Ch3 Archaeology of Region

The role of interaction across long distances has been a persistent theme in archaeological studies in the south-central Andean highlands. The emergence of the Tiwanaku state in the stark and sparsely populated altiplano seems improbable unless one considers the larger geographical setting and the apparent importance of sustained links over long distances that contributed to developments in the Lake Titicaca Basin. Archaeological research that focuses on the Formative Period, the time that preceded Tiwanaku in the Titicaca Basin, has demonstrated that regional centers coalesced out of a multitude of small villages that were articulated through a highly mobile sector of the economy based on camelid caravan transport. This chapter examines the roots of this mobile sector in the economy of the Terminal Archaic and the Formative by focusing on the transport of obsidian and the significance of early evidence of transport in the origins of long distance relationships in the south-central Andes.

This chapter will contextualized the theoretical issues surrounding raw material production and exchange in the south-central Andes. Subsequently, this chapter will explore some models of long distance interaction in the south-central Andes. Next, a review of the regional evidence for the production and circulation of Chivay obsidian from the a few major sites in the highlands will be followed by a temporal survey of Chivay distributions revealed by previous archaeological research. These regional data will be discussed by progressing through time beginning with Chivay obsidian use by early mobile foragers, and continuing in time through to the Inka period. A subsequent section considers broader patterns for use of obsidian in the Andes. This chapter concludes with a discussion of four models for obsidian procurement and circulation from the Chivay source, and the associated material correlates for these models.

A larger set of questions guides this chapter: Do the distributions of Chivay obsidian conform to those of utilitarian products like salt and dried meat, or do the distributions resemble those of prestige goods like spondylous, copper, or gold? Escaping the utilitarian / prestige goods dichotomy, a third group termed "cultural goods" better captures the traditional networks that circulated material like obsidian in the Andes. This chapter will argue that obsidian is an example of a class of material that moves between these simple classifications and that it requires a consideration of context. What pressures were responsible for the rapid change in production and circulation of obsidian during the Terminal Archaic Period? What can obsidian circulation reveal about the regional context and the strategies of early aggrandizers at the beginning of the Formative in the Titicaca Basin?

Developments in obsidian production at the Chivay source intersect with a wide swath of the Andean prehistoric sequence, and these developments will be explored below in three broad cultural periods: the "Archaic Foragers", "Early Agropastoralists", and "Late Prehispanic".

/misc/image004.jpg

Figure 3-1. Chronologies discussed in the text.

"Archaic Foragers" are characterized as mobile foraging groups responsible for the early human use of the obsidian source, exchange, and consumption of the material. Use and exchange of obsidian by Andean mobile foragers during the Early, Middle, and Late Archaic Periods will be considered in light of ethnographic and archaeological evidence.

Next, with the domestication of camelids and chenopodium occurring sometime during the Late Archaic, food production came to dominate the economy by 3300 cal BCE heralding sweeping changes that included early social ranking that first appeared during the Terminal Archaic and early Formative Periods. During this period, referred to here as the Early Agropastoralist period, the production and exchange of obsidian appears to have changed as the material became more widely circulated in the consumption zone and mobility was facilitated by the gradual inception of regular camelid caravan transport routes. Reciprocity-based exchange systems would have coalesced, while personal procurement activities persisted from the earlier period, resulting in distinctive production evidence in the source area. The evolution of expansive states during the Middle Formative and Late Formative through to the Tiwanaku Horizon shifted the regional political structure, and these developments had repercussions in the Chivay source region. As regional centers in the Titicaca Basin emerged and began to dominate the political landscape, the established exchange patterns may have shifted to reflect the role of redistribution believed to have been occurring at the centers. Political forces became increasingly powerful in the Titicaca Basin, as well as expanding out of the site of Wari far to the north in Ayacucho, and the procurers of Chivay obsidian in the Colca area were increasingly working in a border land between powerful polities. What kinds of evidence will be reflected by the combined impacts of local provisioning, reciprocity, and elaborate redistribution systems in the Chivay area and in the Titicaca Basin consumption zone? With the collapse of Middle Horizon polities, the effects of warfare on exchange are apparent in the limited distributions of obsidian during the Late Intermediate Period. Finally, the Inka period was a time of great regional integration and tremendous redistribution of goods, but it appears that obsidian had reduced significance in favor of competing materials like metals. The changing nature of prehispanic obsidian exchange, and the link between exchange theory and the south-central Andean sequence, will be considered in this chapter.

3.1. Andean Economy and Exchange

The Andes present a valuable opportunity for examining anthropological models of economy and exchange in prehistory. Distinctive historical aspects of Andean development, including the emergence of pristine states at high altitude, the administration of vast empires without a formal system of writing, and the wealth of ethnohistoric data provided by Spanish chroniclers, offer important research problems for economic anthropology. This study investigates the procurement and circulation of obsidian from the Chivay source in the south-central Andes during a broad time period that includes major shifts in the economy and in socio-political organization.

Throughout prehistory, the long, narrow Andean cordillera presented distinctive challenges to human groups that were addressed through a variety of technological and social strategies. Here, a focus on lithic raw material procurement and exchange permits obsidian circulation to serve as an indicator for particular types of regional interaction.

Exchange has a complex role in mediating human relationships over distance (see Section2.2.2), and in the south-central Andes, obsidian appears to have served as both a political tool and relatively ordinary aspect of economic activity. In part, the persistence of exchange in herding regions reflects lack of autarky among the dedicated pastoralists; they require vegetative and agricultural products, and such items are widely available in the sierra and foothills. In reference to the early development of long distance caravan networks in the altiplano, David Browman (1981: 413) notes "[t]he trade in consumables is less spectacular than the trade in luxury items, and more difficult to detect archaeologically, but it was much more important to the average altiplano inhabitant." In other words, due to the environmental contrasts in the Andes, relatively mundane consumables like salt, ajipeppers, and even coca leaf could precipitate a low-level but persistent demand for exchange of goods between adjacent ecological zones and, in some cases, across larger distances.

A central point of the following discussion is that wide-ranging exchange networks, apparently organized at the level of the household and facilitated by caravan transport, are a persistent theme in the south-central Andean highlands. These networks do not integrate easily with the exchange typologies presented above (Section 2.2.4), and this form of articulation is sometimes seen as irrelevant "background" reciprocity in models of early competitive leadership. However, this distributed mode of integration may have served as an early foundation for subsequent political organization in the region. This study focuses on obsidian procurement and distribution and infers that other goods were also being transferred along these networks. While the simple assumption that evidence of obsidian circulation is analogous to prehistoric trade in a multitude of other more perishable goods is problematic, the persistence of obsidian exchange in the south-central Andes is compelling evidence of generalized contact over distance.Andean approaches to regional economy are reviewed here in order to examine the distribution of obsidian and other goods through diverse mechanisms of procurement and exchange.

3.1.1. Economic organization and trade in theAndes

In a cross-cultural perspective, Andean exchange relationships throughout prehistory exhibited characteristic organizational traits of societies dwelling in mountain ecological zones; although in the late prehispanic periods distinctive features of inter-zonal control emerged in the south-central Andes. There is general consensus among Andeanists that the mechanisms of merchantilism and market economies - prices reflecting supply and demand - did not exist in the late prehispanic south-central Andes[2].

Polanyi's substantivist economic typology based on peasant households in non-capitalist settings is still widely used in the Andes with some modification.

Reciprocity, redistribution, and non-market trade are the institutional means by which indigenous Andean economies operate. All evidence points to the overriding fact that true market systems did not operate in the central Andes, as they did in central Mexico and in a number of complex societies of the Old World. Exchange did exist on a massive and pervasive scale, however, and the concept of administered trade is the superior means of understanding this phenomenon in the prehispanic central Andes. Trade existed, but it was not one based on market principles. Virtually all cases of trade were administered by some corporate group, constituted along sociological (kinship) or political lines (Stanish 1992: 15).

The major modes of economic interaction will be reviewed below. Finally, the closing section of chapter 3 will explore more specific material expectations of how each of these economic modes may appear in procurement areas such as the Chivay obsidian source.

Direct Access

Direct household procurement of goods is largely structured by the geographic relationships between consumer residence and a given resource. Access to products from complementary ecological zones by consumers who undertake a personal voyage to acquire and gather those goods is a form of direct access. For resources that are widely distributed in ecological areas or along altitudinal bands, direct access is a recurring theme in Andean prehistory. The procurement of unique or unevenly distributed or resources, such as salt or obsidian, is a different configuration entirely in a vertically organized region like the Andes, because distant consumers are forced to articulate with production areas far beyond their immediate and complementary neighbors in an altitudinally stratified exchange relationship. This kind of multiethnic, direct household procurement for salt occurs in the Andes to this day (Concha Contreras 1975: 74-76;Flores Ochoa 1994: 125-127;Oberem 1985 [1974];Varese 2002).

Direct access by foragers was the first mode of procurement in the Andes, and this acquisition mode probably dominated in Early Holocene prior to the population growth that permitted the development of reciprocity networks. The procurement of raw materials in a manner that is incidental to other subsistence activities is a more efficient means of acquiring these goods, an activity described as "embedded procurement" by Binford (1979: 279). Communities, ethnic groups, and even prehistoric states appear to have maintained direct access to resources in other zones, and (as is stipulated by the definition of the direct accessmode) this kind of articulation is for direct consumption or redistribution on the level of the corporate or state entity. This is part of a much celebrated pattern in Andean research, a feature known as vertical complementary (Murra 1972), a topic that will be returned to in more detail below. Direct access by states to unusual sources of raw material such as metals and minerals are well documented in the Andes. These include as Inka mine works, and distinctive Inka artifacts and architecture are commonly encountered in association with the procurement areas. In access between herding and agricultural sectors, there is also an ethnographically documented direct access mode described as "ethnic economies". In this direct access mode, ethnic groups will control parallel strips of vertical land holdings, and sometimes non-contiguous tracts, ranging from lower agricultural zones to high puna that may lie several thousand vertical meters above. This pattern is documented on the eastern slope of the Andes for the Q'eros of Cuzco (Brush 1976;Flores Ochoa, et al. 1994) and several communities in northern Potosí in Bolivia (Harris 1982;Harris 1985). The important concept of the direct access organization is that entities that were consuming the goods were directly responsible for acquiring them. If there is inter-household barter or transfer of any kind, then the arrangement likely belongs to a type of reciprocity relationship.

Reciprocity

The institution of reciprocity is important in all societies, and in the contemporary Andes reciprocal relationships are elaborate and permeate village life. It is an arrangement for the transfer of labor or goods that is organized without coercive authority between entities equal in status, although sometimes disputes are settled by community leaders. Andean labor reciprocity includes agricultural work, roof raising, canal cleaning, terrace building, and other services; as such, reciprocity structures the traditional village economy in the Andes (Alberti and Mayer 1974;Stanish 2003: 67). These kinds of reciprocal arrangements are frequently delayed, although compensation can be accelerated through recompense in products. For example, a herder might bring a caravan down to a farming area in the lower valleys and spend some days contributing labor to the agricultural harvest in exchange for some portion of the yield.

In many premodern economic transactions relationships of balanced reciprocity structured these arrangements. Two forms are likely in the Andes, that include (1) a multitude of small, household exchanges creating "down-the-line" artifact distributions (see Figure 2-2), presumably this type of exchange is responsible for the long distance transmission of small, portable goods for much of the pre-ceramic period. A second form consists of (2) barter exchange relationships with regular long distance caravans that articulated with settlements, and perhaps at periodic fairs, that transported goods over potentially greater distances. This mode would effectively consist of unadministered trade. The institution of reciprocity will be explored in more specific contexts below.

Redistribution

The basis of relations between political elites and non-elites in the prehispanic Andes was shaped by redistribution. In Andes during the later prehispanic period, redistributive mechanisms linked elites to non-elites through the redistribution of consumables like coca, and feasts of food and chichabeer in exchange for labor and political support. "The manipulation of redistributive economic relationships among the elite and their retainers, most notably of exotic goods and commodities, stands at the core of the development of Prehispanic Andean complex societies" (Stanish 2003: 68). Often these surplus goods were produced through efficient mechanisms orchestrated by elites, and the benefits and prestige derived from these surpluses would be accrued disproportionately by political leaders.

The central collection of goods for redistribution or use by the state includes taxation which, in the Inka period, was through mit'alabor. With respect to the exchange of lithic raw materials, Giesso (2003) argues that at Tiwanaku household stone tool production was a form of taxation. Giesso cites ethnohistoric evidence referencing the Inka period and argues that the household knapping of projectile points could have contributed to the provisioning of the state armory. The means by which non-local material arrived in the Tiwanaku homeland is unclear, but in the Inka case the raw material was acquired locally or it was provided by the state (Giesso 2003: 377). Earle (1977: 215) argues that redistributionshould be considered as two major groups with leveling mechanisms on one side and complex institutional mechanisms for wealth accumulation on the other (see Section 2.2.4).

Administered Trade

Non-market administered trade was another major means of transfer in the prehispanic south-central Andes. Building on Polanyi's classic definition of non-capitalist economic types, Charles Stanish (2003: 69) describes a form of elite-administered, non-market trade that was capable of procuring non-local goods and that provided wealth to elites as follows. Garci Diez's Titicaca Basin visita, a 16thcentury Spanish census document (Diez de San Miguel 1964 [1567]), describes how local elites in the Lake Titicaca Basin would have their constituents organize llama caravans for trading expeditions to adjacent regions. In neighboring areas such as the Amazon basin to the east, the Cusco valley, and western slope valleys, agricultural goods such as corn, fruits, and other products were sweeter, faster growing, and more abundant than in the Titicaca Basin. The colonial visitaindicates that, based on the colonial currency, corn in the Titicaca Basin was worth 5.7 to 6.9 times the amount that it was worth in the Sama valley (an area in modern-day southern Bolivia) where it was abundant. Stanish (2003: 69-70) argues that administered trade benefited elites because they were able to appropriate this difference in value, and through feasting and other ceremonial functions, a portion of this wealth was redistributed to commoners. It is important to mention that these same colonial sources indicate that the commoners also organized trading ventures and would take advantage of these elite-organized journeys to conduct private barter exchange on the side. With reference to herders conscripted into elite orchestrated trade caravans "those in Lupaca country 'who had their own cattle [cargo llamas]' (Diez de San Miguel 1964 [1567], f. 13v) went to the coast and to the lomas to barter on their own. …the maize growers on the irrigated coast were eager for the highlander's animals, their wool and meat" (Murra 1965: 201). Thus, elites organized large caravans and apparently possessed the surplus wealth and the camelid caravan animals in advance to initiate the trading expedition, but the herders that they conscripted also engaged in household-level barter. For the elites, their organizational efforts earned them significant wealth and status for a relatively modest outlay of costs. For the herders, it appears that they were able to embed household economic transactions with their mit'alabor service by conducting their side barter activities. Apparently, even the powerful Lupaca elite had to concede some independent trade activity to their caravan drivers. What, then, of the relationship between caravanerosand elites during earlier periods, when elites probably had less consolidated power than during the contact period?

The evidence suggests that "administered trade" was not the first form of long distance caravan exchange. As mentioned above, relationships of balanced reciprocity have long served to articulate herders with those living in complementary ecozones such as sierra agriculturalists, coastal fishers, and residents of the eastern lowlands. But the long distance transport of diffusive goods like obsidian are well-demonstrated and form a continuous network configuration that contrasts with the segmentary, vertically organized exchange between valley and puna (Figure 2-3). Archaeological distributions (Browman 1981;Burger, et al. 2000;Dillehay and Nuñez 1988;Nuñez and Dillehay 1995 [1979]) and contemporary ethnoarchaeological studies (Lecoq 1988;Nielsen 2000;West 1981) attest to the capacity for small scale, household-level organization of multi-week caravan expeditions. This evidence suggests that there were probably at least two major types of long distance caravans operating from the Late Formative onwards. The question then becomes: what was the relationship between household-level caravans and elite-administered trade? Did elites co-opt functions that were previously coordinated on the household level? If elites acquired control of some segment of caravan traffic, what strategies did elites use to wrest control from caravan drivers that, the evidence suggests, were very independently-minded people (Browman 1990: 419-420;Nielsen 2000: 517-520)? These questions concerning the origins and configurations of regional interaction in the south-central Andes are at the center of this discussion of changes in obsidian procurement and the regional circulation of goods in prehistory on the perimeter of the Lake Titicaca Basin.

3.1.2. Economy and exchange in mountain environments

Cross-cultural studies of human adaptation to mountain environments have revealed a number of common features between production strategies employed by people in the Andes, the Himalaya, and the Alps (Funnell and Parish 2001;Guillet 1983;Rhoades and Thompson 1975;Tomka 2001). These commonalities in adaptation to mountain settings include:

  1. Both specialized and mixed procurement systems are found, but there is a predominance of mixed systems.
  2. Specialized procurement systems are interconnected through regional exchange.
  3. Land holdings by a particular social unit can be non-contiguous and distributed across multiple ecological niches.
  4. In extensive procurement zones, such as grazing areas, land tenure is communal, whereas in intensive procurement, such as irrigated farmland, tenure is often family based.

These strategies are responses to characteristics of mountain settings that include altitude-based biotic ecozones, limited productivity in any single zone, and risk to herding and farming in production activities.

These regular features of production in mountain settings provide a comparison against which to evaluate procurement strategies in the Andes. A number of characteristics of production common to mountain environments have been inappropriately conceived as exclusively Andean in an essentialist tradition referred to as lo andino (Starn 1991;Van Buren 1996), while conversely others have sought to impose Andean models on regions where the model does not necessarily apply (Goldstein and Messerschmidt 1980). These models of regional interaction in mountain environments, both in Andean and general geographical models, can be contrasted with regional distributions of raw materials. Obsidian and other raw materials circulated widely in the Andes, and the spatial patterns described by these materials, may be examined in light of other regional patterning like stylistic distributions, as well as economic models of regional interaction.

Vertical complementarity

The contrasting ecological zonation found in low-latitude mountain regions worldwide has resulted in distinctive social configurations that appear to reduce risk, broaden the selection of products available in a given zone, and provide opportunities for strategic advancement by particular individuals or groups. These configurations have been investigated in two broad sets by Andeanists. First, there are a number of scholars who address vertical complementarity as a general process that is comparable with other mountain regions of the world. Second, there is a particular configuration known as "Vertical archipelagos" first described by John Murra (1972) that has been widely discussed in the Andean literature.

Verticality writ large

Vertical complementarity encompasses a variety of strategies for the problems posed by human use of resources at different scales, and by the broad natural diversity across relatively small distances in mountain environments (Aldenderfer 1993;Masuda, et al. 1985). These problems of articulation are addressed through mobility, through direct control of different zones by a single group, by mutualism between residents of different zones, and through a variety of exchange relationships. Vertical organization has been recorded among modern Quechua and Aymara communities (Brush 1976;Brush 1977;Flores Ochoa, et al. 1994;Harris 1982;Harris 1985;Platt 1980).

/misc/image005.jpg

Figure 3-2. Contemporary types of Andean zonation (Brush 1977: 12 ).

Based on contemporary observations, Brush (1977: 11) describes three systems:

  1. the archipelago type, or vertical complementarity, as conceived by Murra;
  2. a compressed type, where a single village can access and control all resources without resorting to means of long distance control;
  3. the extended type that corresponds to horizontal complementarity.

On a more localized scale it is possible to see vertical control strategies within a particular valley in a mixed agropastoral strategy that has been called "compressed archipelago" (Figure 3-2). In the central part of the Colca valley on the western slope of the Andes in Arequipa, Peru, Guillet describes vertical household relations.

To what extent do households integrate both puna pastoralism and valley farming into their production and exchange strategies? First, most households residing in the puna tend to specialize in herding and do not have agricultural fields that they cultivate directly. Similarly, many village households neither belong to family surname groups with access to puna pastures nor count themselves among those who have gained control of communal pastures ( botaderos) on the slopes behind the village. Households that follow such specialized strategies must perforce use the exchange nexus to obtain complementary products (Guillet 1992: 133).

Additional evidence for micro and macro vertical complementarity is discussed in the context of the Colca valley (Casaverde Rojas 1977: 172;Málaga Medina 1977: 112-113;Pease G. Y 1977;Shea 1987: 71).

Vertical complementarity can viewed as an anthropological principle that describes the propensity for social groups in mountain environments, from foragers to state societies, to geographically broaden their social and economic base and reduce risk by exploiting a variety of environmental settings (Aldenderfer 1993;Guillet 1983;Salomon 1985). Salomon (1985: 520) presents complementarity strategies in prehispanic Ecuador as varying in two dimensions.

/misc/image006.jpg

Figure 3-3. Diagram of institutions of Andean complementarity (from Salomon 1985: 520). Numbered modes reference the "Modes of Trade" in the figure from Renfrew (1975: 41-43 ) shown in the previous chapter as Figure 2-2.


One is between decentralized systems based on reciprocity, and the other is based on centralized systems of redistribution. There is an underlying neo-evolutionary correspondence implied in many of these models, as chiefdoms and states are believed to have been responsible for the network convergence perceived in redistributive systems, however it is important to observe that due to the variety in products, social relationships, and economic configurations, it is likely that a great many of the institutions presented in Figure 3-3 occurred simultaneously, and in general there is no direct correlation between confluence and evolutionary typologies. Another important dimension involves the number of political units participating in the interaction ranging from direct access, dyadic relations, exchange systems and open barter. The vertical complementarity literature in the Andes is valuable for considering prehispanic exchange relationships in that it has compelled a number of scholars to explore explicitly the relationship between ecological zonation, production, and social organization.

Vertical archipelagos

The work of John Murra on vertical complementarity has been among the most influential ethnohistoric studies in the Andes. The premise of the vertical archipelagos model is that the rapid altitudinal change along the flanks of the Andes produced a pattern where social groups residing in non-contiguous ecological strata formed distinct communities that developed around intensified production in these strata. Polities and ethnic communities sought to control a variety of these resource pockets following the Andean ideal of self-sufficiency.

Murra's (1972) seminal article distilled observations from ethnohistoric sources, in particular the visitacensus Garci Diez (1964 [1567]) of Chucuito in Puno, conducted only 35 years after the Spanish invasion. Murra showed that late prehispanic altiplano societies obtained direct access to products from a variety of ecological niches through this practice, and that the strategy was a guiding model of organization in some Andean polities (Salomon 1985). According to Murra (1985) the principal characteristics of the vertical archipelago model can be summarized as follows:

  1. Ethnic groups aimed to control numerous "floors" or ecological niches in order to exploit the resources that are found there and maintain self-sufficiency. As a consequence, powerful polities residing in the altiplano had considerable wealth in pastoral herds but also controlled land on both the western and eastern flanks of the Andes.
  2. The population was concentrated on the altiplano and those living in the peripheral 'islands' maintained continuous social and economic contact with the core region.
  3. The institutions of reciprocity and redistribution guaranteed that those living in 'islands' maintained the rights to products available in the center. These rights were defined through kinship ties that were ceremonially reaffirmed in the center.
  4. A single 'island' could be shared by different ethnic groups, though the coexistence could be tense.
  5. The relationships and functions of these colonies changed through time and may have become more exploitative. They are perhaps an antecedent to the mitimaqstrategy of the Inka empire.

While a chiefdom level of organization and centralized power is a principal characteristic of the Late Intermediate polities that practiced vertical complementarity in the period examined by Murra, the concept was been explored and expanded by archaeologists in the subsequent decades. In Murra's original description, verticality referred explicitly to direct control of a diverse resource base without engaging in trade with other ethnic groups, thereby preserving what Murra (1972) has described as the ancient Andean ideal of economic self-sufficiency that permeated Andean society and ideology far beyond his Lupaca case study. Stanish argues that Murra was explicit about excluding exchange processes, in part, because he perceived "a structural linkage between exchange and markets" (1992: 15). As prehispanic market mechanisms were absent in the south-central it was believed that barter and exchange were also of minimal importance and other means of articulation, such as direct control, were emphasized. Murra, however, later modified the definition to include specialized exchange centers, a change which Browman (1989: 324) argues confuses the issue because it subsumes a variety of processes into a single model.

