You are on page 1of 14

Journal of Archaeological Science 38 (2011) 1784e1797

Contents lists available at ScienceDirect

Journal of Archaeological Science


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

Intra-skeletal variability in trace elemental content of Precolumbian Chupicuaro human bones: the record of post-mortem alteration and a tool for palaeodietary reconstruction
A.-F. Maurer a, *, M. Gerard b, A. Person a, I. Barrientos c, P. del Carmen Ruiz c, V. Darras d, C. Durlet e, V. Zeitoun f, M. Renard a, B. Faugre d
a

Laboratoire Biominralisations et Environnements sdimentaires UPMC, ISTeP, UMR 7193, 4 Place Jussieu, 75252 Paris cedex 05, France Institut de Minralogie et Physique des Milieux Condenss IRD, UMR CNRS 7590 UMPMC, 4 Place Jussieu, 75252 Paris cedex 05, France CEMCA Sierra Leona 330 Lomas de Chapultepec, 11000 Mexico, D.F., Mexico d UMR 8096 CNRS-Paris 1, Archologie des Amriques, 21 alle de luniversit, 92023 Nanterre, France e Universit de Bourgogne, UMR CNRS 5561 Biogosciences, 6 bd Gabriel, 21000 Dijon, France f UMR 9993 CNRS-Muse Guimet, 19 avenue dIna, 75116 Paris, France
b c

a r t i c l e i n f o
Article history: Received 10 December 2009 Received in revised form 7 March 2011 Accepted 8 March 2011 Keywords: Chupicuaro Apatite Geochemistry Intra-skeletal variability Diagenesis Diet Hydrothermalism

a b s t r a c t
This study applies an intra-skeletal sampling strategy to examine post-mortem alteration of archaeological human bone from west Mexico, and to reconstruct ancient diet. Human bone from the Chupicuaro culture (Mexico, Preclassic period) constitutes an ideal material with which to examine subsistence strategies because the specic hydrothermal environment in which the population lived would have provided certain food components (hydrothermal waters and carbonates) with distinct signature in Ca, Mg, F, Li, Sr, Mn, V and U values. Four to ten samples were taken from the long bones of six skeletons. Bone trace element content (Ca, P, F, Mn, Mg, Na, Li, V, Zn, Rb, Sr, Ba, Y, La, Ce, Nd, Th, U) and bone alteration parameters (crystallinity, organic matter and secondary calcite content) were analysed at the intra-skeletal level. Stable isotopic signatures (bone d13C and d18Ocarbonate) and histological analyses were also performed on a single bone from each individual. Results indicate that all of the skeletons were affected by post-mortem mineralogical, structural and geochemical transformations. Biological bone d13C values seem preserved for most of the individuals but an increase in crystallinity accompanies depletion in bone d18O values. The combination of bone alteration parameters with bone elemental content shows that in this very specic context, a widespread dissolution-recrystallisation is unlikely. Of the hydrothermal tracers, Sr, F and Li were of particular interest because their retention in living tissues is related to the amount ingested. The intra-skeletal Li content does not reveal any pattern but Li depletion is not excluded. In contrast, Sr and F show a progressive intra-skeletal diagenetic enrichment likely due to gradual diffusioneadsorption processes. The bones with the lowest concentrations in these elements are assumed to yield the best representative ante-mortem values. The signal extracted from each skeleton, a very unusually high bone Sr, F and Li content, is interpreted as reecting the consumption of the local hydrothermal products, which are also enriched in these elements. 2011 Elsevier Ltd. All rights reserved.

1. Introduction The Chupicuaro population was one of the most important of the Preclassic cultures (ca. 600 B.C.eA.D. 250) in Western Mesoamerica. The Chupicuaro settled approximately 170 km northwest

* Corresponding author. Earth System Science Research Center, Department of Applied and Analytical Paleontology, Institute of Geosciences, University of Mainz, Johann-Joachim-Becher-Weg 21, 55128 Mainz, Germany. Tel.: 49 6131 39 23429; fax: 49 6131 39 24768. E-mail address: anmaurer@uni-mainz.de (A.-F. Maurer). 0305-4403/$ e see front matter 2011 Elsevier Ltd. All rights reserved. doi:10.1016/j.jas.2011.03.008

of Mexico City, along the Lerma River in a unique environment where the archaeological settlements are principally located around hydrothermal springs (Darras et al., 1999) (Fig. 1). These hot springs produce raw materials such as iron oxyhydroxides, and some are bordered by voluminous hydrothermal carbonate deposits. Hydrothermal products may have been used for construction and ceramic production, as suggested by the excavation of lime used for mortar and coating, and by the red and white pigments characteristic of the Chupicuaro ceramics (Darras and Faugre, 2001, 2005; Mikrut, 2003). Hydrothermal products, such as water and carbonates may also have been consumed. For

A.-F. Maurer et al. / Journal of Archaeological Science 38 (2011) 1784e1797

1785

Fig. 1. Location of the study area. A) Map of Mexico showing the location of the Acambaro valley. B) Map of the Acambaro Valley indicating spatial distribution of the Preclassic archaeological settlements. Digitization: Rodolfo Avila (CEMCA) C) Close-up of the hydrothermal area with the archaeological site JR24 studied.

example, in Mexico today, carbonates are used in an alkaline processing method (nixtamalization) to prepare the traditional corn dough (masa), to improve its nutritional quality (Bressani et al., 1958; Katz et al., 1974). This dough is then used for making tortillas, which are cooked on a comal (ceramic griddles). Although comal fragments have been discovered in the Basin of Mexico (Middle Preclassic) (Niederberger, 1976), it is unclear whether the Chupicuaro populations used this method. Therefore, the location of the Chupicuaro settlements, together with the close proximity of raw materials available for construction, ceramic production and diet, raise questions about the settlement strategy of this population. To investigate the importance of the hydrothermal springs in the economy of these pre-Columbian groups, an interdisciplinary study, The Chupicuaro Project, combining archaeology, geomorphology, sedimentology and geochemistry was designed. Convergences between geochemical composition of some raw materials and archaeological artefacts (Darras, 2004) indicate the utilisation of carbonates and iron oxides in craft production. The present study examines the geochemical composition of Chupicuaro skeletons to determine whether they yield a signal attributable to the incorporation of some hydrothermal components in their diet. Modern vertebrate bone geochemistry is directly related to the food and water consumption. The bone signature is an average of the geochemical composition of the main foodstuffs ingested during the last 10e30 years before death, according to the turnover rate of the anatomical part considered (Marshall et al., 1973; Tanaka et al., 1981). Major, minor and trace element content (Ca, P, F, Na, Mg, Sr, Ba, Zn, Mn, Li, V, U, Rb, Y, La, Ce, Nd, Th) of Chupicuaro bone mineral (called carbonate-hydroxylapatite or bioapatite

Ca10 [(PO4)6x (CO3)x](OH)2, Chang et al., 1996) and stable isotopic composition (d13C and d18Ocarbonate) are examined in this study. The elemental content of the hydrothermal carbonates and waters of the area is analysed for comparison with the skeletons. Trace elements provided by the diet follow different metabolic pathways depending on whether they are essential (ETE) or not (NETE) (i.e. vital or non-vital) for the organism. ETE (Ca, Mn, Mg, Na, P, F, Zn.) are subject to homeostatic control mechanisms that include regulation of absorption, excretion and tissue retention (Combs, 2005). Bone mineral plays a major role in those processes because of its non-stoichiometric properties (Weiner and Traub, 1992), nano size of the crystals (Posner, 1987), high specic surface, and carbonate content (Neuman and Neuman,1958; Weiner and Traub, 1992). The homeostatic mechanisms are maintained for varying nutrient intakes as long as ETE are ingested within adequate ranges. When ETE ingestion is decient or in excess, homeostatic regulation cannot be properly achieved (WHO/IPCS, 2002). Consequently, hypo or hyper concentration of ETE leads to a wide variety of clinical effects (Tapiero and Tew, 2003; Lindh, 2005). In contrast, NETE are not regulated. Instead, they accumulate in the organs depending on the amount ingested. Among these elements, Sr, Ba and Pb are mainly stored in the skeleton (Bauer et al., 1956) and are therefore commonly used to reconstruct the diet of past populations (Toots and Voorhies, 1965; Sillen and Kavanagh, 1982; Balter et al., 2002; Sponheimer et al., 2005). Bone isotopic composition is also diet-related. The d18O value of biological carbonate-hydroxylapatite mainly records the d18O of drinking water (Longinelli, 1984; Luz and Kolodny, 1985; DAngela and Longinelli, 1990; Bryant et al., 1994; Delgado Huertas et al., 1995) whose isotopic composition is mainly

1786

A.-F. Maurer et al. / Journal of Archaeological Science 38 (2011) 1784e1797

inuenced by geoenvironmental factors (Dansgaard, 1964). Bone apatite d13C records the isotopic composition of all dietary components such as proteins, lipids and carbohydrates (Ambrose and Norr, 1993; Tieszen and Fagre, 1993; Schwarcz, 2000). In continental ecosystems, d13C is driven by the isotopic variability in the photosynthetic pathways of different plants: 13& (ranging from 15& to 10&) for C4 plants (some shrubs, tropical grasses) and 27& (ranging from 35& to 21&) for C3 plants (trees, herbs, major shrubs, and shady grasses) (OLeary, 1981; Ehleringer, 1989). Unfortunately, the dietary signal may be modied by burial. The fossilization process is complex and multiphase (Trueman and Martill, 2002). The degradation of the organic matrix exposes bone crystals to soil solutions with which chemical interactions occur. This results in loss, addition or exchange of some elements acquired in vivo by the bone mineral. These geochemical transformations are accompanied by crystal maturation (improvement of the size of the homogeneous crystallites, elimination of defects) (Weiner and Bar-Yosef, 1990). In addition, diagenetic minerals precipitate in the pore spaces. Paleodietary reconstructions using bone major and trace element content are typically conducted on the diaphysis of long bones. However, long, compact bones may not always be preserved from alteration, as indicated by the highly variable intra-bone elemental content of archaeological skeletons (Keplinger et al., 1986). Here we propose to use the intra-skeletal variability in trace element content in order to control for post-mortem changes. Careful selection of the least altered bones enables us to reconstruct

