You are on page 1of 104
Precolumbian Jade 2 Compositional and Structural Characterization of Maya and Costa Rican Jadeitites RONALD L. BISHOP EDWARD V. SAYRE, JOAN MISHARA. bjects carved in jade were highly prized in both the Maya region of O Mesoamerica and in Costa Rica, as is attested by their artistic embellish- ‘ment and presence in funerary contexts. As jade is limited by its geological conditions of occurrence, the determination of the directions in which it moved through precolumbian societies provides information on ancient social linkages. The term jade has been used to describe a variety of different minerals from the time of the Spanish Conquest (Foshag 1957). In general, it refers to nephrite, an amphibole; or to jadeite, a pyroxene. Nephrite (a compact, inter- felted crystalline aggregate of the tremolite-actinolite series) is not known to occur in Mesoamerica nor has it been documented in Costa Rica, although actinolite was recovered during our exploration of the metamorphic rocks of the Santa Elena Peninsula in Costa Rica. Although a wide variety of stones have been carved in both regions, the finest carving seems to have been de- voted to jadeite and albite. In an earlier investigation of jade in Maya civilization, Hammond and his associates (1977) discussed Rathje’s (r970) views concerning the need for a de- veloping Maya society to have a greater social base from which “elite func tionaries” could be recruited. To move up socially into narrowing elite posi- tions required various forms of “wealth,” signs of which were translated into “wealth” in grave goods. As the Maya elite grew progressively more restricted, grcater quantities of status goods, such as jade, were concentrated within the hands of the members of the upper echelons, Hammond and his co-workers note that this model concerns only the final distribution of jade, without in- corporating information regarding the procurement or exchange of the mater- 3e Ronald L. Bishop, Eduvard V. Sayre, and Joan Mishara ial; further, they comment that the model may be significantly biased by the history of excavation in the Maya region, which focused on major ceremonial centers and ritual or elite contexts (197737) If a growing elite segment exercised greater demands for visible signs of wealth and power, it is reasonable to expect that these consumptive demands were accompanied by increasingly “directed” procurement of raw materials. Control over and organized exploitation of the resource areas for elite goods sould be a reasonable means of maintaining a supply of the signs of status. Knowledge of the specific procurement source or sources of the materials and the extent to which access to the source(s) was restricted to different Maya cen- ‘ers would inform on economic and ritual aspects of Maya society. In order to obtain data that would contribute toward understanding the ature of jade exploitation and movement, we began a compositional investi- gation of Maya jade, employing Instrumental Neutron Activation Analysis (NAA) and focusing especially on predominantly jadeite-bearing source rocks ‘and on artifacts. A geographical focus for source characterization was the Mo- agua Valley of Guatemala, with its known source of jadeite. Ultimately, our sampling of artifactual material was extended beyond the Maya region to in- clude jadeitie specimens from Costa Rica. Boston’s Muscum of Fine Arts and the Department of Chemistry at Brookhaven National Laboratory cooperated in our research; for example, samples taken from materials were studied by X-ray diffraction at either facil- ‘gy. The conservation staff of the Rescarch Laboratory at the Muscum of Fine “Arts was frequently called upon to restore jades that had been sampled. With she exception of the extensive program of X-ray diffraction examination and the microprobe analyses of selected specimens, the majority of the analytical ‘data were in hand in the early 1980s when Bishop and Sayre joined the Con- secvation Analytical Laboratory (CAL) of the Smithsonian Institution. Inter- ‘sittent statistical analysis of the chemical data obtained at Brookhaven made ‘ebvious the fact that we needed additional mineralogical and structural data. “This task was undertaken by Joan Mishara, scientist with CAL. (CHARACTERIZATION acterization of materials has been a fundamental activity in archaeology ‘sed art history, used to address such questions as the existence of extent of rial exchange among societies, the development of technical processes and stylistic traditions, and the nature of the cultural and technical impact of society on another. As used here, characterization refers to the identifica- ‘of those parameters or properties within an object that allow one to at- te itto a particular cultural or geographic source. Different approaches can provide varying types of analytical parameters: example, stylistic, technological, morphological, structural, and composi- | One might think that the last-named approach, compositional analysis, 3 2 PRECOLUMBLAN JADE has a brief history because of its relatively recent intensive use in archacologi- cal investigations. In fact, chemical characterization was used in the 1800s to study the nature of jade celts associated with Celtic monuments (Damour 1865). In a review of archaeological chemistry Harbottle (1982:14) credits Damour as an early advocate of an “interdisciplinary” investigation; he recog nized that linking the jades recovered from tombs to their sources would pro- vide information on the presumed migrations of ancient peoples. Benefiting from developments in analytical instrumentation and computerized data analysis, compositional analysis has become an essential approach in address- ing archaeological problems concerning extraction, production, and distribu- tion of a broad series of material remains: metals including copper, bronze, sil- ver, gold, and lead; pottery and clays; natural products such as amber, seashells, ivory, and bone; glasses and faience; stones and mincrals including, sandstone, marble, limestone, chert and flint, soapstone, galena, turquoise, and obsidian (sec references in Harbottle 1982). Each of these listed materials, natural or man-made, presents its own unique analytical and interpretative problems. One of the “simplest” systems is that of obsidian, a material relatively uniform in composition over an an cient procurement area—and by extension, relatively homogeneous within the artifact sampled. In contrast, jades (jadcitites) from a common procurement area are found to be relatively compositionally heterogeneous. This miner- alogical hetcrogencity obfuscates ready interpretation of obtained chemical data. Even with such heterogeneity, however, we find that in many cases chemical analyses, especially those utilizing data on trace clements, permit the culturally interpretable differentiation of artifacts. This is possible because the variation among specimens from different sources is significantly greater than the variation among specimens from a single source. If we can characterize the raw material from specific procurement areas, the statistical group properties for these areas can stand as compositional refer- ences against which the subsequent comparison of artifacts can be made. In- terpretative strength is given to the attributions of an artifact to composition ally characterized sources through the additional patterning seen in nonchem- ical or nonmineralogical data such as site provenience, form, or style. Although we previously referred to the common distinction of “jade” on the basis of amphibole or pyroxene components—nephrite and jadeite—the situation is not that simple, Elsewhere the concept of social jade has been i voked as a heuristic device to call attention to the fact that it may not have been essential that a particular mineral was used for meeting the social re- quirements evidenced by the presence of an artifact (Lange and Bishop 1988). That is, potentially any of a range of “greenstones” would be socially accept- able as materials to be carved in some representation. Thus we find occurring within a mosaic mask, for example, both jadeite and albite as well as other grcenstones. It appears, however, that for a predominance of the Maya carved artifacts, ranging from light to brilliant green or grecn-black, the finest carv- Ronald L. Bishop, Edward V. Sayre, and Joan Mishara ing scems to have been reserved for the stones that are predominantly jadeite or closely related albite. When one views minerals represented in several of the large Costa Rican collections, one can note the duplication of forms and detailed attention to carving in a vast variety of stones. Whether or not the so- cial preference for jadeite or jadcite-albite was as strong in Costa Rica as it seems to have been in the Maya region is unknown, Understanding the mineralogical and geochemical variation associated swith jadeite is enhanced if one considers not just jadeite but rather jadcitite— ocks in which impure jadeite is found (Duncan 1986:7; Chapter 1), Jadeitites, as found in Guatemala, contain sodium pyroxenes, including jadeite, most fre- quently occurring with other minerals as major constituents—such as albite, muscovite-paragonite, and sphene. Jadeitites evolve in metamorphic environments in which both crystallo- chemical effects and stability in solutions influence the chemical composition of the mineral. Of the two factors, stability in solutions appears to be the most critical (Graf 1977; Morgan and Wandless 1980), with the size of the host cation influencing the rate of replacement. For this reason, the pattern of re placement in jadeite would be similar for the mineral albite with which it is in frequent association; jadeite and albite are often found to be interconvert= able during metamorphism. Within a given solution, therefore, the rare-earth clements, with their generally predictable geochemical behavior, would sub- stitute similarly in both minerals. Such similarity of substitution behavior may be true for other elements. From this standpoint, patterned, demonstra- ble differences