On a theoretical level a further limitation of the original 'verticality archipelago' model lies in its adaptationalist orientation (Earle 2001).Adaptationist models of exchange and regional control have their basis in Service's (1962) proposal that these arrangements arise from environmental diversity, and then chiefs emerge to administer and redistribute goods produced regularly by their retainers.As observed by Van Buren (1996: 346), the archaeological origins and perpetuation of the archipelago pattern was founded on the assumption that groups benefit, as a whole, from the control of multiple tiers and the ecological resources that are produced in those archipelagos. As mentioned above in the discussion of administered trade, colonial documents emphasize the independence of commoners and the ability of subjects to practice subsistence barter, a pattern that leads authors to suggest that the vertical archipelagos pattern may have had more of a political basis than a foundation in ecological and subsistence practices. The ultimate roots of such a system may lie instead in the capacity of the rulers of such groups to organize larger scale trade and convert the value differential between products in the different ecological zones into political prestige through feasting and ceremony (Stanish 2003: 69-70).

Horizontal complementarity

Soon after Murra published his 1972 paper about vertical complementarity, researchers began noting that vertical complementarity is only one of a number of strategies employed in the Andes. A different kind of geographical interaction pattern, one that stays within broad regions such as the altiplano or the littoral, has been called horizontal complementarity. Contrasted with Murra's vertical complementarity model, in horizontal complementarity polities would directly control many parcels in a given niche. Here, the adaptationalist argument is somewhat less obvious since a horizontal complementarity strategy seems redundant unless particular resources were only available in one sector of a given horizontal territory. This phenomenon could reflect a cultural mechanism for the continued social integration of communities distributed widely across a vast region. An analogous situation can be observed in the culturally-constructed Yanamamö trade for ceramic wares, maintained to provide a social catalyst for villages to come together for feasting and marriage-making, despite the risk of conflict and treachery (Chagnon 1968).

Relations of horizontal complementarity between coastal valleys have been observed on the littoral (Netherly 1977;Rostworowski 1977;Shimada 1982), and for the higher altitude altiplano area where this kind of organization is referred to as the "altiplano mode of economic integration" (Browman 1977) and, to a lesser extent, the modern "extended type of Andean zonation" (Brush 1977: 12-16;Gade 1975). The extended type takes place in regions of expansive, contiguous production where exchange forms the basis for circulating products from other zones. In the expansive altiplano where there are fewer impediments to travel and, with the domestication of camelids, relatively low transport costs with cargo animals, materials may have been conveyed over substantial distances.

/Figs_Ch3/region_exchange_types.jpg

Figure 3-4. Subsistence exchange for products by ecozone versus single-source, diffusive goods.

A general schematic of these contrasting relationships is shown in Figure 3-4. While there is a great deal more complexity and variability to economy and exchange than is communicated in Figure 3-4, for example tubers are grown in the high elevation zones and herding does occur below 3800 masl, the purpose is to highlight the contrasting network characteristics. These two characteristics include

(1) Subsistence exchange.The acquisition of products available by zone contrasts

(2) Diffusive.The network for goods that radiate diffusively, such as obsidian and salt.

While the mixed agropastoral strategies are also very common, particularly in ecotone areas such as the puna rim, the subsistence exchange that articulates dedicated pastoralists with agriculturalists has been documented ethnographically in many studies of which only a selection are shown in Figure 3-4 (Bastien 1978;Brush 1977;Custred 1973;Flores Ochoa 1968;Palacios Ríos 1981;Winterhalder and Thomas 1978).

Ecological complementary has long served to transfer goods between ecozones, however the movement of goods laterally through areas of homogeneous resources, such as horizontal complementarity across the puna, implies other mechanisms of transfer. In addition to strategies described previously, it has been proposed that large periodic markets are a possible solution to the problems of social and economic integration among agropastoralists living in widely distributed settlements.

Browman (1990: 405-411) reviews the evidence for prehispanic and colonial period markets in the Peruvian and Bolivian altiplano. The colonial period aggregations in the Andes are products of historical circumstances, however Browman's evidence suggests that periodic fairs in some form were a feature of the prehispanic economy. While evidence concerning the actual goods exchanged at periodic fairs during the prehispanic period is scarce, these events could have been an effective way for obsidian to have been distributed in highland Peru and these fairs will be discussed in more detail below. Seasonal or annual gatherings are also ethnographically known among foragers in arid regions of the world with low population densities (Birdsell 1970: 120;Steward 1938). Social aggregations in some form can be expected to date back to the foraging period and these aggregations probably included, among other activities, the exchange of goods.

Evidence for barter, markets, and marketplaces in theAndes

"They did not sell… nor did they buy… they thought it was a needless operation" and all universally planted what they needed to support their households and thus did not have to sell foodstuffs, nor raise prices nor did they know what high prices were…(Garcilaso 1960 [1609]: 153; cited in Murra 1980: 142).

Was obsidian exchanged through a market system reflecting supply and demand in the prehispanic Andes? Were exchange specialists present in the prehispanic Andean highlands? If so, when did they appear and what did they transport? While extensive exchange has been documented in the prehispanic Andes, it is widely believed by Andeanists that exchange in the prehispanic central Andes was not based on market institutions. Building on the typology developed in Chapter 2 following Renfrew(1975)and Salomon(1985), differing degrees of specialization and independence can be expected to have existed among actors in exchange relationships (Figure 2-2andFigure 3-3). If specialized traders were present in prehispanic Andes, how did they interface with state authority during the Middle Horizon and Late Horizon, and what was their position during periods of regional conflict like the Late Intermediate Period?

Much of the evidence supporting the alleged lack of market-based exchange is derived from studies of Inka economic organization, where ethnohistorical accounts and archaeological datasets converge. A great deal has been written on the topic of Inka economy, and obsidian exchange was relatively diminished during this period, therefore this discussion will be limited to a few relevant issues regard exchange specialization. Several of the sixteenth century chroniclers are clear that while barter was widespread, the barter values of goods did not reflect fluctuations in supply and demand as in a market economy. However, the lack of consensus on the issue of markets and commerce during the Inka period stems from inconsistency in the cronistasthemselves. As reviewed by Murra(1980: 139-152)and LaLone(1982)the denial of a market exchange by Garcilaso de la Vega (quoted above) can be contrasted with numerous accounts of large and small markets, and a long tradition of barter exchange of various types.

The issue of Late Horizon marketplaces and market exchange is explored here because one of the principal questions that may be considered with changing obsidian distributions through time is the possible role of commercialism in prehispanic Andean exchange. The appearance of exchange specialists, such as freelance caravans moving certain commodities and responding to the changes in barter values that result from surpluses and shortages, would have presented a mode of transport distinctive from that of local reciprocal exchange or regional or state redistribution.

The vertical archipelago model functioned as an alternative to trade for goods from neighboring areas because "regional differences in production were, by preference, handled by means of colonization instead of through barter or trade."(Murra 1965: 201). As mentioned, Stanish argues that Murra initially excluded exchange mechanisms because of a perceived association between exchange and market economies. Further it can be argued, building on Appadurai(1986: 33), that in certain contexts of ranked or stratified societies with elaborate redistribution mechanisms, market systems of exchange represent a threat to the centralized ideological power of redistribution. "There is great advantage to leaders who are able to portray their resource-control strategies as reciprocity, redistribution, and generosity. Non-centralized resource-control strategies are, by definition, not 'control' strategies"(LaLone 1982: 296). If centralization is a principal determinant of state control on market exchange, were the peripheries involved in greater numbers of barter transactions?

Merchants in the Andes: Late Prehistoric and ethnohistoric evidence

Coastal Trade

The strongest evidence for merchant specialists in the Andean region comes from relatively peripheral areas of the Inka Empire, from the Pacific coast of what is now Peru, and from the coast and highlands of prehispanic Ecuador. The coastal Late prehispanic traders of Chincha, near the Paracas peninsula, have been described as consisting of 6000 merchants who traded in Cusco, among the Colla (and presumably the Collagua), and in Ecuador, but little is known about how these expeditions were organized(Patterson 1987;Rostworowski 1976;Rostworowski 1977;Sandweiss 1992). LaLone(1982: 308)notes that Rostworowski was not able to connect Chincha traders with marketplaces despite her assertion that these represented commercial exchange. Sandweiss(1992: 10)believes that Chincha trading expanded a great deal under the Inka following the Inka conquest of the Chimu to the north.

Coastal products such as spondylous and other goods were known to have been transported in large balsa rafts. There are numerous contact-period accounts of merchants plying the Pacific littoral beginning with the renowned loaded boat of balsa logs encountered off the coast of Ecuador by Pizarro on his second trip south, several years before the actual Spanish invasion of the Andes(Hemming 1970;Murra 1980: 140). The boat had a crew of 20 and had a small cabin and cotton sails. Murra is confident the boat was an Inka "registry" because the crew knew Quechua and a few were captured by the Pizarro's army who later used them as interpreters. While the Spanish paid particular attention to precious metals, the contents appear to have contained wealth goods including gold and silver ornaments, bracelets and anklets, headdresses and mirrors, and a great deal of cotton, wool and rich embroidery. A small weighing scale as well as a great deal of shell, probably spondylous and strombus, were found on board(Hemming 1970;Murra 1980: 140). Although the activity of this boat was described as trade (the word rescataris used in the text) by the Spanish observer Sámano-Xerez written in 1527-1528(Porras Barrenechea 1937: 21), it is highly likely that this boat, with its cargo of elite goods, was in fact carry ritual offerings from the Inka to some northern destination. The evidence for long distance exchange between Ecuador and Mesoamerica has been long been a topic of interest in New World archaeology(Coe 1960;Zeidler 1977).

Ecuadorian highlands

The strongest evidence for mercantilism and markets in the prehispanic Andean highlands comes from colonial Ecuador(Hartmann 1971;Salomon 1986). Hartmann argues that the Inka economy had a significant market component based the following evidence: (1) the Spanish saw gatherings that they identified as "markets" from the very earliest reports, although as they were coming from Mexico the Spanish used the Nahuatlword " tianguez"; (2) commodities were plentiful and varied, including both staple and luxury items; (3) both Quechua and Aymara had specialized terms for buying and selling; (4) market activity was not suppressed by the Inka authorities, only regulated to suit their interests(Hartmann 1971).

The existence of markets in prehispanic Ecuador is a particularly interesting question because the Quito area was conquered by the Inka only 30 years prior to the Spanish invasion and therefore the region had only recently been absorbed into the Inka Empire. Salomon(1986)examined the ethnohistoric evidence for precolonial and colonial markets in highland Ecuador and found that the contact period evidence provides insights into pre-Inka customs as well as the Inka response. The Quito valley is in a position to serve as a hub for the transfer of products from the Amazon lowlands, the Pacific coast, and the páramohighlands. In this sense, the Quito valley is in a similar geographical configuration, but on a smaller scale and a different ecological zone than the Lake Titicaca Basin. The strong dependence of early Spanish residents on the markets, and the founding of new markets by Spaniards, leads to some uncertainty as to the precolonial importance of markets. However, there is a variety of evidence for a pre-Inka merchant class in central Ecuador called mindaláesthat gathered in stationary markets and controlled trade in cotton, coca, and salt that they would bring from lower-lying regions(Salomon 1986: 203-204). Barter exchange between non-specialized traders occurred as well, typically of household surplus goods, and Salomon argues that both mindalamerchant organization and non-specialized barter were ancient developments in Ecuador. In contrast, Patterson(1987)argues that the merchantile organization was not a long-established system but rather a Late prehispanic period response to opportunities presented on the northern border of Tiwantinsuyu. Vertical archipelago organization is also found in Ecuador both in agricultural production and in targeted procurement communities such as stable colonies for salt production(Oberem 1981 [1976]: 79). However, Salomon shows that archipelagos were, in most cases, a late phenomenon that was introduced by the Inka. In addition, obsidian distributions in Ecuador can provide insights into Andean exchange in a context with functioning markets(Burger, et al. 1994).

As for trade specialization elsewhere in the Andes, LaLone(1982: 307)sees a latitudinal gradient from north to south where markets and freelance traders may have been more abundant in the northern periphery of the Inka empire, and subdued or non-existent in areas full under Inka control. A notable exception to this gradient are the sea traders from Chincha(Rostworowski 1977;Sandweiss 1992). The evidence is far less secure for the southern periphery of the Inka Empire, but the implication of the Ecuadorian data is that solving zonation problems through the vertical archipelago approach was promoted by the Inka in Ecuador following the Inka conquest, which perhaps calls into question the pervasiveness and pre-Inka antiquity of the vertical archipelago strategy in the south-central Andes as well.

Traveling Peddlers in the south-central Andes

In contemporary contexts, peddlers are found with frequency in areas that form boundaries areas between different commercial spheres of interaction and lacking in consistent distribution of goods. Browman (1990: 422) reviews ethnographic evidence for mobile peddlerswho perform bulk-forming and bulk-breaking services in the south-central Andean highlands. The peddlers will provide manufactured items to rural pastoral communities, though often at a substantial mark-up, and will trade for items like hides and wool in time for purchasing fairs in the regional centers. The puna between Lake Titicaca and the western slopes in Arequipa are particularly active with comerciantes ambulanteswho schedule their travel cycles to correspond with patron saint festivals, as well as distributing goods to communities without regular markets (Flores Ochoa 1977: 148 ;Flores Ochoa and Najar Vizcarra 1976) as well as traveling herbalists and related groups (Bastien 1987). In the small community of Cerrillos in southwestern Bolivia near the Argentinian border, Nielsen (2001: 166) reports that peddlers, referred to as cambalacheros, would pedal bicycles from the city of Oruro bringing clothes and metal pots to sell or to barter for hides.

The Collaguas of the Colca valley were frequently visited by itinerant peddlers from Puno, according to Casaverde (1977: 185). These vendors known as polveñosmaintain established compadrerelationshipswith Colca valley households in order to have reliable hosts and potential buyers in the valley. Although transactions frequently take place through barter and the host and other residents are not obliged to trade, the profit motive of the peddlers is understood. These examples illustrate some of the variety in forms of distribution that may have had some basis in the prehispanic economy.

Limitations of ethnohistory and ethnography

Ancient economy in the Andes has been most fruitfully studied by combining archaeological evidence with ethnohistoric and contemporary sources. General models of great importance to Andean studies such as vertical complementarity, the question of the commerce in the prehispanic economy, and a structure of dual organization in the Inka administration are examples of models emerging from ethnohistoric sources but with empirical support from archaeological research. Ethnohistorically based models also have their limitations in that they can unduly influence interpretation much in the way that the "tyranny of ethnography"(Wobst 1978)should not necessarily define the range of archaeological possibility and inference. Ethnographic sources are valuable for the evidence of pastoral patterns, the priorities of caravan drivers, and the articulation between mixed and specialized pastoral and agricultural practices. However, modern features including the presence of the cash economy, markets, truckers plying the highways, and various modern job opportunities impact the structure of exchange relationships and caravan transport(Browman 1990;Nielsen 2000).

Ethnographic studies of contemporary llama caravan drivers are focusing, by necessity, on relatively marginalized communities that are sufficiently conservative to continue to herd llamas despite a variety of often more lucrative alternatives (Lecoq 1988;Nielsen 2000). Scholars have noted, however, that caravan drivers enjoyed relatively high status and autonomy in the late prehispanic and early colonial period which contrasts strongly with the economically marginalized modern day caravanero(Murra 1965). Furthermore, herding has become relatively low status in many regions of the contemporary Andes because the economic focus has moved downslope since Spanish contact due to the growth of the coastal economy and the importation of various low altitude crops and livestock. In other words, the prominence of caravan drivers as central economic agents has been greatly diminished in recent centuries, and ethnographically documented interactions with agriculturalists is one circumstance in which this relative shift in power may result in distorted perceptions of prehispanic interaction patterns. The relatively high status accorded to truck drivers and other purveyors of goods and information in contemporary Andean villages was likely to have been ascribed, instead, to the relatively cosmopolitan drivers of llama caravans during prehispanic times.

3.2. Long distance trade

The transport of goods over distances that exceed immediate complementarity relationships in mountain environments is well-documented archaeologically and ethnographically in the south-central Andes. Point specific resources like obsidian and salt have a distinctive, radiating distribution pattern as compared with subsistence exchange between ecological zones (Figure 3-4). Mechanisms that include direct acquisition and down-the-line exchange (boundary reciprocity) are likely to have been long-term exchange modes that served to disseminate of goods horizontally through a single ecological setting like the Andean altiplano. However, it is known ethnohistorically that long distance transport with the aid of camelid caravans, either with direct procurement and including very few transfers (long distance trade caravans), was a common method for the lateral distribution of goods. When did long distance caravan transport begin to dominate regional exchange in the south-central Andes, and who initiated this form of transfer between far-flung populations? Principal factors that influence the origin and perpetuation of long distance trade routes by highland pastoralists in the south-central Andes include the following features.

(1) Cargo animals:While not exceptionally strong, llamas are effective cargo animals because they are relatively compliant, they are not water tethered, and they can consume a range of grasses found on the altiplano so that they do not have to transport their own fodder.

(2) Topography:By virtue of the open and predominantly low-angle topography of the altiplano, the movement of loaded cargo animals across the altiplano requires lower effort than travel along the eastern or western sierra that are bisected by deep valleys.

(3) Resources:On an inter-regional scale, the altiplano divides complementary resource areas from the Amazon lowlands to the Pacific Ocean and these converged on the altiplano during particular time periods.

These features created circumstances that allowed for the wide-distribution of materials like obsidian and other products in the south-central Andean highlands. Long distance exchange and spatial relationships have been presented as a primary factors in the appearance of early social complexity during the Middle and Late Formative (Bandy 2005;Stanish 2003: 159-164). However, given the antiquity of camelid domestication in the Andes the long distance caravan pattern probably predated the Middle Formative by a millennium or more. The presence of caravans and the transport of complementary goods around the high, flat altiplano are part of a number of characteristics that created the circumstances within which social inequalities evolved. Competition for social power emerged during the Formative from a context that included these features of long distance exchange both in terms of the capacity for regional interaction, and the social institutions that surrounded the organization and scheduling of exchange in the region. These regional exchange mechanisms had long term consequences based the theory of Clark and Blake (1994: 17) who argue that social ranking was the unintended outcome of early political actors, operating within the institutional constraints of their circumstances, pursuing short term prestige goals for themselves and for their supporters. Following this model, the established circulation mechanisms of non-local products during the Early Formative, and perhaps earlier, are likely to have had a significant influence on the strategies pursued by aggrandizers during subsequent periods such as the Middle Formative.

3.2.1. Household-level caravans

Adaptationalist explanations for the origins and significance of long distance caravan networks are unsatisfactory, yet the more explicit aggrandizer models for the rise of elite-administered caravans that are documented ethnohistorically, refer to a regional-scale phenomenon that occurred relatively late in the prehispanic past (Stanish 1992: 14;Stanish 2003: 69). As was discussed above, the evidence suggests that household level organization of long distance caravans should be considered as a possible hypothesis for the long distance circulation of goods prior to the Middle or Late Formative when archaeological evidence for ranked society appeared in the Titicaca Basin.

In terms of the two configurations described above as subsistence exchangeand the single-source diffusivegoods (Figure 3-4X), the capacity and incentive for long distance caravan activity appear to be much older than the evidence for elite administration. In other words, on a geographical level and in terms of archaeological distributions one should consider that long distance caravan exchange was possibly organized on the household level thousands of years before elites were clearly organizing labor for the construction of monuments and the hosting of feasts.

What were the contexts in which individuals and households to began to organize long distance caravans in prehistory? Many of the "diffusive" goods with radiating distributions were not required for subsistence, strictly speaking. While small quantities of salt are biologically necessary for humans, salt is available in low densities in many parts of the altiplano and hunters and pastoralists can actually acquire salt from the consumption of meat and blood. In terms of the need for obsidian, it is evident that obsidian has flaking characteristics not available in other stone materials, but high quality cherts are available in many regions where the archaeological evidence shows that obsidian was imported from relatively large distances. Semi-precious stones like lapis lazuli are found in archaeological contexts such as burials after 3300 BCE in the south-central Andes. A variety of perishable goods probably circulated along such exchange networks as well, and as wild plants and agricultural products from particular regions are renowned to this day, such as hot peppers, sweet corn, herbs, and potent coca leaf, there was likely to have been variability in products available from particular valleys over others (especially during the earlier stages of plant domestication). If these products were transferred by caravans, perhaps in dried form, they would have radiated along distributions that were closer to diffusive sources like obsidian.

At risk of creating a rigid dichotomy, these items that were relatively common-place in some regions may have been closer to "prestige goods" because they were transported substantial distances. Following agency models for early leadership, those who mobilized resources to acquire such goods may have been striving to differentiate themselves from their communities and their neighbors. However, it is important to note that many of these goods, such as obsidian, do not appear to have been used exclusively by a small, restricted segment of the population in the archaeological contexts where they are found. Nevertheless these goods were widely transported and the mere presence of these goods indicates that some effort was expended to acquire the product (and, according to Hayden (1998: 44), these are automatically a form of prestige technology). In terms of the importance of "ordinary goods" (Smith 1999, see section 2.2.2), such products may have had important cultural and ethnic associations that lent value to the acquisition of these goods and prestige to those who could acquire them while simultaneously having widespread availability in a community. Archaeologists often define "prestige goods" based on spatially-delimited contexts that imply restricted or elite contexts of consumption, and this sometimes leads to claims that these alleged prestige goods were the impetus for long distance exchange. Rather than inferring that prestige goods defined the exchange networks and then "everyday goods" followed suit, Smith (1999) notes that the cultural information and these non-local yet widely-consumed materials like obsidian are perhaps better considered as "cultural goods", linked to ethnicity signaling and practice among herders (Nielsen 2000: 521-526), and complemented by other markers of ethnicity the most visible of which were probably textiles. As regional links became established, obsidian may have become a means for demonstrating participation and cultural affiliation with the subsistence level networks that exchanged goods including obsidian. Even if archaeologists are not able to detect exclusivity in the access to such goods in intra-site spatial distributions, as the materials were widely consumed in the community, it seems probable that the ability to acquire these goods reflected positively on the individual or household capable of procuring non-local products.

3.2.2. Incentives for early caravan formation

Down-the-line exchange can account for many archaeological distributions of cultural goods such as obsidian, both in volume and in temporal persistence. In terms of efficiency, down-the-line reciprocity networks require lower travel distance and less risk than is incurred traveling through unfamiliar territory (Renfrew 1975: 44). Furthermore, as Nielsen (2000: 24, 514) notes, caravan driving is hard work and it only becomes worthwhile when distances are large and/or loads are heavy or bulky. These bulky items tend to be transported medium distances (40-120 km), but smaller and more valued items might be transported long distances (Nielsen 2000). Traveling peddlers also carry small and light items without the added responsibility of managing many camelids in a caravan, therefore in some ways caravans and peddlers are complementary alternatives for the independent circulation of goods. Nielsen further observes that ancient caravans probably carried a variety of products in their cargo to reduce risk, and they likely carried anythingthat was worth transporting and that could be traded within their social and cultural parameters.

Given the evidence for increased sedentism from the Terminal Archaic and onwards (Aldenderfer 1998;Craig 2005), down-the-line exchange may be considered as the null hypothesis, with direct procurement and household level trade caravans as alternative hypotheses. Consistent but low levels of down-the-line exchange embedded in social relations form the background against which to examine other exchange modes such as caravans and possible non-market mercantilism. What kinds of possible incentives existed for the development of household-level caravans in order to acquire such goods, given the wide variety of potential trade and consumption patterns in prehistory? Incentives for household level caravan organization may include

(1) The maintenance of a social network and a demonstration of extra-local alliances that exceed the level of basic regional interaction implicated in subsistence-related exchange.