Chupicuaro dietary habits to determine whether or not they were inuenced by their hydrothermal environment. 2. Material and methods 2.1. The Chupicuaro setting: JR24 e La Tronera The JR24 site is located on the right bank of the Lerma River in the Acambaro Valley, near the modern hamlet of Puruagita (state of Guanajuato, Mexico). It is located on an eroded Pliocene lava ow and near three hot springs with hydrothermal carbonate deposits (Fig. 1c). The archaeological excavations performed in 2000/2001/ 2002 uncovered 27 strata and 21 occupation layers composed of sands, silts and carbonates covering a consolidated pyroclastic substrate. The stratigraphic deposits (5 m deep) reveal that the area had been occupied from Early Chupicuaro to Mixtlan phases (600e400 B.C. to A.D. 1e250; Darras and Faugre, 2001, 2005). Nine primary individual burials were uncovered, two of newborns, one of a young child, and six of subadults or adults. The burials consist in simple pits digged in forest soil or anthropogenic sediment. Most of the individuals were buried within small circular funerary structures (Darras and Faugre, 2001) with the exception of individuals S3 and S8. The bone samples used in this study were collected from the adult specimens. In total 48 samples were collected from six different skeletons (S1, S3, S6, S7, S8 and S9) (Table 1). One skeleton (S1) is dated from the Mixtlan phase and the others from the Late Chupicuaro phase (400e100 B.C.). The adult

Table 1 Information for the six adult skeletons studied, S1, S3, S6, S7, S8 and S9. The bones collected from each individual are specied (F: femur, T: tibia, H: humerus, R: radius and U: ulna; L: left, R: right). The archaeological period, position of the body, pathologies, sex and age are given. The depth of the burial pit and the sediment used to ll it are also indicated. Burial 1 Number of samples 10 Bones sampled 2F, 2T, 2H, 2R, 2U Period Mixtlan Depth (cm) 195 Burial ll Clay mixed with gravel and occupational residues Position of the body Left lateral decubitus E-W cranial deformation Pathologies Caries, 3rd grade wear consolidated fracture (L rib) slight exostose (lombar/cervical column) Caries, 1st grade wear slight exostose (lombar column) Caries, 3rd grade wear No wear, no carie inter vertebral disc necrosis growth arrest lines Caries, 3rd grade wear exostose (lombar column) Caries, 3rd grade wear consolidated fractures (R clavicle, R coxal) exostoses (cervical/ lombar column, R ulna and radius) Cifosis osteomyelitis (R ulna and radius) slight periostis (L tibia) dental uorosis Sex/age Woman 35e40 years

R, HD, 2F

Chupicuaro

570

Ashes and calcareous rocks

Extended dorsal decubitus E-W cranial deformation

Woman 30e45 years

6 7

10 4

2F, 2T, 2H, 2R, 2U 2H, 2F

Chupicuaro Chupicuaro

620 530

Forest soil and ashes Forest soil

Extended dorsal decubitus E-W cranial deformation Extended dorsal decubitus E-W cranial deformation

Man 25-30 years Woman 18-20 years

10

2F, 2T, 2H, 2R, 2U

Chupicuaro

590

Forest soil

Extended dorsal decubitus E-W cranial deformation

Man 30-35 years

10

2F, 2T, 2H, 2R, 2U

Chupicuaro

600

Forest soil

Extended dorsal decubitus NE-SW cranial deformation

Woman 35-40 years

A.-F. Maurer et al. / Journal of Archaeological Science 38 (2011) 1784e1797

1787

remains were oriented east/west, with the skeleton lying in dorsal decubitus, and the upper limbs positioned alongside the thorax with the skull to the west or to the east. The exception was skeleton S1 which was lying in dorsal decubitus with exed legs on the left side (Darras and Faugre, 2001). In addition, the following samples were obtained for complementary analyses: - a modern cow bone discovered at the surface of JR24 and used as a reference sample, - hydrothermal and non-hydrothermal samples: water (7 hydrothermal, 3 rivers, 4 springs) collected in the area of Puruagita, hydrothermal carbonates sampled at Puruagita (3 samples) and the carbonate fraction of hydrothermal deposits, 20 km westward, at Aguas Calientes hot springs (2 samples) (Fig. 1).

The cortical bone samples were cut using a drill equipped with a diamond saw. After a mechanical cleaning, they were divided in two aliquots: one for histological and crystallinity examination and the second for chemical analysis. 2.2. Histological analysis Transverse bone sections were embedded into a polyester resin under vacuum, before being cut with a diamond saw, glued on glass slides and polished with cerium oxide. The nal thickness of the thin sections is 30 mm. One thin section was examined for each skeleton (Fig. 2) using optical transmitted light microscopy. The histological index (HI) of Hedges et al. (1995) was used in order to estimate the quality of preservation of the fossil bones. HI varies from 5 for fresh bone (all structures are

Fig. 2. Histological sections of the Chupicuaro bones (skeletons S1, S3, S6, S7, S8 and S9; F: femur, H: humerus; L: left and R: right) under polarized light.

1788

A.-F. Maurer et al. / Journal of Archaeological Science 38 (2011) 1784e1797 Table 2 NIST SRM 1400: recovery data for certied and non-certied analytes. All elements certied in SRM were accurately determined by the proposed method. Concentrations in italics are below the quantication limit. Analyses were performed on 21 samples at different times. Analyte Target Ca % P% Mg % Na % F% Sr mg g1 Ba mg g1 Zn mg g1 Mn mg g1 V mg g1 Li mg g1 U mg g1 Y mg g1 La mg g1 Ce mg g1 Rb mg g1 Nd mg g1 Th mg g1 38.18 0.13 17.91 0.19 0.684 0.013 0.6 0.125 249 7 240 10 181 3 17 0.77 0.09 0.95 0.14 0.066 0.003 0.29 0.03 0.39 0.08 0.82 0.10 0.71 0.13 0.32 0.02 0.12 0.00 Measured 38.51 18.01 0.665 0.61 0.114 252 241 178 15 0.78 1.13 0.060 0.23 0.28 0.56 0.47 0.30 0.11 0.66 0.17 0.014 0.04 0.006 5 7 6 1.0 0.14 0.10 0.005 0.02 0.02 0.07 0.01 0.02 0.01 % Recovery 101 101 97 101 91 101 100 98 86 101 118 92 80 74 68 67 94 87

distinguished) to 0 when less than 5% of the bone is intact. When the thin sections showed differential histological preservation, a mean index was assigned, based on the percentage of intact bone. 2.3. Crystallinity analysis and the secondary calcite content of bone voids For bulk mineralogical analyses, bones samples were ground with an agate mortar and pestle. The powder was pressed in a glass-aluminium sample-holder. A Siemens D500 diffractometer with Ni-ltered CuKa radiation at 40 kV and 30 mA was used for X-ray diffraction analysis (XRD). Samples were scanned from 2 to 70 (2q), with counting for 6 s every 0.02 , in a rotating sample-holder. The CI (Crystallinity Index) was determined using the semi-quantitative method of Person et al. (1996). A semiquantitative estimation of the calcite content (precipitated within the pore structure) was calculated as Surfacecalcite[104]/ Surfaceapatite[002]. 2.4. Chemical analysis 2.4.1. Pretreatments and sample digestion Bone powder was etched in dilute acetic acid (0.1 M) for 1 h. The residual material was rinsed and then used for heat treatment in order to remove the bone organic matter (OM). A low temperature (300  C) heat treatment was applied to maximise the removal of OM while minimising microstructural and ultrastructural changes to the bone mineral. Bone powder, in a platinum beaker, was placed into a closed system furnace under vacuum for 6 h to distil the OM. This was followed by heat treatment in a mufe furnace at 300  C for an additional 20 h to oxidize the OM carbon. This step was achieved by a pure O2 ow for 1 h. This protocol ensures the complete destruction of bone OM and allows an estimation of the bone protein content (bone % weight loss). A 100 mg sample of the resulting ash was dried overnight at 80  C and weighed into a Teon beaker before being digested in 5 ml of 20% nitric acid (Merck, suprapur) and heated at 80  C for 1 h. The resulting solution was then diluted with deionized water to a total volume of 50 ml, stocked in polypropylene or Teon tubes and kept for analysis. 2.4.2. Trace element and stable isotope analysis Bone chemical analyses were performed using a Varian UltraMass ICP-MS for bone Li, V, Zn, Rb, Sr, Ba, Y, La, Ce, Nd, Th and U contents. Bone Ca, Mn, Mg and Na contents were analysed with a Varian ICP-AES. In order to avoid matrix effects from major elements (essentially Ca and P) on bone trace element content, we applied a standard addition method. The standard matrix was prepared with NIST SRM1400 bone ash digested with nitric acid, with a 200-fold dilution for Ca, and a 5-fold dilution for the other elements. The standard calibration was prepared from Spex, 1000 mg/ml certied stock. The quality control and method validation were performed by analysis of NIST SRM1400 bone ash (Table 2). Fluoride concentrations were measured by potentiometry using a specic electrode. PO4 was measured photometrically. Isotopic analyses were performed using a mass spectrometer VG SIRA9. Samples were reacted under vacuum with orthophosphoric acid overnight at 50  C. All water and carbonate analyses were performed by ICP-MS and ICP-AES at the Institut de Recherche pour le Dveloppement (Bondy, Paris) and the Centre de Recherches Ptrographiques et Gochimiques in Nancy (France), respectively.

a a a b b a c a b c c c c c c c c c

a- NIST certied concentrations, b- NIST informational value, c- Hinners et al., 1998.