in elemental concentration suggest that the specimens being analyzed were derived from different hydrothermal environments, To the extent that this holds true, an analytical technique capable of sufficient sensi- tivity to detect very small elemental concentrations, and quantified with ade- quate precision, offers the opportunity for characterizing source and artifac- tual jade sogenesis of the jadeitites or jadeite, as these are found in the works of Dun- can (1986) and Harlow (Chapter 1), among others (Harlow notes additional complexities involved in the formation of jadeitites caused by mechanical mixing of components). Our comments are to call attention to the fact that the mineralogy and geochemistry of jadeitic materials are highly complex and expectedly variable within our understanding of the formation processes. An attempt to characterize the jadeitic rocks of the Motagua River valley and to compare the measured parameters with those obtained from measure- ments made on artifacts must operate within the limits of the variation that exists in nature. One of the major contributions of our work is to have begun, to set limits on the extent of chemical variation within the jadeitites of the Motagua Valley. We hope that geologists will find our data of theoretical in- tcrest for their own work. Our approach to understanding chemical variation in the jadeitites, for the purposes of archaeological jadeitite characterization, however, is empirical. ite materials. It is not our purpose here to provide details of the pet- 33 4 FiouRe2.1 Locations of sampled artifacts, Motagun PRECOLUMBIAN JADE se ary come retideada SELIZE ‘ean. een uarémaa, ade tees MATERIAL SAMPLING The only source of jadeitite in Central America whose location is presently known lies along the Motagua River in southern Guatemala (Figure 2.1). More-detailed discussions of the source region are presented in this volume by Harlow (Chapter 1) and Garza-Valdés (Chapter 7). Bishop collected 75 rock specimens from this source region, of which approximately half were predom- inantly jadeite. These were collected from an area extending some 25 km along the river and north from the river about five kilometers up the valley of a broadly cut, small tributary to the Motagua. Two jadeitite specimens from this same general source region were supplied by carlier collection efforts of Russcll Sykes and Charlotte Thomsen. Approximately 80 additional jadeitic samples were obtained from the extensive stockpiles of JADES, S.A., which is mining the source commercially. Ronald L. Bishop, Edward V. Sayre, and Joan Mishara As acknowledged in our previous discussion of the jade project (Bishop ‘etal. 1985), artifacts from the Maya area were selected primarily on the basis ‘of their availability. From the outset of the project, the Peabody Museum of ‘Harvard University graciously made its collections available for study, and we cooperated with its curators and conservators in the selection of those specimens that would undergo sampling for neutron activation. Similarly, ‘che Department of Archaeology of Belize, through the efforts of the regional archaeological specialists, and the Royal Ontario Museum were particularly clpful. ‘The number of samples analyzed were numerically skewed toward cen- eral and northern Belize, reflecting the availability of specimens from sites in ‘Cacilo, Nohmul, Altun Ha, Cahal Pech, and Santa Rita. Specimens ranged in ‘sme from the Early Preclassic through the Postclassic periods. In addition to he Belizean specimens, samples have been analyzed from the Cenote at Chichén Itz, Yucatin, Mexico; Copin and the El Cajén area of Honduras; ‘end the Salama Valley and Motagua Valley sites in Guatemala, In order to provide greater perspective on jadcitite compositional varia~ ‘Son, the project was eventually expanded to include artifacts from Costa Rica. ‘Peemission was obtained through the Department of Anthropology at the “Masco Nacional de Costa Rica to export jades for analysis from the museum ‘sed from the large collection of the Museo del Jade in the Instituto Nacional “de Seguros. The sample of Costa Rican specimens was selected on the basis of ‘secential for having jadeite as a major component, as determined by visual in- Secction; additional nonjadeitic specimens were exported for mineralogical “Sdeatification. A complete list of all jadeitites analyzed by neutron activation is ‘erovided in Appendices A-F at the end of this volume. ‘The willingness of all institutions involved to have the specimens cu- ‘excd in their collections included in a project that at times required limited “gescructive analysis contributed significantly to whatever knowledge this pro- pect advances. UASALYTICAL METHODS Hammond and others (1977) did in their earlier characterization work, we INAA, which was carried out at Brookhaven National Laboratory. Ss highly sensitive technique, capable of detecting certain elements below a per-million level, is based on the principle that when a target element's is exposed to neutrons, the element may be changed into a form that is ic or radioactive, The unstable atomic nucleus releases energy through ‘emission of radiation; the energy is specific to that atom, and the intensity seclated to the amount of the target element in the sample. We are concerned swith the emission of energy in the form of gamma rays. Although in the- ‘over two-thirds of naturally occurring elements are subject to activation measurement through INAA, their sensitivities vary greatly. Additional 35 38 PRECOLUMBIAN JADE TABLE 2.1 Comparison of Oxide Concentrations from the Light and Dark Fractions of Motagua Jadeitite Sample J S006 NiO $0; Fu,0; Lu,0, Yb;0; HO, Cr,0, Fe0, CoO POT PPM PPM PPM PPM PPM PPM PCT PPM. J S006 1629 1.20 0.029 0.023 0.098 2.15 100079158 JS006 622-955 On31 0.075 343 67.04.29 45.29 PCT ~ percentage; PPM = parts per million from two analyses of different phases from a Motagua region source sample (Table 2.13 Plate 1a). The first analysis, J $006, is from a sample taken in the light green portion of a cut jadeitite slab supplied by JADES, S.A. The sccond, J*8006, is from the green-black region of the slab. Even when analytical error is allowed for, pronounced differences in the elemental concentrations can be noted for a majority of the oxides. In the example of the slab just described, one can know whether or not the sample was taken from a mafic section of the specimen. Greater analytical ambiguity exists for the analysis of an entire specimen—a procedure followed extensively in the program of jade analysis. Although there is no sampling error introduced by the analysis of a whole artifact, the resulting composi- tional profile is a weighted combination of all mineralogical components, whether visible on the surface or not. A basic consideration, therefore, was the extent to which we could repeat- edly sample specimens and still reproduce our measurements. Not all dark re- ree sample exhibited the pronounced compositional differences ble 2.1, although the darker the phase the more likely were large differences in elemental concentration. Significant chemical variation can also be exhibited in specimens for which only a single phase or component is ap- parent, Multiple samplings and subsequent analyses were carried out on sev- ral Motagua source jadeitic rocks. Data from the replicate analyses, expressed as the deviation from the log mean values obtained for multiply sampled spec imens, are presented in Table 2.2, The standard deviation (s.d.), expressed asa percentage, is calculated for the mean deviations and can be used as an overall measure of our ability to reproduce our measurements when taking samples at different locations from a given specimen. PREVIOUS ANALYTICAL FINDINGS Samples that shared a similar compositional profile, defined by use of a matrix of Euclidean distances or visual comparison of elemental profiles, were uscd to form trial groups; formed in this general manner, these trial units were based Ronald L. Bishop, Edward V. Sayre, and Joan Mishara 2.2 Log Deviations from Means Calculated from the Repeated Oxide ‘Determinations by INAA of Maya Jadeitite Artifacts and Motagua Source Samples Nay $ej03 Euj03 Luy0; HO, C03 Fez05 CoO Yb,05 JSIA 0.099 0.375 0,065 -0242 0.830 -0.219 0.089 -0.130, — JSIB -0.100 -0.376 0,066 0.242 0.830 0.219 -0.089 0.130, — JSIBA -0.008 0.025 0.135 0.004 0.045 0.619 -0.039 0.007 0.020 JSISB -0.003 -0.078 -0.097 0.037 -0.045 -0.383 -0,001 -0.013 -0.016 JSISC 0.012 0052 -0.038 -0.040 -0.001 0.236 0.041 0.020 -0.005 JS68C — 0.137 0.090 0.113 -0295 0535 0.000 0.060 0.188 JS68F — 0138 -0.089 0.113 0295-0535 0.000 -0.060, -0.183 J*S1O] 0.004 0.007 0.041 -0.010 -0.007 0.045 0.034 0.022. -0.040 J*S1OT 0.001 0.007 -0.041 0.010 0.007 -0.045 -0.035 -0.022 0.040 JS96 0.002 0.131 -0,071 0.041 0.186 0.363 -0.030 -0.050 -0.008 JS97 0,002. 0.131 0.071 -0.041 -0.186 -0.363 0.030 0.050 0.008 JS3X — 0023 0105 0.009 0301 — — — 0016 JS3¥ — 0023 -0.105 -0.009 0301 — — — -0016 JS103 0.014 -0.068 -0,093 -0,002 0.015 -0.061 -0.059 -0.077 0.048 JS10¢ -0.014 0.069 0.093 0.002 -0.015 0.060 0.058 0.076 -0.044 JSH44 0.010 0.058 0.140 0.025 0.192 0.047 0.034 0.083 0.081 FSIS 0010 0.058 0.140 0.025 0.192 -0.047 -0.035 -0.085 -0.081 JSH2A 0.068 0133 -0.001 0.040 0.205 0319 0.087 0.281 -0.250 JSHB -0.003 0.019 -0.037 0.001 0579 0.295 0,019 0.183 0.005 JSHC 0.014 0.302 0,046 0.080 0256-0593 0.200 -0.360 -0.015 JSII1 0,056 -0.182 -0.143 0.052 0523-0321 0.113 0.172 0.098 JSH2 0.037 -0.153 0.072 -0.009 -0.027 0.041 -0.033 0.108 0.124 JSH3 0.036 0.483 0,062 — 0033. 0.258 0237 0.177 0.037 JSIG -0.016 0.336 0.296 -0.044 0.122 0.023 -0.095 -0.303 0.239 JSU7 0.011 0.413 0.309 — 0.005 0.018 0.008 0.050 -0.281 JSU8 0.005 -0.749 0,014 0.045 0.118 -0.042 0.088 0.253 0.042 JS9 0.000 0.175 -0.285 0.149 0.016 0.180 0.093 0.154 -0.035 JS93 0062 0.216 0.299 0.210 0.126 0.063 0.159 0.213 0.086 JS 0.062 0.392 -0.015 -0359 0.141 0.242 0.253 0.368 -0.052 J*S87 0.022 0.122 -0.115 -0.028 0.030 — — 0.033 -0.042.