(2) Demand for a greater quantity of types, and the number or sizes of products than can be acquired from down the line exchange.

(3) A need or desire to exceed the products available to one's neighbors or to avoid dependency on an entire reciprocity network of neighborly relations.

(4) In a biological adaptationist framework, the risk and potential of caravan driving was a form of costly signaling (Aldenderfer 2006;Craig and Aldenderfer In Press).

Secure evidence of caravan activity, beyond simple evidence of pastoralism from faunal remains and evidence of corrals, may be a challenge for archaeologists and the systematic study of spatial distributions of non-local goods may shed light on the differences between subsistence pastoralism and caravan trade activity. One example of caravan activity from consumption contexts may come in the form of distinctive evidence of pooling of certain goods at particular sites along major transportation corridors, but establishing changes in relative density probably requires representative samples from archaeological collections from contexts that were presumably major caravan networks and those that were more peripheral to principal caravan routes. Sourcing from contemporary household-level assemblages may serve to differentiate such pooling (Figure 2-4), as was attempted with Oaxaca obsidian by Pires-Ferreira (1976). In the south-central Andes, stronger evidence for differentiating caravan trade from down-the-line trade may come from differences in artifact form and context from a variety of non-local products. In short, differentiating caravans from down-the-line exchange will require inference from multiple lines of evidence and from domestic contexts, evidence that is extremely scarce in the south-central Andes.

3.2.3. Exchange between herders and farmers

Ethnographic accounts of the interaction between caravan drivers and agriculturalists may be colored by the comparatively low status of caravan drivers in modern circumstances. Among isolated agricultural valleys in prehispanic Andes caravan drivers likely represented an important link to both information and non-local products. As the owners of the means of transport and the initiators of long distance interaction, caravan drivers were in a strong position to influence trade negotiations with dedicated agriculturalists that did not have their own llama herds or did not have schedules that permitted them to undertake long voyages. Strong links and commitments between herders and particular agriculturalists are, then, in some ways against the interests of herders because in a market context they "held the cards" in terms of negotiating favorable conditions of exchange. From the perspective of regional economic interaction two principal groups that, as per Dillehay (1993: 253), largely complement one another and may result in a relatively stable political and social environment:

(1) Mobile herders with an economic focus on hunting, pastoralism, and who are limited in their scheduling by the needs of the herd and annual cycle caravans.

(2) Relatively sedentary agricultural communities, often with an animal husbandry component as well. These households are largely restricted in their scheduling and long distance travel due to the requirements of agriculture.

In this configuration the mobile herders have greater autonomy, but they depend on articulation with dedicated agriculturalists. The relationship between mobile pastoralists and sedentary agriculturalists in Old World contexts has been the subject of comparative study by Khazanov (1984: 198-227). One common pattern is for trade between nomads and sedentary society to be manipulated for political purposes by administrators and elites in the regional centers. In the fifteenth century, the Chinese state sought to regulate exchange with nomadic tribes occurring at trading posts on the frontier. A recurring pattern occurred where the Chinese government would attempt to control the nomads by restricting trade, and nomads would, in turn, "acquire the right to trade by using arms." (Khazanov 1984: 206). That is, if exchange was curtailed the nomads would resort to violent means as described by Levi-Strauss (1969: 67) and Sahlins (1972: 302) who note the link between exchange and warfare (Section 2.2.5).

In the Near East, nomads were in a more profitable position because often they were the essential link between isolated oases. While administrators may have sought to control nomads trading with their communities, the demand by farmers for the milk and meat products from nomads, and the transportation services offered by nomads, often placed the herders in an advantageous position (Khazanov 1984: 208). In the Andes, a similar pattern linked agricultural valleys like the Colca in the high sierra, the mid-altitudes, and perhaps the littoral, by way of camelid caravans, although by the Late Prehispanic it appears many valley communities had their own large herds pastured in the adjacent puna.

3.2.4. Types of products carried by caravans

Ethnographic and ethnohistoric accounts of the types of goods carried by llama caravans provides a context for discussing the volume of trade and spatial relationships in the prehispanic period. Nielsen (2000: 65-67) observed that caravan drivers transport virtually anything that they think they can transport and will be able to trade, and therefore the argument that certain caravans were transporting purely subsistence goods, while others carried purely prestige goods, is probably unfounded. Nielsen notes that as part of their diversification strategy, caravans in prehistory probably variously transported subsistence goods, cultural items, and prestige goods as they articulated with networks at different levels with a diversity of social associations.

Informative ethnohistoric evidence for caravan transport comes from reports describing the provisioning of the infamous silver mines at Potosí that are reviewed by Browman (1990: 408). These reports state that 40,000 llamas were reserved by Potosí for provisioning and another 60,000 were brought as support for indigenous workers fulfilling their tax obligations through labor. Although with these Potosí data it is difficult to extrapolate from the substantial transportation requirements for mining and ore milling, and the demands of the Spaniards overseeing the mining operations, to a conception of goods that may have circulated during prehispanic times, these data are informative on the variety of items mobilized for the mining effort. Goods included manufacturing items such as cloth, wool, wood and dung as fuel, and building supplies. Subsistence goods included potatoes and ch'uno,meat, maize and chicha, and various fruits and vegetables. Goods that could be classified as cultural / prestige items included herbs, medicines, stimulants including quantities of coca leaf, and hallucinogenics like ayahuasca(1990: 408).

In Mesoamerica, where cargo animals were not available, human bearers carried goods for hundreds of kilometers and canoe transport was used extensively. Drennan (1984: 110) observes that textiles may have represented a significant portion of the goods being transported long distance in the prehispanic period. Similarly, woolen textiles in the Andes were an important trade good for highland pastoralists and probably represented a substantial part of the goods offered for barter between pastoralist caravan drivers and agriculturalists from the beginning of the mutualistic relationship between pastoralists and agriculturalists. This demand for textiles would have been especially strong among agriculturalists living outside of the cotton-producing coastal area. The domestication of camelids for cargo transport sometime in the latter part of the Archaic allowed for more efficient transport of goods throughout the remainder of the prehispanic period. However, this is not to suggest that daily staples were transported, as in modern circumstances where fruits are trucked to the altiplano markets where they are bought for daily consumption.

3.2.5. Caravan travel distances and speeds

Ethnoarchaeological studies provide details on the more immediate decision making practices of caravan drivers including the daily routine, the rate of travel, and the scheduling of rest days. References in the ethnographic and ethnohistoric literature concerning to the velocity and capacity of llama caravans provides benchmarks for estimating the rate of travel in prehispanic times. There is some variability in the reported weights, speeds, and distances in the ethnographic and ethnohistoric literature (Bonavia 1996: 501-515).

Mode

Weight

Distance

Time

Reference

Caravan distance calculated by coca chews

3 km level, 2 km uphill

Approximately 40 min per "cocada", 6-8 per day.

(Zahm 1911: 180)

Llama

Up to 40-45 kg

(Tschopik Jr. 1946: 533)

Llama caravan (ethnohistoric)

75 - 100 lb loads (34.1 - 45.4 kg)

10-12 miles (16-19 km) / day

(Murra 1965: 185)

Llama caravan

25-30 kg

(< 40 kg)

15-20 km / day

Ten hour marches. A long trip can last 30 days.

(Flores Ochoa 1968: 118, 130)

Llama caravan

Approximately
35 kg

25 km / day (150 km journey)

From 8-9am to nearly 4 pm, or 8 hrs per day

(Custred 1974: 277)

Llama caravan (ethnohistoric)

11 miles (17.7 km) / day

1 day: from daybreak to noon

(Murra 1980: 48)

Llama caravan

Approximately
25 kg

20-25 km / day

1 day: 7 am to 2-3 pm.

(Lecoq 1987: 8, 20)

Llama caravan

Approximately
25 kg

15-20 km / day

Journey: 2-3 months

(Flannery, et al. 1989: 106, 115)

Llama caravan (ethnohistoric)

10-20 km / day

1 day: From dawn until early afternoon

(Hyslop 1984: 294-298, 302)

Llama caravan (model for long trips)

30 kg

20 km / day

6 days a week of travel

(D'Altroy 1992: 85)

Llama caravan

15-20 km / day

1 day: From dawn until early afternoon.

(Browman 1990: 403)

Llama: Lighter loads

25-35 kg

300-400 km journey

Journey: 2-3 months

(Browman 1990: 403)

Llama: Heavy loads

50-60 kg

Short distance

Short duration

(Browman 1990: 403)

Llama caravan (salt blocks)

23 kg

15-25 km / day

From 4-5am to 2-4pm, or 6-9 hrs/day, no stops

(Nielsen 2001: 169-176)

Table 3-1. Reported llama caravan loads, distances, and times.

Caravan drivers generally arise at first light and begin preparing for the journey and loading animals for an early departure. Caravans often travel until early afternoon when camp is established and the animals are allowed to graze. As camelids do not pasture at night (whereas Equusdo) ample time must be provided for animals to feed during the afternoon in order to avoid stressing the animals (Nielsen 2000: 446-449). Rest days are taken regularly on caravan routes that exceed six days, with Nielsen (2000: 461) reporting one rest day for every three to five days of travel (Lecoq 1988: 185-186;West 1981: 70). The top priority with respect to nightly campsite selection is the needs of the herd animals. Quality pasture is sought for the animals, the next priority is sufficient water, and additionally the emotional condition of the llamas is considered as reported the llamas can be restless in certain camps. Subsequently the needs of humans are considered including hunting opportunities, trade opportunities, and the comfort of the camp. Thus, while economic and social demands frame the larger scale decisions of caravan routes and products to transport for trade, the needs of the herd animals dominate in short term decision making (Nielsen 2000: 490).

Male llamas are larger and caravan animals are typically castrated males, based on some reports, but "left intact" according to others. Flores Ochoa (1968: 118), reports that castrated llamas produce better meat and wool, but non-castrated llamas make better caravan animals. According to most other reports caravan llamas are castrated because they are stronger and tamer, and this practice allows herders to manage mixed herds (Browman 1990: 398;Nielsen 2001: 168;West 1981: 66).

Estimating travel velocity

A computer model for estimating travel speed based on topography, where velocity over a segment of trail is calculated as function of slope, could be derived using the function presented by Tobler (1993) for hiking and horseback riding (the source of cost paths shown in Figure 3-5). While these topography based calculations have serious limitations (Connolly and Lake 2006: 252-255), for general estimates over larger regions with measurable changes in terrain steepness these estimates are superior to the simple use of slope for estimating velocity. Such models would preferably be derived using original data from fieldwork, perhaps based on a contemporary study that takes into account the size and weight of the cargo animal, the amount and type of cargo, and the performance of the cargo animals based on trail conditions mapped using GPS receivers.

In the course of ethnoarchaeological fieldwork accompanying a llama caravan in Bolivia, Nielsen (2000: 449;2001: 184) notes that there were differences in the going and returning portions of a 2-3 month caravan journey that had implications for the overall travel speed. The out-going trip involved visits with companions and a variety of rituals at propitious locations along the route. During the return journey, in contrast, no rituals were performed but the animals moved more slowly because they were carrying large loads of produce uphill from the eastern lowlands; and they had been traveling for months and reportedly their feet hurt. Thus the lack of ritual performance and visiting and bartering saved time, but the walking speeds were slower, breaks were longer, and layover rest days were more frequent and longer.

Evidence of relative travel distances in the region

In the Mantaro region of the central Andes, Earle (2001: 310) notes that long distance exchange (classified as taking place over a distance of 50 km or greater, to lands beyond those held by the local ethnic group), rarely includes staple items. During the Late Intermediate and Late Horizon, items that were found in the Mantaro area include items of metal and shell described as "wealth items", emphasizing the political nature of long distance exchange in this period.

While Colca valley obsidian was primarily conveyed throughout the Titicaca Basin during prehispanic times by unspecified groups, during the colonial period Colca valley polities are known to have been dominant caravan operators in the region. Residents of the Colca valley owned large camelid herds and they were responsible for extended caravan journeys to the altiplano during the Colonial period. Documentary evidence shows that the Collaguas from the main Colca valley initiated caravans to transport products, such as Arequipa wine to Cusco and Canas, and Colca corn to "wherever they would like to sell it" (Crespo 1977: 56). A testament to the role of the Collaguas, and the Colca valley in Arequipa more generally, comes from Toledo's (1924 [1570-1575]) visita generaldescribing populations of Colca community members resettled as farmers in the new city of Arequipa, Yanahuara, and environs. Toledo states that these Collaguas mitimaesgroups should produce wine in these lower-altitude areas for transport to Cusco and to the mines of Potosí in Bolivia over 700 km to the south-east (Málaga Medina 1977: 114).

(1) Patio exchange

Information exchange concerning the relative barter values for goods are subsumed by the intimacy of the exchange context. In ethnographic accounts of exchange the negations occur predominantly in the privacy of domestic patio areas, and not in a public forum, such as a marketplace, where prices and equivalencies are public knowledge (Humphrey and Hugh-Jones 1992).

(2) Temporality of caravan travel

Caravan travel involves negotiating long distances with unforeseen delays. Due to the difficulty in scheduling encounters between caravans as highly mobile segments of the population, settled villages probably served as important nodes in the fluid regional interaction network (Dillehay and Nuñez 1988). The manner in which these temporal factors intersect with the diachronic nature of reciprocal exchange is counter to the synchronic expectations of establishing immediate, market-based equivalences for goods.

When caravan drivers do not have established trade partners in a settlement they may initiate a new trade relationship with unknown partners upon entering the community. Casaverde Rojas (1977: 177) describes how women in Cabanaconde, in the lower Colca valley, would besiege the first caravans to arrive seasonally at the village entrance with offers. The women would attempt to establish trade partner relationships by negotiating favorable terms for agricultural goods and a place to corral the caravan animals in exchange for labor and pastoral products. Nielsen (2001: 183) states that when market prices fluctuate then contemporary caravanners in Oruro, Bolivia, will sometimes avoid established trade partners in order to better pursue profit opportunities with the greater variety offered by modern market forces; the compadrazgoinstitution seems to be waning. As Danby (2002) and M. E. Smith (2004) both argue, commercialization and alienability of products in all ancient economies should perhaps best be considered in terms of degrees and not in absolutes.

3.2.6. Circuit mobility and role of the periphery

The circuit mobility model of Nuñez and Dillehay (Dillehay and Nuñez 1988;1995 [1979]) conceives of the development of regional interaction in the south-central Andes in terms of decentralized "circuits" traveled by regular camelid caravans. This articulation between far-flung communities is envisioned as beginning in "Late Archaic" (rather, during the "Terminal Archaic" using the terminology in this dissertation) and Formative times.

Herder-caravan societies moved in fixed spiral-like transhumance paths between two or more axis settlements either along a puna-to-puna vector, a puna-to-coastal vector, or a puna-to-selva vector… Continuity and stability was given to the circuit herder-caravan movement by settlements at both ends of its pathway. For this movement to have maintained equilibrium, its pathway must have been balanced by relatively homogeneous, fixed axis settlements which offered multiple resources and services from their particular ecological zone and by ferias (or fairs) where goods were exchanged (Dillehay and Nuñez 1988: 611).

This historical model is of importance here because it highlights the limitations of a core-periphery focus for addressing certain regionally distributed processes like the emergence of control over regional exchange routes during the Formative in the Lake Titicaca Basin (Nielsen 2000: 88-92;Yacobaccio, et al. 2002: 171-172). In models of Late Formative complexity in the Titicaca Basin presented by Browman (1981) and, to a lesser extent, Kolata (1993: 274), caravans and long distance trade play a prominent role. In Browman's formulation, the core areas of regional centers became increasingly powerful due to craft specialization guilds and other institutions that reached their apex at Tiwanaku, albeit many of these expectations for guilds have not been borne out in more recent research (Isbell 2004: 216;Rivera Casanovas 2003). Kolata holds it was the productivity of raised-field agriculture that formed the principal economic mechanism behind Tiwanaku's florescence, with caravan based articulation being a secondary component. Both scholars emphasize the dominant position of the core areas of regional centers, a position that on a regional scale was ultimately attained by the Tiwanaku state. In contrast to these models that emphasize centralization, the Nunez and Dillehay model holds that it is integration by way of caravan trade routes themselves, and that these routes developed into "leading circuits" when they served to connect important centers. Further, as traditions became established, the relationships between circuits and principal settlements along these routes provided temporal continuity and stability to a system that is otherwise fluid and mobile. Their model gives more autonomy and influence to this integrating, caravan-based element in society such that "the sedentary (or axis) settlements of the population were maintained (and often created) and controlled by the mobile sectors." (Dillehay and Nuñez 1988: 621).

This model also contains weaker points, such as an adaptationalist underpinning for the origins and incentives for participation in these caravan exchange networks, and the model ascribes the emerging dominance of Titicaca Basin centers as resulting from their environmental characteristics. The system is described as "harmonious and cohesive" (Dillehay and Nuñez 1988: 620) as caravans are used to efficiently spread patchy resources over a wider region. The adaptationalist interpretation of harmony is based on a lack of archaeological evidence for warfare in the Formative, with the Andean tradition of reciprocity serving as the cohesive force. Further, the system segregates discrete "highland" and "lowland" communities, while archaeological evidence supports a gradient with no clear demarcation.

A more current theoretical orientation for this model would focus on the motivations of traders, the influence of non-local goods in axis communities, and the status acquired by caravanners based on the importance of their role on a regional scale. The emergence of this distinctive, dispersed economic form that appears simulatenously with beginnings of social inequality during Terminal Archaic and Formative may shed light on the underpinnings of social inequalities that developed during the Formative. Further, a greater exploration of the strategic relationship between aggrandizers in Titicaca Basin regional centers and these long distance traders may provide evidence for how the first elites in early polities were able to coordinate labor and garner resources from the region. Despite the historical and adaptationalist focus, the Nuñez and Dillehay model brought a focus on the role decentralized networks in connected emerging settlements that grew to have regional influence. This emphasis on the mobile sector of society highlights the potential influence in both resources and political support that could come from second-tier communities dispersed across the large expanses of the altiplano in the political strategies of early elites in Titicaca Basin centers.

3.2.7. Compadrazgo relationships and commerce

A trade relationship termed compadrazgo,or simply conocidos,is reported between regular trade partners; typically between those who live in complementary ecological areas (Browman 1990: 404-405;Flores Ochoa 1968; but see ;Nielsen 2000: 437-438;Nielsen 2001: 182-183). For example, if a llama caravan driver from a particular area of the puna and a farmer in a mid-altitude valley with a variety of products have regularly exchanged goods over the years, and then they teach their children of the relationship using fictive-kin terms; a tradition of mutualism is established between herder and farmer that can potentially last for generations. The relationship offers stability and predictability to both sides of the exchange in barter rates, types of goods, and quantities to be exchanged.

The nature of this encounter is critical to understanding Andean reciprocity relationships and the degree of alienability of goods being exchanged (Burchard 1974;Mayer 1971). As mentioned above, the caravan driver has mobility and choices in terms of travel routes and communities where to partake in exchange. The maintenance of long term exchange partners through compadrazgois therefore a constraint on caravan autonomy. Two characteristics of compadrazgo relationships appear to underscore the embeddedness of the interaction:

Yapa and over-reciprocation

Barter relationships are often cemented with a yapa:a little bonus given to the buyer that takes the form of an over-reciprocation to assure future transactions (Browman 1990: 421;Sahlins 1972: 303). The magical powers attributed to the yapanotwithstanding, Sahlins' (1972: 308-314) develops a functional economic explanation for over-reciprocation where it serves a similar mechanism to price fluctuations in market-based societies. When an over abundance of product A relative to product B exists in a barter situation based on traditional equivalencies (hence, a lower value for A may result, in a market system), the provider of product A may over-reciprocate and thus, based on the morality of reciprocal arrangements, guarantee future compensation from the trade partner.

The primitive trade partnership is a functional counterpart of the market's price mechanism. A current supply-demand imbalance is resolved by pressure on trade partners rather than exchange rates. Where in the market this equilibrium is effected by a change in price, here the social side of the transaction, the partnership, absorbs the economic pressure. The rate of exchange remains undisturbed - although the temporal rate of certain transactions may be retarded (Sahlins 1972: 311).

The discrepancy that must be resolved synchronically in neoclassical market economics is resolved diachronically in reciprocal arrangements (Danby 2002). Browman (1990: 421) does not believe the over-reciprocation device described by Sahlins is in evidence in the Andes. As Browman observes, there is ethnographic evidence that suggests that barter rates do, in fact, fluctuate in response to supply and demand. The arrangement described above is one possible configuration that occurred in prehistoric circumstances, however, and it is a possible means of assuring the long term persistence of exchange relationships (Burchard 1974;Mayer 1971).

Seasonal fairs and the temporality of caravans

If seasonal fairs and aggregations were a feature of the prehispanic altiplano, as discussed by Browman (1990), interactions may have taken notably different forms in those contexts. Seasonal fairs may have had the significance of religious festivals in the contemporary Andes where the devout sometimes travel for weeks in order to arrive at auspicious times. Fairs and cultural occasions may, then, have been blended with economic transactions.

Scheduled festivals with elaborate dances, music, and costumes are a major cultural contribution in contemporary altiplano communities like Paratía (Flores Ochoa 1968) and despite the lack of simple material correlates for archaeological study, cultural items like song and dance were probably significant features in a variety of prehispanic reciprocal exchange contexts (J. Flores Ochoa 2005, pers. comm. July 2005). Despite the relative marginalization of altiplano cities in the modern economy (or perhaps a reflection of this marginalization), traditional festivals endure as important cultural features in the Titicaca Basin. Citing early twentieth-century sources, Browman (1990: 409) reports that at major shrine at Copacabana, Bolivia, between 40,000 and 50,000 "traders" would converge at times scheduled to coincide with ceremonies at the shrine.

If economic transactions occurred in association with these festivals in prehispanic times, either as a central feature or relegated to the periphery of the cultural events, the transactions may have assumed certain characteristics of marketplace exchange. These characteristics would have included public knowledge of barter equivalences and perhaps more immediate, synchronic exchange due to the short time period of convergence at the festival. As mentioned, however, marketplace concentrations do not necessarily imply true "market economies" with fluctuating prices reflecting supply and demand (LaLone 1982). Assuming that economic transactions that may have occurred at these fairs did not create moral conflict (by debasing sacred ceremonies with lowly economic transactions, in a Euro-American perspective) they would have created an excellent context for the transfer of both cultural goods and prestige items, and for the control of certain exchange practices by administrators or elites. Nevertheless the problem remains that dedicated agriculturalists with harvest goods for exchange would have been absent from these fairs on the altiplano because dedicated agriculturalists would not have the schedule or the herd demographics that would have permitted them to initiate long distance caravans. Therefore a variety of strategies probably developed to allow the transfer of products with the emergence of caravans that traveled, on the large scale, according to schedules dictated by seasonal gatherings, harvest schedules, and other economic and cultural circumstances. These developments imply the emergence of something of a continuum between the more alienable exchange that occurred in seasonal gatherings, and more inalienable barter that occurred in the intimate exchange context of compadrazgorelationships.

3.2.8. Discussion

The environmental and cultural context of the south-central Andean highlands framed the circumstances in which emerged the long distance traffic in various products during the prehispanic period. The domestication of camelids sometime in the past 6000 years allowed for more efficient transport of bulky goods. There remain many unknown aspects to the network that articulated dedicated agriculturalists with pastoralists in the prehispanic past, however archaeological and ethnographic evidence allows for inference regarding the following major points:

(1) Network configuration.Products available by ecozone were transported in numerous, segmentary articulations between ecozones while other products, only available in a few locations, were apparently conveyed diffusively and were between transported ecozones and across homogeneous terrain like the altiplano (XFigure 3-4X).

(2) Motivations for change in mode of interaction.The domestication of animals and plants, changes in sedentism, and the development of social inequalities were some of the factors that contributed to development of long distance caravans. It is evident that the original modes of interaction: Direct acquisition and down-the-line exchange, were supplemented by household organized caravan trade, and finally administered caravan trade, but the timing of these changes is difficult to establish with precision.

