in the
considered selection of sites and samples.
2009 Elsevier Ltd. All rights reserved.

Keywords:
Residue analysis
Micro-remains
Preservation
Conservation
Decay
Taphonomy
Stone tools
Experimental archaeology
Replication

1. Introduction
It has been argued that researchers are unable to describe the
mechanism of residue decay (Grace, 1996; Odell, 2001). This paper
aims to investigate some of these taphonomic issues. Grace
observes that residue analysts . simply record a phenomenon
(.) and then make functional interpretations by analogy without
having explained the process that gave rise to the phenomenon.
(Grace, 1996, 214). Odell argues that residue analysts are inclined
to consider what they see, not what is missing (Odell, 2001, 63).
Although these issues are acknowledged and in some cases
examined (e.g. Barton, 2009; Barton et al., 1998; Fullagar, 1988,
1998; Haslam, 2004; Jones, 2009; Lombard and Wadley, 2007; Lu,
2006; Rots and Williamson, 2004; Wadley and Lombard, 2007),
they remain essentially unresolved. Potentially the results from
residue studies may be oversimplied, or worse inaccurate.
Stone tool micro-residue analysis aims to identify microscopic
remains or traces that are left on a tools surface after use. Analysts
have identied organic plant and animal remains and inorganic
deposits. Residue analysis is generally conducted with a direct or
indirect light microscope, using a range of magnications (between
50 and 800 times), an experimental comparative collection and
sometimes an ethno-archaeological comparative collection

E-mail address: geeske.langejans@wits.ac.za


0305-4403/$ see front matter
doi:10.1016/j.jas.2009.11.030

2009 Elsevier Ltd. All rights reserved.

(Anderson, 1980; Bruier, 1976; Fullagar et al., 1996, 1999; Loy, 1997;
Miller, 1979; Rots and Williamson, 2004; Williamson, 2000a,b).
The identication of processed materials is of relevance to
a wide range of research questions (Bruier, 1976; DominguezRodrigo et al., 2001; Fullagar, 2006; Fullagar and Jones, 2004; Hardy
et al., 2008; Horrocks et al., 2007; Hurst et al., 2002; Lombard,
2005; Piperno et al., 2004; Staller and Thompson, 2002). Residues
themselves may also serve as a source for further investigation, for
example DNA analysis (e.g. Hardy et al., 1997; Nelson et al., 1986;
Williamson, 2000b) and they can be used to study rare artefacts for
additional information (e.g. Barton, 2007; Loy, 1998; Loy and Dixon,
1998).
As yet, it is unclear to which extent residues on tools are the
result of use or of taphonomic processes. As residues are expected
to be inuenced by taphonomy, direct analogies between use and
remains are problematic (Grace, 1996; Haslam, 2006; Odell, 2001),
especially when quantitative spatial analyses are not conducted on
representative samples (e.g. Lombard, 2005, 2007; Wadley and
Lombard, 2007). When analysts are able to describe the mechanism
of residue preservation and predict under which circumstances
(what) residues preserve, residue analysis has great potential.
There have been many investigations into the molecular decay
of blood and fat (e.g. Cattaneo et al., 1993, 1991; Eisele et al., 1995;
Evershed, 2008; Evershed et al., 2001; Evershed and Tuross, 1996;
Gurnkel and Franklin, 1988; Kooyman et al., 1992; Tuross et al.,
1996). The decay and taphonomy of blood and starch has been