-0.005 JrS88 0.022 0.122 0.114 0.029 0.030 0.000 0.033 0.042 0.006 JXA — 0028 0.152 0.072 0.101 -0.004 0.001 -0.019 0.169 JXH — 0028-0151 0.071 0.102 0.004 -0.001 0.019 -0.169 Sse eae a3 teers ss esl 017 O11 (0070.19 0.21 00702 0.08 Percentage 6.12 47.88 2846 17.35 5363 60.66 16.24 30.93 19.49 Standard deviation 0.04 024 014 O1L 0290.28 0090.16 OIL Percentage 9.50 74.89 37.24 29.24 94.83 92.35 44.09 29.5 Missing data or data paired with missing data are excluded these are indicated by — in the cel. 39 e PRECOLUMBIAN JADE. ‘Through an iterative process involving the formation of trial groups, the removal of specimens lying outside acceptable confidence limits, the addition of samples found to have acceptable probability of belonging, and the succes- sive recalculation of the group's elemental correlational characteristics, ten ref erence groups were determined. These are primarily refinements of the previ- ous data summary—refinements, however, that permit new interpretations to be made, The formation and evaluation process required that each group be distiner from every other group in terms of a stated confidence interval. It is reasonable to belicve that several of the groups relate to one another in some manner. For example, two of the groups that derived from the analyses of light-colored Costa Rican samples, though compositionally distinct using the criteria discussed above, are merged to form a larger Costa Rican Light refer- ence group that deviates from the other characterized groups in a consistent manner (analytical and descriptive listings arranged by reference group are siven in Appendices A and B). Our carly attention was to the source materials recovered from the Mo- tagua Valley, our ability to characterize the source areas, and the extent to which we could attribute any artifacts to a Motagua-focused compositional reference group. As previously noted, a fundamental distinction was observed between a group of compositionally similar source specimens and matching artifacts that have been designated Motagua Light, after their light green color, and Motagua Dark, which refers to its characteristic green-black color. The composition of either of these two groups, numerically biased toward Motagua River source specimens, is easily differentiated chemically from the group of samples designated Maya Green. Samples from the Maya Green group, with their significantly higher chromium values (Table 2.3), are an emerald green color, and they are more similar to the jades popularly associ- ated with the high status of the Maya hierarchy. The relationships of the three ‘groups to one another are shown in Figure 2.2 where the mean concentrations are expressed as ratios to the mean concentrations of the Motagua Light group. For example, the mean cobalt concentration in Maya Green specimens was about twice that of Motagua Light specimens, and the mean cobalt con- centration in Motagua Dark was four times greater than that of Motagua Light. X-ray diffractometry provided data on the mineralogical components of a subset of the chemically analyzed specimens. The majority of the samples ana- lyzed were the hollow-drill-extracted cores that had cooled for two to three years after being analyzed by instrumental neutron activation, The small cores, or fragments extracted from source materials or artifacts, were in- spected for multiple phases or minerals, using a binocular microscope. Miner- als that were obviously different were separated. If the specimen appeared to be homogeneous, it was used “as is.” All samples were ground to a fine pow- der in an iron mortar and mounted on a Philips low-background quartz, holder that was designed to accommodate small quantities of samples. Ifa suf- Jane Stevenson Day nacaste, Rosales Zoned Engraved, Guinea Incised, and Tola Trichrome bear relationships in iconography and designs to jades. In addition, Carrillo and Galo Polychromes, which begin late in the same period but which reach their height during the Early Polychrome period, .D. 500-800, are also connected with jade imagery. In contrast, pottery from the Atlantic Watershed (with the exception of one type, Africa Tripods [Snarskis 198]) shows little relationship with jade artifacts, Guanacaste Ceramics and Jade The burial ceramics of Guanacaste are often exoti¢ items themselves, In com- paring them with jade items of the same period I looked primarily for two sets of relationships: similar design elements, and related themes or iconography. I speculated that if these objects were indeed associated with each other as ritual items at the time of death, they should reveal similar iconography, forms, and designs ROSALES ZONED ENGRAVED. The finest of the Zoned Bichrome pottery types is Rosales Zoned Engraved. It is a well-made ceramic that may be slipped in red, black, or a light brown and is engraved with fine line incisions that often, but not always, outline a design in a second color. Examination of examples of this ceramic type and of jade ornaments produced some interesting com- parisons A massive piece of Rosales pottery in the form of a seated male figure is illustrated in Figure 22.2. The figure wears a headdress in the form of a band with a rectangular object bound into it on the left side of the forehead. Several thousand plain celts of similar form were recovered by Hartman (1907349) at the Zoned Bichrome cemetery of Las Guacas on the Nicoya Peninsula, and plain celts or axes are recorded from other Guanacaste sites as well (Day 1984). ‘Most of these celts are of various kinds of greenstone and very carefully made and polished (Figure 22.3). There seems little doubt, based on Hartman's doc- umentation of the mass of similar celts at Las Guacas, that the rectangular headband object on the figure is one of these celts, which indicates that the polished axes retained an ancient ritual significance over time. In their basic form the abundant jade axe gods of Guanacaste appear to have been cut from such celts. One of their most interesting aspects is that the decoration utilizes only the upper half of the celt form, leaving what was obviously the blade sec~ tion as a smooth, polished element of ornament (Figure 22.4). This seems to suggest that the ancient celt form itself was still an important aspect of this rit- ual object during the Zoned Bichrome period and that it had a special asso tion with the iconography of elitism and jade. This association of polished celts with status is also seen in Olmec iconography, which suggests that a pri- mary source of inspiration for the earliest Costa Rican jade imagery and tech- niques (perhaps as early as 500 B.C.) may lie in Preclassic mesoamerican cule tures to the north (see Easby 1968; Balser 1961). 292 PRECOLUMPIAN JADE FIGURE 222 Rosales Zoned Engraved ceramie Figure with cet headdress (Jan and Frederick R. Mayer collection, hereafter indicated by JPM) Fiouee 22. Jade ba axe god with upturned claw thu Fine-line incising (TPM) Jane Stevenson Day 203 FiGURE22.5 Rosales Zoned Engraved figure with upturned claw thumb and fine-line incising after Tillett 1988) FIGURE 227 Clay stamps with headless figures (JFM) Froune 228 Stone mace head depicting skeletal human head FM) 204 PRECOLUMBIAN JADE A second resemblance between the jades and Rosales pottery is in the use of the technique of fine-line incising. Numbers of pieces of jade are decorated with curvilinear incising similar to that on the body of Rosales figurines (com- pare Figures 22.4 and 22.5). Related to this is an unusual rendition of the thumb that occurs both on jade axe gods depicting bats and on Rosales fig- urines. The hands, usually held at waist level with fingers almost touching, are depicted in both media with a distinctive upturned clawlike thumb. A third motif or theme relating Rosales pottery to objects of jade can be seen on a vessel from the Pacific area (Baudez. 1970: Plate 16). The piece is dec- orated with a standing headless human figure in black on a red background; fine-line incising outlines the figure. Such headless figures are also found in jade. Several are depicted in the catalog of Costa Rica's Instituto Nacional de Seguros (1980:107), and one standing headless figure in jade is in the Jan and Frederick R. Mayer collection in Denver, Colorado (Figure 22.6). Both the jades and the vessel are thematically related to a group of ceramic stamps or seals that also depict headless standing human figures (Figure 22.7), It is unique that these three sets of Zoned Bichrome burial objects (jades, ceramics, and seals) portray the body rather than the more usual decapitated head of a vietim. However, they appear iconographically related to a trophy-head cult with great time depth in Costa Rica, a cult whose origins probably lay in northern South America (Stone 1977; Leibsohn 1685). Stone mace heads from ‘Zoned Bichrome burials depicting human or skeletal heads (Figure 22.8) fur- ther suggest the relationship of these Zoned Bichrome artifacts with a cult concerned with headhunting and decapitation. GUINEA INCISED. ‘The most elaborate incised monochrome pottery of the pe- riod is Guinea Incised. On many of these ceramic vessels we find clear rela- Uonships with the jade objects known generically as axe gods (Figures 22.4, 22.9). First, there is a relationship in symbolic forms. Both media depict human figures, anthropomorphic bats (Figures 22.9, 22.10), owls, and alliga- tors as major icons, Second, textile/basketry designs appear incised on objects in both media. These include guilloches, frets, interlocked L-forms, mat pat- terns, chevrons, and triangles. On jades the designs most often appear as tex- tile bands worn in headdresses or as a decorated tonguclike element (see Fig- ure 22.9); on the ceramic vessels they are most often used as rim bands or deco- tative panels (Figure 22.11). We have to be aware that geometric design cle- ments can take only a limited number of forms; nevertheless, the similarity of designs on objects in these two media and their base in textile patterns scems ‘more than coincidental. The designs on the pottery are more extensive and elaborate than those on the jades, but this is probably because it is easier to in- cise