(3) Means of trade.Features of the Andean barter economy such as enduring trade relationships between households in complementary zones cemented by institutions like over-reciprocation and fictive kinship, are well demonstrated ethnographically in the region. However, seasonal market-like gatherings are also reported in the region and exchange of goods in those contexts may have been more alienable, and transactions may have been more synchronic in nature. Such gatherings may also have had evolutionary significance because they potentially relate to the development of early leadership in regional centers including ceremony, feasting, the use of monumental architecture, and centralized control of trade in certain goods.

The persistent themes in mountain agropastoral settings of non-autarkic economies and risk reduction through mixed subsistence strategies serve as a reminder that variability was probably the rule in exchange relations as well. Absolutes in exchange patterns were probably rare, and a degree of both self-interested trade and elements of embedded, social and symbolic fraternity likely existed between trade partners since early antiquity. While a number of plausible models have been proposed for the later Prehispanic periods where both household-level and elites-administered trade caravans appear to have transported a variety of goods in the region, the initial contexts for caravan trade remain largely unexplained. These initial contexts are particularly important because this mode of organization contributed to the regional context and institutional base in which early leaders in the Titicaca Basin had to operate in order to begin the process of expanding their influence in access to labor, resources, and ideology of their communities.

If enduring regional interaction had persisted since the early days of the pastoral economy during the preceramic, this may be connected to the factors that lead to an increasing consolidation of power in the Titicaca Basin during the Formative Period. These questions are central to understanding the foundations of regional integration that emerged in the Lake Titicaca Basin during the Middle and Late Formative Periods. The above discussion has sought to elaborate upon a possible context for early household-level caravan organization that is principally based on the "circuit mobility" model of Nunez and Dillehay (Dillehay and Nuñez 1988;1995 [1979]) but without following the adaptationalist approach, and with more specific empirical contexts for early caravans. In the ensuing discussion of obsidian procurement and circulation in the south-central Andes existing evidence from obsidian circulation in the region is presented that provides the context for examining obsidian production in more detail.

3.3. Regional patterns and major sources of obsidian

Sources of tool-quality obsidian occur in discrete locations along the Western Cordillera in the south-central Andes. The spatial pattern of obsidian sources in southern Peru, spaced at an interval of approximately 100-300 km along an arc trending north-west to south-east, is evident from maps of these source distributions. It is volcanic processes at a continental scale that have resulted in emplacement of obsidian at 3000 and 5000 meters above sea level. As will be discussed in more detail in Chapter 4, obsidian sources appear to be regularly spaced paralleling the spine of the Western Cordillera, however geologists believe that "the complex tectonic and lithospheric variables at continental subduction zones preclude the development of any regular pattern of volcano spacing" (Clapperton 1993: 77-80). Nevertheless, volcanoes do occur at major crustal fractures and at fracture intersections.

/Figs_Ch3/andes_3obs_sources.jpg

Figure 3-5. Known south-central Andean obsidian sources used in prehistory with least cost paths (Tobler's function on SRTM DEM data) from the three major sources to sites in the region.

In the prehispanic central and south-central Andes obsidian from three sources dominated the archaeological use of this material, followed by obsidian from a number of minor, local sources that were only used for tool production in the area of geological origin (Figure 3-5). Finally, there were countless sources of geological obsidian that were used negligibly, or not at all, by humans apparently due to the undesirable attributes of the glass. Outside of the study region, additional major Andean obsidian sources have been studied to the south in northern Argentina and surrounding terrain (Yacobaccio, et al. 2004;Yacobaccio, et al. 2002) and to the north in Ecuador (Burger 2003;Burger, et al. 1994).

The three most widely-circulated obsidian sources in Peru and north-west Bolivia are the Quispisisa, Alca, and Chivay types, and these three types will be investigated here in more detail. Burger et al. (2000: 348) argue that it is the large nodule size and the homogeneity of the glass that is the primary determinant of how widely obsidian was circulated in prehistory and by these criteria these three sources are exceptional in the Andes.

In the course of this research a relational GIS database has been assembled using published data on obsidian sourcing in the south-central Andes (Burger, et al. 2000;Craig 2005: 908-916;Frye, et al. 1998;Giesso 2000) and using these data a few summaries have been generated for the purposes of this study. The trends in these data show broad patterns in the use of obsidian by chemical type, but given the early stages of this kind of research, these values also reflect a sampling bias based in the history of research and the collections available for sampling. With further development of sourcing technologies, including the more widespread use of portable XRF units, obsidian from a wider variety of sites will be characterized in coming years.

Type

Samples

Sites

Ave. Distance
(Km) *

Ave. Time
(Hrs) **

Max. Distance
(Km) *

Max. Time

(Hrs )**

Chivay

531

48

171.0

40.0

553.6

120.8

Alca

305

50

249.6

66.0

988.7

260.7

Quispisisa

525

63

260.3

66.7

846.7

208.5

Table 3-2. Three major Peruvian obsidian sources showing average and maximum distances and times. * based on straight-line distance measure, ** Least-cost path walking time based on Tobler's(1993) hiking function.

/misc/image007.jpg

Figure 3-6. Cumulative frequency graph showing three major Peruvian obsidian sources by consumption site altitude.

As will be discussed below, the dominant pattern that has emerged for both the Chivay and the Alca obsidian types is one of highland interaction. Throughout the prehispanic period, artifacts of Chivay obsidian have never been found below an altitude of 1280 masl, and that was in the context of a Tiwanaku colony, at the site of Omo in Moquegua (Burger, et al. 2000: 338;Goldstein 1989). Alca type obsidian artifacts are similarly found mostly at high altitude; the coastal Quebrada Jaguay paleoindian finds are anomalous for Alca distributions. The next lowest altitude context for Alca is also at Middle Horizon Omo. Quispisisa, on the other hand, is commonly found at coastal sites from the Nasca area and northward, beginning with a strong representation at the cotton preceramic site of San Nicolas (Burger and Asaro 1978;Vaughn 2005).

One possible explanation for the lack of Chivay and Alca obsidian in low elevation contexts is that camelid caravans were less common in the low elevation coastal areas of southern Peru and northern Chile. While there is strong evidence of camelid corrals in northern coastal Peru (Shimada 1982), the examples of pastoral sites in the coastal areas of the south-central Andes are rare. Stanish (1992: 57) states that there no sites with corrals below 1,500 masl in the Moquegua drainage, and he presents other evidence for the scarcity of pastoral sites in southern coastal Peru. However, Wheeler et al. (1995) examine the remains of 26 camelids, both llamas and alpacas, at the Chiribaya site of El Yaral at 1000 masl in Moquegua. Moving northward in the Pacific coastal drainages, both Christina Conlee (2000) and Kevin Vaughn (2005) report evidence of domesticated camelids at low altitude in the Nasca area. Vaughn (2005: 92) describes a corral-like structure at the site of Marcaya at 1000 masl. As mentioned, the Nasca zone is rich in Quispisisa obsidian. A variety of factors may have contributed to the pattern of Quispisisa obsidian consumption on the coast and Alca and Chivay consumption in the highlands including the greater breadth of the altiplano in the Titicaca Area, the position of geological source relative to the coast, and the severe aridity in the Atacama area of southern Peru and northern Chile.

3.3.1. Obsidian and larger geographical associations

The specific cultural and temporal associations of these distributions are explored in a series of publications by Richard Burger, particularly in Burger et al. (2000). There appear to have been two overarching geographical associations that cut across the various distributions of obsidian in space and time.

Availability, demand, and mobility

Chivay was largely conveyed into the Lake Titicaca area by Titicaca Basin communities and polities that were organized around the lake and the adjacent broad altiplano. For sites in the Lake Titicaca Basin, the database shows that 453 obsidian artifacts have been sampled from all time periods, and 408 (90%) of the samples were of the Chivay obsidian type. Another interesting aspect to this pattern is that the proportion of Chivay obsidian holds true for the north Titicaca Basin as well as the south when the two areas are considered separately. Using Bandy's (2005: 92) division of the Titicaca Basin where the site of Camata and northward are considered the North Basin, and with the South Basin extending as far south as Khonkho Wankane and south-west to Qillqatani, all obsidian samples in the region were aggregated from the entire prehispanic period. The LIP site of Cutimbo, though slightly south of Camata, was joined with the North Basin sites while Incatunahuiri and the Ilave valley sites were counted with the South Basin.

Samples

Sites

Chivay Type

Alca Type

Other Types

NorthTiticaca Basin

284

16

261

91.9%

20

7.0%

3

1.1%

SouthTiticaca Basin

231

13

205

88.7%

5

2.2%

21

9.1%

Total

515

29

466

90.5%

25

4.9%

24

4.7%

Table 3-3. Obsidian in the north and south Titicaca Basin by counts and percents.

These patterns reveal that Chivay obsidian accounts for approximately 90% of obsidian from the Titicaca Basin, with a greater presence of Alca obsidian in the North Basin. It seems that the presence of obsidian from alternative sources, whether it was the Tumuku source (still unlocated) or other obsidian sources further to the south, counter balances the access to Alca type in the North Basin. This pattern, with heavy use of Chivay obsidian in the South Basin, reflects overall integration in the Titicaca Basin through prehistory as well as the lack of high quality obsidian sources with large nodules south of the Chivay source, until one arrives at the Zapaleri source close to the frontier of Bolivia, Argentina and Chile, located 700 km to the south of Lake Titicaca (Yacobaccio, et al. 2004: Fig. 1). It has been noted that, in cultural terms, the Chivay distributions are especially linked to areas integrated by the Tiwanaku state, the Pukara polity, and with their predecessors: the Middle Formative communities linked by Yaya-Mama stylistic features (Burger, et al. 2000). However, what is particularly compelling about the evidence from obsidian is that these social and economic links in obsidian exchange appear to predate the cultural-stylistic links and become defined in the Terminal Archaic roughly 1000 years earlier than regional evidence of the Yaya-Mama stylistic attributes that first appear at Chiripa in the southern Lake Titicaca Basin.

The distribution of Alca material is confined by the availability of Quispisisa on one side and Chivay on the other, but nevertheless Alca obsidian was transported the furthest with material during the Middle Horizon being conveyed as far as the Wari sites of Huamachuco and Marca Huamachuco near Cajamarca (Burger, et al. 2000: 336). Alca material was also transported the furthest distance at an early date, as it was conveyed 769 km to Chavín de Huantar during the Early Horizon where it was found in both elite and commoner contexts (Burger, et al. 2000: 313-314). Quispisisa predominates in Wari assemblages but it also circulated widely before the Middle Horizon, such as the 589 km to Chavín de Huantar during the Early Horizon.

Terrain adjacent to source

The altitudes of obsidian consumption sites (XFigure 3-6X) also reflect the terrain adjacent to each of the sources. The Quispisisa source is in an area with deep river valleys and a number of the sites sampled are from lower elevation or coastal contexts. The Alca source is similarly close to deep river valleys, but it also borders the altiplano and the much of the Alca samples came from the highlands of the Department of Cusco. The Chivay source is on the periphery of the broad altiplano and with no major altitude loss it was possible to transport material to a variety of sites that include Pikicallepata to the northeast and Qillqatani to the southeast. Further detail on these obsidian distributions are described below, and are published elsewhere (Burger, et al. 2000; cited in Craig 2005: 908-916;Frye, et al. 1998;Giesso 2000).

3.4. Chivay Obsidian Consumption Contexts

Archaeological research conducted at four sites in the south-central Andean highlands, the consumption zone for Chivay obsidian, provide evidence for the changing use of Chivay obsidian over long time periods. This section will begin by focusing on the four sites because they important here as the sites have provided stratified evidence of obsidian consumption during the Archaic and Formative in well-dated archaeological contexts. This review of the sites will be relevant in subsequent discussions of change in obsidian use over time in the region, and for interpreting activities at the Chivay source.

3.4.1. Asana

The site of Asana is at 3435 masl in the Osmore drainage of Moquegua, and it lies 185 km or 50 hours by the hiking model to the south-east of the Chivay source. Excavations at Asana were directed by Aldenderfer(1998: 76-80)between 1986 and 1991 and in the lower levels of the deeply-stratified site obsidian was found in low concentrations.

Titicaca Chronology (BCE)

14C yr bp

Lab Code (Beta-)

Cal.BCE

14CSample Provenience

Obsidian Provenience

Obsidian Type

Obsidian

Artifacts

Early Form.

2000-1300

3640±80

23364

2300-1750

II (sup), TU1

II

Aconcahua (visual ID)

n=3, 0.4%

Late Archaic

5000-3300

6040±90

24634

5210-4720

XIVb, I25b

XIV

n=1, 2%

Middle Archaic

7000-5000

6550±110

24629

5680-5300

XVIIa, G24b

KF4a: J29D-8; XVIIA

Chivay

6550±110

24629

5680-5300

XVIIa, G24b

KF4b: J29D-8; XVIIA

Chivay

Early Archaic

9000-7000

8620±110

47057

8200-7450

PXIX, P38b

8780±90

43920

8250-7600

PXX, V51a

KF5: F28D-5; PXXi

Chivay

8720±110

33303

8250-7550

PXXI, R43c

KF6: P38D-W; PXXI

Chivay

8720±110

33303

8250-7550

PXXI, R43c

KF2: W41A-4; PXXII

Chivay

8720±120

43922

8250-7550

PXXIII, X36c

8720±110

35599

8250-7550

PXXIV, T34a

n=11, 0.36%

KF3: T42B-7;F28D-5

Chivay

n=1, 0.08%

KF1: U41D-6; PXXV

Chivay

9820±150

40063

9900-8700

PXXXIII, S38c

Table 3-4. Asana obsidian samples, collections, and associated14C samples by level (Aldenderfer 1998: 131, 157, 163, 209, 268;Frye, et al. 1998).

These data from Asana show that low frequencies of Chivay obsidian occurred regularly through Early and Middle Archaic levels at the site despite the presence of high quality cherts nearby, and low quality obsidian from the Aconcahua source located about 50 km away. During the Late Archaic, obsidian disappears from the site and in a level that is transitional between the Terminal Archaic and Early Formative, obsidian returns in the form of lower-quality Aconcahua material. These data will be discussed in more detail below as obsidian use is explored by time period.

3.4.2. Qillqatani rock shelter

The rock shelter of Qillqatani is at 4400 masl along the Río Chila to the southwest of Mazo Cruz, and it is 221 km or 50.3 hours walking time, by the hiking model, across the puna from the Chivay obsidian source. Qillqatani lies on the southwestern perimeter of the Lake Titicaca Basin in the headwaters of the ríoHuenque drainage in the Department of Puno, and it is not far from the headwaters of the Osmore drainage on the western slope in the Department of Moquegua. The shelter has a strong pastoral component, but excavations revealed occupations dating back to the Middle Archaic. The shelter is also known for elaborate rock art panels. Qillqatani was excavated in 1990 by Aldenderfer in two adjacent blocks, labeled East Block and West Block (1999;2005;in prep.). The East Block measured 20 m2in horizontal extent, while the West Block measured 60 m2.


(a) Titicaca chronology,

calibrated dates

(b) Qillqatani periods,

calibrated dates

(c) Uncalibrated14C dates in r.c.y.b.p.

(d) Qillqatani provenience
of obsidian

(e) Obsidian sample

(f) Block

LH AD1476 - 1532

LH 1476 - 1532

WIII / EIII

Level D11d-507/WIII:

400±50bp(B118156)

     

LIP AD1100 - 1476

LIP 1100 - 1476

WIV - WV / EIV

 

H12C-3; WV; LIP

Chivay

West

Tiwanaku Horizon

AD400 - 1100

Tiwanaku

AD 500 - 1100

WVI - WVII / EVI

Level D9a/F7/WVIII:

2210±60bp(B93354)

E9A-5; WVII; F-3

Aconc.

West

E6A-4; WVII

Chivay

West

Late Formative

400 BCE - AD500

WVIII - WIX /
EVII - EVIII

 

D6C-8; WX; F-5; 1.45

Chivay

West

Late Formative

500 BCE - AD 400

D9A-8; WX; F-5; 1.45

Chivay

West

G13A-10; WX

Chivay

West

G13A-10; WX

Chivay

West

D6C-9; WXI; F-6; 1.6

Chivay

West

H12D-10; WXI; F-12; .75

Chivay

West

H12D-10; WXI; F-12; .75

Chivay

West

I12D-8; WXI; F-13; 2.00

Tumuku

West

Formative C

900 - 400 BCE

WXII - WXIV /
EIX - EXIV

East Block dates

Level D25d/F10/EVI:

2210±50bp(B93355)

Level F23a/F4/EXI:

2620±60bp(B93357)


West Block date

Level G7c-513/WXII:

2550±80bp(B43924)

E25B-10; EIX; F-7

Chivay

East

Middle Formative

1300 - 500 BCE

D25D-6; EIX; F-7

Chivay

East

D26B-10; EXI; F-12

Alca

East

F23B-11; EXI; F-3

Chivay

East

F24B-9; EXI; F-13

Aconc.

East

D6D-11; WXIII; F-8; .75

Chivay

West

F11A-3; WXIII; 1.65

Chivay

West

D26D-13; EXIV; F-10

Tumuku

East

D26C-13; EXIV

Chivay

East

D26C-13; EXIV

Tumuku

East

Formative B

1500 - 900 BCE

WXV - WXIX

 

D8C-12; WXVIII; F-11

Chivay

West

D8C-12; WXVIII; F-11

Chivay

West

G7D-11; WXVIII; 5.5

Chivay

West

Early Formative

2000 - 1300 BCE

Level D6/F7/WXVI:

3000±70bp(B43929)
Level D6a-522/WXIX:2940±70bp(B43925)

E9D-12; WXVIII; F-9

Chivay

West

Formative A

2000 - 1500 BCE

WXX - WXXIII

E6C-14; WXX; F-7; 1.4

Chivay

West

 

E9A-15; WXX; 2.4

Chivay

West

H13A-15; WXX; 4.6

Chivay

West

Terminal Archaic

3300 - 2000 BCE

Terminal Archaic

3300 - 2000 BCE

WXXIV

Level F9a-532/WXXIV:3660±60bp(B43926)

F28B-19; WXXIV; HF-1

Chivay

West

F28B-19; WXXIV; HF-1

Chivay

West

F28B-19; WXXIV; HF-1

Chivay

West

G9E-19; WXXIV; 98.70

Chivay

West

G9E-19; WXXIV; 98.20

Chivay

West

G10B-20; WXXIV; HF-2

Chivay

West

Late Archaic

5000 - 3300 BCE

Late Archaic

5000-3300 B.C.E

WXXV-WXXIX

Level E9d-541/WXXX:
5620±120bp(B43927)

E9D-18; WXXX

Chivay

West

Middle Archaic

7000 - 5000 BCE

Middle Archaic

7000-5000 B.C.E
WXXX - WXXXVI

Below level WXXX:7100±130bp(B18926)

7250±170bp(B18927)

FAC-25; WXXXIV

Source unknown

West

Table 3-5. Qillqatani excavation levels, radiocarbon dates, and obsidian samples. (a) Titicaca chronology, (b-c) levels and14C dates from Qillqatani, all dates are on charcoal and were analyzed by Beta Analytic, from Aldenderfer (1999). (d-f) Obsidian from Qillqatani chemically provenienced at MURR from data presented by Frye, Aldenderfer, and Glascock (1998), with "Tumuku" type replacing reference to "possible Alca-2", and Aconcahua type abbreviated to "Aconc."


/misc/image008.gif

Figure 3-7. Qillqatani data showing percentage of bifacially flaked tools and percentages of debris made from obsidian per assemblage by count (Aldenderfer 1999;Aldenderfer in prep.).


 

TOOLS

DEBRIS

 

Obsidian

Non-Obsidian

 

Obsidian

Non-Obsidian

 

Qillqatani Period

No.

Ave Wt.

No.

Ave Wt.

Total

No.

Ave Wt.

No.

Ave Wt.

Total

Late Horizon

AD1476-1532

   

13

0.84

13

10

0.68

128

1.57

138

LIP

AD1100-1476

2

0.75

11

3.35

13

22

0.46

308

1.81

330

Tiwanaku

AD500-1100

3

0.80

37

1.49

40

50

0.54

565

2.23

615

Late Formative

400BCE-AD500

9

0.84

102

2.19

111

91

0.47

859

2.19

950

Formative C

900-400 BCE

19

1.21

272

2.44

291

144

0.46

1157

1.39

1301

Formative B

1500-900 BCE

   

2

 

2

85

0.58

525

1.73

610

Formative A

2000-1500 BCE

6

0.93

32

1.49

38

160

0.53

1057

1.37

1217

Term. Archaic

3300-2000 BCE

12

0.99

58

3.00

70

79

0.79

403

1.79

482

Late Archaic

5000-3300 BCE

5

1.57

36

3.57

41

18

0.56

147

1.78

165

Middle Archaic

7000-5000 BCE

2

0.20

33

7.26

35

37

0.60

674

2.76

711

Total

58

1.05

596

2.87

654

696

0.55

5823

1.84

6519

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Table 3-6. Qillqatani periods by tools (bifacially flaked) and debitage (all other lithics), in obsidian and non-obsidian categories.

 


Excavations in the two blocks at Qillqatani revealed a switch from hunting and gathering to a predominantly pastoral economy between the occupation in level 25, and that of level 24 that was stratigraphically superior (Aldenderfer 2005: 20;). The occupation in Level 24 appears to have been the first residential occupation, and it is considerably thicker than were previous levels that appeared to have been logistical in nature. The ovoid structures of Level 24 are slightly larger and are free standing, while in the preceding levels the structures are against the rear wall of the rock shelter. Level 24 is the first level with ceramics, but it is being considered Terminal Archaic in this analysis due to economic evidence and the exceptionally early ceramics at Qillqatani.

Qillqatani Period

Weight in Grams

Total Count

0 - 1 g

1 - 2 g

2 - 3 g

3 - 4 g

4 - 5 g

Late Horizon

9

 

1

   

10

LIP

22

       

22

Tiwanaku

44

5

1

   

50

Late Formative

79

8

4

   

91

Formative C

129

11

2

2

 

144

Formative B

71

8

4

1

1

85

Formative A

135

17

5

1

2

160

Terminal Archaic

58

12

5

4

 

79

Late Archaic

13

4

1

   

18

Middle Archaic

29

4

2

2

 

37

Total Count

589

69

25

10

3

696

Table 3-7. Counts of obsidian debitage at Qillqatani by weight (g).


The earliest obsidian at Qillqatani is a Middle Archaic sample from an unknown source. In the Late Archaic, Chivay obsidian begins to appear at the rock shelter. Subsequently, during the Terminal Archaic obsidian from Chivay occurs in relatively large numbers and it persists until the end of the Early Formative (the Qillqatani Formative Cperiod) where chemical testing of ten obsidian samples reveals there is greater heterogeneity in obsidian procurement. In this period, Chivay obsidian makes up 60% of the obsidian tested by the Missouri University Research Reactor (MURR), the Tumuku type makes up 20%, and the Aconcahua type and Alca type make up the remaining 10% (one sample of each).

Three obsidian artifacts from two proveniences at Qillqatani, West level XI and East level XIV, were found to belong to the Tumuku chemical group, a group that derives from an as-yet-unlocated source that probably lies in the southern department of Puno. This chemical group has not yet been well characterized, as no source samples are available and the samples analyzed by NAA in the 1970s at Lawrence Berkeley Labs (Browman 1998: 309-311; Burger and Asaro 1977) are not easily comparable with more recent NAA results from MURR. As a consequence, in initial studies of these three Qillqatani artifacts it was proposed that the artifacts belong to a subgroup from the Alca obsidian source in northern Arequipa, a subgroup that was referred to as "Alca-2" (Frye, et al. 1998) and "Alca-Z" (Jennings and Glascock 2002: 111). More recently, Michael Glascock at MURR has stated that these three Qillqatani samples probably derive from the Tumuku source (M. Glascock, March 2006, pers. comm.).