3. Results 3.1. Hydrothermal products Trace element data from analysis of hydrothermal waters and their associated recent carbonate caps, rivers and wells from Puruagita, as well as carbonates from Aguas Calientes, are presented in Table 3. Puruagita hydrothermal springs are slightly acidic (mean pH 6.3) and warm (mean T C 39.3  C). The nonhydrothermal waters sampled from adjacent rivers and wells (referred to as cold waters below) are more basic (mean pH 7.5 and 7, respectively) and the wells are colder (mean T C 25.4  C). The hydrothermal waters are highly enriched in Cl, Na, Li, Sr and Rb (over 10 times) and enriched in HCO3, Ca, Mg, F, Ba, Mn, Cr, Cu, Y, Nd and La compared to the cold waters. In contrast, hydrothermal waters are highly depleted in U and SO42 and slightly depleted in V compared to cold waters. Carbonates from Puruagita hot springs are highly enriched in Sr but highly depleted in Rb, V, Y, Nd, La, U, Ce compared to the other hydrothermal deposits sampled at Aguas Calientes. Mg, Ca, Ba, Li and Mn enrichment and Cu, Cr and Na depletion in carbonates from Puruagita are also signicant. None of the hydrothermal products (water and carbonate) analysed exceeds the Tolerable Upper Intake Levels (Food and Nutrition Board, Institute of Medicine, 1997, 2001, 2004) considered for a normal diet (i.e. 1 or 2 L of water and several g of carbonates ingested per day). According to the composition of hydrothermal products (waters and carbonates), Ca, Mg, F, Li, Sr, Ba, Mn, V and U contents can be used as tracers of hydrothermalism in the area of Puruagita. Their content in Chupicuaro bones is therefore of particular interest. Ba however will not be discussed further. Its incorporation into bone, like Sr, is related to the amount of calcium ingested (Elias et al., 1982) and although Ba/Ca differentiates hydrothermal deposits from Puruagita and Aguas Calientes, it is not a good tracer for the hydrothermal waters analysed (Table 3). Conversely, Sr/Ca is signicantly higher in these potential dietary components. 3.2. Human bones from La Tronera Trace elements, OM, calcite contents, stable isotope composition, CI and HI of Chupicuaro human bones are summarized in

A.-F. Maurer et al. / Journal of Archaeological Science 38 (2011) 1784e1797

1789

Table 3 Composition of hydrothermal products (waters and carbonates, grey zones). The average value and the standard deviation of the samples are given. The number of samples analysed is given in brackets. Acceptable concentrations in water as well as recommended or tolerable intakes are presented for most elements/species. Font style and symbols (bold, italic, * and **) correspond to individual references listed below the table.

ELEMENTS

WATERS

GUIDELINES FOR HEALTHY LEVELS

HYDROTHERMAL DEPOSITS

hydrothermal
(7)

rivers
(3)

wells
(4)

in water

intake DRI: Dietary Reference Intakes RDA AI UL

Puruagita
(3)

Aguas Calientes
(2)

mean pH mean TC
g.l-1

6.3 39.3
g.l-1

7.5
g.l-1

7 25.4
g.l-1 g.j-1 g.j-1 g.g-1 g.g-1

HCO3ClNa Ca Mg FNO3SO42Li Sr Rb Ba Mn Cr V Cu Y Nd La U Ce

989 730 669 113 14 3.6 2.4 0.6 3366 2428 347 338 205 9 1.9 1.5 63 14 9 8 5

52 46 47 12 4 0.4 0.2

.103 .103 .103 .103 .103 .103 .103 .103

209 16 44 24 8 1.8 2.5 8 54 252 30 110 63 2 5.2 0.2

34 6 11 2 1 1.6 11 56 28 7 83 94 1.3 4.9 0.1 14 1 2

.103 .103 .103 .103 .103 .103 .103 .103

94 5 18 13 4 0.7 7 11 11 127 31 36 0.5 1 5.1 0.1

48 5 7 9 2 0.6 3.5 9 8 94 15 35 0.3 0.5 3.4 0.1 3 1 1

.103 .103 .103 .103 .103 .103 .103 .103

250 200

.103 .103

4-1.5 10 500

.103 .103 .103

2300 .103 1500 .103 1000 .103 310-420 .103 3 4 .103

3600 .103 2300 .103 2500 .103 350 .103 10 .103 ND .103

9 5 338 6 11 2

.103 .103 .103

15 0.7 .103 109 13 .103 3 0.8 .103

62 142 13 62 31 3.5 0.9 0.8 26 10 5 8 3 .10-3 .10-3 .10-3 .10-3 .10-3

*1000

**500000

2000 1000 50 100-50 unregulated 1300 1000


.10-3 .10-3 .10-3

1800-2300 25-35 900

11000 ND 1800 10000

184 5163 13 902 0.5 2 1 2.4 703 411 499 34 1022

103 694 6 211 0.2 2.2 0.9 2

37 140 122 133

1 30 4 29

13 3 22 6 22 7

26 2 4 444 10

.10-3 .10-3 .10-3

419 .10-3 5 .10-3

9 1 2 394 12

2000 * .10-3 1050000 * .10-3 2000 * .10-3 30000 -20000 .10-3 2000 * .10-3

301 .10-3 11 .10-3

166 .10-3 22550 780 .10-3 163 .10-3 10800 0 .10-3 210 .10-3 16350 1202 .10-3 30 .10-3 3825 21 .10-3 197 .10-3 34500 2263 .10-3

Ba/Ca Sr/Ca

0.0030 0.0004 0.022 0.001

0.0045 0.0033 0.010 0.001

0.0025 0.0013 0.009 0.002

0.0027 0.0006 0.015 0.0022

0.0012 0.0004 0.0013 0.0004

MCL: Maximum Contaminant Level (US EPA, 2009) MAC: Maximum Acceptable Concentrations or OG: Operational Guidance value (FPT Committee on Drinking Water, 2008)

RDA: Recommended Dietary Allowances; AI: Adequate Intakes UL: Tolerable Upper Intake Levels from Food and Nutrition Board, Institute of Medicine, National Academies, 1997, 2001, 2004

* iAC: indicative Admissible drinking water Concentrations (de Boer et al., 1996) * from Schrauzer, 2002 ** from Anke, 1993

Table 4 (see Appendix A for more detail). Complementary data from a modern cow bone are also presented. 3.2.1. Mineralogical and structural parameters Skeletal Ca/P ratios vary between 2.19 (S8) and 2.28 (S9) with an average value of 2.24 0.03. CI ranges from 0.07 (S1) to 0.28 (S9) with an average value of 0.20 0.07. Skeletal OM content varies between 10% (S7) and 27% (S1), with a mean value of 17 6%. The thin sections show that all skeletons are affected by micro-bacterial attacks with HI ranging from 1 to 3 (Fig. 2). Secondary calcite is present in all bones except one (S1). 3.2.2. Geochemical composition 3.2.2.1. Trace element content. Trace element content of the human skeletons differs from fresh modern cow bone, with higher F, Sr, Ba, Zn and Li. The Chupicuaro bones also display higher Mn, V, U values

compared to the modern cow bone for which these elements do not exceed the quantication limit. Although around the quantication limit, Rb, Y, Ce, La, Nd and Th content of the human bones are slightly higher (except Rb) than that of the modern reference sample. In contrast, Na and Mg content of the archaeological human bones are lower than the content analysed in the modern animal bone. Most elements display an intra-skeletal coefcient of variation of between 10% and 25% (F, Na, Mg, Sr, Ba, Zn, Li, V, U). This variability is higher (27%e63%) for Rb, Y, La, Ce, Nd and Th content, which are around the quantication limit. Bone Mn content displays the highest intra-skeletal variability (80%). The intraskeletal variability of the elements that are used as tracers for hydrothermalism is mainly driven by the upper limb bones that typically display higher F, Sr and Mn concentrations (Fig. 3) and lower Mg content (except for S8). For individual S9, F and Sr enrichment is higher in the tibias. No general intra-skeletal pattern

1790 Table 4 Geochemical composition (trace elements, stable isotopes), mineralogical and structural parameters (organic matter OM, Crystallinity Index CI, calcite content and Histological Index HI) of the studied human skeletons. The mean skeletal values, as well as the intra and inter-skeletal coefcient of variations (cv) are given. Data from a modern cow bone (non-pretreated with acetic acid) from site JR24 are presented for comparison. Data in italics are below the quantication limit. The bold was used to highlight the skeleton number.