982

G.H.J. Langejans / Journal of Archaeological Science 37 (2010) 971985

residues is needed. Preservation may be inuenced by the season of


deposition. For example, if a tool with residues was deposited
during the dry season at Sibudu it would quickly become desiccated
and thus protected from biological decay. If a tool was deposited
during the wet season, the residues would have taken longer to dry
out and they would have been prone to deterioration during this
time. Rapid burial or prolonged exposure could also inuence
preservation (also see Barton, 2009). Sedimentation rates can also
be dependent on seasonality.
. This means that in a conserving environment residues are
equally well preserved in the four-week and one-year experiments and that if circumstances remain stable residues might
preserve for long periods of time.
The rst part of the prediction is falsied. Even under the benign
circumstances in the caves, residues decay. The second part of the
predictions appears true.
After one-year, few or no residues preserved outside the caves at
Sibudu and Sterkfontein. The equation in Fig. 14 predicts that a
numbers of A (the residue) react (this may be a chemical reaction or
a biological reaction) with b numbers of B (water/oxygen) to
create c numbers of C (by-product) and d numbers of D (carbon
dioxide). At equilibrium as much A and B changes in C and D and
vice versa. In the beginning only A and B are present and the
reaction is fast, but as A and B and are consumed the reaction slows
down (Raiswell, 2001). Therefore, residues deteriorate quickly
during initial deposition and later, when equilibrium is reached, the
amount of residue present stabilises (also see Barton et al., 1998).
However, with biological decay of residues, equilibrium may
never be reached because in most environments, the by-product
carbon dioxide is lost as a gas (Cronyn, 1990; Raiswell, 2001;
Watkinson, 2001) and the organic residues are eventually
completely consumed. When just deposited decay is fast because
there are many residues and substrate for microbes. As the residues
diminish there is less foodstuff for micro-organisms. The microbes
probably diminish in number as a result (also see Barton, 2009).
This appears as stabilisation of the decay process or equilibrium,
but it is the result of fewer residues that are available for decay.
Decay may thus follow a more exponential trend than a linear one.
Biological decay can be extremely slow due to inhibiting factors,
such as low temperatures. Slow decay may be mistaken for
permanent preservation. However, chemically and sometimes
visually such residues do undergo change. Slow decay might take
tens of thousands of years, provided that the burial environment
remains stable. Changes in the environment can cause these
inhibiting factors to change and rapid decay could set in. Once
excavated, archaeological tools are at danger of losing their
potential residues.
Chemical decay could reach equilibrium, but only when the
secondary products (C and D in the equation) remain in the environment. In permeable sediments, for example, C and D leach out
and decay continues until all residues are consumed. Therefore, in
an unstable environment residues are more likely to deteriorate
rapidly.
The coarse-grained slides at Zelhem are likely to preserve residues better than ne- and medium-grained ones. The irregular
surface of coarse-grained stones provides micro-relief in which
residues may be protected.
This prediction is incorrect. In the Zelhem fat experiments, the
ne and medium grained sandpaper preserved residue best. The
sandpaper slide experiments also demonstrate that residues
preserve badly on glass-type surfaces and this has implications for
the study of artefacts made on quartz and other glass-like surfaces.
Because the surface of the ne-and medium-grained paper is larger

than the surface of the course-grained sandpaper, residues are


probably better preserved on the rst. A large, irregular surface is
best for residue preservation. It is possible that residue preservation is inuenced by the pH of the raw material. Porous stone may
also preserve residues by absorbing them. Residues, however, may
become unidentiable (stains) as a result. More experimentation is
needed to investigate this.
Some organic materials are hardier than others.
Some residue types will preserve better than others. After oneyear plant tissue and bres preserved better than any of the other
experimental residues outside the caves. The cell walls of plant
tissue do not readily decompose and plant tissue contains sugars
that decay in a more complicated way than those of starch. When
examining the one-year experiments outside the caves, starch was
vulnerable when unprotected and did often not preserve. Starch
has no cell walls and is easy and rewarding for micro-organisms to
break down. They are, however, abundant in plant material and
their sheer quantity might explain their presence in the archaeological record. Additionally, the dry circumstances inside Sibudu
may have ensured excellent preservation. After one year muscle
tissue preserved poorly outside the caves. Animal cells do not have
cell walls and they are easy to break down. Blood also appears to be
fragile. After a year muscle tissue preserved better inside than
outside the caves. Bone consists of more mineral components and is
more complex to decompose than other animal components.
Although bone preserved well after one-year, particularly in the dry
cave settings, bone working leaves few residues (personal observation and also see Hardy et al., 2001), decreasing the residues
survival chances. Fat it is a complex molecule and difcult for
microbes to break down. Therefore, good fat preservation is
generally expected, but because fat can get absorbed in the rock and
in the sediment, it may not be observed during analysis.
In general, the experimental residues preserved best in the dry
settings of caves or rock shelters, which probably inhibit (biological) microbial decay. PH, temperature and burial appear to play
a less important role. A protective micro-environment may have
preserved otherwise fragile residues from mechanical and perhaps
chemical and biological decay. It is no surprise that the experiments
conrm that micro-remains decay similarly to macro-remains.
Although some residue types are more resilient than others, the
environment can also lead to differential preservation. Different
residues have different preservation optima and processed materials can potentially be misrepresented as some residue types may
be missing. To avoid misinterpretation, the preservation conditions
per residue type should be carefully charted. Fig. 15 is a rst attempt
to do so. It summarises the variables inuencing preservation and
predictions for four main residue types; it is based on the literature
review and the results from the experiments.
7. Conclusion
In the Introduction I discussed a research lacuna in residue
analysis (Grace, 1996; Odell, 2001): residue specialists make
a direct analogy between the residues they observe and materials
that were processed. Consequently, there is a danger that the
results of residue studies are oversimplied. Thus, to enhance
interpretative frameworks for past human activity and to identify
potential methodological shortcomings, it is necessary to gain
understanding of the relevant (post-) depositional processes.
The residue decay mechanism appears to be mainly driven by
biological decomposition. Tools from stable sites with no, or low,
bioactivity are best for residue analysis. In practice this excludes
most open-air sites. Dry and desiccated sites, waterlogged,
extremely acidic or alkaline sites, a close proximity to heavy metals