Flake sizes at Qillqatani indicate that relatively large flakes were being discarded (Table 3-7), particularly during the Terminal Archaic and Early to Middle Formative. A central question concerning obsidian circulation is to what degree were flakes used for shearing and butchering tasks? Were flakes sufficiently large to be hand-held for these activities? Length measurements from these flakes from Qillqatani are not available, but an estimate of the size of those flakes based on their weight is provided by comparing similar obsidian flakes from the Chivay source dataset. An obsidian flake weighing 4g from the Chivay dataset had the following dimensions: 45 x 13 x 7mm. A smaller obsidian flake weighing 3g measured 31 x 14 x 6mm. These flakes appear to have been sufficiently large for shearing and butchering tasks. Further technical studies, as well as use wear and residue analysis, may shed more light on the use of large obsidian flakes in the regional consumption zone.

Small quantities of material and regional models of exchange

The actual volume or weight of obsidian at Qillqatani is relatively small. Obsidian from the Chivay source appears in small quantities in nearly every stratum in the approximately 8000 year sequence. The most informative data on long distance exchange come not from the total mass of obsidian, but from the variability in the relative percentage of obsidian flaked and obsidian tools in each lithic assemblage through the sequence. Nevertheless, the weight of obsidian tools at Qillqatani totals 43 g, and the sum of the weight of obsidian flakes is 382 g out of a total of 11,085 g of lithic material. Thus, the entire quantity of obsidian excavated at Qillqatani is only about 1.5% of the cargo that could be carried by a single llama.

These figures are meant to highlight a central issue with obsidian studies overall: the actual quantities of obsidian encountered and evaluated from the consumption sites throughout the south-central Andes are relatively low. The significance of obsidian circulation over the larger region is not matter of weight or volume, but rather a question of consistency and changes in the proportions of particular sources utilized over time. These data from Qillqatani underscore this issue. A comparison of the persistence of non-local obsidian at Qillqatani as of level 24 (X

Table 3-6XandXTable 3-7X) with the non-local obsidian at Asana (XTable 3-4X) during earlier periods of the Archaic, suggests that on a regional scale mechanisms of exchange, or direct acquisition, were more intermittent during the earlier Asana phases. In contrast, Qillqatani has high quality cherts and low quality Aconcahua type obsidian available in the immediate proximity of the rock shelter. Thus, the question becomes: why transport Chivay obsidian 200 km when alternative materials are locally available? Further, why was the transmission of Chivay obsidian so consistent through time? In this perspective the low but consistent quantities of Chivay obsidian that were conveyed throughout the south-central Andean highlands during the prehispanic period can be seen as a gauge of highland interaction and horizontal complementarity, with wider implications for exchange.

3.4.3. Sumbay

The work of Máximo Neira Avendaño (1990;1968) is distinguished as being the earliest systematic Archaic Period excavations in the Arequipa highlands. In the mid-1960s Neira excavated at seven preceramic sites close to the Sumbay train station, including several rock shelters, known as Sumbay-1 through Sumbay-7. The sites were later renamed "Ccollpa-Sumbay" by Eloy Lináres Malaga in 1984 (1990;1992) as part of the CONCYTEC survey.

The Sumbay area lies 42 km to the south-east of the Chivay source across the pampa and it is 9.3 hours from the obsidian source area by the hiking model. The main cavern, SU-3, is a rock shelter that is concealed in narrow canyon that cuts through the puna. This rock shelter measuring 15m across, 11m deep, and 6m in height at the dripline, was badly looted in the 1930s, and Neira's excavations focused on the remaining intact portions of the cave. The cave is known for exceptional rock art panels featuring both petroglyphs, ochre pictographs, and a variety of animal species including elongated camelids, a hunting puma, humanoid figures, and surithe Andean ostrich.

Seven excavation pits were placed in the main shelter, SU-3. Pit 5 produced obsidian samples in every level along with a large number of tools made from pitchstone, a dull vitreous material formed from weathered obsidian that has absorbed water from the environment. Pitchstone has 4-10% water while obsidian has 0.1 - 0.5% water (M. Glascock, 2006 pers. comm.). The source of this pitchstone is probably somewhere near Sumbay and further study of this pitchstone may provide an interesting complement to the chemical characterization of obsidian.

/misc/image009.jpg

Figure 3-8. Sumbay pitchstone projectile points.

Titicaca Chronology

Sumbay

Stratum

14C yr bp

Lab # (Bonn)

Cal.BCE

Obsidian Type

Artifacts

1

None

Many very small flakes.

Late Archaic

5000-3300 BCE

2

5350±90

1559

4350-3980

One pitchstone knife, 5 broken pitchstone points, 1 quartzite scraper, 100 flakes.

3

6160±120

1558

5400-4750

Two obsidian samples (Chivay)

One pitchstone foliate point with concave base, 4 pitchstone scrapers, 6 broken pitchstone points, 1 broken obsidian point, 62 flakes.

Middle Archaic

7000-5000 BCE

4

None

One obsidian sample (Chivay)

Six pitchstone points: 2 incomplete stemmed with concave base, 1 almost complete, 3 foliate points, one with broken tip and broken base. Seven incomplete pitchstone points, 2 incomplete quartzite points. One pitchstone knife, 110 flakes, 1 worked bone, 5 bone concentration.

Table 3-8. Sumbay, SU-3 Pit 5. Strata with obsidian samples and associated artifacts(Neira Avendaño 1990: 32-33).

Twenty-five obsidian samples were analyzed from the Sumbay area. Twenty samples from the surface and terreplain of the rock shelter SU-2, two from the surface of the rock shelter SU-3, and 3 from excavations in SU-3, Pit 5 (Burger, et al. 1998: 209;Burger, et al. 2000: 278). All samples turned out to be of the Chivay type. Since stratum 3 was dated to the Middle - Late Archaic transition, the stratum 4 obsidian sample is stratigraphically below than that, so it is probably Middle Archaic in date.

3.4.4. The Ilave Valley and Jiskairumoko

Located in the Department of Puno in the western Lake Titicaca Basin, the Ilave river valley is an open, terraced river valley and it forms the highest volume river drainage to flow into Lake Titicaca from the west. Jiskairumoko (95-189) is the largest multicomponent Archaic/Formative site identified during survey and it lies 200 km from the Chivay obsidian source or 44 hours travel time across the expansive puna grasslands. Mark Aldenderfer conducted research in the Ilave valley from 1994 to 2002 including a pedestrian survey, a testing program, and excavation work (Aldenderfer 1997;Aldenderfer 1998). In 1997 and 1998 Cindy Klink conducted research in the Huenque drainage, a principal tributary that joins the Ilave from the south (Klink 2005;Klink 2006;Klink and Aldenderfer 1996). Nathan Craig excavated at the site of Jiskairumoko and tested at numerous other sites between 1999 and 2002 (Craig 1999;Craig 2005;Craig and Aldenderfer 2002;Craig and Aldenderfer In Press). Tripcevich conducted a viewshed analysis of sites in the Ilave drainage using data from the 1994 and 1995 Ilave valley survey (Tripcevich 2002).

IlaveValley

In the Ilave valley, obsidian was scarce prior to 3300 BCE, but it appears in a number of Terminal Archaic and Formative contexts after that date. Diagnostic projectile points from the Ilave area reveal a dramatic change in material type with the Terminal Archaic.

/misc/image010.jpg

Figure 3-9. Comparison of projectile point counts in the Ilave Valley and the Upper Colca.

Using only diagnostic projectile points from the Klink and Aldenderfer (2005) point typology, these data reveal a shift to greater use of obsidian with series 5 projectile points. Following the point typology, all series 1 through 4 points are diagnostic to the Terminal Archaic or earlier, except for types 4C and 4E which have been excluded from this analysis. The lower part ofXFigure 3-9Xreveals data classified into the same three groups, but with data from the Chivay obsidian source area resulting from the 2003 Upper Colca project. Comparing these two datasets reveals that with the advent of Series 5 points in the Terminal Archaic after 3300 BCE there is a sudden upswing in obsidian use in the Ilave consumption zone that corresponds perfectly with a dramatic shift in projectile point production at the Chivay source area itself. There appear to have been pan-regional changes occurring with the widespread adoption of series 5 style projectile points, the use of obsidian, and these changes are possibly related to the greater use of the bow and arrow.

Jiskairumoko (95-189)

The site of Jiskairumoko was excavated using a broad, horizontal décapage technique between the years 1999 and 2002, and Craig's (2005) dissertation forms the site archive. Radiocarbon dates fromexcavation work at Jiskairumoko range from the Late-Terminal Archaic transition with a date of 4562±73bp(AA58476; 3520-3020 BCE)through to the end of the Early Formative, with a date of3240±70bp (Beta-97321; 1690-1390 BCE). Note that Craig (2005) has the Terminal Archaic dating to 3000-1500 BCE, while in the Upper Colca Project and in this dissertation the period dates to 3300 - 2000 BCE All references to the "Terminal Archaic" have been adjusted to the latter time range in this document unless otherwise noted.

Craig and Aldenderfer chemically characterized 68 bifacially-flaked obsidian tools from excavated contexts at Jiskairumoko (96% of all obsidian tools from Jiskairumoko) by sending these items to M. Steven Shackley at the UC Berkeley Archaeological XRF Lab(Craig 2005: 513, 908-916). The samples were compared with four geological samples that provided to Shackley from nodules that I had collected from Chivay source in 2003. Three of the nodules provided to Shackley were from the Maymeja area and one was from east of Cerro Hornillo. The XRF results showed that 97% of the Jiskairumoko obsidian tools (n = 66) were from the Chivay source and 3% (n = 2) were from the Alca source. These values diverge slightly from obsidian type distributions in the Titicaca Basin more generally (XTable 3-3X), where 90.5% (n = 466) of all of the obsidian artifacts tested to date from the Titicaca Basin are Chivay type. These data indicate that the people at Jiskairumoko were more intensively using Chivay obsidian in the Terminal Archaic and Early Formative than were Titicaca Basin residents over all, especially in contrast to the variety in obsidian types that emerge from samples dating to later times.

Obsidian artifacts from excavations at Jiskairumoko reveal that obsidian was used in similar proportions for projectile points from excavated contexts as for points from surface contexts in the Ilave valley. Craig(2005: 685)observes that 18% (n = 54) of the excavated Jiskairumoko projectile points are made from obsidian. He also notes that 50% of these have edge modification in the form of serration or denticulation while only 15% of non-obsidian points have edge treatment, a statistically very significant pattern that he attributes to the symbolic importance of this material. Obsidian was found, along with other lithics and a carved camelid effigy, in Burial 1 at Jiskairumoko in the grave of an elderly adult female with cranial deformation, a primary deposit dating to the Terminal Archaic or 3019-2859 cal BCE(Craig 2005: 680-682). Graves excavated in the Ilave valley area also contain a variety of non-local grave goods including obsidian, lapis, and gold, but non-local grave goods only occur in contexts dating to the Terminal Archaic and later.

3.5. Andean Obsidian Distributions through Time

While obsidian spatial distributions are well demonstrated, it is more difficult to ascertain the temporalpatternsbecause many of the obsidian artifacts from the south-central Andes that have been sourced are from an uncertain temporal provenience. The Andean obsidian sourcing literature, organized and published primarily by Richard Burger, has provided a temporal context for obsidian samples wherever possible. Temporal control for obsidian artifacts fall into three categories

(1) excavated contexts with a temporal association.

(2) surface materials spatially associated with temporally diagnostic artifacts.

(3) direct chemical sourcing of temporally-sensitive diagnostic projectile points.

Good temporal control from excavated contexts is preferred, but a number of the samples are from the number 2group, above, because the samples have weak temporal control and are from sites that are assumed to be single component based solely on diagnostic artifacts. Finally, a fourth method, obsidian hydration dating, has been used in the Andes but with mixed results. The high diurnal and seasonal temperature variation in the Andean highlands suggests that hydration dating can only be reliably used as a relative dating method in conjunction with14C dates (Ridings 1996).

Chronologies based on stages on the one hand, and periods or horizons on the other, are used in Andean archaeology. Some pan-Andean research, such as studies of regional obsidian exchange, have used the horizons absolute chronology developed by John H. Rowe (1967) in collaboration with Dorothy Menzel and based on the Ica ceramic "master sequence". This approach allows for a maximum of compatibility between regional datasets and it has been used widely, and even extended into the Titicaca Basin (Burger, et al. 2000). However, archaeological research conducted in the Lake Titicaca Basin has largely followed a chronology based on evolutionary stages that differs from the horizon chronology primarily during the Formative Period and the Tiwanaku Period (Hastorf 1999;Kolata 2003;Stanish 2003: 85-90). As the bulk of Chivay type obsidian artifacts are found in the Lake Titicaca Basin, this dissertation will review Andean obsidian distributions using a Titicaca Basin chronology, and it will follow the Formative Period temporal divisions used by Stanish (2003). These regional data will be presented in terms of three major temporal periods in order to be consistent with the framework adopted with the new data reported for this dissertation.

 

Titicaca Chronology

Years cal. AD/BCE

Late Prehispanic

Late Horizon

A.D. 1476 - 1532

Late Intermediate

A.D. 1100 - 1476

Tiwanaku

A.D. 400 - 1100

Early Agropastoralists

Late Formative

500 BCE - AD400

Middle Formative

1300 - 500 BCE

Early Formative

2000 - 1300 BCE

Terminal Archaic

3300 - 2000 BCE

Archaic Foragers

Late Archaic

5000 - 3300 BCE

Middle Archaic

7000 - 5000 BCE

Early Archaic

9000 - 7000 BCE

Table 3-9. Temporal organization of data.

These three major temporal divisions serve to differentiate the major forces that appear to have structured obsidian procurement and distribution. Rather than focusing on the "preceramic" and "ceramic" dichotomy, which is only peripherally related to obsidian use, a decision was made to present data in terms of these three major divisions. Furthermore, socio-political changes that accompanied obsidian circulation in the Terminal Archaic are better considered in conjunction with the later developments in the Formative than together with the earlier periods of the Archaic.

The use of the term "Archaic Foragers" is not meant to suggest that no pastoralism occurred prior to 3300 BCE Many scholars place the initial domestication sometime between the Middle and the Late Archaic in the chronological terms used here (Browman 1989;Kadwell, et al. 2001;Wheeler 1983;Wing 1986). However, initial domestication and adoption of a dedicated pastoralist economy are distinct events, and the evidence from a number of sites in the south-central Andean highlands suggests that the process of shifting to a food producing economy in the Andean highlands featuring camelid pastoralism and agriculture based largely on seed-plant and tuber cultivation was complete by circa 3300 BCE in the Upper Colca region. Likewise, these divisions between "forager" and "pastoralist" are not meant to suggest that foraging activities ceased in the Terminal Archaic, merely that foraging diminished in importance in terms of caloric intake and food security.

The Early, Middle, and Late Archaic Periods have the additional benefit of being differentiated by relatively consistent changes in projectile point morphology through prehistory. A projectile point typology has recently been developed by Klink and Aldenderfer (2005) that synthesizes data from previously developed point typologies with new evidence from stratified Archaic and Formative excavations in the region. This typology has proven to be extremely useful for investigating surface distributions of projectile points, and it will was used throughout this dissertation both for assessing the age of obsidian points on a regional scale, and for assigning temporal categories to surface sites identified in the course of our 2003 survey work.

/Figs_Ch3/Proj_pt_chron.jpg

Figure 3-10. Projectile Point Typology with Titicaca Basin chronology (Klink and Aldenderfer 2005)

 

The temporal groups shown in Figure 3-10 visibly depict the changes in temporal control provided by projectile point evidence between the "Archaic Foragers" and the "Early Agropastoralists". With the advent of series 5 projectile point styles in the Early Agropastoralists time, the forms of projectile points became a far less effective means of differentiating time periods because projectiles do not change with as much regularity. Series 5 projectile points, except for type 5a, are diagnostic only to the time after the domestication of camelids and the dominance of food producing, pastoral economy. The following review will examine only a few of the most significant aspects of these distributions by time period, with a focus on insights from obsidian exchange on the preceramic and Early Agropastoralist stages of Andean prehistory.

3.5.1.Archaic Foragers in the South-Central Andes

To date, relatively few Archaic archaeological sites have been excavated in the area of the south-central Andean highlands, and therefore the regional knowledge of this important time period is limited. A brief overview of each period will be provided here, with an emphasis on archaeological sites containing Chivay type obsidian, in order to contextualize the regional relationships and interactions that were occurring during these preceramic periods. This review will focus on the western slope of the Andes and the Lake Titicaca Basin, the region where Chivay type obsidian abounds. Much of the Archaic Foragers obsidian in the literature that has been sourced was derived from undated, surface contexts from sites that are assumed to date to the Archaic based on projectile point styles or other inference by archaeologists. Klink and Aldenderfer's (2005) projectile point typology permits the classification of site occupations by time period with greater certainty during the Archaic when projectile point styles change with regularity. In some cases, the projectile point typology can be used to assign a date range to temporally diagnostic obsidian projectile points that have been directly analyzed for chemical provenience.

In Burger et al.'s (2000: 275-288) review, procurement and exchange during the preceramic is evaluated through chemical provenancing data from 87 obsidian samples, 70 of which are from the Chivay source. Unfortunately, of the thirteen sites where the samples were collected in the south-central highlands, only three sites (Chamaqta, Asana, and Sumbay) contain excavated contexts placing the samples in preceramic levels. The other samples are assigned to the preceramic group through inference from projectile points, or because they are from a-ceramic sites. The difficulty in this arises from the fact that obsidian was used with much greater frequency in the beginning of the pastoralist periods, as was mentioned above in the case of the Ilave valley (Section 3.4.4 ).

Obsidian materials from both the Chivay and Alca sources were transported relatively long distances from the Early Archaic onward, indicating that material from these obsidian sources has properties that were prized at a relatively early date. Sites such as Asana and Qillqatani, both over 200 linear km from the Chivay source, contain non-local obsidian through much of the preceramic sequence despite the immediate availability of high quality cherts, as well as the lower-quality Aconcahua obsidian, in the local area. This evidence underscores the nature of obsidian provisioning in the Archaic Period. As compared with later pastoralist use of obsidian, Archaic foragers were relatively selective in their use of raw materials. This may reflect the priority placed on dependable hunting technology as compared with the shearing and butchering needs of pastoralists, or the different social and symbolic priorities attached to obsidian by hunters versus herders.

/Figs_Ch3/ArchForagers_region.jpg

Figure 3-11. Chivay type obsidian distributions during the "Archaic Foragers" time
(circa 10,000 - 3,300 BCE).

Early Archaic (circa 10,000 - 7,000 BCE)

Evidence from the Early Archaic in the south-central Andes is slim, but it is accumulating gradually. The available data suggest that during this time period human groups went from initial exploration of the highlands to more stable, year around occupation of the sierra and puna. The best evidence from the highlands comes from a handful of stratified sites in the region with Early Archaic components, as well as distributional evidence from archaeological survey work(Aldenderfer 1997;1998;1999;2003;Craig 2005: 452-484;Klink 2005;Nuñez, et al. 2002;Santoro and Núñez 1987;).

A regional study of projectile points from excavated contexts by Klink and Aldenderfer(2005: 31, 53)noted that projectile point styles diagnostic to the Early Archaic (Series 1 and Types 4a, 4b) and the Early-Middle Archaic transition (Type 2a) have a broad geographical distribution. These types are found throughout a region that includes the littoral, western sierra, and puna areas of extreme northern Chile and southern Peru, and as far north as Pachamachay cave in the Junin puna(Rick 1980). Given the low population densities of that time period, this wide regional stylistic distribution suggests that a high degree of mobility was being practiced in the Early Archaic and first part of the Middle Archaic. Further to the north, in contrast, raw material types in the Junin puna of central Peru suggest to Rick(1980)that very low forager mobility was occurring throughout the early preceramic period. Early Archaic obsidian distribution data largely confirm with the model of high mobility and show that humans found good sources of lithic raw material in the sierra early, and then transported the material or exchanged it widely from a relatively early date.

Current paleoclimate evidence indicates that glaciers advanced during the late glacial between 11,280 and 10,99014C yr bp, and then, despite cool temperatures, the glaciers receded rapidly, perhaps as a result of reduced precipitation(Rodbell and Seltzer 2000). Glaciers were in approximately modern positions in southern Peru by 10,90014C yr bp (circa 11,000 cal BCE). The Early Archaic Period, with reference to diagnostic projectile point styles(Klink and Aldenderfer 2005), begins at circa 9000 cal BCE During the Early Holocene and the first part of the Early Archaic conditions were wetter and cooler than modern conditions. Subsequently, during the latter part of the Early Archaic, the climate began an episode of long-term aridity that lasted through the Late Archaic(Abbott, et al. 1997;Argollo and Mourguiart 2000: 43;Baied and Wheeler 1993;Paduano, et al. 2003: 272). Opportunities in the highlands for human foraging groups were created by a number of new ecological niches for plants and animals that opened up during the Early Holocene. These niches represented a resource pull for mobile foragers that countered the increased difficulty of subsistence in the hypoxic, high altitude environment faced by the non-adapted early settlers in the highlands(Aldenderfer 1999). A review of the evidence for forest cover on the altiplano during the Archaic, with deforestation occurring as a result of pastoral intensification, is provided by Gade(1999: 42-74).

During the Late Pleistocene lacustrine period, paleo-Lake Titicaca (Lake Tauca) was much larger but only a few meters higher than is modern Lake Titicaca (Clapperton 1993: 498-501). Radiocarbon dates on shells show that Lake Tauca was probably still present as late as 10,08014C yr bp (or 9900 cal BCE). Sediment cores from Lake Titicaca indicate that there was an increase in sub-puna vegetation and fire from vegetation prior to 9000 BP (Paduano, et al. 2003: 272). Evidence from paleoclimate records and fluvial geomorphology point to a time of increased aridity and salinity in Lake Titicaca, with short, episodic moist spells beginning around 8000 cal BCE and continuing through the Middle Archaic and Late Archaic Periods (Rigsby, et al. 2003;Wirrmann, et al. 1992). The climatic data for this time suggest that with deglaciation in the Early Holocene a resource niche opened up that exerted a pull on plant and animal species towards the high altitude regions, and that early human groups responded to these opportunities by colonizing the high Andes.

Chivay obsidian in the Early Archaic

Evidence of human use of the Chivay Source beginning in the Early Archaic Period comes from survey work adjacent to the source and from excavations at the site of Asana, 200 km away in Moquegua. As will be discussed in Chapter 7, during the course of survey work in the area of the Chivay source in 2003 the Upper Colca team collected several dozen Early Archaic type projectile points, the majority of them made from obsidian. Evidence for the regional consumption of Chivay obsidian begins with the site ofXAsanaXthat were described earlier in SectionX3.4.1XandXTable 3-4X. Aldenderfer(1998: 157, 163;2000: 383-384)encountered small quantities of obsidian in two levels belonging to the Asana II/Khituña Phase, placing them in the Early Archaic Period. In level PXXIX just one obsidian flake from Chivay was identified in an assemblage consisting of 1,152 lithic artifacts weighing 746 g. This level could not be dated directly, but it is assessed at 9400 uncal bp as it lies stratigraphically above a14C sample from level XXXIII dating to 9820±15014C yr bp (Beta-40063; 10,000-8700 BCE).

Obsidian from Chivay appeared in greater quantity in level PXXIV where eleven flakes of Chivay obsidian made up 0.36% of the lithic assemblage by count. These flakes were all small tertariary flakes, chunks, and shatter, suggesting to Aldenderfer(1998: 163)that a bifacial core of Chivay material was reduced on site. A14C sample from level PXXIV dated to 8720±120 bp (Beta-35599; 8250-7550 BCE). The Khituña phase has been interpreted by Aldenderfer(1998: 172-173)as representing a residential base in the high sierra and the beginnings of permanent settlement above 2500m in elevation. This interpretation is based on the presence of high sierra and puna lithic raw materials, but no coastal materials.