ceramic than hard greenstone. TOLA TRIGHROME, Tola Trichrome is a ceramic type that appears to be transi- tional between the fine Rosales Bichrome tradition and the polychrome Jane Stevenson Day 295 FIGURE22.10 Guinea Incised crane bowl with incied FicURE 229 Jade pendane with bat Figure and textile panera panel JEM) layered headdress JEM) FIGURE 22.11 Tola Trichrome bat effigy jae (ater Tillett 1988) Fioure 22.12 arvillo Polychrome bat effigy tripod jae GEM) Ronald L. Bishop, Edward V. Sayre, and Joan Mishara 1 MOTAGUA LIGHT mi MOTAGUA DARK © MAYA” GREEN RELATIVE FACTOR Na Sc Eu Lu Yb Hf Cr Fe Co ELEMENTS. ficient amount of sample was available, a normal sample holder was used. Separable components of a sample were analyzed individually and then re- combined; a random sample of the total fraction was run from the combina- tion, Progressing from jadeite to omphacite shifts in diffraction-peak position and higher angles reflects expansion of the erystal lattice due to substitution of heavier elements. This displacement was monitored through the use of inter- nal standard, NBS SRM 675 mica. All results were reproducible; that is, each of the peaks associated with the major mineral types—jadeite, albite, omphacite, and so on—gave consistent patterns when repeated. Although variable, data that serve to distinguish among the two Motagua groups and that of the Maya Green are found in the mineralogical patterns re- vealed by the X-ray diffraction analysis. The Motagua Light samples can be characterized as consisting of major abundances of jadeite and albite, with oc- currences of paragonite and analcite (see Appendix C). This mineralogical pattern is consistent with that described by Harlow (Chapter 1) for the Mo- tagua-I source, The Motagua Dark specimens contain less abundant jadeite, major abundances of omphacite, and variable amounts of analcite. In contrast, the analyzed Maya Green samples possess abundant jadcite, trace omphacite, and relatively low abundances of albite, muscovite, and analcite. Thus, the mineralogy of the three groups supports the neutron activation data and sets the Maya Green group apart from both the characterized Motagua Light and Motagua Dark groups, As with the chemical data, X-ray diffraction serves 10 differentiate samples of Maya Green from those of the Motagua Light refer- ence unit, as revealed in the lattice spacings (Table 2.4). Because of the sub- 8 Four 22 Concentrations felative © Motagua Light means for Motagua Light, Motagu Dark, and Maya (Green reference groups (ee) VSL GE) OE OZ) GTI (ES) GORD (eZ) IE (os) o€+ — oz) BET) LI 9) HET (io) s09 MID 429 9) “98 (I9E DTT — (OB) ORST. IE) “BOE eet BST OUI) 'tIS (ATHY ZT FD (65) s+ (GS) 09? = BS)LVED. zDD IED 202) 6280 (Gh) OTT 8) (os) o¢t TOF GD s+ ELSSTO. OZ) IFTO. GE) EVE (OLD.Z9S0 (oc) civo—(&S) 608) FIED. 65) ITO FORA (osteo (100) 6H» I) «FR HY STL (IE). 1e6 OD) LIZ GOH) HE (HH) S909) OW FOF on wT (86) EO.) SHE «OD GTE (GED AHLO (ea) 962 (66) 6SZ BL) - FOR — OO). uz FO (Or) gsc Gs) cE (oh) aoe CED SSO. Ue) eee (on act — (669680 GE) SSF OZ) TI FOS NOTIN wa Suva (or) 909 (OE) oct BI) BRT fe) 90TO. AI) 9 (+) go) SORD GE) OFT — (LO) 6s60 FoF 2) ee 6) CZ) WR) OE) ST) Sth ET) EL) 6B). BOHN ey us-) ) = 9) (r=) dee) (i=) wee) a=) ect 18 ea 187 01) woot qpuopdaoxg ara 181 ueany ee trong e103 omy stv oop wp enteoyy exSroyg sdnoig souss2joy auMig]Y pute arntape{ axp jo (seBrauso19q) suor esac] PrEpUTIG PUL SME HONENUDIUD apIKC, FZ FTAVL Ronald L. Bishop, Bdward V. Sayre, and Joan Mishara 45 Tape 24 Shift in Lattice Spacings of Isostructural Pyroxene Minerals in Mo- tagua Light, Motagua Dark, and Maya Green Jadeitites Moracua Lic Maxa Gace’ Moracua Dark Jaerrrre, Japerrere Ompuacrre. (n= 8) (n=7) (n=5) 5891 S75 403 538 411 246 200 +06 144 £16 a 107 403 2 #10 197 478 £0.18 10.36 £0.59 129 3.84 40.29 746 41.81 Lol 1.20 £0.06 232 40.72 0.02 032 40.15 0.10, £0.10 0.019 £0.01 0.014 0.005 0.038 0.007 ool 0,052 +£0.09 0.069 40.019 9.98 98.43 94.2 stitution of heavy ions into the crystal lattice, there is a significant shift to higher lattice spacings, which differentiates Motagua Light from Maya Green, and Maya Green from Motagua Dark, Electron microprobe analyses were carried out for several of the samples analyzed by instrumental neutron activation and X-ray diffraction in order to be certain that we were dealing with the presence of omphacite and not chloromelanite in many of the samples. Prepared samples were analyzed on the ARL microprobe at the Mineral Sciences Department, Smithsonian Institu- tion, using a 1.5 kev accelerating potential, 30 microamp beam current, and 10 second counting period. A 50 micron defocused beam was used to analyze five locations on each sample. All elemental concentrations were standardized against Kakanui hornblende. ‘The determinations for each sample, as well as the data recalculated to Si .00 per six oxygens, are listed on Table 2.5. Different schemes have been advanced to address problems associated with analyzing the microprobe data of pyroxenes to determine the relative contributions of the acmite-augite- jadeite system (Carpenter 1979). Following inspection of the data in Table 2.4, wwe observed that the number of cations per two silica of Ca and Mg are ap- proximately equal (within analytical error). Because in augite (CalFe,Mg]Si0,) the sum of iron and magnesium ions should equal the calcium ions, and be~ cause in cach case the magnesium ion concentration alone slightly exceeded the calcium ion concentration, we assumed there was no iron present in augite, and assigned all iron ions to acmite. Accordingly, we defined the pres- ence of augite as being equal to the mean ionic concentration of Mg and Ca, and the presence of acmite (NaFeSi,Q,) as being equal to the iron ionic con- centration. Finally, following Carpenter (1979), we let the cation abundance of Al represent the amount of jadeite (NaAISi,0,) in the sample. In principle, TaBLe 2.5 Electron Beam Microprobe Analyses of Jadeitites and Recalculation to the Acmite-Augite-Jadeite System A024 JA013 JAIL J AL8S JA368 —-JAMol—-JAISB A202 F209 AIO = JA203 Resneence Grovr MAYG MAYG = MAYG = MAYG = MAYG = MAYG-—CRL. cRL. CRE CRL cR 5792 5750 57.29 57.77 5732 57.32 59.81 58.66, 5242 58.3262 19.66 2023 20.67 22.68 20.22 19.17 24.73 2B, 267 15138 (25.89 22 121 9 1 121 128 116 135, 097 146 091 5.06 3.82 352 257 3.89 429 118 124 087 0.96 0.94 4.79 492 452 327 488 5.0 15 22 1 176 Lis 1044 10.54 11.2 1058 1025 nn 1329 1218 1238 1209 99,3 98.1 977 985 98.1 979 1004 99.9 1002 101.0 Canons ren 281 2.00 2.00 2.00 2.00 2.00 2.0 20 2.00 2.00 2.00 083 0.85 093 083 079 097 0.93 02 102 0.04 0.03 0.03 0.04 0.04 043 0.04 0.03 0.04 0.03 020 018 03 020 022 0.06 0.06 ot 0.05 0.05 018 017 12 018 021 0.06 0.08 0.04 0.06 0.4 070 a7 075 072 069 076 0.88 079 0.82 078 Tol 3.97 3.95 3.95 3.96 397 395 389 3.99 3.92 3.99 391 Aenre- verte Jadeite 76.00 me 80.00 86.00 78.00 76.00 91.00 89.00 93.00 91.00 93.00 Acmite 6.0 4.00 4.00 3.00) 4.00 4.00 3.00 4.00 3.00) 4.00 3.00 Augie 18.00 18.00 18.00 11.00 18.00 20.00 600 7.0 4.00 5.00 4.00 so Moya, eames = PUES 109915 UOHPMID = OIG any ¥1999 = YO AMET VEDNY EOD = “TAO ozr we on'ts 86 ro +O oso 800 “so oz £%6 Is zor we oz oz ais oor ove, ozs 80 “0 veo oso 800 oso 007 £86 sos sear 168 sz ave ozs ove 006 o0'2s we oro seo ovo oro 0 oz rs 99 096 owe sce wt S09 o0'8e oor oss seussis 66 seo ovo ero soo 90 07 796 +9 zvou OL svt seo Ios ovec os onze w-susOnY-LLNIE +e zo 1z0 a) sro wo wz ise wie swouvg 06 oe, Ivs oF ist oat ors ae 890 1z0 zo 00 £0 007 686 agar ws ag oat 8685 80 +80 10 oro 800 +80 wz $66 soz we 161 ore e207 irs qo, FN or ow oa folv ‘ors 4B PRECOLUMBIAN JADE. the sodium-ion concentration should be equal to the sum of the jadeite and acmite concentrations; hence in this case equal to the sum of the aluminum, plus iron ion concentrations, However, in each specimen the sodium ion amounts were less than the aluminum alone. Recognizing the inherent diffi- culty of determining sodium by an X-ray emission technique, we attribute the low sodium values to analytical error and ignore them in deriving the molecu- lar ratios. Following usual procedure, the molecular ratios were finally nor- malized to add up to 100 percent. The cation ordering of these data for the acmite-augite-jadeite system is plotted in Figure 2.3 against the compositional ranges given by Essenc and Fyfe (1967). Because several specimens of two of the reference groups were run, separate symbols have been assigned to the data points belonging to the Maya Green, Costa Rican Light, and remaining samples. Included in Table 25 is an analysis of jadeite (number 106442) from the Mineral Sciences collee- tion of the National Museum of Natural History; this sample was collected by Foshag from the Motagua Valley. All samples occur low relative to the aemite apex, within the jadeite- augite range, and well below the limits given for chloromelanite. For the sam- ples showing greater concentration of omphacite, the data are roughly similar to that of Harlow and Olds (1987: Table 2). Although analyses arc limited, the microprobe data reveal separation of the members of the Maya Green and Costa Rican Light specimens, but overlap between a Chichén Green and the Costa Rican Light specimen. This suggests that the microprobe is sufficiently sensitive to differentiate among some of the jadeitites containing various pro- portions of related minerals. The Costa Rican Light specimens all lic near the pure jadeite corner, and the Maya Green are close to the jadeite-omphacite boundary. Of the four specimens lying well within the omphacite range, one is a dark Motagua source specimen and two are artifaets from the Motagua site of Quirigua. We feel that this compositional range will be found to typify the “omphacite of the Motagua Valley. The samples of the Maya Green group represent relatively the most ho- mogencous and one of the most easily differentiated of the defined reference groups. For this reason, we have opted to use the standardized characteristic vectors derived from the variance-covariance properties of this group to illus- trate the separation