A.-F. Maurer et al. / Journal of Archaeological Science 38 (2011) 1784e1797

Calcite

0.15 8 0.45 114 0.41 21 0.43 68 0.31 72 0.29 57 62 0

is observed for Li (except in S6), V (except in S6 and S1) and U content. Inter-individual variability varies between 20 and 30% for Na, Mg, Sr, Zn and Li, 30 and 40% for F, Ba, V, U and Th, between 40 and 50% for Rb and by more than 50% for Mn, Nd, La, Ce and Y. 3.2.2.2. Stable isotope composition. Bone d13C varies between 3.2& (S1) and 1.1& (S8) with a mean skeletal value of 1.9 0.8&. Bone d18O ranges from 5.6& (S3) to 4.4& (S1) with an average value of 5.0 0.4&. 4. Discussion 4.1. Patterns of bone alteration The increase in CI, the precipitation of secondary calcite and the degradation of OM are common in archaeological bones (Person et al., 1995; Salige et al., 1995; Nielsen-Marsh and Hedges, 2000; Trueman and Tuross, 2002; Reiche et al., 2003; Smith et al., 2007). The precipitation of calcite indicates contact of bones with groundwater. The increase in crystallinity is often referred to as dissolutionerecrystallisation process (Trueman and Tuross, 2002) which would delete the dietary signal. However, bones submitted to a heat treatment to monitor diagenetic effects (Person et al., 1996; Munro et al., 2007, 2008; Zazzo et al., 2009) show that loss of CO3 radicals in the carbonate-hydroxylapatite also improves the crystallinity without necessarily implying a dissolution-recrystallisation process (Person et al., 1996). The intra and inter-skeletal variability of these parameters (Fig. 4ab) show gradual changes of bone tissues during burial within the same archaeological site for contemporaneous (except S1) human skeletons. Bacterial attack (Fig. 2), which is an early diagenetic process that starts soon after death (Yoshino et al., 1991; Bell et al., 1996) appears to be independent of these changes (age, crystallinity, OM and calcite content). The analysis of these bone structural changes gives a rst estimation of the skeletons state of preservation. Individual S1 seems to be the least altered of the skeletons, with a high OM content, a low CI and no calcite. This skeleton is more recent (Mixtlan phase) than the other Chupicuaro skeletons. A relationship between crystallinity increase and age is not usually observed except perhaps during the early diagenesis phase (Person et al., 1996; Sillen and Parkington, 1996). Skeleton S1, together with skeletons S3, S8 and S9 represent the general diagenetic trajectories. The slight deviance of skeletons S6 and S7 from these general trends may indicate different taphonomic histories. Skeleton S6 shows a high calcite content in spite of a fairly high OM content, suggesting an important contact with groundwater. In contrast, the increase in CI of skeleton S7 (except for the right humerus) seems to have been delayed, in spite of its low OM content. The analysis of bone d13C and d18Ocarbonate was conducted in order to further investigate Chupicuaro diet. However, bone CI is correlated with bone d18Ocarbonate. In contrast, bone d13C does not show any apparent correlation with bone CI (Fig. 5). An increase in crystallinity associated with a severe depletion of bone d18Ocarbonate without any signicant modication of bone d13C, was also observed during experimental heating (Munro et al., 2008). The abrupt shift in bone d13C (w1.5&) at around CI 0.10e0.14 separates the skeletons studied into two main groups 1) containing only two bones from the individual S1 and 2) containing bones from individuals S6, S7, S8 and S3, whose d13C values are very similar (average 1.42 0.24&). The bone d13C value of individual S9 lies between the two groups (d13C 2.33&) and has the highest CI (0.28). It is difcult to know whether or not it results from diagenetic exchange with carbonates of the diagenetic calcite. Bone trace element content varies at the intra and inter-skeletal level. Oxides may be responsible for the large intra (83%) and inter-

OM %

Ca/P

Th mg g1

Nd mg g1

Ce mg g1

La mg g1

Rb mg g1

U mg g1

V mg g1

Li mg g1

Mn mg g1

Zn mg g1

Ba mg g1

mg g1

0.51 23 0.49 18 0.61 43 0.73 18 0.39 19 0.99 14 0.62 23 35 0.00 Mean S1 cv (%) Mean S3 cv (%) Mean S6 cv (%) Mean S7 cv (%) Mean S8 cv (%) Mean S9 cv (%) Mean Intra-skeletal cv (%) Inter-skeletal cv (%) Modern cow bone
a

F %

0.61 7 0.45 5 0.84 92 0.46 7 0.43 8 0.48 18 0.54 23 29 0.79 Hedges et al., 1995.

Skeletons JR24

Na %

0.26 13 0.18 6 0.14 8 0.19 8 0.19 13 0.18 12 0.19 10 21 0.45

Mg %

3271 10 2133 9 1848 10 2572 6 2023 13 2242 7 2348 9 22 832

Sr

1492 20 1065 13 664 20 1352 13 766 18 1120 14 1076 16 30 156

186 13 213 13 293 33 175 17 209 19 193 39 212 22 20 132

6 53 23 72 39 103 14 128 20 64 40 77 24 83 58 1.7

30 22 38 16 24 15 45 4 34 9 23 11 32 13 26 1

15 17 26 16 32 19 27 8 24 15 39 15 27 15 30 0

17 38 54 20 57 23 48 20 36 15 28 18 40 23 39 0

0.23 24 0.17 44 0.16 23 0.33 61 0.11 10 0.33 77 0.22 40 41 0.46

0.46 34 0.32 93 0.77 108 0.13 53 0.30 33 1.51 56 0.58 63 87 0.04

Y mg g1

0.28 35 0.14 37 0.22 55 0.10 42 0.11 21 0.44 25 0.21 36 60 0.03

0.46 34 0.30 78 0.32 55 0.14 59 0.21 37 0.90 26 0.39 48 70 0.04

0.31 16 0.19 38 0.27 48 0.13 38 0.15 19 0.48 26 0.26 31 51 0.05

0.08 21 0.05 28 0.07 24 0.06 40 0.04 19 0.10 32 0.06 27 33 0.02

41 3.41

L3.2

L1.3

L1.7

L1.6

L1.1

L2.3

L1.9

d13C

&

9 2.91

L4.4

L5.6

L5.0

L4.8

L5.2

L5.3

L5.0

d18O

&

2.22 1 2.27 1 2.22 2 2.24 1 2.19 2 2.28 2 2.24 2 1 2.19

0.07 38 0.22 18 0.16 27 0.23 26 0.25 19 0.28 11 0.20 23 37 0

CI

27 11 17 18 21 13 10 10 14 16 16 16 17 14 34 37

41 5a

HI

A.-F. Maurer et al. / Journal of Archaeological Science 38 (2011) 1784e1797

1791

Fig. 3. Intra-skeletal trace element cartography. Content of the hydrothermal tracers in bones (Mg, F, Sr, Li, V, U and Mn) are reported for each part of the skeleton. The grey zones highlight mostly the upper limb bones (S1, S6 and S8) and the tibias in the case of S9.

individual (58%) variability in bone Mn content. They are commonly found in archaeological bone pore structure (Williams and Potts, 1988). Mg is an essential trace element and is therefore regulated by homeostatic mechanisms. However, although the intra and inter-skeletal variability in bone Mg content is fairly low (10% and 21%, respectively), all of the human skeletons studied display a Mg depletion, compared to modern bones, which is probably related to the decay of the collagen matrix (Fig. 4c). S6 is particularly affected as it shows a signicant Mg depletion despite substantial OM content. This depletion, together with this skeletons calcite content, argues for an important leaching mechanism that mainly affected the upper limbs. The S6 upper limb bones also display the minimum intra-skeletal Li content, although no general intra-individual pattern is observed for this element (Fig. 3). Li depletion during burial is therefore not excluded when considering a unidirectional diagenetic trajectory for this element (i.e. not preceded by early post-mortem enrichment). In that case, the high Li values found in Chupicuaro bones would not be attributable to diagenetic enrichment. Bone Mg leaching is also accompanied by a diagenetic enrichment in F, attested by the very high values of the human skeletons compared to the modern cow bone, the very high intra-skeletal variability of individual S6s F concentrations (43%), the correlation between bone F and calcite content and the correlation between bone F and V content (Fig. 6a and b). S8 and S1 also display correlations between F and calcite and between F and V content respectively (Fig. 6c and d). All of the skeletons show postmortem F addition regardless of their skeletal F content and intraskeletal variability. As Sr concentrations are related to those of F for each skeleton (Fig. 7), bone Sr content also results from a postmortem enrichment process, in spite of the generally low intraskeletal variability (mean value 9%). In most cases, the upper limbs yield the highest Sr and F content. Trace element uptake mechanisms remain obscure (Kohn, 2008) and it is therefore difcult to explain the exact mechanism for such Sr, F and V enrichment in the Chupicuaro human bones. It is worth

noting that with only two exceptions (S9 left humerus and right tibia), all of the skeletons display a Ca/P ratio that is in the biological range 2e2.3 (Trueman and Tuross, 2002). This result strengthens the idea that this parameter is not a relevant diagenetic proxy (Hubert et al., 1996; Fabig and Herrmann, 2002). However, other parameters do show that dissolutionerecrystallisation processes seem unlikely: i) Chupicuaro bones REE content is close to 1 ppm, typical concentrations for in vivo bone (Trueman and Tuross, 2002); ii) Bone F, Sr and V content are not associated with an increase in crystallinity at the intra-skeletal scale, nor at the population level; iii) The U diffusioneadsorption process operating on bone crystallites (Badone and Farquhar, 1982; Millard and Hedges, 1995, 1996; Simon et al., 2008) may explain the relatively homogeneous intra-skeletal U content, the signicant human bone U content, and the absence of any relationship between U and other elements or mineralogical parameters. The intra-skeletal variability observed in this study cannot be attributed to a biological repartition, which would imply similar patterns for the skeletons studied. It is therefore due to taphonomic processes and probably related to differential kinetic diffusion parameters depending on the chemical elements, the part of the skeleton and the individuals. The skeletal parts least affected by the progressive enrichment of Sr, F and V concentrations (generally lower limb bones) should therefore be considered the most representative with regard to the in vivo signature of the Chupicuaro bone. 4.2. Consumption of hydrothermal products Of the tracers of Puruagita hydrothermalism (cf. 3.1.), Li, F and Sr are of particular interest because their retention in living tissues