In surface contexts in the Ilave valley, surveys directed by Aldenderfer(1997)located three obsidian projectile points in forms that are possibly diagnostic to the Early and Early-Middle Archaic transitional period (types 1A, 3A, 3B). These three points were analyzed using a portable XRF unit in 2005 and all three were found to be of the Chivay type(Craig and Aldenderfer In Press).

Alca obsidian at Quebrada Jaguay

The earliest obsidian identified in the south-central Andes comes from one of the oldest confirmed sites in South America, the Paleo-Indian site of Quebrada Jaguay 280 near the coast north of Camaná, Arequipa. Sandweiss et al.(1998)report that Alca type obsidian was the dominant lithic material at the site in the Terminal Pleistocene and Early Holocene phases that were identified. Twenty-six Alca type obsidian flakes came from the older occupation that occurred during the Terminal Pleistocene and the context is dated by twelve14C dates on charcoal falling between 11,105±260 bp (BGS-1942; 11,700-10,400 BCE) and 9850±170 bp (BGS-1956; 10,100-8700 BCE). A later occupation contained one flake of Alca obsidian and it belonged to the Early Holocene II phase dated by four14C samples falling between 8053±115 (BGS-1944; 7350-6650 BCE) and 7500±130 (BGS-1700; 6600-6050 BCE). Three other obsidian flakes from the Alca source were collected but they could not be assigned to a temporal context.

Sandweiss et al. observe that the Chivay source is "less than 20 km further; the absence of Chivay obsidian at QJ-280 may indicate that this source was covered by glacial readvance during the Younger Dryas (circa 11,000 to 10,00014C yr B.P.)"(Sandweiss, et al. 1998: 1832). The Maymeja volcanic depression of the Chivay obsidian source indeed shows notable evidence of recent glaciation, but obsidian matching the "Chivay Type" is also found in smaller nodules on adjacent slopes outside the Maymeja zone as well as in the streambed below Maymeja having eroded through alluvial erosion (see Section4.5.1for maps and a discussion of the Chivay obsidian source geography). Procurement of obsidian from Chivay did not necessarily require entering the Maymeja area, however better source material can be obtained from the within the area.

Regional scale GIS analyses show Quebrada Jaguay to be 154 linear km from the Alca source, or 35.7 hours by the hiking function. In contrast, the trip to Quebrada Jaguay from the Chivay source is 170 linear km and 41.5 hours by the hiking function. Despite this relatively difference in distance, no Chivay obsidian was found at Quebrada Jaguay. It is worth noting that, despite this early find of Alca obsidian on coast, both the Alca and the Chivayobsidian types were predominantly found at high altitude throughout prehistory based on current evidence from obsidian sourcing studies. In the bigger picture, the coastal Quebrada Jaguay paleoindian finds are anomalous for Alca distributions. The next lowest altitude context for Alca to date is at the site of Omo in Moquegua, a Tiwanaku colony site at 1250 masl dating to the Middle Horizon.

Recent work by Kurt Rademaker and colleagues at the Alca source has shown that bedrock outcrops of obsidian are found as high as 4800 masl, and there are large pieces found at 3,800 masl that appear to have been transported downslope by glacial action and colluviation (K. Rademaker 2005, pers. comm.). Deposits of Alca obsidian at much lower elevations have been reported that are probably the result of pyroclastic flows(Burger, et al. 1998;Jennings and Glascock 2002). These deposits found close to the floor of the Cotahuasi valley at 2500 masl were probably available throughout the Younger Dryas and perhaps further sub-sourcing of Alca obsidian will better answer the question of which Alca deposits were being consumed during the Terminal Pleistocene by the occupants of Quebrada Jaguay.

Quispisisa Obsidian in the Early Archaic

Another early use of obsidian in the Andes is the evidence of Quispisisa obsidian in levels dated to ~9000 uncalBP or 8500-7750 cal BCE (Burger and Asaro 1977;Burger and Asaro 1978;Burger and Glascock 2000;MacNeish, et al. 1980) from Jaywamachay a.k.a. "Pepper cave" (Ac335) located at 3300 masl in Ayacucho in the central Peruvian highlands. As was recently explained by Burger and Glascock (2000;2002), for nearly three decades the Quispisisa obsidian source was mistakenly believed to have been located in the Department of Huancavelica, but in the late 1990s the true Quispisisa source was finally discovered in the Department of Ayacucho. Jaywamachay is the closest of the sites excavated by MacNeish's team to the newly located Quispisisa source. It is 81.3 linear km or 22.0 hours walking calculated using the hiking function, while the bulk of the sites excavated by MacNeish were in the Ayacucho valley approximately 120 km from the relocated Quispisisa source. The Puzolana (Ayacucho) obsidian source is located close to the Ayacucho valley and high-quality, knappable obsidian is available at this source, but the maximum size for these nodules is 3-4 cm, significantly limiting the potential artifact size for pieces made from this source (Burger and Glascock 2001). It is interesting to note that, in several cases in the south-central Andes, the Archaic Period foragers seem to have ignored small sized or low-quality obsidian sources that later pastoralists ended up exploiting. The Aconcahua source near Mazo Cruz, Puno, previously described, was not used at the adjacent Qillqatani rock shelter until the Middle Formative (Frye, et al. 1998), and it was not used at Asana until the Terminal Archaic (Aldenderfer 2000: 383-384).

Burger and Asaro(1977: 22;1978: 64-65 )chemically analyzed a projectile point of Quispisisa obsidian found at La Cumbre in the Moche Valley that was reportedly from a context dated to 8585±280 bp or 8500-6800 cal BCE(Ossa and Moseley 1971). The investigator, Paul Ossa, now doubts the Early Archaic context and instead believes the point may be Middle Horizon in date (Richard Burger, 11 March 2006, personal communication), but it is still a noteworthy case of long distance transport within the Wari Empire, and it perhaps involved the use of coastal maritime transport. Using the relocated Quispisisa source in the Department of Ayacucho the linear distance is 846.7 km and by the hiking function the distance is calculated as 199.8 hours, primarily confined to the coast.

Discussion

Regional relationships indicated by both stylistic criteria and obsidian distributions point to high mobility during the Early Archaic Period. Recent evidence of projectile point similarities from northern Chile suggest that mobility was high along the Pacific littoral, and between the sierra and puna during the Terminal Pleistocene and Early Holocene around 25° S latitude(Grosjean, et al. 2005;Nuñez, et al. 2002). During the Early Holocene, foraging groups resided around high altitude paleo-lakes on the altiplano of northern Chile between 20° S and 25° S latitude that persisted longer than did the paleo-lakes at the latitude of Lake Titicaca. These lakes subsequently dried up during the Middle Archaic Period and occupations declined until the end of the Late Archaic Period but this distributions point to the mobility and early high altitude occupation of the adjacent puna of Atacama.

On the whole, obsidian proveniencing and analysis has shed light on human activities during the Paleoindian period and Early Archaic at a regional scale. Chemical analysis techniques, non-destructive XRF analysis in particular, are becoming more pervasive because the methods are being refined and the equipment is becoming more portable. Further field research will provide a greater understanding of this time period, but evidence will also likely to emerge from existing collections as obsidian projectile points diagnostic to the Early Archaic are systematically provenienced.

Middle Archaic (7000 - 5000 BCE)

The Middle Archaic Period, as with the preceding period, is still poorly understood in the region as very few highland sites containing stratified deposits have been studied from this period. In the Lake Titicaca Basin paleo-climatic evidence points to significant aridity, a lake level approximately 15m below the current stand, and high salinity in Lake Titicaca that is comparable to that of modern day Lake Poopó(Abbott, et al. 1997;Wirrmann and Mourguiart 1995). Human occupation during the Middle Archaic in the Lake Titicaca Basin was notably higher than in the preceding period, but research shows that settlement continued to occur in the upper reaches of the tributary rivers and not adjacent to the shores of the saline Lake Titicaca(Aldenderfer 1997;Klink 2005).

Obsidian distributions during the Middle Archaic

At the site of Asana on the western slope of the Andes, Aldenderfer(1998: 223)observed architectural features that suggest that a longer residential occupation of the site by entire coresidential groups was occurring in the latter part of the Middle Archaic Muruq'uta phase.One flake of obsidian from the Chivay source was found in the upper levels of theMuruq'uta phase (XTable 3-4X).This occupation dates to the Middle Archaic -Late Archaic transition as the flake was stratigraphically above a14Csample that dated to 6040 ± 90 bp (Beta-24634; 5210-4720 BCE).

At the rock shelter of Sumbay SU-3 obsidian was recovered from excavation levels that date to as far back as the transition between the Middle and Late Archaic (SectionX3.4.3X).

Discussion

Relatively little is known about the Middle Archaic in the south-central Andean highlands. Elsewhere in the south-central Andes, archaeologists have noted an absence of settlement during the mid-Holocene timeframe corresponding to the Middle and Late Archaic Periods. In the dry and salt puna areas of northern Chile, and along the south coast of Peru, a significant decline or absence of mid-Holocene sites has led investigators to refer to this period as the silencio arqueológico(Nuñez, et al. 2002;Núñez and Santoro 1988;Sandweiss 2003;Sandweiss, et al. 1998). This designation apparently does not apply to the Titicaca Basin or to the sierra areas of the Osmore drainage where no Middle Archaic occupation hiatus has been observed.

With gradual population increases and adaptation to the puna, social networks extending across the altiplano and connecting communities and their resources residing in lower elevations with puna dwellers, were probably beginning to take form. Exchange of resources, including obsidian, between neighboring groups may have been in the context of both maintaining access to resources and risk reduction. From a subsistence perspective, Spielmann (1986: 281) describes these as buffering, a means of alleviating period food shortages by physically accessing them directly in neighboring areas, and mutualism,where complementary foods that are procured or produced are exchanged on a regular basis. Another likely context for obsidian distribution during the Archaic Period is at periodic aggregations. Seasonal aggregations have been well-documented among foragers living in low population densities, where gatherings are the occasion for trade, consumption of surplus food, encountering mates, and the maintenance of social ties and ceremonial obligations (Birdsell 1970: 120;Steward 1938). If analogous gatherings occurred among early foragers in the south-central Andes it have would created an excellent context for the distribution of raw materials, particularly a highly visible material like obsidian that was irregularly available in the landscape.

Late Archaic (5000 - 3300 BCE)

The Late Archaic in the south-central Andes signals the beginning of economic changes in the lead-up to food production, and furthermore the first signs of incipient social and political differentiation are evident in a few archaeological contexts. Some of the obsidian samples reported by Burger et al(2000: 275-288)with possible Late Archaic affiliations are from surface contexts at multicomponent sites and it is therefore difficult to confidently assign these samples to any particular period of the Archaic.

Chivay obsidian during the Late Archaic

Evidence of Late Archaic obsidian use comes from excavations at the previously discussed sites of Asana, Qillqatani, and Sumbay. Several obsidian samples in the Burger et al. (2000) study were of portions of diagnostic projectile points that resemble types in Klink and Aldenderfer's (2005) point chronology. From illustrations and text(Burger, et al. 2000: 279, 281)the Qaqachupa sample appears to belong to the Late Archaic. Burger et al. describe this point as resembling a Type 7 point from Toquepala in Ravines'(1973)classification, and Klink and Aldenderfer(2005: 44)mention that their Type 4D point strongly resembles the Toquepala Type 7, making it diagnostic to the Late Archaic.

Asana

In excavated data from Asana, in Moquegua, one sample of Chivay obsidian was found in a Muruq'uta phase occupation in Level XIV-West lying above a palimpsest dated to 6040±90 (Beta-24634; 5210-4766 BCE) (Aldenderfer 1998: 269), as shown above in SectionX3.4.1X. This level lies on the Middle Archaic / Late Archaic transition. Interestingly, Chivay obsidian disappears at Asana subsequent to this time period. Furthermore, evidence of lower raw material diversity in the Late Archaic lithic assemblage point to greater geographical circumscription.

Qillqatani

Chivay obsidian at Qillqatani is first found in the Late Archaic level WXXX dated to 5620±120 (Beta-43927; 4800-4200 BCE). The Chivay material is the second oldest obsidian fragment identified from Qillqatani (SectionX3.4.2X), the oldest obsidian is from an as-yet unknown source. Notably, the assemblages from the Qillqatani excavations do not begin to contain obsidian from the Chivay source until considerably later date than did the excavations from Asana.XFigure 3-7Xreveals that both obsidian tools and debris increase as a percentage of the assemblage in the Late Archaic. The counts of obsidian, however, are still relatively low.

Qillqatani is slightly further away than Asana than from the Chivay source, and it is at a higher altitude and further to the east. It has been suggested that perhaps the obsidian was transported via a coastal route at this early date (Frye, et al. 1998). As mentioned, Alca obsidian was also found on the coast in the Terminal Pleistocene levels although, like Chivay, Alca obsidian distributions conform over the long term to a highlands orientation. As no Chivay obsidian as ever been found below the 1250 masl (at Omo), and all Archaic Period Chivay obsidian is found above 3000 masl, the littoral route between Chivay and Asana seems improbable.

IlaveValley

Craig and Aldenderfer(In Press)report that two obsidian projectile points from the Ilave valley in a type 3f form, diagnostic to the Late - Terminal Archaic were analyzed in 2005 with a portable XRF unit and were found to be of the Chivay type.

Sumbay

At Sumbay SU-3,three obsidian samples were analyzed from excavated contexts and all three turned out to be of the Chivay type(Burger, et al. 1998: 209;Burger, et al. 2000: 278), as shown in SectionX3.4.3X. Two14C dates were run and returned dates from the early part of the Late Archaic(Ravines 1982: 180-181). One sample was from stratum 3 and it was dated to 6160±120 (BONN-1558; 5400-4750 BCE). Another sample was from above it in stratum 2 and it dated to 5350±90 (BONN-1559; 4350-3980 BCE). One of the three obsidian samples came from Stratum 4 of unit 5, while the other two samples came from higher levels. The Stratum 4 sample probably dates to the Middle Archaic.

Discussion

Obsidian from securely dated Late Archaic contexts show something of a reduction in regional distribution and a greater focus on locally available lithic material, suggesting a reduction in mobility or exchange in the Late Archaic. Similarly, projectile point styles became increasingly more limited in spatial distribution, with greater local variability during the Late Archaic implying reduced mobility(Klink and Aldenderfer 2005: 53). This is consistent with Aldenderfer's(1998: 260-261)observations about reduced mobility during the Late Archaic Qhuna Phase occupation at Asana when the occupants ceased to use non-local lithic raw materials. During this phase, Aldenderfer also describes increasingly formalized use of space at Asana, evidence of a ceremonial complex, and greater investment in seed grinding. In short, during this time a circumscribed population with reduced mobility was probably living in higher densities and exhibiting signs of ceremonial activity that are consistent with the scalar stress model for the emergence of leadership(Johnson 1982).

It is also worth considering the impact that scarcity may have on valuation. The lack of discarded obsidian signifies that it was not being knapped or resharpened and it was probably not abundant, but that does not mean that obsidian was not known in the larger consumption zone during this period. In the subsequent time period, the Terminal Archaic, obsidian becomes abundant on a regional scale at the same time as a host of other social and economic changes were occurring. This period, and the previously discussed Middle Archaic, correspond with what was referred to as the silencio arqueologico(Núñez and Santoro 1988)due to a dearth of archaeological data observed by investigators working in Northern Chile. The reduced evidence of circulation of obsidian from Chivay appears to correlate with a reduction in archaeological evidence regionally.

Thereis strong representation of Chivay obsidian in Arequipa at Sumbay, and it is likely that Late Archaic projectile point forms are found in the North Titicaca Basin as reported by Burger et al. (2000). However, at Asana there is little obsidian from the Colca. Possibly these reduced distributions of obsidian reflect the reduced mobility and more complex architectural investment in Late Archaic contexts at Asana(Aldenderfer 1998), and prior to the development of extensive, long distance exchange that were potentially initiated by early caravan networks during the Terminal Archaic.These conclusions, however, on not based on particularly robust data, as the sample of sites for this time period is relatively small.

Discussion and review of the Archaic Foragers period

The transport of obsidian from both the Chivay and Alca sources fluctuated during the Archaic Foragers period, and the distribution remained confined to the sierra and altiplano areas of the south-central Andes. The small and irregular quantities of obsidian consumed throughout the region could have been procured and transported by mobile foragers, or conveyed through down-the-line exchange networks. Conceivably exchange could have taken place a seasonal gatherings as well, based on analogy from other foraging groups worldwide, although there is no direct empirical support for such gatherings in the south-central Andean Archaic. At Asana, the most well-stratified highland Archaic site investigated to date, evidence of non-local obsidian was intermittent and consisted of very small quantities of material in a pattern that is perhaps exemplary of the discontinuous nature of regional exchange during this time.

Obsidian from the Quispisisa source (340 linear km to the north-west of the Chivay source), was likewise used primarily in highland contexts during the Archaic Foragers period, particularly in rock shelters excavated by MacNeish and colleagues in the central highlands. While Quispisisa material was circulated at a number of preceramic coastal sites(Burger and Asaro 1978), current evidence suggests that those contexts post-date 3300 cal BCE (i.e., cotton preceramic) and are therefore Terminal Archaic in the terms used here. It appears that while Quispisisa material was used widely on the coasts subsequent to the Archaic Foragers period, material from Chivay and Alca were virtually always confined in the highlands with the exception of the earliest obsidian evidence of all: Alca at Quebrada Jaguay. In the Archaic Foragers period, as in later periods, the evidence from obsidian is contrary to the widely-discussed Andean vertical complementarity models, because obsidian distributions suggest that the movement of people or products between ecologically complementary zones was not widespread.

3.5.2.Early Agropastoralist obsidian distributions

The "Early Agropastoralists" period (3,300 BCE - A.D. 400) begins with what appears to have been a shift to a chiefly pastoral lifeway, greater regional interaction, and a more intensive production and circulation of obsidian from the Chivay source area in the Terminal Archaic. In this discussion, the Early Agropastoralists period continues through the Late Formative and subsequently beginning AD400, with the ascendancy of Tiwanaku, the "Late Prehispanic" time block begins. The changes during the Terminal Archaic that mark the beginning of the Early Agropastoralist time include the growing importance of food production, the expanded production and circulation of obsidian, and socio-political differentiation that began to appear in the Terminal Archaic, phenomena of greater interest to this research than the presence or absence of pottery.

/Figs_Ch3/Early_AgroPast_region.jpg

Figure 3-12. Chivay type obsidian distributions during the "Early Agropastoralists" time (3,300 BCE - AD 400).

Why an "Early Agropastoralists" time block?

In lieu of the traditional preceramic / ceramic divide in archaeology, the "Early Agropastoralist" time block was defined for this project because, in many ways, the conditions during the final millennium of the Archaic had more in common with events in the Formative than with the preceding Archaic Periods. Differentiating sites as "Early Agropastoralist" beginning in the Terminal Archaic requires that one delimits a firm boundary on a process that was millennia in the making. Considering the Terminal Archaic as the beginnings of agropastoralism, and thereby linking it to the attendant settlement distribution that includes series 5 projectile points, connects cultural and economic evidence from throughout the region with a generalized estimate of when food production took hold in terms of scheduling, mobility, and prioritizing the needs of agricultural planting and herding. As mentioned, the term "Early Pastoral" is not meant to imply that no food production occurred before this date and hunting and gathering did not persist after this date. Indeed the current evidence suggests that camelids were first domesticated sometime during the Middle or Late Archaic, probably after 5000 BCE (Browman 1989;Kadwell, et al. 2001;Wheeler 1983;Wing 1986). The "Early Agropastoralist" period refers to a period after 3300 BCE that features a predominantly food producing economy in the highlands including intensive pastoralism, as well as seed-plant and tuber cultivation. These changes in food production may have been linked to the accumulating evidence of increased rainfall and shortening of the dry season beginning circa 2500 cal BCE (Baker, et al. 2001;Marchant and Hooghiemstra 2004;Mourguiart 2000). These changes in economy are joined by evidence of increased sedentism, widening stylistic distributions pointing to long distance cultural integration, and the beginnings of social differentiation apparent in architecture and grave goods. The circulation of obsidian was linked to these phenomena because increased regional interaction was perhaps articulated through camelid caravans, although the presence of caravans in the Terminal Archaic is being explored, not assumed, in this study.

In the Titicaca Basin, the changes incurred between the Terminal Archaic and the Late Formative are monumental. However, from the perspective of the Chivay source the intensification revealed archaeologically, both at the source and in the consumption zone, is notable beginning in the Terminal Archaic, and the intensification does not change notably throughout the Formative Period. Defining the ending of the "Early Agropastoralist" period as the end of the Late Formative is similarly problematic. By the Late Formative, the economic influence of Pukara and other regional centers probably impacted the peoples of the Upper Colca, but direct evidence returning from Pukara, only 140 km away, is scarce in the Colca. In contrast, during the Tiwanaku times, regional states with socio-economic impacts over hundreds of kilometers dominated both the circulation of goods like obsidian, and regional settlement organization.

Terminal Archaic (3300 - 2000 BCE)

The Terminal Archaic ushers in a suite of social and economic changes in the south-central Andes and, consistent with these developments, obsidian begins to circulate in significantly greater quantities during this period. Obsidian is increasingly used for projectile points during the Terminal Archaic, a trend that continues into the Formative Period (Burger, et al. 2000: 294).

Regional patterns in Terminal Archaic sites are somewhat difficult to assess from surface finds because the Terminal Archaic is lacking in exclusive diagnostic artifacts in both the lithic or ceramic artifact classes. As is shown inXFigure 3-10X, the most common projectile point styles that belong to the Terminal Archaic, such as Types 5B and 5D, also persist through the ceramic periods, leaving only type 5A and part of type 4F as diagnostic to exclusively the Terminal Archaic (Klink and Aldenderfer 2005: 48). Furthermore, by definition, the Terminal Archaic is a preceramic period, which precludes a ceramic means of assigning chronology. From site organization characteristics, a site might be considered to be a Terminal Archaic site if it has pastoralist attributes but it is aceramic and has series 5 projectile point types represented that belong, at least partially, to the Terminal Archaic.

The beginnings of the Andean agropastoral strategy are apparent during this period at sites like Asana where seasonal residential movement occurred between the high sierra and the puna (Aldenderfer 1998: 261-275;Kuznar 1995). Another attribute of the Terminal Archaic at Asana is a disappearance of ceremonial features. Evidence for a shift from hunting to pastoralism comes primarily from evidence of corrals (Aldenderfer 1998), from changes in the ratio of deer to camelid remains, and from the ratio of camelid neonate to adult remains. At sites where the process has been observed, the transition to full pastoralism at various sites in the Andes is usually perceived as a gradual process dating to sometime between 3300 to 1500 BCE At Qillqatani, however, the transition was relatively abrupt and it occurred in level WXXIV that dates to approximately 2210-1880 BCE (XTable 3-1X).

Chivay obsidian during the Terminal Archaic

While the number of excavated Terminal Archaic sites is relatively small, general processes are apparent from recent work at several sites in the south-central Andean highlands.

At Ch'uxuqullu on the Island of the Sun, Stanish et al.(2002)report three obsidian samples from the Chivay source in Preceramic levels. Eight obsidian flakes were found in aceramic levels and were described as being from a middle stage of manufacture, and three of these were sourced with NAA. Two samples came from levels with a14C date of 3780±100bp (Teledyne-I-18, 314; 2500-1900 BCE) and a third Chivay obsidian sample comes from an aceramic level that immediately predates the first ceramics level which occurred at 3110±45bp (AMS-NSF; 1460-1260 BCE). Citing paleoclimate data, the authors observe that boat travel was required to access the Island of the Sun at this time.