G.H.J. Langejans / Journal of Archaeological Science 37 (2010) 971985

and cave sites potentially preserve residues best. Different residues


preserve differentially. Of the tested materials, blood, muscle tissue
and starch are fragile and bone, fat and woody plant tissue are more
durable. Residues always undergo decomposition, but with slow
decay residues give the appearance of being permanently
preserved. Slow decay can take tens of thousands of years, provided
that the environment remains stable.
On the basis of these results it is possible to make predictions
about residue preservation at archaeological sites. These predictions may be used to select or exclude sites from analysis and to
interpret the identied use-related remains in terms of differential
preservation. It also suggests caution regarding the decontextualised interpretation of residues observed on archaeological tools.
Unfortunately the taphonomy of many sites precludes the preservation of ancient residues, but when sites and artefacts are carefully
selected, based on feasibility studies, residue analysis can be
a valuable tool for the reconstruction of processed materials and
past activities.
Acknowledgement
This research was sponsored by the National Research Foundation (NRF) and the Palaeontological Scientic Trust (PAST), South
Africa. Special thanks go to Lyn Wadley for her continued advice,
support and commenting on drafts of this document and to
Christopher Henshilwood for sponsoring my post-doc at the
University of the Witwatersrand. Gerrit Dusseldorp, Tom Hazenberg, Marco Langbroek, Marlize Lombard, Annelou van Gijn and
Kartsen Wentink helped with the experiments and logistics, or
provided advice, guidance and comments on drafts of this document; their contributions are much appreciated.
References
Adu, J.K., Oades, J.M., 1978a. Physical factors inuencing decomposition of organic
materials in soil aggregates. Soil Biology and Biochemistry 10, 109115.
Adu, J.K., Oades, J.M., 1978b. Utilization of organic materials in soil aggregates by
bacteria and fungi. Soil Biology and Biochemistry 10, 117122.
AGIS, 2007. Agricultural Geo-Referenced Information System: Gauteng Soil Survey
2009. http://www.agis.agric.za/agisweb/gauteng_soil_survey.
Anderson, P., 1980. Testimony of prehistoric tasks: diagnostic residues on stone tool
working edges. World Archaeology 12, 181194.
Babot, M.D.P., Apella, M.C., 2003. Maize and bone: residues of grinding in northwestern Argentina. Archaeometry 45, 121132.
Barton, H., 2007. Starch residues on museum artefacts: implications for determining
tool use. Journal of Archaeological Science 34, 17521762.
Barton, H., 2009. Starch granule taphonomy: the results of a two year eld
experiment. In: Haslam, M., Robertson, G., Crowther, A., Nugent, S.,
Kirkwood, L. (Eds.), Archaeological Science under a Microscope. ANUE Press,
Canberra, pp. 129140.
Barton, H., Matthews, P.J., 2006. Taphonomy. In: Torrence, R., Barton, H. (Eds.),
Ancient Starch Research. Left Coast Press, Walnut Creek, pp. 7594.
Barton, H., Torrence, R., Fullagar, R., 1998. Clues to stone tool function re-examined:
comparing starch grain frequencies on used and unused obsidian artefacts.
Journal of Archaeological Science 25, 12311238.
Bell, L.S., Boye, A., Jones, J.S., 1991. Diagenic alteration to teeth in situ illustrated by
backscattered electron imaging. Scanning 13, 173183.
Berendsen, H.J.A., 1997. Landschappelijk Nederland. Fysische Geograe van Nederland, Van Gorcum, Assen.
Blanchette, R.A., 2000. A review of microbial deterioration found in archaeological
wood from different environments. International Biodeterioration & Biodegradation 46, 189204.
Blanchette, R.A., Obst, J.R., Timell, T.E., 1994. Biodegradation of compression wood
and tension wood by white and brown rot fungi. Holzforschung 48, 3443.
Bocherens, H., Tresset, A., Wiedemann, F., Giligny, F., Lafage, F., Lanchon, Y.,
Mariotti, A., 1997. Diagenetic evolution of mammal bones in two French
Neolithic sites. Bulletin de la Societe Geologique de France 168, 555564.
Bruier, F.L., 1976. New clues to stone tool function; plant and animal residues.
American Antiquity 41, 478484.
Bull, I.D., Simpson, I.A., Dockrill, S.J., Evershed, R.P., 1999. Organic geochemical
evidence for the origin of ancient anthropogenic soil deposits at Tofts Ness,
Sanday, Orkney. Organic Geochemistry 30, 535556.
Carter, D.O., Tibbett, M., 2005. Forensic taphonomy: the adaptation of the soil
microbial decomposer community to soft tissue burial. In: Kars, H., Burke, E.