among our reference groups. In the plots of the projec- tions of samples onto Maya Green's standardized eigenvectors, we are merely showing the sample data points relative to a normalized space. Accordingly, specimens belonging to a well-characterized comparative group will cluster together in this space. Unlike the practice when using techniques such as dis- criminant analysis, there has been no “pooling” of the groups or imposition of a structure on the data by attempting to maximize the separation among the compared groups (sce Bishop and Neff 1988). In Figure 2.4, for example, Motagua Light and Motagua Dark are shown with the samples of Maya Green, In this and similar figures an 80 percent con- Ronald L. Bishop, Edward V. Sayre, and Joan Mishara ACMITE t CHLOROMEL ANITE YADEITE OMPHACITE AUGITE * ! BE, A” atet = COSTA AICAN LIGHT mM MOTAGUA ARK @ MISC A NAYA GREEN of CHICHEN GREEN fidence ellipse surrounds a given group and was calculated as though the group werc of infinite size. The ellipse about the base reference group, Maya Green, is spherical in Figure 2.4, refleeting that all of the elemental correlation hhas been removed, and that the new axes, constituted from linear combina- tions of the original measurements, are mutually independent. The contribu- tions of the original measurements to the axes, shown in Figure 2.4 and Fig- ures 2.6-2.10, are given in Table 2.6. Not all of the “green” jadeitite artifacts that have been grouped belong to the Maya Green reference units. In our earlier data summary (Bishop ct al. 1985), we discussed a group we designated Chichén Green because of the pres- cence of bright to light green artifacts that had been recovered from the Cenote at Chichén Itza in Yucatan, Actually, from our first attempts, two groups with a strong pattern of a Chichén Itza provenience were noted. The relation of the Chichén Green to the other groups or to the Motagua source region, however, was unclear. Additional neutron activation analyses supplemented by X-ray diffraction data now suggest that though the newly constituted Chichén Green reference units are demonstrably separable from the other character- ized groups, itis best considered to be related to the Motagua source area. Re- cently analyzed samples from the Motagua locations of San Agustin Acasaguastlan (Smith and Kidder 1943) and Robert Terzuola (Feldman et al.1975) sites occur in the group, as does a Motagua source sample from San 9 Fioure23 Plot of electron ‘beam microprobe analyses of jadeittes recalculated to the acmite-augitejadeite system clawification of Exsene and Pyfe (1967) 50 FIGURE 24 Motagua Light, ‘Motagua Dark, and Maya Green reference groups shown relative to Vectors and 8 derived from the Maya Green reference ‘group. Ellipses indicate 80 percent confidence interval. PRECOLUMBIAN JADE 0.0084 0.0083 | 0.00x2 | 0.0015 | o.oo | 0.0009 0.0007 0.0005 0.0005 B 6,000 0.0002 - 0.000 - Maya Green) 0.0003 4 ° 0.0002 | -0.0003 | 0.0004 wo.oois 0.0014" 0.0010" -0.00068 "0.0002 "0.0002 "0.0006 vector 5 Agustin Acasaguastlan that was submitted to us by Edwin Shook. These Mo- tagua specimens not only relate to the original Chichén Green specimens in composition but also have the same light green color. ‘Although it is statistically separable from the specimens in the Motagua Light group, Chichén Green is compositionally similar to the samples of Mo- tagua Light (Figure 2.5). A notable difference in the chromium values that in- Aluence the group's slight divergence from the Motagua Light specimens is il- lustrated in Figure 2.6, In terms of mineralogy, the analyzed Chichén Green samples are similar to the mineralogical suite characteristic of the Motagua area jadeitites: abundant jadeite and albite, with the presence of paragonite, muscovite, and analeite (see Appendix C; Harlow and Olds 1987: Table 13 and Chapter 1). ‘Another group, designated Motagua Exceptional, is present in Figure 2.6, along with the Chichén Green and other Motagua groups. The samples in this group are not numerous, and few have been analyzed by X-ray diffraction; for those that have been diffracted, their predominant jadcite composition may be significant. All members of this group are specimens that were obtained from the stockpiles of the commercial jade factory JADES, S.A., and reflect specific Motagua sources that were being mined in the early 1970s; included in this ‘group are specimens from a jadeite “pod” that was mined out of a surround ing layer of albite. Lower temperature and higher pressure favor the formation of jadeitite over albitites; the two usually occur in association. It was not surprising, there- fore, to find albite along with jadeite artifacts and frequently to find that ‘green albite represents the dominant mineral in some of the greenstone arti- facts. Similar to the Motagua specimens, the compositional characterization of ost oe 429 wr oft orl 0 “vo ¥0 Notun LLNOD 40 BOKANCD loaes7o —te-aneso—zoazes0 wanes = waLeo woassl —ourascro «= go-aevD—eo-aOHE savwanaory @ aro 0) sige eer) BIO) HDA) HD ETO. OBL” ETO. —-C!D. ©) 900° () wre 0) se) OTH HOO = aetD =] TO. zero, WN ue Fo (GH 0160] sro" gore eV (G80) eo BIO =) OO) 9500 FO4D @® sero GD) FeFO Torr 9) o6L0- =) cov) LTO" ( sro) 9770) seo" FOL © ero (ED) ser 400 AD LED) HD) HOY OH) SSO” = LOD. ~—®) LTO FORK, () too 2) erro” bo" (cD coro (Om) ogve- (HI) bee or ¢60 @ of =) +00 FORT ( e900 GD sego- oer ) 96r0 F670 (el) e9¢0 (2) sero (be) 98. ) ceco- foeng ® zro GN) S860) e980 =) 900 ) elo —@) sou") 920) 90 leo (0) cor =) 1570) 1900) SOW GD EO OD EO OH SD. STO’ DON 6 2 9 ‘ ‘ f z 1 sow dnorg souss9yoy wda19 vkeyy 2Yp Jo ssoIV9, onsUDDEIeYD pozipsepuEIg 9-7 TAY I, 32 Fioure25 Concentrations relative to Motagua Light means for Motagua Light, Motagua Exceptional, and Chichén Green reference soups FiGuRE2.6 Motagua Light, Motagua Dark, Motagua Exceptional, and Chichén Green reference groups shown relative ro Vectors2 and 7 derived from the Maya Green reference group. Ellipse as in Figure 24 PRECOLUMBIAN JADE RELATIVE FACTOR vector 128 64 32 16 8 4 Z 1 1/2 1/4 1/8 1/16 1/32 1 MOTAGUA LIGHT © MOTAGUA EXCEPTIONAL © CHICHEN GREEN eb oe Na Se Eu Lu Yb Hf. ELEMENTS 0.0009 .0008 - .0008 .0008 | o.000+ 0.0008 0.0002 0.0001 0.0001 0.0002 0.0004 0.0008 + 0.0006 Motogve Light Motogua Dork Motogue Exceptional 0.003 0001 (0.007 Vector 2 0.003 Ronald L. Bishop, Edward V. Sayre, and Joan Mishara albitites divides into light and dark groups (Figures 2.7 and 2.8). (The rela- tionship of the albitites to the Motagua source arca is discussed below.) At this point we have discussed the compositional patterns for source and artifactual groups from the Maya region—the region for which a jadeitite source is known in the Motagua Valley. Farther to the south, in lower Central America, jadeitic artifaets are commonly associated with the Pacific and At- antic watersheds of Costa Rica; whether or not there is a local source for the lower Central American material is unknown. What is readily apparent is that a diversity of greenstone—and nongreenstone—was used. Interpretative difficulties are exacerbated by the fact that so few of the specimens have re- sulted from controlled excavation, Under the working hypothesis that a lower Central American source ex- isted that supplied the raw material for the region's needs, we extended our sampling to include Costa Rican jadcitite artifacts from Costa Rican institu- tional collections. If a jade source separate from the known Motagua source was exploited in precolumbian times, then the analytical patterns observed in the artifacts would contribute to our understanding of jadeitite variation. Our recent data analysis has revealed, as did our earlier effort, that once again dark and light compositional groups could be observed. The Costa Rican artifacts are compared to characterized groups from the Maya region in Figures 2.9 and 2.10. The Costa Rican groups are multivariately separable from those of the north, but aside from being very low in chromium abundance fail to reveal any single key elemental difference from the Maya specimens. Mineralogically, some differences can be noted between specimens of the Costa Rican Light and the Maya samples, notably in the low occurrence of paragonite and muscovite. The occasional presence of hornblende as an acces- sory in the light and dark divisions differentiates some of the samples from specimens in the Maya groups. Upon detailed inspection, the Costa Rican specimens ean be subdivided, each division being multivariately separable from the other, using the criteria we have employed for the other groups in this study. The mean concentrations and stan- dard deviations of the two Costa Rican Light groups are listed in Table 27, and their subgroup affiliation is noted in the case listing (see Appendix A). Treated as, a single group, the samples can be observed to diverge from the other character- ized groups in a patterned manner, as shown in Figures 2.9 and 2.10. Combining, the two groups, however, or ineluding “possible” group members that lie just outside of the confidence interval cutoff broadens the characteristies of the group. Including too much variation in a group will permit almost all of the samples from our investigation to have significant probabilities of belonging to the group. This so-called floodgate phenomenon is one reason that our grouping of the samples is conservative; we focused on only the major patterns of the data set— the demonstrably separable groups of presumably related samples—rather than attempting to be exhaustive. As a consequence more than half of the analyzed specimens are not assigned to a reference group (sce Appendices b, #, and F). 