1792

A.-F. Maurer et al. / Journal of Archaeological Science 38 (2011) 1784e1797

0.45 0.35

-1 -1.5

Mg (%)

C
13

0.25 0.15 0.05 2.00 1.60

-2 -2.5 -3 -3.5 -4 -4.4

Bone calcite content

O= -4.08 - 4.84 * CI R = 0.89

18

1.20 0.80 0.40

Oc
18

-4.8 -5.2 -5.6 -6 0 0.05 0.10 0.15 0.20 0.25 0.30 S1 S3 S 6 JR24 S7 S8 S9

b
0.00 0.30 0.25 S1 S3 S 6 JR 24 S7 S8 S9 modern cow bone

CI
Fig. 5. Relationships between bone crystallinity (CI) and bone isotopic composition (d18Ocarbonate and d13C). Two samples were analysed from skeleton S1 (right humerus and left tibia) and one sample was analysed for each of the other skeletons: the left femur of S7 and S3, the left humerus of S6 and S8 and the left tibia of S9.

CI

0.20 0.15 0.10 0.05 0.00 5 10 15 20 25 CI = 0.43 - 0.01 * OM R = 0.86

a
30 35 40

OM (%)
Fig. 4. Relationships between bone organic matter content (OM) and (a) bone crystallinity (CI), (b), bone calcite content, and (c) bone apatite magnesium content. The results are given for all of the bones analysed from each human skeleton from site JR24 Data from a modern cow bone are given for comparison. The magnesium content of all of the bone samples (archaeological and modern) was obtained after acetic acid pretreatment.

(and especially in bones for F and Sr) is determined by the amount ingested. The concentration of these elements in hydrothermal products is therefore compared to the data extracted from each skeleton (i.e. the best representative concentrations of the in vivo signature). Except for its therapeutic role in treating bipolar affective disorders (Schou, 1968), Li does not appear to be an essential element for life (Lonard et al., 1995) and its biochemical function in vital components of the body is not known (Anke et al., 2005). Plasma Li concentrations vary according to the Li intakes (Zaldivar, 1980; Schrauzer, 2002). Following ingestion, non-excreted Li reaches different target organs such as cerebellum, cerebrum, kidneys, hair and liver (Schrauzer, 2002; Anke et al., 2005). Li is also retained and released by bone (Schrauzer, 2002; Kosla and Anke, 2005). The lack of human bone Li concentration data associated with known ingested concentrations makes it impossible to calculate the Li ingested from the measured Li content of the Chupicuaro bones. However, Li content has been measured in modern Japanese ribs as less than 0.08 mg g1 (Yoshinaga et al., 1995). Assuming that this is related to a daily dietary Li intake of around 0.06 mg in Japan (Shiraishi, 2005), the very high Li content of the Chupicuaro skeletons (mean 32 mg g1) can be hypothesized to be the result of the

ingestion of hydrothermal waters, which would supply at least 3.3 mg of Li to the Chupicuaro per day from water intake alone. F is considered essential, or at least potentially essential, in trace amounts (Lindh, 2005). It is mainly incorporated in the skeleton (Cerklewski, 1997). Homeostasis occurs particularly during the growth period, through the combined effects of skeletal uptake and urinary excretion (Cerklewski, 1997). However, F absorption and plasma F concentrations are not regulated (Whitford, 1996; Cerklewski, 1997). Plasma F content, and in turn bone, is partly related to the amount of F ingested (Ekstrand et al., 1977). A study conducted on the iliac crest of modern Americans who were exposed to variable water F content (Zipkin et al., 1958) illustrates this relationship (Fig. 8). The Chupicuaro skeletal F concentrations are in agreement with the intake of drinking water reaching an F content of between 2.6 and 4 ppm. These concentrations are compatible with those of the hydrothermal waters (3.6 ppm), close to the maximum contaminant level (MCL) dened by the Environmental Protection Agency (cf. Table 3). The MCL was established in order to protect against crippling skeletal uorosis, the most severe stage of the disease. Fluorosis affects teeth and bones when the F intake exceeds a toxic level, which is not particularly well-dened. Skeletal uorosis is composed of four stages (preclinic, I, II, III or crippling), which range from weak symptoms to a signicant alteration of mobility. Bone F concentrations and the levels at which skeletal uorosis occurs vary widely (Fig. 8). Male individuals S6 and S8 are unlikely to have been affected by the disease. Conversely, female individuals 1, 3, and 7 may have experienced uorosis, especially the early stage of the disease. The female individual S9 is the only one who would have been strongly affected by the disease, falling within stages II and III. Dental uorosis was detected on S9s teeth but no skeletal uorosis was observed (Barrientos and del Carmen Ruiz, 2009). The very high skeletal F content of S9 can therefore be attributed to a post-mortem enrichment. The numerous pathologies (periostite, osteomyelite, enthesopathie) and fractures (Garcia, 2002) that affected this individual during her life very likely enhanced post-mortem F addition.

A.-F. Maurer et al. / Journal of Archaeological Science 38 (2011) 1784e1797

1793

Bone fluoride content (%)


1 0 .2 5 0 .5 0 .2 1 .1
ulnas radius humerus tibias femurs

Bone calcite content

S8
R = 0.83

S6
R = 0.89

LU

0 20

d
S1
R = 0.85

a
S6
R = 0.85

0 44

Bone V g/g

11 0 .3

c
0 .7 5 0 .2

b
1 .1

22

Fig. 6. Intra-skeletal relationships between bone uoride and bone calcite content (a and d) and bone uoride and vanadium content (b and c) for three skeletons S1 (c), S6 (a and b) and S8 (d). The correlation coefcient, R, between bone uoride and bone calcite content of skeleton S6 was calculated after exclusion of the left ulna (LU).

Bone fluoride content (%)


4100 2400

S1
R = 0.71

S3
R = 0.96
ulnas radius humerus tibias femurs

2700 0.3 0.75 0.38

1900 2800

0.58

Bone Sr content g/g

2200

S6
R = 0.93

S7
R = 0.97

1400 2700

0.2

1.1 0.55

0.9

2350 2500

S9
R = 0.70

S8
R = 0.75

RT 1900 0.7 1.25 0.25 0.5 1500

Fig. 7. Intra-skeletal relationships between bone uoride and strontium content for all skeletons S1, S3, S6, S7, S8 and S9. The correlation coefcient, R, of skeleton S8 was obtained after exclusion of the right tibia (RT).

1.4 1.2 % F human bones 1

actual population USA Zipkin et al., 1958

1.4 Chupicuaro population


ante mortem values post mortem enrichment

skeletal fluorosis stages 1.2 1


max

cold waters 0.8

hot waters

JR24
S9

0.8
mean

0.6 S7 0.4 0.2 0 0 1 2 3 4 ppm F drinking water modern S6 S8 cow bone S1 S3

0.6 0.4 0.2


preclinic I II III

min

Fig. 8. Comparison of the uoride content of Chupicuaro bones and modern bones (USA). Grey circles represent uoride content from individuals whose drinking waters uoride content and time residence are known (Zipkin et al., 1958). Fluoride content of the waters of Puruagita (cold and hot) is indicated. Skeletal uorosis stages associated with iliac crest or pelvis uoride content are also shown (see a review of the Committee on Fluoride in Drinking Water/National Research Council, 2006). It is worth noting that the lowest value in stage III was known to be associated to hypocalcemia or secondary hyperparathyroidism.

0.0070

bone Sr/Ca pure hydrothermal product consumer calculated for 0.2< ORSr <0.3
0.3

0.0108

0.0074

0.0077

S1
ante mortem values post mortem enrichment

S7
hydrothermal water
0.0060

max

S9 S3
0.3

mean

bone Sr / Ca

0.0050

min

0.0040

hydrothermal carbonates Puruaguita

0.2

S6 S8

0.3
river water
0.0030

0.2

modern cow bone


bone Sr/Ca measured in the human skeletons and a modern cow bone from site JR24

0.0020

wells

0.2
0.0003/4

Fig. 9. Comparison of Sr/Ca ratios from Chupicuaro bones and a known pure hydrothermal product consumer. Ratios have been calculated for ORSr (Observed Ratio) values comprised between 0.2 and 0.3 and taking into account the average strontium and calcium content of each hydrothermal product (see Table 3).