The Ilave Valley and Jiskairumoko

Terminal Archaic projectile points in the Ilave valley demonstrate the dramatic shift in the frequency and use of obsidian that occurred in this time period (SectionX3.4.4X). XRF analysis of 68 obsidian artifacts excavated from Jiskairumoko show that the Chivay source was used with particular intensity in this period. The XRF study found that 97% of obsidian bifaces from Terminal Archaic and Early Formative levels at Jiskairumoko were from the Chivay source, however published obsidian studies from all time periods show that typically 90% of all obsidian analyzed from Titicaca Basin sites are from the Chivay source (XTable 3-3X).

Qillqatani

The percentages of obsidian tools and debris remain generally similar to those in the Late Archaic level except that the counts are much higher (XFigure 3-7X). By count, obsidian tools are doubled, and obsidian debris is 4.3 times greater, and all the obsidian has visual characteristics of the Chivay type (Aldenderfer 1999).

Based on the density of obsidian, the relationship with the Chivay source area 221 km away seems to be well-established by this time, and it is a relationship that becomes even more well-developed in the Early Formative. Six samples of obsidian were analyzed at the MURR facility from Terminal Archaic contexts that are associated with a radiocarbon date of 3660±120 (Beta-43926; 2210-1880 BCE). All six obsidian samples were from the Chivay source.

Asana

At the site of Asana, obsidian reappears at the end of the Terminal Archaic during the Awati Phase dated to3640±80 (Beta-23364; 2300-1750 BCE)where it makes up 0.4% of lithic materials, but this obsidian was not from the Chivay source. It was judged from distinctive visual characteristics to have come from the Aconcagua obsidian source only 84 km to the east of Asana, near the town of Mazo Cruz(Aldenderfer 2000). Aconcahua type obsidian has characteristics that are less desirable for knapping due to fractures and perlitic veins that cross cut the material (see Appendix B.1), and while it was possible to derive sharp flakes for shearing and butchering functions, the material was probably not used for projectile point production(Frye, et al. 1998).

The shift to Aconcahua obsidian in the Awati phase at Asana is particularly puzzling given the evidence for Chivay obsidian circulation at this time period. It is precisely at the end of the Terminal Archaic that a dramatic spike in the use of Chivay obsidian at Qillqatani (SeeXTable 3-5XandXFigure 3-7X) took place. One may ask: Why is it that when the occupants of Qillqatani are importing Chivay obsidian in unprecedented quantities, the people of Asana are getting only small quantities of low-quality obsidian? In addition, this low-quality obsidian comes from Aconcahua, a source adjacent to Qillqatani?

Given the pattern of early Chivay obsidian at Asana, these Terminal Archaic distributions suggest that the high sierra residents at Asana were not participating in an altiplano-based circulation of goods as the Qillqatani residents. The residents of Asana never again participate in the circulation and consumption of Chivay obsidian, while at Qillqatani the consumption of Chivay material continues strongly for another one thousand years.

Other obsidian types during the Terminal Archaic

Obsidian from alternative sources in the circum-Titicaca region, including the unlocated sources of Tumuku and Chumbivilcas types, are used in greater quantity during the Terminal Archaic judging from associated projectile point evidence provided in Burger et al.(2000: 280-284).

Evidence from close to the Alca obsidian source provides new information about long distance interaction during Terminal Archaic. At the site of Waynuña at 3600 masl(Jennings 2002: 540-546)and less than one day's travel from the Alca obsidian source, recent investigations have uncovered a residential structure with evidence from starch grains resulting from the processing of corn as well as starch from arrowroot, a plant necessarily procured in the Amazon basin(Perry, et al. 2006). Given the long distance transport of arrowroot, it is conceivable that the plant material arrived as a form of reciprocation or direct transport from travelers moving between the Amazon and the Alca obsidian source. The Cotahuasi valley also has major salt source and other minerals that would potentially draw people procuring such materials. The starch samples were found in a structure on a floor dated by two14C samples. One sample was dated to 3431±45 (BGS-2576) 1880-1620 BCE, and another was 3745±65 (BGS-2573) 2350-1950 BCE

Further north, the Quispisisa type obsidian was particularly abundant during the Terminal Archaic at the preceramic coastal shell mound site of San Nicolas, along the Nasca coast, in a context associated with early cotton(Burger and Asaro 1977;Burger and Asaro 1978: 63-65). Quantitative data on the consumption of obsidian at San Nicolas are unavailable and the temporal control is weak because the "cotton preceramic" date is derived from association with cotton and no ceramics, not from direct14C dating.

Discussion

The distribution of obsidian from all three major Andean obsidian sources: Chivay, Alca, and Quispisisa, expanded considerably during the Terminal Archaic. It is notable that both the Chivay and Alca sources expanded, but the distribution remained confined to the sierra and altiplano areas of the south-central Andes. In comparison, obsidian from the Quispisisa source (340 linear km to the north-west of the Chivay source), has been found in significant quantities in Ica on the coast of Peru. While many of the early coastal obsidian samples have weak chronological control, the quantities of Quispisisa obsidian found in possible Archaic contexts is noteworthy. The fact that Chivay obsidian has never been found in coastal areas, and Alca is not found on the coast after the Paleoindian period, is remarkable considering the extensive evidence of coastal use of Quispisisa obsidian beginning in the Terminal Archaic.

Early Formative Period (2000 - 1300 BCE)

In this time period, new socio-economic patterns became well established in the south-central Andes while the distribution of Chivay obsidian was at its maximum both in geographic extent and in variability of site types. During this time period, societies were characterized by sedentism, demographic growth, increased specialization, and early evidence of social ranking(Aldenderfer 1989;Craig 2005;Stanish 2003: 99-109).

At sites like Jiskairumoko in the Titicaca Basin, architecture from the Terminal Archaic consisted of pithouses with internal storage, ground-stone, and interments with grave goods that included non-local items (like obsidian) that were presumably of value(Craig 2005). During the Early Formative around 1500 BCE this architectural pattern gives way to larger, rectangular, above-ground structures that lacked internal storage.

As reviewed by Burger et al. (2000: 288-296) using the Ica "Master sequence" chronology under the roughly contemporaneous "Initial Period" (though the Early Formative ends 300 calendar years earlier), the distributions of obsidian are notable in their extension both north and south from the Chivay source, and in their concentrations at early centers like Qaluyu. Obsidian from Chivay also persists on the Island of the Sun in the Early Formative(Stanish, et al. 2002).

Chivay obsidian distributions

At Qillqatani, obsidian flakes are available in quantity during this Titicaca time period which roughly overlap with the Formative A and B (XTable 3-5X). Seven obsidian samples were analyzed at MURR that all corresponded with the Chivay source. These samples were found in levels adjacent to a context that dated to2940±70bp (Beta-43925; 1380-970 BCE).

It appears that obsidian was being obtained as nodules or blanks and being knapped down to projectile points. Obsidian flakes represent 15% by count (n = 160) of all the flakes in the period "Formative A", and obsidian projectile points represent 12% by count of all points in this level(Aldenderfer 1999).

Obsidian is first found at Qaluyu dated to ~3250 uncal bp which calibrates to 1640-1420 BCE(Burger, et al. 2000: 291-296), or the end of the Early Formative where Burger et al. report evidence of nodules or blanks arriving for further reduction at the source. Some of the Chivay obsidian found at Qaluyu were medium sized pebbles; one example Burger et al. analyzed had the dimensions 1.9 x 1.4 x 1.2 cm. Qaluyu is 140 linear km from the Chivay source, or 34.6 hours by the hiking model, so it is roughly 70% of the distance of Qillqatani from the source. There are substantial quantities of obsidian at Qaluyu during the subsequent Middle Formative.

Discussion

Paleoclimatic studies have documented a decline in abundance of arboreal species in the Lake Titicaca Basin and an increase in open-ground weed species reflecting disturbed soils after ca. 1150 cal BCE(Paduano, et al. 2003: 274). This has been interpreted as reflecting intensification in food production and population that occurred during the Early Formative. Some argue that the grasslands of the altiplano are an anthropogenic artifact of a human induced lowering of the treeline due in part to the expansion of camelid pastoralism(Gade 1999: 42-74). The increasing sedentism, cultivated plants, expanding camelid herds and the growing influence of prominent settlements like Qaluyu in regional organization signify that significant socio-economic dynamism was underway by this period. The evidence of hierarchy takes the form of "very moderate social rank" acquired by certain individuals, such as religious specialists(Stanish 2003: 108). These traits are first suggested by jewelry, grave goods, and non-residential structures in the Terminal Archaic, but they become further elaborated during the Early Formative after 2000 BCE.

There is no evidence of political ranking in the structure of settlement distributions between Early Formative sites in the Lake Titicaca Basin, and all the sites are less than one hectare in size(Stanish 2003: 108), but the evidence of obsidian circulation from Qillqatani and other pastoral sites suggests of regional exchange system not yet focused around Lake Titicaca. Evidence acquired from the earliest occupational levels at Titicaca Basin regional centers suggest that regional interaction was low. At the site of Chiripa, in the southern Lake Titicaca Basin, Matthew Bandy describes traces of sodalite beads, obsidian, and sea shell in excavations from Early and Middle Formative contexts.

It is clear, then, that long distance trade in mortuary and prestige items took place as early as the Early Chiripa phase [1500-1000 BCE]. Equally clear, however, is that this early exchange involved very small quantities of the objects in question. This trading would seem to have been very sporadic and infrequent. There is no evidence in the Early Formative Period for the sort of regular caravan trade postulated by Browman (his "altiplano mode"; see Browman 1981: 414-415) (Bandy 2001: 141).

However, the evidence for obsidian distributions at Qillqatani that are discussed here point to routine exchange, perhaps by caravan trade, along on the Western Cordillera in the Early Formative. It appears that future regional centers like Chiripa were not yet participants in these exchange networks.

The dynamic nature of interregional interaction that occurred during the Early Formative is evident from the diversity in obsidian types at sites, and the lack of geographical restrictions that appear to structure exchange during later periods. While settled communities were evident, the lack of hierarchy and centralization suggests that they were not integrated by supra-local organization. Yet the economic basis for long distance relationships appears to have taken form by the Early Formative. The decentralized and variable nature of exchange in this period, which is abundantly evident at Qillqatani and yet not evident at sites like Chiripa, implies that another form of integration was linking communities, like the residents of Qillqatani, with the Chivay source 221 km to the north-west.

Middle Formative Period (1300 - 500 BCE)

During the Middle Formative Period, social ranking became established in the Titicaca Basin. The changes are most evident in the settlement structure as some regional centers grew to become far larger than their neighbors and feature sunken courts, mounds, and specialized stone and ceramic traditions. Stanish(2003: 109-110)interprets these changes in terms of an ability of elites to mobilize labor beyond the household level.

The stylistic evidence suggests that during the Middle Formative the north and south Titicaca Basin were relatively separate spheres, with Qaluyu pottery in the north and fiber-tempered Chiripa ceramics in the south extending only as far north as the Ilave river. However, Chivay obsidian is encountered in both the North and South Basin. Christine Hastorf (2005: 75)suggests that by the end of this period (the Early Upper Formative) evidence of ethnic identity and ritual activity is supported by ritual architectural construction and non-local exchange goods. It is further inferred that "Plants such as coca ( Erythroxylumsp.), Anadenanthera ( A. colubrine, A. peregrine), and tobacco (Nicotiana rustica)surely would have been present in the Basin by this time, perhaps associated with snuff trays…"(Hastorf 2005: 75). An increase in long distance exchange is commonly found as part of a complex of features associated with ideological and social power during the Middle Formative in the region, and it appears that existing exchange routes, such as the one along the Western Cordillera connecting Chivay with Qillqatani, were increasingly routed towards the Titicaca Basin regional centers during this time.

Chivay obsidian distributions

Chivay obsidian occurs in small quantities at a number of Middle Formative sites in the southern Basin including Chiripa and Tumatumani, and it persists at Ch'uxuqullu on the Island of the Sun. At Tumatumani, 3% of the projectile points are made from obsidian(Stanish and Steadman 1994). Bandy(2001: 141)reports that at Chiripa they recovered only small quantities of obsidian in the time spanning 1500-200 BCE and these were in the form of finished bifaces. For the entire time span, after four excavation seasons, they report only 87.1g of obsidian. Tumuku type obsidian was identified in Chiripa in Condori 1B component circa 1500-1000 cal BCE levels(Browman 1998: 310, dates calibrated). At the site of Camata, on the lakeshore south of the modern city of Puno, four obsidian samples were analyzed from contexts that range from circa 1500 - 500 cal BCE and all four were of Chivay type obsidian(Frye, et al. 1998;Steadman 1995).

The evidence from the North Titicaca Basin regional centers is even more intriguing as there is a significant presence of Alca type obsidian during the early part of the Middle Formative, and Chivay obsidian is found in the Cusco Basin during this time. Qaluyu, Pikicallepata, and Marcavalle all contain both Chivay and Alca obsidian between approximately 1100 - 800 cal. BCE, the early part of the Middle Formative(Burger, et al. 2000: 292;Chávez 1980: 249-253). Subsequent to this overlap in obsidian use, there appears to have been significant overlap in other stylistic attributes as well. These similarities include common traits in ceramic vessel forms between Chanapata vessels in the Cusco area and Qaluyu vessels in the North Titicaca Basin(Burger, et al. 2000: 292).

However, during the latter part of the Middle Formative after 800 BCE the Yaya-Mama religious tradition first emerges at the site of Chiripa(Bandy 2004: 330;Chávez 1988), a tradition that eventually unifies the north and south areas of the Titicaca Basin during the Late Formative. As noted by Burger et al.(2000: 311-314), with the appearance of the Yaya-Mama tradition the Alca and Chivay obsidian distributions become more asymmetrical. Alca obsidian makes up 16% (n = 9) of the obsidian in a pre-Pukara context at the site of Taraco on the Titicaca lake edge in the North Basin, however while Chivay obsidian was found at Marcavalle and other Cusco sites previously during the Early Formative, obsidian from the Chivay source is absent during the Middle Formative and it does not re-appear in the Cusco region again until the Inka period. Alca obsidian, on the other hand, expands outward during this period as it is found in the Titicaca Basin to the south-east, and it is also is transported a great distance to Chavín de Huantar. Both of these examples of long distance transport have been attributed to religious pilgrimage(Burger, et al. 2000: 314).

Qillqatani

At Qillqatani, the Middle Formative comprises the bulk of Formative B layers and all of the Formative C layers (see Qillqatani data in Section X3.4.2X). In Formative B layers, Chivay obsidian is the only type represented in the four samples that were analyzed and the lithic assemblage suggests that formal tools were not being produced at the site as no evidence of obsidian tools were found in these levels. Obsidian flakes, however, persist as 18% of the lithic assemblage from that level. Subsequently, in the Formative C level that begins around 900 BCE and corresponds approximately with the latter half of the Middle Formative as well as the rise of the Yaya-Mama tradition in the south Titicaca Basin, there is a distinctive shift in the use of obsidian at Qillqatani. Whereas all prior obsidian samples from Qillqatani were Chivay after the initial Middle Archaic sample, the obsidian samples in Qillqatani Formative C levels are only 60% from the Chivay source.

The other samples come from Aconcahua, a source of lower-quality obsidian that is near the Qillqatani shelter, and from Tumuku, an as-yet undiscovered source that may be located close to the three-way border between Peru, Bolivia, and Chile, and finally Alca obsidian occurs for the first and only time at Qillqatani in these levels. Given that Alca obsidian also occurs at Pukara, Incatunahuiri (surface) and at Taraco in quantity (16% of assemblage, n = 9), the presence of Alca obsidian at Qillqatani is consistent with the abundance of Alca material in circulation in that time. A sample of Alca obsidian has also been found on the Island of the Sun (Frye, et al. 1998), though it was from a surface context.

Table 3-6Xshows that obsidian tools in Formative C levels at Qillqatani are abundant (n = 19) and relatively large on average (1.21 g), and the non-obsidian tools were also very abundant (n = 187) for this level.

Discussion

Middle Formative obsidian distributions appear to demonstrate the emergence of a distinctive Titicaca Basin exchange sphere. One could argue that the emerging elites that mobilized labor to build the initial mounds and courts, and sponsored specialized artistry in stone and ceramics, may have precipitated a demand for greater exotic exchange goods as a source of prestige. Stanish(2003: 162)believes that this is the process that occurs later, during the Late Formative, when he argues that this process is connected to wealth generation for sponsoring feasts and other activities, though he admits the data are sparse. Evidence of long distance exchange from contexts belonging to the Early Formative and first half of the Middle Formative (1500-1000 BCE) at sites like Chiripa are sparse, irregular, and generally involve very small, portable goods; however the evidence from Qillqatani supports other models of more regular interaction along established exchange routes.

Late Formative Period (500 BCE - AD 400)

During the Late Formative, significant social ranking developed and dominated the socio-political landscape in the Lake Titicaca Basin. The complex polities that emerged on either end of Lake Titicaca were distinguished by architecture, stoneworking, and ceramic traditions. A three-tiered site size hierarchy is evident in the Late Formative, and the construction of prominent terraced mounds with sunken courts occurred in a few major sites at this time. The elite ceramics and stoneworking, and the construction of elaborate mounds as a venue for large-scale feasts and human sacrifice at sites like Pukara and early Tiwanaku, can be interpreted as a means of demonstrating the large-scale organization of labor by elites (Stanish 2003: 143).

The patterns of obsidian circulation that emerged at the end of the Middle Formative became very well established in the Late Formative. Excavations at centers in the North Basin have found a reduction in the presence of Alca obsidian in the Titicaca Basin, but the Alca material is still present although it is found in minor quantities at the larger sites as compared with Chivay obsidian. The diversity of obsidian types in the Titicaca Basin samples reported by Burger et al. (2000: 306-308) for this time span is low, as compared with the diversity of types used in the Cusco area, because in the Titicaca Basin it is virtually all Chivay obsidian.

In recent excavation work at Pukara on the central pampa at the base of the Qalasaya, Elizabeth Klarich (2005: 255-256) found that obsidian was generally available and obsidian use was not associated with any intrasite status differences. Probable representations of obsidian points and knives appear in Pukara iconography, and small discs have been found at Pukara and Taraco that were possible ceramic inlays (Burger, et al. 2000: 320-321). The Late Formative Period on the southern end of the Titicaca Basin has small amounts of Chivay obsidian at Chiripa and Kallamarka. InGiesso's (2000: 167-168)review of lithic evidence from several Formative Period sites he found no obsidian use in these collections except for samples from Khonkho Wankane which recent research directed by John Janusek has revealed to be principally a Late Formative center. The furthest confirmed examples of transport of Chivay obsidian are these examples encountered in southern Titicaca Basin sites. The two furthest confirmed contexts for Chivay obsidian transport are represented bytwo samples from a Late Formative context at Kallamarka (Burger, et al. 2000: 308, 319, 323) and six samples found at Khonkho Wankane (Giesso 2000: 346), both about 325 km from the Chivay source, or 72 hours by the walking model.

At Qillqatani, the obsidian returns to being primarily Chivay type during the Late Formative although 1 out of the 9 obsidian artifacts analyzed from Late Formative levels was of the Tumuku type (SectionX3.4.2X). During the Late Formative there is a steady decline in the percentage of debitage made from obsidian at Qillqatani, a trend that continues in the Tiwanaku period.

In the Colca Valley, there are few traces of the Late Formative consumers of Chivay obsidian. As will be further explored in this research project, there is very little Qaluyu or Pukara material diagnostic to the Middle or Late Formative in the Chivay obsidian source region. Steven Wernke found a diagnostic classic Pukara sherd with a post-fire incised zoomorphic motif that resembles a camelid-foot from a Formative site above the town of Yanque in the Colca (Wernke 2003: 137-138).

Pukara materials are known to have circulated, probably through trade links, throughout the south-central Andes. Pukara sherds have been found in the valley of Arequipa in association with the local Formative Socabaya ceramics (Cardona Rosas 2002: 55). In the Moquegua valley, both Chiripa-related and Pukara pottery have been found (Feldman 1989), and a Pukara textile was found in a possible elite grave context in the Ica Valley (Conklin 1983).

Discussion

The decline in forest cover in the Lake Titicaca Basin beginning during the Early Formative intensified throughout the Formative Period. Between 2500 - 800 cal. BCE a decline in fine particulate charcoal was detected in a lakecore drawn from the southern end of Lake Titicaca's Lago Grande(Paduano, et al. 2003: 274). The pollen and charcoal record indicates thatforest cover finally disappeared around AD0, coincident with population increases and resource pressure associated with Late Formative Period socio-economic intensification.

The obsidian circulation during the Late Formative shows distinct patterns in the Titicaca Basin and in Cusco. In the Titicaca Basin, the diversity of sources is reduced, with virtually all the material coming from Chivay and Alca, and the Alca samples are primarily affiliated with a single period at Taraco. Even at the site of Qillqatani, the diversity is reduced as compared with the previous Middle Formative level (Qillqatani Formative C). Current evidence suggests that the economic circulation was more integrated and that it was probably under some form of control by the dominant regional centers of this time. The furthest confirmed evidence of Chivay obsidian transport is from among Late Formative contexts at Kallamarka and Khonkho Wankane.Yet, diagnostic evidence from Titicaca Basin Late Formative polities in the Colca valley area is scarce. Part of the difficulty in understanding the Formative at the Chivay obsidian source area is that the Formative ceramic sequence in the Arequipa highlands is still being refined.

Discussion and review of "Early Agropastoralists" obsidian

From the Terminal Archaic through the Formative Periods the circulation and consumption of obsidian expanded dramatically. Chivay and Alca obsidian is found in a wide variety of sites from isolated rock shelters to early regional centers, and it appears in consistent quantities that differ considerably from the intermittent nature of regional obsidian supplies in preceding periods. The increased circulation of obsidian is part of a spectrum of changes that began in the Terminal Archaic, but obsidian distribution patterns are particularly useful because they are quantifiable.

In some ways obsidian appears most closely linked to tasks associated with camelid pastoralism, such as shearing and butchering. However, in the central Andean littoral, Quispisisa obsidian is found in density in "cotton preceramic" and Initial Period (i.e., Terminal Archaic and Early Formative) coastal sites where presumably pastoralism, if present, was a very minor part of the economy. Furthermore, in the south-central Andean highlands obsidian is found in pastoralist rock shelters but also in regional centers, and the small discarded obsidian flakes and small projectile points do not appear to have served as adequate shearing implements due to their size.

Social complexity during first part of the "Early Agropastoralists" period is manifested most prominently along the Pacific littoral. In the coastal context of what is now the Department of Lima, in central Peru, monumental architecture dating back to 3000 BCE is well-established at the site of Aspero, and the transition to yet more monumental preceramic construction at sites slightly inland has been documented in recent research(Haas, et al. 2005;Shady Solis, et al. 2001). The shift inland is argued to be related, among other things, to the increased importance of harnessing labor surpluses through the production of cotton for anchovy nets and textiles, and to competitive monument building between elites. Evidence of exchange with highland and Amazonian groups is apparent in the form of tropical feathers and other non-local prestige goods. While these developments occurred over one thousand km to the north of the Chivay area along the Pacific coast, early complex organization is also reported in northern coastal Chile in the Chinchorro II and III traditions spanning the Late Archaic through the Middle Formative(Rivera 1991). A long history of cotton production for nets and textiles, imported wool textiles, elaborate burial traditions, and long distance exchange with the highlands and the Amazon characterize the later Chinchorro tradition. These regional patterns underscore the wide scope of the changes that occurred during the Early Agropastoralists time. It has been argued that the beginning of social inequality in the highland Andes may have been stimulated indirectly by demand for wool from aggrandizers in coastal societies(Aldenderfer 1999).