33 54 Ploure2.7 Albice eference groups shown eclative to Vectors | and 7 derived From the Maya Green reference group. Flipses as in Figure 2. Picuae 2.8 Comparison of| albiteand Motagua reference groups relative to Vectors 1 and 7 derived from the Maya Green reference group. Ellipses as in Figure 24. PRECOLUMBIAN JADE 001s oot 0013 oo oon 4 0009 0008 007 4 0008 0005; 004 0003 002 ‘001 vector 7 -0.0001 9.0002 -0.0003 Arbite Dank Naya Green aaite Ligne 0.0004 “0.008 -0.00 Vester 7 bite Dork _¥ veotor 4 otogua Derk 0.901 Vector 1 0.008 0.003 0.005 Abite Light (0.008 0.0004 0.0003 + 0.0002 0.0001 -2.0001 0.0002 -9.0003 0.0004 ‘9.0005 2 0.0006 0.0007 -0.0008 0.0009 9.0010 -0.0081 0.0012 -0.0013 -o.00t4 0.0015 0.0008 0.0008 0.0004 0.0002 0.0002 ~0.0004 vector 7 0.0006 -0.0008 0.0010 0012 -0.0014 0036 Ronald L. Bishop, Edward V. Sayre, and Joan Mishara aye Green Gosta Rican Light Coste Rican Dark 0.0008 0.0015 0.0025 vector 2 Notagua Light Gosta Rican Lignt Wotaaus Dark 5 costa Rican Dari ~0.008 “0.00% 0.001 0.003 vector 2 5 FiouRe2.9. Comparison of| ‘Costa Rican Light and Dark reference groups relative to Veetors 2 and derived From the Maya (Green reference group. EEllipses as in Figure 24 FIGURE 2.10 Comparison of Motagua and Costa Rican reference groups relative to Vectors? and 7 derived from the Maya Green reference group. Ellipses as in Figure 24 56 PRECOLUMBIAN JADE ‘TABLE 2.7 Comparison of Subdivisions of the Costa Rican Light Reference Group Costa Rican Costa Rican Licirr? (a= 12) NaO Bo 8) Rt ® S20; 349 (43) 341) EujO5 0.228 (106) 0.136 67) 14,05 389 (266) 07 a9) Yb;0; 0213 @o4y 0283 (111) HO, 3.08 (8) 183 @5) 2,0, 637) 758 (189) Fe,0; 124 G2) 140 42) Coo 3.83 @7) 360 (58) Ina few cases among the ungrouped samples we can indicate that a sam- ple may relate to a reference group, although it is not a member of the group. ‘These are samples that may have had a sufficiently low likelihood of group membership that we excluded them; others may have had some data missing, thus making their attribution less secure. (Samples falling into the category of relating but not belonging to a reference group are differentiated in Appendix D by the group to which they relate [noted in brackets}.) Variation in the Motagua source area is expressed in several of the refer- ence groups that have been formed: Motagua Light; Motagua Dark; Motagua Exceptional; Chichén Green; Albite Light; and probably Albite Dark. These groups, through their inclusion of Motagua source samples and of fragments of worked or unworked jadeitites from sites within the Motagua Valley (San Agustin Acasaguastlan or Terzuola), serve to limit the elemental composition that one would be likely to encounter in the jadeitites from the valley. It is, possible, of course, that additional sources of jadcite might be discovered even- tually in the Motagua Valley, ones with compositions that would include even the Maya Green samples. ‘The Motagua Light, Motagua Dark, and Motagua Exceptional groups in- clude the majority of the Motagua Valley source jadeitites. Represented in these groups are artifacts recovered from Belize (Altun Ha, Cuello, Cerros, Nohmu), Copan, and the Salama Valley of the Baja Verapaz. Although the Chichén Green group contains 16 of the 24 specimens from Chichén Itzé, it also includes samples from Altun Ha, the Salama Valley, and from Motagua locations. Representatives of different compositional reference groups are present in PRECOLUMBIAN JADE served compositional ties to the resources of the Motagua Valley. None of the jade specimens was of the high chromium or emerald green variety (Hirth 1988). If we are correct in our linkage of the albitites of the albite geoups and, by extension, of the numerous jades from El Caj6n to the Motagua Valley, would it not be reasonable to expect that examples of the Maya Green compo- sitional group would have been recovered—if the Motagua were the source for the emerald green type of jadeitites? The absence at El Cajén of the emer- ald green jadeitites may be an indication that the source of the Maya Green jadeitites was not within the characterized Motagua source area ‘The separation of the Costa Rican Light and Costa Rican Dark groups from the characterized groups from the Maya region has been achieved chemi- cally (see Plate 2c: Costa Rican Dark; Plates 2d, 3a and di Costa Rican Light). ‘The validity of their chemical distinctiveness from the Motagua samples finds mineralogical support in the virtual absence of mica and the low albite content in the Costa Rican specimens (see Appendix C). Although the color of the Costa Rican Light jadeitic specimens ranges widely, there is a tendency toward a “bluish to sea green” hue (Easby 1968:15). Both Costa Rican groups contain samples predominantly reported to be from the Pacific side of Costa Rica rather than from the Atlantic side, but this distribution is primarily because of our exclusion of the string-sawed, more completely decorated specimens that are more typical of Costa Rica's Atlantic Watershed. Absent from the charac- terized reference units but attributed to the Costa Rican Dark group during the attribution of the nongrouped samples was the finely carved jadeite “Olmecoid” spoon (Costa Rica, INS 1980:37, top; Lange, Bishop, and van Zelst 1981: Figure 45), which had previously been a member of our Costa Rican ref- erence unit but was excluded during our recent refinement of the group. Not attributed to any of the reference groups was the Costa Rican jadeitite pendant bearing inscribed Maya glyphs (J A264) (sec Balser 1974: Plate X; and Plate 3b). ‘The proveniences of specimens in the Costa Rican Light group reveal that several group members were recovered outside of Costa Rica, especially from Belize. Particularly interesting is the celt fragment, | A423 (see Plate 30), from Cerros. It falls unambiguously within the Costa Rican Light group and is of the bluish hue characteristic of the Costa Rican jadeitites. Similarly, other samples from the Belizean sites of Cuello and Cerros also occur within the group. Clearly the pattern indicates that the ancient inhabitants of Belize drew upon some of the same sources that supplied material for the jadcitites of Pa- cifie Costa Rica. The interpretation of compositional group affiliation is not as easy for some of the Costa Rican Light members. From a mineralogical standpoint, the inclusion of specimen jA222 from Cuello is questionable, for it has the basic ‘mineralogical components of the Motagua-focused specimens; it is included in the Costa Rican Light group, however, based on its probability of membership in the group. And, the Belize specimens aside, the remaining members of the ‘group with proveniences to the north are all undecorated celts. Ronald L. Bishop, Edward V. Sayre, and Joan Mishara Our discussion up to now has dealt with the geographic distribution of specimens in the reference groups. From a temporal perspective, the presence of the emerald green specimens in archaeological contexts spans almost 1,500 years, from the Middle Formative burials at Cuello (Hammond et al. 1979:903 Hammond 1982:134~135; Andrews and Hammond 1990) into the Postelassic burials at Santa Rita (Chase and Chase 1988). Given the relative homogeneity within the characterized Maya Green group, the same general source of the emerald green jadeite apparently was exploited from the earliest Maya occu- pation into Spanish times. Ic is more difficult to comment on the time period during which albitites may have been more extensively used in the Maya region. The characterized Albite Light group is strongly oriented toward the El Cajén region; only a sin- gle Belizean specimen is included. Among the specimens that are not attrib- uted toa compositional reference group, a strong Classic period use of albitites isindicated. ‘The carving of jadeitites may be slightly earlier in the Maya region than in Costa Rica, where it is reported from Zoned Bichrome burials perhaps as early as 500 8.C. Greenstone carving in Costa Rica continues through the ‘Zoned Bichrome period (A.D. 500) but ends approximately A.D. 700. The lack of specimens recovered from controlled excavation in the country prevents an assessment of whether or not the use of jadeitites and albitites was restricted to any subperiods. Ie is clear that there was simultaneous use of jadeitites and albitites as well as quartz, serpentine, glaucophane, and others. What was the relative value of the different stones? Does the presence of several emerald green beads in one context represent the accumulation of more status or the expression of more control than finely carved pendants in albite? Perhaps it was a situation where the emerald green jadeitite would be used if available. These are questions we cannot answer here any more than we can indicate a specific source for the materials of the Maya Green or Costa Rican compositional groups. We have been able to observe statistically definable chemically based groups and to see their mineralogical correlates. Discussion has centered around the reference groups and has not addressed the totality of the samples analyzed. Color may be a reasonable first approximation to our groups, rang- ing from the emerald green of the Maya Green to the gray-green of the ichén Green and to the bluish green of the Costa Rican. To the extent that we are correct in our attribution of the Chichén Green to the Motagua, the question of the source of the Maya Green jadeitites remains prominent. If more work were carried out in the future, perhaps our existing groups would again be refined, and some of the unclassified samples attributed to a reference group, It is clear, however, that chemical characterization of jadeitites should be carried out in conjunction with extensive structural and mineralogical analysis. Whether we are considering the emerald green jadeitites or the light 60 PRECOLUMBIAN JADE green albitites, long-distance exploitation of resource areas is indicated for most of the sites included in this study. Societal needs could be met at times through the use of more-local materials. For example, some of the materials, recovered in the Salama Valley of the Baja Verapaz may have been obtained from the Motagua Valley, whereas other materials, consisting of schists or feldspars, were more likely obtained from the region's local metamorphic en- vironment (Bishop 1988). Future investigators of “jade” utilization in the pre- columbian New World must consider a range of possible materials and an ap- propriate matching of analytical