983

(Eds.), Proceedings of the 33rd International Symposium on Archaeometry, 22


26 April 2002, Amsterdam. Vrije Universiteit, Amsterdam, pp. 453455.
Cattaneo, C., Gelsthorpe, K., Phillips, P., Sokol, R.J., 1993. Blood residues on stone
tools: indoor and outdoor experiments. World Archaeology 25, 2943.
Cattaneo, C., Gelsthorpe, K., Phillips, P., Sokol, R.J., Smillie, D., 1991. Identication of
ancient blood and tissue ELISA and DNA analysis. Antiquity 65, 878881.
Cheshire, M.V., Mundie, C.M., Shepherd, H., 1969. Transformation of 14C glucose and
starch in soil. Soil Biology and Biochemistry 1, 117130.
Chesworth, W., 1992. Weathering systems. In: Martini, I.P., Chesworth, W. (Eds.),
Weathering, Soils, and Paleosols. Elsevier, New York, pp. 1940.
Child, A.M., 1995. Towards an understanding of the microbial decomposition of
archaeological bone in the burial environment. Journal of Archaeological
Science 22, 165174.
Clausen, C.A., 1996. Bacterial associations with decaying wood: a review. International Biodeterioration & Biodegradation 37, 101107.
Cronyn, J.M., 1990. The Elements of Archaeological Conservation. Routledge,
London.
Cronyn, J.M., 2001. The deterioration of organic materials. In: Brothwell, D.R.,
Pollard, A.M. (Eds.), Handbook of Archaeological Sciences. John Wiley & Sons,
Chichester, pp. 627636.
Crow, P., 2008. Mineral weathering in forest soils and its relevance to the preservation of the buried archaeological resource. Journal of Archaeological Science
35, 22622273.
Deacon, J.W., 1997. Modern Mycology, third ed. Blackwell Science, Oxford.
Dominguez- Rodrigo, M., Serrallonga, J., Juan-Tresserras, J., Alcala, L., Luque, L., 2001.
Woodworking activities by early humans: a plant residue analysis on Acheulian
stone tools from Peninj (Tanzania). Journal of Human Evolution 40, 289299.
Eglinton, G., Logan, G.A., 1991. Molecular preservation. Philosophical Transactions:
Biological Sciences 333, 315328.
Eisele, J.A., Fowler, D.D., Haynes, G., Lewis, R.A., 1995. Survival and detection of
blood residues on stone tools. Antiquity 69, p. 36(11).
Evershed, R.P., 1992. Chemical composition of a bog body adipocere. Archaeometry
34, 253265.
Evershed, R.P., 2008. Experimental approaches to the interpretation of absorbed
organic residues in archaeological ceramics. World Archaeology 40, 2647.
Evershed, R.P., Dudd, S.N., Lockheart, M.J., Jim, S., 2001. Lipids in archaeology. In:
Brothwell, D.R., Pollard, A.M. (Eds.), Handbook of Archaeological Sciences. John
Wiley and Sons, Chichester, pp. 331349.
Evershed, R.P., Turner-Walker, G., Hedges, R.E.M., Tuross, N., Leyden, A., 1995.
Preliminary results for the analysis of lipids in ancient bone. Journal of
Archaeological Science 22, 277290.
Evershed, R.P., Tuross, N., 1996. Proteinaceous material from potsherds and