hydrothermal deposits Aguas Calientes

% F human bones

A.-F. Maurer et al. / Journal of Archaeological Science 38 (2011) 1784e1797

1795

Sr is a non-essential element that is also mainly stored in the skeleton (Staub et al., 2003). With chemical properties similar to those of Ca, it tends to follow the same biological pathways. The preferential absorption and retention of Ca compared to Sr through the gastro-intestinal tract and the kidneys (Comar et al., 1957) results in the reduction of bone Sr/Ca through the trophic chain at each tier (Burton et al., 1999). This reduction is quantied by an Observed Ratio ORSr (Sr/Cabone/Sr/Cadiet). Using ORSr it is possible to predict the bone Sr/Ca content for a consumer of pure hydrothermal products, taking into account the average Sr and Ca value of each ingredient (Fig. 9). ORSr has been determined for modern mammals and in laboratory controlled experiments, where it generally ranges between 0.20 and 0.30 (Comar et al., 1957; Sillen and Kavanagh, 1982; Price et al., 1985, 1986; Balter, 2004). Reported values in humans are typically between 0.25/0.23 (Bryant and Loutit, 1961; Burton and Mercer, 1962), 0.18 (Alexander et al., 1956; Rivera and Harley, 1965; Sillen and Kavanagh, 1982) and 0.12 (Tanaka et al., 1981). Using ORSr between 0.2e0.3, most of the Chupicuaro skeletons display bones Sr/Ca ratios compatible with the consumption of hydrothermal products from Puruagita (water and carbonates) (Fig. 9). Conversely, hydrothermal deposits from Aguas Calientes and cold waters are completely excluded from the Chupicuaro diet. A distinct behaviour is observed in skeletons S7 and S1, whose higher Sr/Ca ratios do not appear to be related to the intake of hydrothermal products. 4.3. Variations within the Chupicuaro population Despite the limited number of individuals studied, three groups can be distinguished: the two males S8 and S6; the females S3 and S9 and the two other females S1 and S7. Males S6 and S8 display the lowest bone Sr and F values (Figs. 8 and 9). This is likely due to physiological differences between the sexes rather than diet. Sr and F are mainly stored in bones and the heavier skeleton weight of the

males (Warren and Maples, 1997) could explain the lower F and Sr values. Such differences (for F) have already been observed in a modern population (Arnala et al., 1985). Furthermore, pregnancy and breastfeeding can cause ORSr variations (Kostial et al., 1969; Blanusa et al., 1970; Blakely, 1989). No sex related differences were observed in the Li content of the bones. Among the females, the skeletons S1 and S7 display particularly interesting patterns. Bones from individual S1 yield the highest Sr/ Ca ratios (Fig. 9) and the lowest d13C value (Table 4). A temporocultural or individual dietary preference is inferred from S1, dated to the more recent Mixtlan phase. Its carbon isotope signature (3.2&) suggests a moderate addition of C3 plant material in the diet compared to the other individuals (1.6& in average), who relied almost exclusively on C4 plants when considering a mean value of 13& for C4 plants and an isotopic spacing between bone and diet of around 12e11& (Hare et al., 1991; Howland et al., 2003). A more important consumption of the common bean could have lowered S1 bone d13C value compared to the other individuals. This plant is known to have been present for at least 2300 years at Tehuacan (Kaplan and Lynch, 1999). The more important consumption of C3-eating animals, as suggested by the numerous faunal remains, worked bone industry and obsidian projectile points in the Mixtlan archaeological levels (Darras and Faugre, 2000, 2001, 2010), would also decrease bone d13C value. However, as S1 is the only skeleton available for this period, the evolution of dietary habits from the Chupicuaro to Mixtlan times cannot be conclusively determined. The S7 skeleton is also characterized by high F, Sr and Li content (Figs. 8 and 9, Table 4), with a similar bone d13C to that of the other individuals. Since diagenesis does not seem to have signicantly affected this skeleton, it is difcult to explain this pattern. Lines of arrested growth were detected on the tibias (Fig. 10). Episodic consumption of hydrothermal products may account for this signature, assuming there is an association between the growth arrest lines and the bone elemental content. However, bones record the average of the overall dietary signal during the last years of life, diluting individual dietary episodes. Furthermore, the Chupicuaro were constantly exposed to hydrothermal products and if they included these products in their diet, it was very likely on a daily basis. If this is the case, the lines of arrested growth would therefore be independent of the bone elemental content of this skeleton, whose composition remains enigmatic. 5. Conclusion Archaeochemistry is useful to recognize human practices and way of life of past populations. However, before exploring such interests, post-mortem chemical changes must be estimated. Bone mineralogical and structural analyses (CI, Ca/P, OM content, secondary calcite and histology) are essential for understanding post-mortem changes, although these proxies alone, are not sufcient to evaluate a skeletons state of geochemical preservation. All of the individuals analysed in this study were affected by mineralogical and structural post-mortem changes which vary at the intra and inter-skeletal scales. All of the skeletons were also affected by geochemical post-mortem modications, but a widespread dissolution-recrystallisation process seems unlikely in this context. A thorough study of the diagenetic skeletal histories of the hydrothermal tracers, F, Sr and Li (whose retention in living tissues is determined by the amount ingested) shows that the progressive intra-skeletal post-mortem Sr and F enrichment is probably due to diffusion-adsorption. We assume that some skeletal parts are better preserved than others. Therefore, the lowest F and Sr intraskeletal concentrations, best represent the biological values. In contrast, it is unlikely that the high Li concentrations found are due to diagenetic addition.

Fig. 10. Two horizontal lines on the femoral diaphysis of the individual S7 that evidence stress event or metabolic disequilibrium.

1796

A.-F. Maurer et al. / Journal of Archaeological Science 38 (2011) 1784e1797 Blakely, R.L., 1989. Bone strontium in pregnant and lactating females from archaeological samples. American Journal of Physical Anthropology 80, 173e185. Blanusa, M., Harmut, M., Kostial, K., 1970. Comparative strontium and calcium metabolism in lactation. Archives of Industrial Hygiene and Toxicology 21, 125e127. Bressani, R., Paz y Paz, R., Scrimshaw, N.S., 1958. Chemical changes in corn during preparation of tortillas. Journal of Food Chemistry 6, 770e774. Bryant, F.J., Loutit, J.F., 1961. Human Bone Metabolism Deduced from Strontium Assays United Kingdom AEA Document, AERE-R-3718. Bryant, J.D., Luz, B., Froelich, P.N., 1994. Oxygen isotopic composition of fossil horse tooth phosphate as a record of continental paleoclimate. Palaeogeography, Palaeoclimatology, Palaeoecology 107, 303e316. Burton, J.D., Mercer, E.R., 1962. Discrimination between strontium and calcium in their passage from diet to the bone of adult man. Nature 193, 846e847. Burton, J.H., Price, T.D., Middleton, W.D., 1999. Correlation of bone Ba/Ca and Sr/Ca due to biological purication of calcium. Journal of Archaeological Science 26, 609e616. Cerklewski, F.L., 1997. Fluoride bioavailability. Nutritional and clinical aspects. Nutrition Research 17, 907e929. Chang, L.L.Y., Howie, R.A., Zussman, J., 1996. Rock Forming Minerals. Volume 5B. Non-silicates: Sulphates, Carbonates, Phosphates, Halides, second ed. Geological Society of London. Comar, C.L., Russell, R.S., Wasserman, R.H., 1957. Strontiumecalcium movement from soil to man. Science 126, 485e492. Combs, G.F.J., 2005. Geological impacts on nutrition. In: Selinus, O., Alloway, B., Centeno, J.A., Finkelman, R.B., Fuge, R., Lindh, U., Smedley, P. (Eds.), Essentials of Medical Geology. Elsevier, pp. 161e177. Committee on Fluoride in Drinking Water/National Research Council, 2006. Fluoride in Drinking Water: A Scientic Review of EPAs Standards. National Academic Press, Washington D.C. DAngela, D., Longinelli, A., 1990. Oxygen isotopes in living mammals bone phosphate: further results. Chemical Geology 86, 75e82. Dansgaard, W., 1964. Stable isotopes in precipitation. Tellus 16, 436e468. Darras, V., 2004. Etude de sources thermales prcolombiennes: incidences conomiques et physiologiques dun complexe hydrothermal sur les populations de Chupicuaro, Guanajuato, Mexique, (priodes prclassique moyenne et rcente, 600 avt. J.C.-300 ap. J.C.). Final Report, UMR 8096-ATIP Jeunes Chercheurs, SHSCNRS. Darras, V., Faugre, B., 2000. Proyecto Dinmicas Culturales en el Bajo, Guanajuato e CHUPICUARO. Informe sobre los trabajos arqueolgicos realizados en febrero y marzo de 2000 en los sitios JR 24, JR 15, JR 33 y JR 17. MAE/CNRS. Presentado al Consejo de Arqueologa del Instituto Nacional de Antropologa. Darras, V., Faugre, B., 2001. Proyecto y Dinmicas Culturales en el Bajo, Guanajuato e CHUPICUARO. Informe sobre los trabajos realizados durante los meses de jenero, marzo y abril de 2001 en la zona de Puruagita, Municipio de Jercuaro, Guanajuato MAE/CNRS. Darras, V., Faugre, B., 2005. Cronologa de la cultura Chupcuaro: Estudio del sitio La Tronera, Puruaguita, Guanajuato. In: Williams, E., Weigand, P., Lopez Mestas, L., Grove, D. (Eds.), El Antiguo Occidente de Mxico. Colegio de Michoacn, pp. 255e282. Darras, V., Faugre, B., 2010. Reacomodos culturales en el valle de Acmbaro al nal del Formativo: La fase Mixtln y su signicado local y global, en El eje LermaSantiago durante el Formativo Terminal y el Clsico Temprano. In: Memoria del Segundo Seminario Taller sobre Problemticas Regionales. INAH, Mxico, pp. 269e301. Darras, V., Faugre-Kalfon, B., Durlet, C., Liot, C., Reveles, J., Berumen, R., Cervantes, O., Caillaud, C., David, C., 1999. Nouvelles recherches sur la culture de Chupicuaro (Guanajuato, Mexique). Journal de la Socit des Amricanistes 85, 343e351. de Boer, J.L.M., Verweij, W., van der Velde-Koerts, T., Mennes, W., 1996. Levels of rare earth elements in Dutch drinking water and its sources. Determination by inductively coupled plasma mass spectrometry and toxicological implications. A pilot study. Water Research 30, 190e198. Delgado Huertas, A., Iacumin, P., Stenni, B., Chilln, B.S., Longinelli, A., 1995. Oxygen isotope variations of phosphate in mammalian bone and tooth enamel. Geochimica et Cosmochimica Acta 59, 4299e4305. Ehleringer, J.R., 1989. Carbon isotope ratios and physiological processes in aridland plants. In: Rundel, P.W., Ehleringer, J.R., Nagy, K.A. (Eds.), Stable Isotopes in Ecological Research. Springer, New York, pp. 41e54. Ekstrand, J., Alvn, G., Borus, L., Norlin, A., 1977. Pharmacokinetics of uoride in man after single and multiple oral doses. European Journal of Clinical Pharmacology 12, 311e317. Elias, R.W., Hirao, Y., Patterson, C.C., 1982. The circumvention of the natural biopurication of calcium along nutrient pathways by atmospheric inputs of industrial lead. Geochimica et Cosmochimica Acta 46, 2561e2580. Fabig, A., Herrmann, B., 2002. Trace elements in buried human bones: intra-population variability of Sr/Ca and Ba/Ca ratios e diet or diagenesis? Naturwissenschaften 89, 115e119. Food and Nutrition Board, Institute of Medicine, 1997. Dietary Reference Intakes for Calcium, Phosphorus, Magnesium, Vitamin D, and Fluoride. National Academy Press, Washington, D.C. Food and Nutrition Board, Institute of Medicine, 2001. Dietary Reference Intakes for Vitamin A, Vitamin K, Arsenic, Boron, Chromium, Copper, Iodine, Iron, Manganese, Molybdenum, Nickel, Silicon, Vanadium, and Zinc. National Academy Press, Washington, D.C.