approach. The refinement of social models involving exploitation and trade of raw materials will proceed once we have a surer understanding of the variation in the materials recovered. To formulate social or economic models using inferences based upon the occurrence of “jade” is an unacceptable simplification. Nore The research discussed here is part of the Maya Jade and Ceramics Project initiated in the Research Laboratory of the Museum of Fine Arts, Boston, and in cooperation with the Department of Chemistry, Brookhaven National Laboratory; analyses were con ducted at Brookhaven under the auspices of the U.S. Department of Energy. Support in the laboratory and in the field was provided by Landon T.. Clay of Boston, Contin- ued statistical analysis ofthe analytical data and extensive additional X-ray diffraction analysis of selected specimens were carried out at the Conservation Analytical Labora- tory of the Smithsonian Institution, Washington, D.C. Release time for Mishara to par- ticipate in the project as well as computational resources to summarize the previously ‘obtained analytical data was given by CAL director Lambertus van Zelst. ‘This project would not have been possible without the help given to us by many individuals. We acknowledge the assistance, loan of samples, or suggestions by Jay and Mary Lou Ridinger of JADES, S.A., Robert Terauola, David Sedat, Edwin M. Shook, and Dell Huelle in Guatemala; H. W. Topsey, archaeological commissioner, and Mark. Gutchen in Belize; the Musco Nacional de Costa Rica, its executive board, and Héctor Gamboa; Ricardo Monge O., past executive president, Zulay Soto de Andrade, and Raiil Castallanos of the Instituto Nacional de Seguros; Oscar Fonseca Z. of the Universidad ‘de Costa Rica; Rolando Castillo M. of the Corporacién Costarricense de Desarrollo (CODESA) in Costa Rica; the Royal Ontario Muscum and David Pendergast in Canada; and our colleagues in the United States who made the effort to supply samples and who have waited patiently (some!) while we tricd to make sense out of scemingly endless tables of numbers: Wendy Ashmore, Arlen F. Chase, Diane Chase, Elizabeth Kennedy Easby, David A. Freidel, James F. Garber, Norman Hammond, Kenneth G. Hirth, Tatiana Proskouriakoff, Prederick W. Lange, Robert J. Sharer, and Gordon R. Willey. Special appreciation goes to the Peabody Muscum and its director, C. C. Lam- berg-Karlovsky, who had the courage and imagination to make the museum's collec tions available for incorporation during the evolution of our analytical studies. ee a = PLATE 1A. Motagua jadetite specimen | 006 from JADES, SA. PLATE 18 Necklace from Altun Ha, Belize; Royal Ontario ‘Museum catalog number 964,216.18. Analytical numbers 1 A387, 388, 389, PLATE 1¢ Carved pendant plaque with glyphs from Altun Ha, Belize; Royal Ontario Museum catalog number L966.10.1 Analytical number | 386 PLATE ID Mosaic miniature mask from Alsun Hi, Belize; Royal Ontario Museum catalog number 956.159.153, Analytical numbers) A395, 396, 397, 398, PLATE 2A Beaded necklace -PLATE2n Human-figure from Altun Hs, Belize; pendant from Altun Ha, Royal Ontario Museum Belize; Royal Oneatio catalog number 1955:.77 Maseutm catalog number Analytical numbers 4324, L966.10.2, Analytical 325, 326, 327,328,459 number 1 A383, PLATE 2e Pendant reported tobe from Costa Rica, Instituto Nacional de Seguros catalog number 6456, Analytical ‘number | 4240, Pua Pendant reported tobe from Costa Rica Instituto Nacional de Seguros catalog number 2244. Analytical number | A265. PLATE3A Pendant reported to be from Guacimo, Costa Rica. Instieuto Nacional de Seguros catalog number 6461. Analytical number 14260. PLATE. 3¢ Celt Fragment from Certos, Belize. Field number Cwl-SP-225 Analytical number ) 423 PLATE 3m Pendane with lyphs reported to be from the Baguces area, Costa Rica, Instituto National de Seguros catalog number 4437, Analytical number 1264 PLATE 30 Pendant reported tobe from Costa Rica, Instituto Nacional de Seguros catalog number 6482, Analytical number 1x61 3 A Possible Source of Raw Material for the Costa Rican Lapidary Industry MARGARITA REYNOARD DE RUENES he search for a raw-material source of Costan Rican artifacts was moti- vated by carlier archaeological studies directed by Carlos Aguilar. Be- fore 1980 most studies dealt with the problems of style and iconography. Inter- disciplinary investigation of the problem of jade sources in Costa Rica, using both geological and archaeological data, is relatively new, and we have only recently begun to search seriously for geological sources in the Pacific region, and associated artifact assemblages. Greenstones abound in the oceans and rivers of Costa Rica, and samples were collected from areas where accumulations of pebbles have been reported, such as Barranca, Orotina, Tamarindo, and the Reventaz6n Valley. The greenstones that have the greatest similarity to precolumbian jades, however, are those from the Reynoard site, The materials analysis in this chapter, al- though limited, contributes descriptive analytical data for the first definition and location of a possible jade source in the Atlantic Watershed of Costa Rica. GEOGRAPHICAL AND CULTURAL OVERVIEW OF CosTa RICA Costa Rica can be divided into two major regions: (1) the Pacific coastal Low- lands and the central Highlands; and (2) the Atlantic coastal Lowlands. This division is a result of a continuous chain of volcanoes and mountains extend- ing from north to south. The major geological conformation is concentrated ‘on the Central Valley and the Talamanca Valley. Here, we focus on the Atlantic region, a plain with exuberant vegetation, high precipitation, and warm temperatures. The region is drained by many rivers flowing from the Central and Talamanca valleys. The Limén Basin is approximately 30-60 kms wide on the Caribbean Coast, extending approxi- mately 50 kms from Limén City, and bordered by Bocas del Toro to the south. The jade artifacts found as mortuary offerings, and often associated with 6 6a PRECOLUMBIAN JADE metates, mace heads, and ceramic vessels, are an important component focus of the Costa Rican archaeological data base (see Figure 2 in Introduction). Jade artifacts occur archaeologically primarily in sites dating to the latter part of the Zoned Bichrome period (300 8.0.~A.D. 5p0) and the Early Polychrome period (a,b. 500-800) in Guanacaste, as well as in the roughly coeval El Bosque and La Selva phases in the Atlantic region (300 ©. collections of such objects in Costa Rica and abroad; however, most of them come from plundering by pothunters. Chronological data and data on the geographical distribution of jade in Costa Rica, primarily on the Pacific Coast, have been accumulated gradually through systematic archacology (Hartman 1907; Stirling r969; Fonseca and Scaglion 1978; Snarskis 1975, 1978, 1979, 1984; Lange 1976; Stone and Balser 19655 Stone 19773 Herta 1975; Guerrero 1986). With the exception of the few specimens that are identified as being from the Motagua Valley region of Guatemala, the origin(s) of the raw mat used to manufacture jade artifacts in Costa Rica is unknown; however, the abundance of jade items suggests a local origin (sce Chapter 2). In contrast to carlier rescarchers’ efforts on the Pacific Coast, we searched for and will here describe a possible jade source in the Atlantic region of Costa Rica, Green- stones from this possible source are highly similar to those materials that were utilized to manufacture precolumbian jades on the Pacific Coast. A.D. 700). There are countless THE SOCIAL. AND ECONOMIC IMPORTANCE OF JADE Generally, the Spanish chronicles state that jade was considered the most pre- cious material by the prehistoric inhabitants of Mesoamerica and lower Cen- tral America. Jade was sacred, a symbol of high social status, and also related to belief in the afterlife. This concept was not understood by the European conqucrors, whose world vision was completely different from that of the In- dians. The Spanish conquerors converted jade into a talisman to cure kidney discases; this is why jade was named nephrite, or piedra de hijar. Once interest in jade objects was forgotten, stories about their source(s) were also left be- hind. Not until the early twentieth century did new interest focus on jade ob- jects, as a result of the development of archacological studics. This cst in jade was based on its artistic and iconographic characteristics, and infer- ences of long-distance trade and elite foreign influence. The source of jade in Costa Rica still is a great mystery. Some believed (Ekholm 1964; Heine-Geldern 1959) that jade was exchanged between pre- columbian groups of America and cultures from another continent, over sea routes that left no distinct trail. The diffusionist theory relates these artifacts with the Chinese culture and even with the Maori culture of New Zealand. There is neither geological nor contextual proof of transpacific contacts. In Mesoamerica jade sources were unknown for a long time; it was not until 1955 that a discovery of a jade deposit was reported on the Motagua ime, inter- Margarita Reynoard de Ruenes River basin in Guatemala (Foshag 1957:11). However, a North American ge- “clogist who was working in petroleum surveys reported discoveries during soto. He found some nuclei and fragments of “esmergadit” and “nephrite” ‘within the Oro and Balsas River basins, located in the state of Guerrero in “Mexico, The finding was reported to Mexican geologist José G. Aguilera, who ‘published the news in the Excelsior newspaper (Mena 1927). The discovery ‘was not well known, perhaps because the Mexican Revolution was going on Guring that time, and confirmation of the geological characterizations is not ‘possible. THE GEOLOGICAL CONTEXT OF JADE AND OTHER GREENSTONES IN ‘Costa RICA Prccolumbian Indians worked different kinds of semiprecious greenstones in -sddition to jade. The chemical composition of these materials varies, but they ‘arc mainly jadcites, calcareous stones, quartz, chloromelanites, opals, serpen- ‘snes, pedernals, and other kinds of soft and hard stones. The color of these “scones varies from milky white to light brown and dark brown; other stones ‘ace various shades of green, from light green to emerald green, moss green, ‘aed shades of blue. Hardness varies according to the composition of the stone. Until now, Costa Rican jade has heen recognized in archacological litera~ sere as a foreign material, perhaps because jades with Olmecoid and Maya ‘Sstures are emphasized in muscum collections, publications, and exhibitions. “A tomb near Tibas (Central Valley of Costa Rica) revealed a jade that may re- ect the Olmec influence (Snarskis 1979395), based on its blue coloration that ss rypical of Costa Rican jades (Ed. note: however, sce Parsons, Chapter 19). ‘Secently, some authors recognized a possible jade source in the Nicoya region “Lange et al. 1981:170); however, no additional information about its specific “Jecation has been reported (see Chapter 2). An abundance of jade artifacts in any region suggests that a possible local -: must be located in or near where these artifacts are found. After find- some greenstones apparently similar in composition to the Costa Rican artifacts (Figure 3.1) in a riverbed in the Atlantic region, and given the ‘chat jades have a widespread distribution there, I suggest the presence of a Atlantic source. |A brief overview of the geological sequence of Atlantic Costa Rica will & understanding the geological formations of the Talamanca Valley and Basin region, where we find the river that forms the beach where the 1rd site is located. Jn the Talamanca Valley there are two series of interest: (1) the Plutonic a series of granite stocks along the valley; and (2) the Cogmatic series, in intrusions are observed along the mountain chain, probably as a result ary magma, consisting of grandiorites, quartziferous diorites, gabros, and granites. Lithologic differences might be caused by intrusive 6 u 3 aoe Margarita Reynoard de Ruenes rocks. The Cogmatic intrusive group also includes adelamitas. ‘The Aguacate complex and the Doan formation of the central and Pacific regions are con- temporancous and related to the Cogmatic intrusive group (K/Ar method, 11.5 mya) and with less volume than the acid phases (9.3 mya). In the Limén Basin we find the Uscari formation, consisting of clays, sand, and mud. This material is covered by a marine sequence from the Mid- dle Miocene denominated Gatun, characterized by green lutites. The Uscari formation is dated by the inferior Miocene formation (Sprechmann et al 1979:69-96, 104). The Reynoard site includes a small beach and sediments from a small river originating in the Talamanca Valley and flowing south to north from the Limén Basin, Detailed geological information about this region is not available, Descriptively, to the cast a small beach contains grcenstones of dif- fercat coloration. To the west a headland is covered by thick tropical vegeta- tion. In the south an accumulation of river eobbles blocks the upper part of the river flow. The exact location of the site is not provided, to prevent destruction by pothunters. We visited the river only during summer, when access to its bank is pos- sible, Three systematic surveys were done, and'all involved the collection of greenstone samples. Because of their striking green color, the greenstones are easily seen through the wet beach sand. This location is considered an ideal potential source of precolumbian jade because the river is close to archacologi- cal sites in which a high concentration of jade was found during the construc tion of a railroad early in the twentieth century ‘THE MATERIAL GATHERED AND TESTED Our first samples were collected from a path in a banana field in the Atlantic region; they looked like green flints. These flints were cut and polished under the supervision of researchers from the Geology Department at the University of Costa Rica. The samples were later compared to precolumbian necklace beads and showed a strong resemblance to them. Michael Snarskis suggested that we show the materials to William Melson of the Smithsonian Institution, who was visiting Costa Rica at that time. Because of the great similarity be- tween the raw materials and the precolumbian items, Dr. Melson recom- mended a hardness test, a specific gravity test, and an X-ray analysis for both precolumbian and raw-material specimens. So far [have been able to carry out just one hardness test, using a stainless steel blade (see Table 3.1). A gravity test has been applied to approximately 40 pebbles. The majority of pebbles had a specific gravity between 2.1 and 2.6, which indicates a high concentration of quartziferous stones very similar in appearance to some jades. However, in addition, four small stones had specific gravities of 3.7, 36, 3.0,and 3.6, indicating a close relation to jade, which has a specific gravity between 3.3 and 3.5. Other kinds of materials associated with 65 PRECOLUMBIAN JADE TABLE 3.1 Hardness Test to Steel and Specific Gravity of the Sample SAMPLE owner HaRDNess AiR Wares I + 4703, 2838 Ib : 42.08 24.00 2 : 59.10 36.85 ey : 30.50 1691 3 Z 2.46 1450 4 3 Baz 1420 5 + 21.70 1352 6 + 47.64 29.45 i 5 1470 In 8 g 34.60 192i 9 E 404 10 = 350 u : 035 Ha : 4.87 2 : 134 B _ 30.28 16.28 14 a 739 3.90 15 : 2.05 120 16 + 82 0.60 16a \ 245 129 7 + 18.00 965 1a E 291 159 18 : 087 050 19 3 60 0.94 19a z 75.80 B46 20 + 065) 2a + 045 n 045 23 r 290 24 > 25 3 5 170 26 - 640 7 - 0.95 28 - 095 28a 150 28b B16 29 : 129 30 + 8.80 31 : 3.95 32 + 940 3 + 647 34 5 20.70 nao 35 _ 1124 5.93 36 - 730 430 37 2 15.70 940 370 + 225 150 38 : 245 Lo 39 - 250 135 B + 430 3.20 aw) 18.65 18.08 22.25 1359 896 927 818 18.19 698 1539 425 3.42 025 4.08 ass 14.00 349 085 022 Lis 835 122 037 086 3034 047 029 025 280 171 136 440 030 057 036 884 0.96 665 3.70 640 458 9.60 531 3.00 630 075 090 095 1.60 Srecirtc Gravery 25 23 27 23) 26 25 ae 26 19 18 21 2.0 14 18 16 18. 21 24 37 En 21 22 23 20 24 23 26 28 20 22 23 24 21 27 a7 24 23 23 20 24 24 21 24 24 24 30 27 26 30 Margarita Reynoard de Ruenes these stones were also collected. Ie is very likely that the river carried most of the metamorphic stones from the mountain chain, the sedimentary stones from the valley, and the voleanic stones from some fault. Evidently many jade or other gemstone ornaments were made of locally available pebbles and cobbles on the Atlantic Coast. The Reynoard site has a greenstone source with characteristics that might correspond to most of the precolumbian material obtained from the Atlantic region of Costa Rica. Such material is considered either jadeite or jade. However, further research is nec- essary to determine whether jade sources exist in this area. It is of great interest for mesoamerican archaeologists to define possible sources of jade: these areas were considered places of particular importance for the economy and the interregional relationships among the precolumbian populations. Discovery of jade sources would allow archaeologists to recog- nize foreign jade materials and enable them to trace the exchange routes that, existed. These exchange networks were vital for the support and development of precolumbian cultural systems. Since Lange (1984b) has posited exchange from Pacific sources (as yet un- located) for the presence of true jade artifacts in archaeological sites on the At- antic Watershed, further research on potential Atlantic sources is of critical importance. Nore Deepest appreciation is owed to the inspirational figure of Carlos Aguilar P., who founded the Laboratory of Archacology at the University of Costa Rica and carried out research throughout the country. His knowledge and energy have stimulated contribu- tions by three generations of students. A people is like a man, ‘who when he vanishes leaves nothing of himself unless he has taken the (precaution to samp his print into the stones of his path. —Francois-Félix Paure et eo) id 4 Jades in the Jade Museum, Instituto Nacional de Seguros, San José, Costa Rica ZULAY SOTO he precolumbian communities of Costa Rica left behind them beautiful works in a stone that we know by the nam jade.” OF the collections in the Jade Museum, the 2,134 jade and “social jade” specimens are the most im- portant, primarily because of their beauty. These specimens demonstrate out- standing artistic achievement with a material that was difficult to work with prehistoric technology. While serving as the basis for exhibition and inform- ing museum visitors about the importance of prehistoric Costa Rica, the jade collections have also served as the basis for research by international scholars (Balser 1980; Lange et al. 1981; Bishop et al. 1985; Bishop and Lange 1988), as have the ceramic collections (Snarskis 1982). We recognize the importance of jade to many New World cultures, in addition to those of Costa Rica. The earliest jade artifacts were made around 800 b.c. by the Olmec in the Gulf Coast and the South of Mexico. It has been suggested by many authors (Easby 1968; Balser 1974, 1980; Lépez-Calleja 1980; and Snarskis 1984) that Olmec groups from Mexico came to Costa Rica to look for blue-green jade as well as jadeite. So far there is no archaeological confirmation of these speculations. Thus far, no centers of jade production have been identified in Costa Rica. However, a geological study of the jade collections in the Instituto Na- cional de Seguras (INS), combined with the “criterion of abundance” (Bishop et al, 1982), suggests that a source, or multiple sources, remain to be found, Two geologists from the University of Costa Rica (Jorge Laguna and Siegfried Kussmaul) studied the jade pieces in the Ns collection and classified the 2,134 pieces into a number of mineralogical and geological categories (Figure 4.1). It is clear that a wide range of raw materials were used to produce artifacts, ‘The two geologists agreed that jade was most commonly present in small Clusters inside serpentine boulders, In Costa Rica these serpentine deposits are present only in the Santa Elena Peninsula (see Figure 2 in the Introduction). In reality these are not true serpentines formed by regional metamorphosis, 68

You might also like