associated soils. Journal of Archaeological Science 23, 429436.
Forbes, S.L., Dent, B.B., Stuart, B.H., 2005a. The effect of soil type on adipocere
formation. Forensic Science International 154, 3543.
Forbes, S.L., Dent, B.B., Stuart, B.H., 2005b. The effect of the burial environment on
adipocere formation. Forensic Sience International 154, 2434.
Forbes, S.L., Stuart, B.H., Dent, B.B., 2005c. The effect of the method of burial on
adipocere formation. Forensic Science International 154, 4452.
Fullagar, R., 1988. Recent developments in Australian use-wear and residue studies.
In: Beyries, S. (Ed.), Industries Lithiques, Traceologie et Technologie. Archaeopress, Oxford, pp. 133145.
Fullagar, R. (Ed.), 1998. A Closer Look: Recent Australian Studies of Stone Tools.
University of Sydney, Sydney.
Fullagar, R., 2006. Residues and usewear. In: Balme, J., Paterson, A. (Eds.), Archaeology in Practice; A Student Guide to Archaeological Analyses. Blackwell
Publishing, Oxford, pp. 207233.
Fullagar, R., Field, J., Denham, T., Lentfer, C., 2006. Early and mid Holocene tool-use
and processing of taro (Colocasia esculenta), yam (Dioscorea sp.) and other
plants at Kuk Swamp in the highlands of Papua New Guinea. Journal of
Archaeological Science 33, 595614.
Fullagar, R., Furby, J., Hardy, B.L., 1996. Residues on stone artefacts: state of
a scientic art. Antiquity 70, 740745.
Fullagar, R., Jones, R., 2004. Usewear and residue analysis of stone artefacts from the
Enclosed Chamber, Rocky Cape, Tasmania. Archaeology in Oceania 39, 7993.
Fullagar, R., Meehan, B., Jones, R., 1999. Residue analysis of ethnographic plantworking and other tools from northern Australia. In: Anderson, P. (Ed.),
Prehistory of Agriculture: New Experimental and Ethnographic Approaches.
Institute of Archaeology, UCLA, Los Angeles, pp. 1523.
Geiger, G., Livingston, M.P., Funk, F., Schulin, R., 1998. b-glucosidase activity in the
presence of copper and goethite. European Journal of Soil Science 49, 1723.
Gernaey, A.M., Waite, E.R., Collins, M.J., Craig, O.E., Sokol, R.J., 2001. Survival and
interpretation of archaeological proteins. In: Brothwell, D.R., Pollard, A.M.
(Eds.), Handbook of Archaeological Sciences. John Wiley & Sons, Chichester, pp.
323329.
Goldberg, P., Macphail, R.I., 2006. Practical and Theoretical Geoarchaeology. Blackwell Publishing, Oxford.
Gott, B., Barton, H., Samuel, D., Torrence, R., 2006. Biology of starch. In:
Torrence, R., Barton, H. (Eds.), Ancient Starch Research. Left Coast Press, Walnut
Creek, pp. 3545.
Grace, R., 1996. Use-wear analysis: the state of the art. Archaeometry 38, 209229.
Greaves, H., 1969. Micromorphology of the bacterial attack of wood. Wood Science
and Technology 3, 150166.
Grupe, G., 1995. Preservation of collagen in bone from dry, sandy soil. Journal of
Archaeological Science 22, 193199.