This intra-skeletal study is useful to reconstruct the individual history, ante and post-mortem of individuals. Furthermore, the comparison of values extracted from the skeletons with the elemental content of hydrothermal waters and carbonates available in the Chupicuaro environment, demonstrates that the Chupicuaro human skeletons record a hydrothermal signal very likely to be due to the consumption of some of these products. However, it is impossible to know whether this hydrothermal diet was consumed directly or indirectly and if this diet was a result of cultural habits or environmental resource availability. Acknowledgements This work was funded by the ACI 67053 and APN 2000 (SHSCNRS). The Chupicuaro Archaeological Project was supported by the French Ministry of Foreign and European Affairs (MAEE), the Centre dtudes mexicaines et centramricaines (CEMCA) in Mexico, and the Centre National de la Recherche Scientique (CNRS). We wish to thank the Institut de Recherche pour le Dveloppement (IRD, Bondy Paris) for the use of their facilities. Many thanks are also due to N. Labourdette (Paris VI), F. Delbes (Paris VI) and F. Lecornec (IRD Bondy) for their technical assistance and to J-F. Salige (Paris VI) and Miranda Jans (VU Amsterdam) for the fruitful discussions about thermal pre-treatment and histology, respectively. A. Zazzo (Musum National dHistoire Naturelle, MNHN, Paris), E. Nunn (University of Mainz) and M. Elliott (MAE, Nanterre) are also greatly acknowledged for their constructive comments. Lastly, we thank an anonymous reviewer for suggestions for improvement of the manuscript. Appendix A Supplementary data related to this article can be found online at doi:10.1016/j.jas.2011.03.008. References
Alexander, G.V., Nusbaum, R.E., MacDonald, N.S., 1956. The relative retention of strontium and calcium in bone tissue. Journal of Biological Chemistry 218, 911e919. Ambrose, S.H., Norr, L., 1993. Experimental evidence for the relationship of the carbon isotope ratios of whole diet and dietary protein to those of bone collagen and carbonate. In: Lambert, J.B., Grupe, G. (Eds.), Prehistoric Human Bone: Archaeology at the Molecular Level. Springer-Verlag, Berlin, pp. 1e37. Anke, M., 1993. Lithium. In: Macrae, R., Robinson, R.K., Sadler, M.J. (Eds.), Encyclopaedia of Food Science, Food Technology and Nutrition. Academic Press, London, pp. 2779e2782. Anke, M., Angelow, L., Arnhold, W., Mller, R., Schfer, U., 2005. Lithium and rubidium in the food chain, intake by man, essentiality and toxicity. In: Satellite Symposium on Lithium, 5th International Symposium on Trace Elements in Human: New Perspectives, Athens, Greece. Arnala, I., Alhava, E.M., Kauranen, P., 1985. Effects of uoride on bone in Finland. Histomorphometry of cadaver bone from low and high uoride areas. Acta Orthopaedica Scandinavica 56, 161e166. Badone, E., Farquhar, R.M., 1982. Application of neutron activation analysis to the study of element concentration and exchange in fossil bones. Journal of Radioanalytical and Nuclear Chemistry 69, 291e311. Balter, V., 2004. Allometric constraints on Sr/Ca and Ba/Ca partitioning in terrestrial mammalian trophic chains. Oecologia 139, 83e88. Balter, V., Bocherens, H., Person, A., Labourdette, N., Renard, M., Vandermeersch, B., 2002. Ecological and physiological variability of Sr/Ca and Ba/Ca in mammals of West European mid-Wrmian food webs. Palaeogeography, Palaeoclimatology, Palaeoecology 186, 127e143. Barrientos, I., del Carmen Ruiz, P., 2009. Aproximacion a la historia biologica de las poblaciones Chupicuaro. Analisis bioantropologicos, huellas de actividad, y paleopatologias (sitios JR24 y TR6). In: 53 ICA International congress of Americanists. Bauer, G.C.H., Carlsson, A., Lindquist, B., 1956. A comparative study on the metabolism of 140Ba and 45Ca in rats. Biochemical Journal 63, 535e542. Bell, L.S., Skinner, M.F., Jones, S.J., 1996. The speed of post mortem change to the human skeleton and its taphonomic signicance. Forensic Science International 82, 129e140.