984

G.H.J. Langejans / Journal of Archaeological Science 37 (2010) 971985

Gurnkel, D.M., Franklin, U.M., 1988. A study of the feasibility of detecting blood
residue on artifacts. Journal of Archaeological Science 15, 8397.
Hackett, C.J., 1981. Microscopical focal destruction (tunnels) in exhumated human
bones. Medicine Science and Law 21, 243265.
Hardy, B.L., 2004. Neanderthal behaviour and stone tool function at the Middle
Palaeolithic site of La Quina, France. Antiquity 78, 547565.
Hardy, B.L., Bolus, M., Conard, N.J., 2008. Hammer or crescent wrench? Stone-tool
form and function in the Aurignacian of southwest Germany. Journal of Human
Evolution 54, 648662.
Hardy, B.L., Kay, M., Marks, A.E., Monigal, K., 2001. Stone tool function at the
Paleolithic sites of Starosele en Buran Kaya III, Crimea: behavioral implications.
Proceedings of the National Academy of Sciences 98, 1097210977.
Hardy, B.L., Raff, R.A., Raman, V., 1997. Recovery of mammalian DNA from Middle
Paleolithic stone tools. Journal of Archaeological Science 24, 601611.
Haslam, M., 2004. The decomposition of starch grains in soils: implications for
archaeological residue analyses. Journal of Archaeological Science 31,
17151734.
Haslam, M., 2006. An archaeology of the instant? Action and narrative in
microscopic archaeological residue analyses. Journal of Social Archaeology 6,
402424.
Hogberg, A., Puseman, K., Yost, C., 2009. Integration of use-wear with protein
residue analysis a study of tool use and function in the south Scandinavian
Early Neolithic. Journal of Archaeological Science 36, 17251737.
Horrocks, M., 2006. Starch residues in coprolites. In: Torrence, R., Barton, H. (Eds.),
Ancient Starch Research. Left Coast Press, Walnut Creek, p. 78.
Horrocks, M., Campbell, M., Gumbley, W., 2007. A short note on starch and xylem of
Ipomoea batatas (sweet potato) in archaeological deposits from northern New
Zealand. Journal of Archaeological Science 34, 14411448.
Hortola`, P., 2002. Red blood cell haemotaphonomy of experimental human bloodstains on techno-prehistoric lithic raw materials. Journal of Archaeological
Science 29, 733739.
Hurst, W.J., Tarka, S.M., Powis, T.G., Valdez, F., Hester, T.R., 2002. Archaeology: cacao
usage by the earliest Maya civilization. Nature 418, 289290.
Jackes, M., Sherburne, R., Lubell, D., Barker, C., Wayman, M., 2001. Destruction of
microstructure in archaeological bone: a case study from Portugal. International Journal of Osteoarchaeology 11, 415432.
Janaway, R.C., 1985. Dust to dust: the preservation of textile materials in metal
artefacts corrosion products with reference to inhumation graves. Science and
Archaeology 25, 3138.
Janaway, R.C., 1987. The preservation of organic materials in association with
metal artefacts deposited in inhumation graves. In: Boddington, A.,
Garland, A.N., Janaway, R.C. (Eds.), Death, Decay and Reconstruction.
Approaches to Archaeology and Forensic Science. Manchester University Press,
Manchester, pp. 127148.
Janaway, R.C., 1996. The decay of human remains and their associated materials. In:
Hunter, J., Roberts, A., Martin, A. (Eds.), Studies in Crime: An Introduction to
Forensic Archaeology. Routledge, London.
Jans, M.M.E., 2003. Bot. In: Kars, H., Smit, A. (Eds.), Handleiding Fysiek Behoud
Archeologisch Erfgoed. Vrije Universiteit, Amsterdam, pp. 5158.
Jans, M.M.E., Nielsen-Marsh, C.M., Smith, C.I., Collins, M.J., Kars, H., 2004. Characterisation of microbial attack on archaeological bone. Journal of Archaeological
Science 31, 8795.
Jones, P.J., 2009. A microstrategraphic investigation into the longevity of archaeological residues. In: Haslam, M., Robertson, G., Crowther, A., Nugent, S.,
Kirkwood, L. (Eds.), Archaeological Science under a Microscope. ANU E Press,
Canberra, pp. 2946.
Joshi, S.R., Sharma, G.D., Mishra, R.R., 1993. Microbial enzyme activities related to
litter decomposition near a highway in a sub-tropical forest of North East India.
Soil Biology and Biochemistry 25, 17631770.
Kars, E.A.K., Kars, H., 2002. The Degradation of Bone as an Indicator for the Deterioration of the European Archaeological Property. Rijksdienst voor Oudheidkundig Bodemonderzoek, Amersfoort.
Kars, H., 2003. Het bodemarchief is kwetsbaar. In: Kars, H., Smit, A. (Eds.), Handleiding Fysiek Behoud Archeologisch Erfgoed. Vrije Universiteit, Amsterdam,
pp. 16.
Kelso, G.K., Mrozowski, S.A., Currie, D., Edwards, A.C., Brown III, M.R.A., Horning, J.,
Brown, G.J., Dandoy, J.R., 1995. Differential pollen preservation in a seventeenth-century refuse pit, Jamestown Island, Virginia. Historical Archaeology
29, 4354.
Kelso, G.K., Ritchie, D., Misso, N., 2000. Pollen record preservation processes in the
Salem Neck sewage plant shell midden (19-ES-471), Salem, Massachusetts,
U.S.A. Journal of Archaeological Science 27, 235240.
KNMI 2002. Klimaatatlas van Nederland. De Normaalperiode 19702000. Elmar,
Rijswijk.
Kooyman, B., Newman, M.E., Ceri, H., 1992. Verifying the reliability of blood residue
analysis on archaeological tools. Journal of Archaeological Science 19, 265269.
Kuksis, A., Child, P., Marai, J.J., Yousef, I.M., Lewin, P.K., 1978. Bile acids of a 3200year-old Egyptian mummy. Canadian Journal of Biochemistry 56, 11411148.
Langejans, G.H.J., 2009. Testing Residues An Experimental Approach. Ph.D Thesis,
University of the Witwatersrand, Johannesburg.
Lehmann, U., Robin, F., 2007. Slowly digestible starch its structure and health
implications: a review. Trends in Food Science & Technology 18, 346355.
Lentfer, C., Therin, M., Torrence, R., 2002. Starch grains and environmental reconstruction: a modern test case from West New Britain, Papua New Guinea.
Journal of Archaeological Science 29, 687698.

Lombard, M., 2004. Distribution patterns of organic residues on Middle Stone Age
points from Sibudu Cave, Kwazulu-Natal, South Africa. South African Archaeological Bulletin 59, 3744.
Lombard, M., 2005. Evidence of hunting and hafting during the Middle Stone Age at
Sibidu Cave, KwaZulu-Natal, South Africa: a multianalytical approach. Journal of
Human Evolution 48, 279300.
Lombard, M., 2007. The gripping nature of ochre: the association of ochre with
Howiesons Poort adhesives and Later Stone Age mastics from South Africa.
Journal of Human Evolution 53, 406419.
Lombard, M., Wadley, L., 2007. The morphological identication of micro-residues
on stone tools using light microscopy: progress and difculties based on blind
tests. Journal of Archaeological Science 34, 155165.
Loy, T.H., 1997. Methods in the Analysis of Archaeological Tool-Use Residues.
University of Queensland, Brisbane (unpublished syllabus).
Loy, T.H., 1998. Blood on the axe. New Scientist 159, 4043.
Loy, T.H., Dixon, E.J., 1998. Blood residues on uted points from eastern Beringia.
American Antiquity 63, 2146.
Loy, T.H., Hardy, B.L., 1992. Blood residue analysis of 90,000-year-old stone tools
from Tabun Cave, Israel. Antiquity 66, 2435.
Lu, T., 2003. The survival of starch residue in a subtropical environment. In:
Hart, D.M., Wallis, L.A. (Eds.), Phytolith and Starch Research in the
Australian-Pacic-Asian Regions: The State of the Art. Pandus Books, Canberra, pp. 119126.
Lu, T., 2006. The survival of starch residues in a subtropical environment. In: Torrence, R., Barton, H. (Eds.), Ancient Starch Research. Left Coast Press, Walnut
Creek, pp. 8081.
Mant, A.K., 1987. Knowledge from post-war exhumations. In: Bodington, A.,
Garland, A.N., Janaway, R.C. (Eds.), Death, Decay and Reconstruction.
Approaches to Archaeology and Forensic Science. Manchester University Press,
Manchester, pp. 6580.
Marchiafava, V., Bonucci, E., Ascenzi, A., 1974. Fungal osteoclasia: a model of dead
bone resorption. Calcied Tissue Research 14, 195210.
Martin, J.P., 1971. Decomposition and binding action of polysaccharides in soil. Soil
Biology and Biochemistry 3, 3341.
Martins, R.F., Davids, W., Al-Soud, W.A., Levander, F., Radstrom, P., Hatti-Kaul, R.,
2001. Starch hydrolyzing bacteria from Ethiopian soda lakes. Extremophiles 5,
135144.
Miller, T., 1979. Stone work of the Xeta Indians of Brazil. In: Hayden, B. (Ed.), Lithic
Use-wear Analysis. Academic Press, New York, pp. 401407.
Nelson, D.E., Loy, T., Vogel, J., Southon, J., 1986. Radiocarbon dating blood residues
on prehistoric stone tools. Radiocarbon 28, 170174.
Nicholson, R.A., 1996. Bone degradation, burial medium and species representation:
debunking the myths, an experiment-based approach. Journal of Archaeological
Science 23, 513533.
Nicholson, R.A., 1998. Bone degradation in a compost heap. Journal of Archaeological Science 25, 393403.
Nielsen-Marsh, C.M., Smith, C.I., Jans, M.M.E., Nord, A., Kars, H., Collins, M.J., 2007.
Bone diagenesis in the European Holocene II: taphonomic and environmental
considerations. Journal of Archaeological Science 34, 15231531.
Odell, G.H., 2001. Stone tool research at the end of the millennium: classication,
function, and behavior. Journal of Archaeological Research 9, 45100.
Painter, T.J., 1991. Lindow man, Tollund man and other peat-bog bodies: the
preservative and antimicrobial action of Sphagnan, a reactive glycuronoglycan
with tanning and sequestering properties. Carbohydrate Polymers 15, 123142.
Painter, T.J., 1998. Carbohydrate polymers in food preservation: an integrated view
of the Maillard reaction with special reference to discoveries of preserved foods
in Sphagnum-dominated peat bogs. Carbohydrate Polymers 36, 335347.
Painter, T.J., 2003. Concerning the wound-healing properties of Sphagnum holocellulose: the Maillard reaction in pharmacology. Journal of Ethnopharmacology 88, 145148.
Parr, J.F., 2003. The identication of Xanthorrhoea resins by starch morphology:
prospects for archaeological and taxonomic applications. Economic Botany 56,
260270.
Pickering, R., 2006. Regional geology, setting and sedimentology of Sibudu Cave.
Southern African Humanities 18, 123129.
Piperno, D.R., Holst, I., 1998. The presence of starch grains on prehistoric stone tools
from the humid neotropics: indications of early tuber use and agriculture in
Panama. Journal of Archaeological Science 25, 765776.
Piperno, D.R., Ranere, A.J., Holst, I., Hansell, P., 2000. Starch grains reveal early root
crop horticulture in the Panamanian tropical forest. Nature 407, 894897.
Piperno, D.R., Ranere, A.J., Holst, I., Iriarted, J., Dickauc, R., 2009. Starch grain and
phytolith evidence for early ninth millennium B.P. maize from the Central
Balsas River Valley, Mexico. Proceedings of the National Academy of Sciences of
the United States of America 106, 50195024.
Piperno, D.R., Weiss, E., Holst, I., Nadel, D., 2004. Processing of wild cereal grains in
the Upper Palaeolithic revealed by starch grain analysis. Nature 430, 670673.
Raiswell, R., 2001. Dening the burial environment. In: Brothwell, D.R.,
Pollard, A.M. (Eds.), Handbook of Archaeological Sciences. John Wiley & Sons,
Chichester, pp. 595603.
Ross, D.J., 1983. Invertase and amylase activities as inuenced by clay minerals, soilclay fractions and topsoils under grassland. Soil Biology and Biochemistry 15,
287293.
Rots, V., Williamson, B.S., 2004. Microwear and residue analyses in perspective: the
contribution of ethnoarchaeological evidence. Journal of Archaeological Science
31, 12871299.