A.-F. Maurer et al. / Journal of Archaeological Science 38 (2011) 1784e1797 Food and Nutrition Board, Institute of Medicine, 2004. Dietary Reference Intakes for Water, Potassium, Sodium, Chloride, and Sulfate. National Academy Press, Washington, D.C. F.P.T. Committee on Drinking Water, 2008. Guidelines for Canadian Drinking Water Quality e Summary Table. www.healthcanada.gc.ca/waterquality. Garcia, J., 2002. Informe sobre la revisin del material seo procedente del sitio La Tronera (municipio de Jercuaro, Guanjuato). Estudio antropolgico y de las patologas. Proyecto arqueolgico Dinmicas Culturales en el Bajio e Chupicuaro. CEMCA, Mxico. Hare, P., Fogel, M., Stafford, T., Mitchell, A., Hoering, T., 1991. The isotopic composition of carbon and nitrogen in individual amino acids isolated from modern and fossil proteins. Journal of Archaeological Science 18, 277e292. Hedges, R.E.M., Millard, A.R., Pike, A.W.G., 1995. Measurements and relationships of diagenetic alteration of bone from three archaeological sites. Journal of Archaeological Science 22, 201e209. Hinners, T.A., Hugues, R., Outridge, P.M., Davis, W.J., Simon, K., Woolard, D.R., 1998. Interlaboratory comparison of mass spectrometric methods for lead isotopes and trace elements in NIST SRM 1400 bone ash. Journal of Analytical Atomic Spectrometry 13, 963e970. Howland, M., Corr, L., Young, S., Jones, V., Jim, S., van der Merwe, N., Mitchell, A., Evershed, R., 2003. Expression of dietary isotope signal in the compoundspecic d13C values of pig bone lipids and amino acids. International Journal of Osteoarchaeology 13, 54e65. Hubert, J.F., Panish, P.T., Chure, D.J., Prostak, K.S., 1996. Chemistry, microstructure, petrology and diagenetic model of Jurassic dinosaur bones, dinosaur national monument, Utah. Journal of Sedimentary Research 66, 531e547. Kaplan, L., Lynch, T., 1999. Phaseolus (Fabaceae) in archaeology: AMS radiocarbon dates and their signicance for pre-Columbian agriculture. Economic Botany 53, 261e272. Katz, S.H., Hediger, M.L., Valleroy, L.A., 1974. Traditional maize processing techniques in the new world. Science 184, 765e773. Keplinger, L.L., Kuhn, J.K., Williams, W.S., 1986. An elemental analysis of archaeological bone from Sicily as a test of predictability of diagenetic change. American Journal of Physical Anthropology 70, 325e331. Kohn, M.J., 2008. Models of diffusion-limited uptake of trace elements in fossils and rates of fossilization. Geochimica et Cosmochimica Acta 72, 3758e3770. Kosla, T., Anke, M., 2005. Lithium content in the environment of Poland and its transmission in the food chain: soileplanteanimal tissues. Satellite Symposium on Lithium, Athens, Greece, pp. 176e186. Kostial, K., Gruden, N., Durakovic, A., 1969. Intestinal absorption of calcium-47 and strontium-85 in lactating rats. Calcied Tissue Research 4, 13e19. Lonard, A., Hantson, P., Gerber, G.B., 1995. Mutagenicity, carcinogenicity and teratogenicity of lithium compounds. Mutation Research: Reviews in Genetic Toxicology 339, 131e137. Lindh, U., 2005. Uptake of elements from a biological point of view. In: Selinus, O., Alloway, B., Centeno, J.A., Finkelman, R.B., Fuge, R., Lindh, U., Smedley, P. (Eds.), Essentials of Medical Geology. Elsevier, pp. 87e114. Longinelli, A., 1984. Oxygen isotopes in mammal bone phosphate: a new tool for paleohydrological and paleoclimatological research? Geochimica et Cosmochimica Acta 48, 385e390. Luz, B., Kolodny, Y., 1985. Oxygen isotope variations in phosphate of biogenic apatites. IV. Mammal teeth and bones. Earth and Planetary Science Letters 75, 29e36. Marshall, J.H., Liniecki, J., Lloyd, E.L., Mariotti, G., Mays, C.W., Rundo, J., Sissons, H.A., Snyder, W.S., 1973. Alkaline earth metabolism in adult man. Health Physics 24, 125e221. Mikrut, D., 2003. Provenance dobjets archologiques? Recherches sur lenvironnement hydrothermal comme signature dun site de production (Culture Chupicuaro, Mexique), Master thesis, University Paris VI. Millard, A.R., Hedges, R.E.M., 1995. The role of the environment in uranium uptake by buried bone. Journal of Archaeological Science 22, 239e250. Millard, A.R., Hedges, R.E.M., 1996. A diffusion-adsorption model of uranium uptake by archaeological bone. Geochimica et Cosmochimica Acta 60, 2139e2152. Munro, L.E., Longstaffe, F.J., White, C.D., 2007. Burning and boiling of modern deer bone: effects on crystallinity and oxygen isotope composition of bioapatite phosphate. Palaeogeography, Palaeoclimatology, Palaeoecology 249, 90e102. Munro, L.E., Longstaffe, F.J., White, C.D., 2008. Effects of heating on the carbon and oxygen-isotope compositions of structural carbonate in bioapatite from modern deer bone. Palaeogeography, Palaeoclimatology, Palaeoecology 266, 142e150. Neuman, W.F., Neuman, M.W., 1958. The Chemical Dynamics of Bone Mineral. The University of Chicago Press, Chicago. Niederberger, C., 1976. Zohapilco. Cinco milenios de ocupacin humana en un sitio lacustre de la cuenca de Mxico. In: Coleccin Cientca, vol. 30. INAH, Mxico. Nielsen-Marsh, C.M., Hedges, R.E.M., 2000. Patterns of diagenesis in bone I: the effects of site environments. Journal of Archaeological Science 27, 1139e1150. OLeary, M.H., 1981. Carbon isotope fractionation in plants. Phytochemistry 20, 553e567. Person, A., Bocherens, H., Saliege, J.F., Paris, F., Zeitoun, V., Gerard, M., 1995. Early diagenetic evolution of bone phosphates: an X-ray diffractometry analysis. Journal of Archaeological Science 22, 211e221. Person, A., Bocherens, H., Mariotti, A., Renard, M., 1996. Diagenetic evolution and experimental heating of bone phosphate. Palaeogeography, Palaeoclimatology, Palaeoecology 126, 135e149. Posner, A., 1987. Bone mineral and the mineralization process. In: Peck, W.A. (Ed.), Bone and Mineral Research, pp. 65e116.

1797

Price, T.D., Connor, M., Parsen, J.D., 1985. Bone chemistry and the reconstruction of diet: strontium discrimination in white-tailed deer. Journal of Archaeological Science 12, 419e442. Price, T.D., Swick, R.W., Chase, E.P., 1986. Bone chemistry and prehistoric diet: strontium studies of laboratory rats. American Journal of Physical Anthropology 70, 365e375. Reiche, I., Favre-Quattropani, L., Vignaud, C., Bocherens, H., Charlet, L., Menu, M., 2003. A multi-analytical study of bone diagenesis: the Neolithic site of Bercy (Paris, France). Measurement Science and Technology 14, 1608e1619. Rivera, J., Harley, J.H., 1965. The HASL Bone Program: 1961e1964. United States Atomic Energy Commission Health and Safety Laboratory Report, vol. 163. Salige, J.F., Person, A., Paris, F., 1995. Preservation of 13C/12C original ratio and 14C dating of the mineral fraction of human bones from Saharan tombs, Niger. Journal of Archaeological Science 22, 301e312. Schou, M., 1968. Lithium in psychiatric therapy and prophylaxis. Journal of Psychiatric Research 6, 67e95. Schrauzer, G.N., 2002. Lithium: occurrence, dietary intakes, nutritional essentiality. Journal of the American College of Nutrition 21, 14e21. Schwarcz, H.P., 2000. Some biochemical aspects of carbon isotope paleodiet studies. In: Ambrose, S., Katzenberg, M.A. (Eds.), Biogeochemical Approaches to Paloedietary Analysis. Kluwer/Plenum, New York, pp. 189e208. Shiraishi, K., 2005. Dietary intakes of eighteen elements and 40 K in eighteen food categories by Japanese subjects. Journal of Radioanalytical and Nuclear Chemistry 266, 61e69. Sillen, A., Kavanagh, M., 1982. Strontium and paleodietary research: a review. Yearbook of Physical Anthropology 25, 67e90. Sillen, A., Parkington, J.E., 1996. Diagenesis of bones from Elands Bay Cave. Journal of Archaeological Science 23, 535e542. Simon, F.G., Biermann, V., Peplinski, B., 2008. Uranium removal from groundwater using hydroxyapatite. Applied Geochemistry 23, 2137e2145. Smith, C.I., Nielsen-Marsh, C.M., Jans, M.M.E., Collins, M.J., 2007. Bone diagenesis in the European Holocene I: patterns and mechanisms. Journal of Archaeological Science 34, 1485e1493. Sponheimer, M., de Ruiter, D, Lee-Thorp, J., Spth, A., 2005. Sr/Ca and early hominin diets revisited: new data from modern and fossil tooth enamel. Journal of Human Evolution 48, 147e156. Staub, J.F., Foos, E., Courtin, B., Jochemsen, R., Perault-Staub, A.M., 2003. A nonlinear compartmental model of Sr metabolism. II. Its physiological relevance for Ca metabolism. American Journal of Physiology e Regulatory, Integrative and Comparative Physiology 284, R835eR852. Tanaka, G., Kawamura, H., Nomura, E., 1981. Reference Japanese man-II. Distribution of strontium in the skeleton and in the mass of mineralized bone. Health Physics 40, 601e614. Tapiero, H., Tew, K.D., 2003. Trace elements in human physiology and pathology: zinc and metallothioneins. Biomedicine and Pharmacotherapy 57, 399e411. Tieszen, L., Fagre, T., 1993. Effect of diet quality and composition on the isotopic composition of respiratory CO2 bone collagen, bioapatite, and soft tissues. In: Lambert, J.B., Grupe, G. (Eds.), Prehistoric Human Bone: Archaeology at the Molecular Level. Springer-Verlag, Berlin, pp. 121e155. Toots, H., Voorhies, M.R., 1965. Strontium in fossil bones and the reconstruction of food chains. Science 149, 854e855. Trueman, C.N., Martill, D.M., 2002. The long-term survival of bone: the role of bioerosion. Archaeometry 44, 371e382. Trueman, C.N., Tuross, N., 2002. Trace elements in recent and fossil bone apatite. In: Kohn, M.J., Rakovan, J., Hugues, K.M. (Eds.), Phosphates: Geochemical, Geobiological and Materials Importance. Mineralogical Society of America, vol. 48, pp. 489e521. US EPA Environmental Protection Agency, 2009. National Primary Drinking Water Standards. 816-F-09e004. www.epa.gov/safewater/contaminants/index.html. Warren, M.W., Maples, W.R., 1997. The anthropometry of contemporary commercial cremation. Journal of Forensic Sciences 42, 417e423. Weiner, S., Bar-Yosef, O., 1990. States of preservation of bones from prehistoric sites in the near east: a survey. Journal of Archaeological Science 17, 187e196. Weiner, S., Traub, W., 1992. Bone structure: from angstroms to microns. The FASEB Journal 6, 879e885. Whitford, G.M., 1996. The Metabolism and Toxicity of Fluoride, second ed. Karger, Basel. WHO/IPCS, 2002. Principles and methods for the assessment of risk from essential trace elements. Environmental Health Criteria 228. Williams, C.T., Potts, P.J., 1988. Element distribution maps in fossil bones. Archaeometry 30, 237e247. Yoshinaga, Y., Suzuki, T., Morita, M., Hayakawa, M., 1995. Trace elements in ribs of elderly people and elemental variation in the presence of chronic diseases. The Science of the Total Environment 162, 239e252. Yoshino, M., Kimijima, T., Miyasaka, S., Sato, H., Seta, S., 1991. Microscopical study on estimation of time since death in skeletal remains. Forensic Science International 49, 143e158. Zaldivar, R., 1980. High lithium concentrations in drinking water and plasma of exposed subjects. Archives of Toxicology 46, 319e320. Zazzo, A., Salige, J.F., Person, A., Boucher, H., 2009. Radiocarbon dating of cremated bones: where does the carbon come from? Radiocarbon 51, 601e611. Zipkin, I., McClure, F.J., Leone, N.C., Lee, W.A., 1958. Fluoride deposition in human bones after prolonged ingestion of uoride in drinking water. Public Health Reports 73, 732e740.

You might also like