G.H.J. Langejans / Journal of Archaeological Science 37 (2010) 971985


Sajilata, M.G., Singhal, R.S., Kulkarni, P.R., 2006. Resistant starch a review.
Comprehensive Reviews in Food Science and Food Safety 5, 117.
Samuel, D., 2006. Human digestion of starch. In: Torrence, R., Barton, H. (Eds.),
Ancient Starch Research. Left Coast Press, Walnut Creek, pp. 7677.
Simpson, I.A., Dockrill, S.J., Bull, I.D., Evershed, R.P., 1998. Early anthropogenic soil
formation at Tofts Ness, Sanday, Orkney. Journal of Archaeological Science 25,
729746.
Smit, A., 2003a. Hout. In: Kars, H., Smit, A. (Eds.), Handleiding Fysiek Behoud
Archeologisch Erfgoed. Vrije Universiteit, Amsterdam, pp. 3542.
Smit, A., 2003b. Leer. In: Kars, H., Smit, A. (Eds.), Handleiding Fysiek Behoud
Archeologisch Erfgoed. Vrije Universiteit, Amsterdam, pp. 5962.
Smit, A., 2003c. Textiel. In: Kars, H., Smit, A. (Eds.), Handleiding Fysiek Behoud
Archeologisch Erfgoed. Vrije Universiteit, Amsterdam, pp. 6368.
Smith, P.R., Wilson, M.T., 2001. Blood residues in archaeology. In: Brothwell, D.R.,
Pollard, A.M. (Eds.), Handbook of Archaeological Sciences. John Wiley and Sons,
Chichester, pp. 313322.
Staller, J.E., Thompson, R.G., 2002. A multidisciplinary approach to understanding
the initial introduction of maize into coastal Ecuador. Journal of Archaeological
Science 29, 3350.
Stiner, M.C., Kuhn, S.L., Weiner, S., Bar-Yosef, O., 1995. Differential burning, recrystallization, and fragmentation of archaeological bone. Journal of Archaeological
Science 22, 223237.
Therin, M., Fulagar, R., Torrence, R., 1999. Starch in sediments: a new approach to
the study of subsistence and land use in Papua New Guinea. In: Godsen, C.,
Hather, J. (Eds.), The Prehistory of Food Appetites for Change. Routledge, London, pp. 438464.
Tibbett, M., Carter, D.O., 2003. Mushrooms and taphonomy: the fungi that mark
woodland graves. Mycologist 17, 2024.
Turner, B., Wiltshire, P., 1999. Experimental validation of forensic evidence: a study
of the decomposition of buried pigs in a heavy clay soil. Forensic Science
International 101, 113122.
Tuross, N., Barnes, I., Potts, R., 1996. Protein identication of blood residues
on experimental stone tools. Journal of Archaeological Science 23,
289296.

985

Tuross, N., Dillehay, T.D., 1995. The mechanism of organic preservation at Monte
Verde, Chile, and one use of biomolecules in archaeological interpretation.
Journal of Field Archaeology 22, 97110.
Ungent, D., Pozorski, S., Pozorski, T., 1981. Prehistoric remains from the sweet potato
from the Casma Valley, Peru. Phytologia 49, 401415.
Ungent, D., Pozorski, S., Pozorski, T., 1986. Archaeological manioc (manihot) from
coastal Peru. Economic Botany 40, 78102.
Vane, C.H., Trick, J.K., 2005. Evidence of adipocere in a burial pit from the foot and
mouth epidemic of 1967 using gas chromatography-mass spectrometry.
Forensic Science International 154, 1923.
Wadley, L., Lombard, M., 2007. Small things in perspective: the contribution of our
blind tests to micro-residue studies on archaeological stone tools. Journal of
Archaeological Science 34, 10011010.
Wadley, L., Lombard, M., Williamson, B., 2004. The rst residue analysis blind tests:
results and lessons learnt. Journal of Archaeological Science 31, 14911501.
Warren, R.A.J., 1996. Microbial hydrolysis of polysaccharides. Annual Review of
Microbiology 50, 183212.
Watkinson, D., 2001. Maximizing the life span of archaeological objects. In:
Brothwell, D.R., Pollard, A.M. (Eds.), Handbook of Archaeological Sciences. John
Wiley & Sons, Chichester, pp. 649659.
Weather-Bureau, 1986. Climate statistics up to 1984. Climate of South Africa WB40,
1474.
Williamson, B.S., 2000a. Direct testing of rock painting pigments for traces of
haemoglobin at Rose Cottage Cave, South Africa. Journal of Archaeological
Science 27, 755762.
Williamson, B.S., 2000b. Prehistoric Stone Tool Residue Analysis from Rose Cottage
Cave and other southern African Sites. Ph.D Thesis, University of the Witwatersrand, Johannesburg.
Williamson, B.S., 2006. Investigation of potential contamination on stone tools. In:
Torrence, R., Barton, H. (Eds.), Ancient Starch Research. Left Coast Press, Walnut
Creek, pp. 8990.
Wilson, A.S., Janaway, R.C., Holland, A.D., Dodson, H.I., Baran, E., Pollard, A.M.,
Tobin, D.J., 2007. Modelling the buried human body environment in upland climes
using three contrasting eld sites. Forensic Science International 169, 618.

Вам также может понравиться