Returning to evidence from highland obsidian distributions, the widespread circulation of obsidian during the Terminal Archaic and Formative is perhaps best discussed in the terminology used in the historical model of Nuñez and Dillehay(Dillehay and Nuñez 1988;1995 [1979]). While this adaptationalist model has theoretical limitations, it serves as a useful alternative to evolutionary chiefdom models that assign a paramount role to central-places in exchange, despite the decline of central-place models in geography(Smith 1976: 24). Following the Nuñez and Dillehay model, it is possible that what is occurring in the Terminal Archaic and Early Formative is the emergence of regular exchange pattern between Qillqatani and Chivay that took the form of a caravan trade "axis" along the western Cordillera. This exchange pattern is among the earliest systematic and demonstrable cases of the circulation of diffusive items through largely homogenous altiplano terrain (XFigure 3-4X) in the pattern that has also been described as horizontal complementarity or the "Altiplano mode"(Browman 1981). Subsequently, during the Middle and Late Formative in the Titicaca Basin, former "axis settlements" like Qillqatani, and the Western Cordillera axis more generally, became relatively less important in long distance exchange relationships. Regional centers in the Titicaca Basin like Taraco, Pukara, Chiripa, Khonkho Wankane, early Tiwanaku, and other centers, expanded in influence during this period of peer-polity competition. In these circumstances, the acquisition and ceremonial use of exotic goods appears to have been an important part of the competitive strategies of aggrandizers. Other evidence for Titicaca Basin-based exchange dynamics include the possible production of hoes of Incatunahuiri olivine basalt at Camata around 850-650 BCE that were then transported and used in southern Titicaca Basin sites(Bandy 2005: 96;Frye and Steadman 2001). While the nature of the relationship between emerging elites in Titicaca Basin polities and caravan drivers that provided links between settlements throughout the region is difficult to describe with precision, it appears that interregional articulation became considerably more elaborate by the end of the Early Agropastoralist period.

3.5.3.The Late Prehispanic

During the Late Prehispanic period the circulation of Chivay obsidian was subjected to pan-regional forces by expansive Tiwanaku and Wari and again during the Inka period. While obsidian appears to have had fairly consistent use in both the Tiwanaku and Wari domains, the use of obsidian relative to other goods appears to have declined.

/Figs_Ch3/region_LatePrehispanic.jpg

Figure 3-13. Chivay type obsidian distributions during the "Late Prehispanic" time (AD 400 - 1532).

Tiwanaku and Wari (AD400 - 1100)

Following the Titicaca Basin chronology the Tiwanaku period begins with the emergence of the Tiwanaku state during the time that has been called "Tiwanaku IV" beginning around AD400 (Stanish 2003). This is centuries earlier than the start of the Middle Horizon, as defined by the presence of Wari in Ica. Wari in Ica spans the period from approximately AD 750 - 1000.

Obsidian consumption at Tiwanaku

Research by Martín Giesso (2000;2003) on the Tiwanaku lithic industry has revealed patterns in the differential use of lithic material in the region. First, it should be noted that obsidian is not abundant in the Tiwanaku core area in terms of the capital of the expansive state. Giesso (2003: 365-366) notes that obsidian artifacts make up only 0.8% of the lithics by count (n=86), of the collections at Tiwanaku. Yet the Tiwanaku core lies 315 km from the Chivay source in Euclidean distance, or 70 hours by the hiking function and therefore while obsidian is relatively scarce, it has been conveyed a considerable distance. The spatial distribution and the artifact form of obsidian at Tiwanaku are more analytically relevant than the total count or weight of obsidian which was quite small. Intrasite data on the contrasts in temporal and spatial use of obsidian at Tiwanaku are critical because, among other things, these data would provide a gauge of the relative persistence of access to the Chivay source as Tiwanaku's regional influence expanded.

Giesso's comprehensive source sampling showed that ten different types of obsidian were in use in the Tiwanaku heartland, although of these ten samples only four samples derive from known source locations. He notes that of the ten obsidian types, only Chivay (Cotallaulli) type is transparent and rest are described as "opaque". Giesso's spatial assessment of the distribution of lithic production activities in the core region showed that obsidian microdebitage was concentrated at certain mounds and in residential sectors as compared with chert, quartzite, and other locally available products. Giesso also organized a sourcing study of basalt artifacts in the Tiwanaku area, and he examined the contexts of production at the Querimita quarry located on the southwestern shores of Lake Poopó in Bolivia (Giesso 2003: 369) just over 300 km south of Tiwanaku. The evidence from Querimita on the production and consumption of basalt provides an interesting regional contrast to the spatial patterning of Chivay obsidian. There are both Wankarani (Formative) and Tiwanaku sites in the vicinity of the Querimita basalt source in the direction of the shores of Lake Poopó, but diagnostic Tiwanaku materials at the quarry itself are not reported (M. Giesso 2006, pers. comm.).

At Tiwanaku, quantities of obsidian and quartz have been found in the construction fill at the ceremonial Mollo Kontu mound, a structure that has served as a local fertility shrine. Nicole Couture argues that "quartz and obsidian fragments also served as mountain icons, in the way that exotic crystals, minerals, and rock candies are used today by Aymara yatiris, or ritual specialists, to represent mountains and thunder in rites to promote agricultural and social fertility" (Couture 2003: 225). Obsidian from Mollo Kontu has been traced to six different sources: Quispisisa, Chivay, Sora Sora, Cerro Zapaleri, and two unlocated sources (Giesso 2000). "The high density of quartz and obsidian artifacts, often five to ten pieces per excavation level, indicates that they were not accidentally included as part of redeposited rubbish, but rather were deliberately added to the clay fill" (Couture 2003: 215). The diversity in the Tiwanaku urban core was such that Chivay material represented a relatively low percentage (76%) of the material, although the ratio of Chivay material in Giesso and Glascock's study on the whole was 90% Chivay, which is typical for the Titicaca Basin. Interestingly, Giesso found that the 19 obsidian samples from the Akapana and Putuni were all transparent samples from the Chivay source, a pattern supported by the ceramics assemblage from those sites that was entirely local. This pattern suggests to Giesso (2003: 368) that a cultural strategy of "ideological purity" was taking place at that location at Tiwanaku due to exclusive use of those particular materials that are perceived as "local" to the Titicaca Basin. Giesso notes that obsidian is found in the form of standardized projectile points, type 4E according to Klink and Aldenderfer (2005), and that evidence of production is found in commoner residential contexts, whereas finished points are often associated with elite ritual contexts. This suggests to Giesso that labor contributions for some segment of the Tiwanaku heartland commoner population may have taken the form of projectile point manufacture.

A comparison of the circulation and use of Chivay obsidian with Querimita basalt strongly suggests to Giesso that the source areas, production, and transport of both exotic materials were controlled by the Tiwanaku state. If so, research at the Chivay source should reveal some evidence of Tiwanaku materials as it did at Querimita, and perhaps research will reveal state mandated standardization of production activities. However, if no evidence of Tiwanaku presence is found at the Chivay source, then this suggests that the relationship between raw material procurement, long distance transport, and state sponsored activities involved a more nuanced relationship between the state and the peripheral economy.

Burger et al. (2000) also analyzed obsidian from Tiwanaku and in their study, from a collection of 18 samples, all from surface contexts, the results were all of the Chivay type. The original Burger and Asaro (1977) study of Bolivian obsidian revealed that three samples purportly from the site of "Sora Sora" were from the Titicaca Basin source, now known as the Chivay source. At 554 km from the Colca Valley this is the furthest reported transport of Chivay type obsidian to date. However, some doubts have arisen as to the spatial origin of these three samples supposedly from the site of Sora Sora (Burger, Dec 2006, pers. comm.), and given the anomalously high transport distance, these three samples are suspect unless additional supporting data become available. The second furthest reported conveyance of Chivay obsidian is 325 km to the Late Formative contexts of Kallamarka and to Khonkho Wankane in Bolivia, as mentioned above.

Obsidian consumption at Wari

At the imperial capital of Wari, obsidian flakes were likewise found in domestic contexts and have been interpreted as "kitchen waste" as no specialized obsidian production sites were found at the site of Wari (Isbell, et al. 1991: 48; contra ;Stone 1983). In the funerary area of Wari known as the Cheqo Wasi sector the obsidian artifacts include a broken knife, 79 flakes (both utilized and unutilized), and two triangular fragments of obsidian with flat surfaces that may have been polished mirrors (Benavides 1991: 64).

At the Wari site of Pikillacta, in the valley of Cusco, evidence from obsidian flakes found in household compounds suggests that obsidian production took place at the household level. In the Haycuchina and Waska Waskan residential clusters, evidence of craft production consists of "waste flakes" of turquoise-colored stone, obsidian, and broken marine shell all collected from the surface (McEwan 1991: 99). Obsidian flakes were found in middens at Pikillacta along with both utilitarian and elite ceramics.

Qillqatani

In Tiwanaku levels at Qillqatani there is a notable decline in the use of obsidian, and of the two specimens analyzed one was of the local low-quality Aconcahua type and the other was Chivay.XFigure 3-7Xreveals that there is a reduction in the percentage of obsidian in the assemblage from this level. In the "debris" category, obsidian falls below 10% for the first time since the Middle Archaic Period. Ceramics were clearly Tiwanaku influenced, but they were relatively poorly made (Aldenderfer in prep.). One may speculate that with the establishment of Tiwanaku colonies in Moquegua, the dominant caravan traffic patterns shifted to an east-west pattern in the Qillqatani area. It is also possible that if the Tiwanaku economic sphere was dominating the circulation of Chivay obsidian, then the material would probably have been conveyed more directly into the Basin and the western cordillera exchange would have been relatively diminished.

The Colca area during the Tiwanaku Period

Relatively little is known about the Colca valley in time periods prior to the Late Intermediate period. The political context of the Colca during this time is intriguing and somewhat ambiguous because the Colca valley lay on the frontier between Wari and Tiwanaku. External cultural influences in the Colca valley proper during the Tiwanaku times appear to have been entirely from Wari, yet Chivay obsidian is not found in Wari sites except in sites with a Tiwanaku component in Moquegua. Given the predominance of Chivay obsidian at Tiwanaku one might expect to encounter diagnostic Tiwanaku material somewhere close to the Chivay source. A number of scholars have commented on the surprising lack of Tiwanaku material in the Colca area given the obsidian distributions in the Tiwanaku heartland (Brooks 1998: 311-313, 454-459;Brooks, et al. 1997;Burger, et al. 1998: 211-212;Burger, et al. 2000: 340-342;).

The geographically closest Tiwanaku sites to the Colca in the department of Arequipa are found on the south-west edge of the city of Arequipa (85 km due south), with the best evidence coming from the site of Sonqonata (Cardona Rosas 2002: 78-87). The Arequipa highlands are largely unstudied, and the closest known Tiwanaku site to the Chivay source on the altiplano appears to be 147 km to the east in the department of Puno at the site of Maravillas just north of Juliaca (Stanish 2003: 189). Tiwanaku sites in this region are generally associated with lakeside agriculture or are found at cross-roads along major travel corridors (C. Stanish March 2006, pers. comm.), and to date no Tiwanaku sites are known in the largely pastoral periphery of the northwestern Lake Titicaca Basin in the direction of the Chivay source. Tiwanaku pottery has been found in Cuzco at Batan 'Urqo (Glowacki 1996: 245) and a Tiwanaku snuff tablet was found at La Real in the lower Majes valley of coastal Arequipa (García Márquez and Bustamante Montoro 1990: 28), but these appear to have been examples of trade goods along the Tiwanaku-Wari frontier.

In the main Colca valley close to the Chivay source, Wari-related ceramics have been excavated from a site just 4 km downstream from the town of Chivay and approximately 10 km from the Chivay obsidian source. A trench excavation exposed red-slipped wares from a domestic context with Tiwanaku Horizon dates in archaeological work associated with the William Denevan's Río Colca Abandoned Terrace Project at the site of Chijra (Malpass 1987: 61;Malpass and De la Vera Cruz 1986: 209, 216;Malpass and De la Vera Cruz 1990: 44-46, 57). Radiocarbon dates associated with these red-slipped wares came from a hearth in a house terrace and produced dates of 1140 ±80 (WIS-1713; AD680-1030) and 1290±90 (QL-4015; AD600-900) (Malpass 1987: 61). This ceramic style has recently been investigated in detail by Wernke as part of his elaboration of the Colca valley ceramic sequence (Wernke 2003: 466-477). Other Wari-influenced sites in the Colca Valley include a large, recently located site in on the north side of the Colca river named Charasuta close to the town of Lari (Doutriaux 2004: 212-223) and the site of Achachiwa near Cabanaconde (de la Vera Cruz 1987;de la Vera Cruz 1988;Doutriaux 2004: 202-207). Curiously, Chivay obsidian is not known to have circulated in the Wari sphere at all despite the proximity of these sites to the Chivay source.

In the main Colca Valley, indications of Wari ideological and stylistic influence are the strongest evidence for external links during the Tiwanaku Horizon. In addition to Wari influences at Chijra and Charasuta, the site of Achachiwa provides intriguing evidence of exotic obsidian intruding into the Colca valley. Achachiwa is a large site adjacent to the modern town of Cabanaconde that has a large Middle Horizon component that appears to be Wari influenced, as well as components belonging to a the local LIP occupation and an Inka occupation (de la Vera Cruz 1987;de la Vera Cruz 1988;Doutriaux 2004: 202-207). Brooks reports that she collected seven obsidian flakes for analysis from Achachiwa that were visually distinct from obsidian she had encountered elsewhere in the Colca (Brooks 1998: 447). Of these seven flakes, none were of Chivay type obsidian although the site is only 46 km downstream of the source. Her analysis showed that six of the flakes were from the Alca source (96 linear km away) and one was from the Quispisisa source (300 km away), a strongly asymmetrical pattern that is non-the-less consistent with Middle Horizon obsidian distributions in southern Peru.

Discussion

Chivay obsidian distributions during the Tiwanaku period are somewhat of an enigma. As Tiwanaku persisted for longer than Wari one might expect Tiwanaku evidence at the Chivay source either preceding or simultaneous with Wari presence in the Colca valley. However, the nature of procurement and distribution of Chivay obsidian during the Tiwanaku period was such that diagnostic materials from the consumption zone do not appear at the Chivay source. There appears to have been considerable nuance in the relationships between state leaders, corporate integration, and ethnic local kin-based groups during the period of Tiwanaku hegemony. Groups living in the Tiwanaku peripheral areas were perhaps consistent with Dillehay's characterization as"…a patchwork of overlapping, geographically disparate, and apparently politically semi-autonomous core valleys, oases, and plateaus or foci of cultural development, each of which primarily exploited its own immediate peer area"(Dillehay 1993: 247). Along these lines,Stanish (2002: 188) notes that early states appear to have "selectively incorporated certain areas around the basin" rather than attempt comprehensive control. John Janusek summarizes a variety of evidence demonstrating that "Tiwanaku was an incorporativemore than it was a transformativestate, simultaneously employing multiple strategies of regional control and influence" (Janusek 2004: 162), along the lines of corporate political strategies described by Blanton et al.(1996). Accordingly, a direct correlation between the predominance of Chivay obsidian in the Tiwanaku economy and clear, material evidence of incorporation in the archaeological remains of communities in the Chivay source area should not be expected. The political affiliation of Colca valley communities was probably made more complex due to the presence of the Wari frontier during the Middle Horizon.

The asymmetrical, export-only exploitation is consistent with a pattern that has been observed at a number of prehispanic obsidian sources in the Andes. Obsidian sources generally have few diagnostic artifacts or architecture in association with quarrying. It was mentioned earlier that Formative distributions of Alca obsidian were strikingly asymmetric with export to long distance consumption sites but no corresponding diagnostics from those consumers back at the Alca source (Burger, et al. 2000: 314, 323). Similarly, the Quispisisa source has little diagnostic evidence the immediate vicinity linking the source to Wari or any other known group, despite the long history of use of that obsidian type in Wari sites (Burger and Glascock 2000;Burger and Glascock 2002). The asymmetrical nature of Chivay obsidian use during the time of Tiwanaku and Wari is, therefore, consistent with a pattern apparent at raw material sources elsewhere in the Andes where obsidian procurement does not involve reciprocation or discard of diagnostic artifacts from the consumption zone.

Late Intermediate Period (AD1100 - 1476)

Current evidence suggests that following the collapse of the Tiwanaku state, a prolonged drought occurred until AD1200 after which time Aymara chiefdoms emerged in the region referred to as Collasuyu by the Inka. A central question of this time period concerns the extent to which obsidian distributions can reveal whether Tiwanaku period interaction patterns persisted into the LIP in the forms assumed by economic organization and long distance exchange. While the LIP is known as the auca runaor the "time of strife" when fortified hilltop refuges " pukaras"were constructed in abundance in the Titicaca Basin and adjacent territories, the weapons used in these conflicts appear to have primarily been percussion weapons like slings and clubs, and not obsidian tipped-projectiles.

Chivay obsidian during the LIP

In a review that considers the LIP and Late Horizon together in one discussion, Burger et al.(2000)note the declining presence of obsidian in most LIP and LH contexts in the region. Many of the chemically provenienced obsidian samples from the Titicaca Basin come from surface contexts with either LH or LIP associations, and the pattern revealed in the Titicaca Basin from these samples is one of almost exclusive use of Chivay type material. One of the reasons for the strong presence of Chivay obsidian over Alca obsidian may have resulted from cultural affinities between Aymara groups(Browman 1994). The Colla in the North Titicaca Basin and the Collagua in the upper Colca area share many traits including the construction of pukaras(Wernke 2003: 262-263), mortuary features including chulpaburial structures and fiber-wrapped mummy encasings(de la Vega, et al. 2005;Wernke 2003: 225-234), and other commonalities.

In the northern Lake Titicaca Basin, Arkush(2005: 247)notes that two-thirds of the occupied pukaras throughout her survey area contained obsidian on the surface, and that chert flakes and blades are common. In one instance, at Calvario de Asillo (AS1) prepared obsidian cores and a concentration of flakes was encountered. Obsidian projectile points were small, triangular base-notched points (series 5), consistent with the Klink and Aldenderfer point typology for the Terminal Archaic and onwards(Arkush 2005: 709-711). Metal objects are common as well, with copper and copper-alloy pendants and tupus(long pins) frequently found at pukaras.

Six obsidian samples from the mesa-top pukaraand chulpaburial tower complex of Cutimbo were analyzed by Frye et al.(1998)and the samples were 100% from the Chivay source. Burger et al.(2000: 343-344)indicate that one obsidian sample from the renowned chulpacomplex of Sillustani was also from the Chivay source. Burger et al. observe that while a single Alca flake was excavated at Taraco from an LIP context, and another was found on the surface at the LIP or LH site of Kolkeparke near Ayaviri(Burger, et al. 2000: 343), the Titicaca Basin is otherwise entirely supplied from the Chivay source during the LIP.

At the rock shelter of Qillqatani (SectionX3.4.2X) a large percentage (15.4%) of the tools from LIP levels were made from obsidian, although the count is very low (n = 2) and the tools were extremely small. The single obsidian sample that was analyzed proved to be from the Chivay source.

In the south-western Lake Titicaca area, Hyslop(1976: 118-119)found that the Lupaqa sites he encountered in the course of his road system survey contained primarily basalt and quartzite flaked stone. More recent survey work in the Ilave and Huenque drainages(Klink and Aldenderfer 1996)found that high-quality cherts are abundant in the region, which suggests that the prominent use of basalt at the sites Hyslop encountered was by choice, not by necessity, and was perhaps a reflection of lakeshore agricultural activities.

Late Horizon (AD1476-1532)

The expansion of the Inka Empire during the Late Horizon resulted in a restructuring of the long distance movement of goods. The ability of the Inka state to transport stone is vividly demonstrated by the transport of hundreds of andesite ashlars weighing up to 700 kg apiece the 1600 km distance from Cusco to Saraguro, Ecuador (Ogburn 2004;Ogburn 2004). The ability to move people and goods over long distances was an integral part of state apparatus; however, it is evident from obsidian distributions that not all goods were more widely distributed during the Late Horizon.

It appears that if the Inka had a particular demand for a substance it could be acquired from over great distances. The Late Horizon was referred to as the "tin horizon" by Lechtman (1976) due to the lengths that the Inka would go to procure tin for copper production. In the Mantaro Valley, tin was not present in Late Intermediate Period coppers but it was found in all seven copper implements from Late Horizon levels where it made up, on average, 5% of the metal composition (Earle 2001: 311;Owen 2001: Tables 11.1-11.3). This tin is thought to have come from mines in Bolivia or southern Peru, and this exchange was facilitated by the Inka state (Lechtman 1976). Thus tin was transported from southern Peru, perhaps from the vicinity of obsidian rich lands of Arequipa, but apparently little obsidian was transported along those same Inka transportation routes.

Ethnohistoric accounts indicate that during the Late Horizon the control of natural resources sometimes occurred through restricted access to raw material sources in the Andes. Access to tunnels leading to particularly rich gold deposits at Inka gold mining operations were restricted at the tunnel mouths (Burger and Glascock 2002: 364;Sancho de La Hoz 1968 [1534]: ch. XVIII: 332). The deposits of other natural resources, such as obsidian, are distributed across the geological landscape such that directly controlling access would have been difficult or impossible.

Burger et al.(2000: 343-347)report very little Chivay obsidian circulating in contexts that are definitely Late Horizon in date. Small quantities of obsidian were found in Cusco, and these turned out to be entirely of the Alca type except for one notable case of Chivay material at Machu Picchu. In one of the only cases of Chivay obsidian use in the department of Cusco since the Late Formative, several unmodified nodules of Chivay obsidian were identified from a collection of small obsidian pebbles excavated by Hiram Bingham in 1912 at the gateway to Machu Picchu. Burger et al. (2000) note the significance of the unmodified state of these small obsidian pebbles, as the small size of these nodules suggest that it was not the tool-making potential, but rather the natural glass itself that was "suitable in its apparently natural state as an offering or sacred object comparable to quartz crystals"(Burger, et al. 2000: 347;Rowe 1946: 297). This view is consistent with observations about the possible significance of the 'essence' of obsidian as natural glass in Andean cosmology that will be explored in more detailed below.

Discussion and review of "Late Prehispanic" obsidian

Obsidian distributions during the Late Prehispanic period reflect the dramatic changes in economy and ideology imposed by powerful states in the region. In the Tiwanaku and Wari spheres of influence obsidian circulation achieved its largest known extent. Particularly intriguing are the examples of non-local obsidian consumption in close proximity to rival obsidian sources, a phenomenon that occurs twice during the Middle Horizon. Three samples of Chivay obsidian were found close to the Sora Sora obsidian source in western Bolivia in a possibly Tiwanaku context (Burger, et al. 2000: 340), and Quispisisa and Alca obsidian were found at the Wari-influenced site of Achachiwa, only 46 km downstream of the Chivay source, in the lower Colca valley (Brooks 1998: 447). During the LIP this pattern of high mobility and sprawling interregional contact is reversed as regional circulation of obsidian is curtailed and the Chivay regional pattern returns to a distribution confined to the Colca region and the North Titicaca Basin. Finally, during the Inka period, obsidian appears to have had a largely diminished significance, particularly in contrast to other materials that were circulated prodigious distances by the Inka. This discussion of the spatial distributions of prehispanic obsidian will be complemented by a review of the uses and forms that obsidian artifacts assume in the archaeological and ethnohistoric record.

3.6. Obsidian Use in the South-Central Andes

Obsidian was knapped into relatively few artifact forms in the south-central Andes. The most common formal stone implement was a bifacially flaked projectile point, but other bifacial tools such as knives and scrapers were also commonly produced throughout the consumption region. The other major technical class for obsidian artifacts were simple flakes. As a sharp, but fragile, cutting implement, a freshly struck obsidian flake was potentially useful for butchery and wool shearing purposes. Dransart (2002: 108-109) reports that in rural communities modern wool shearing is accomplished either with the lid of a tin can that has been folded over so as not to cut the user's hand, or with a broken piece of bottle glass. Contemporary herders in the Colca region report that obsidian flakes, and sometimes broken glass vessels, are used for castrating animals because it was explained that as non-metal tools do not oxidize, they are less likely to introduce infection into the animal (T. Valdevia 2003, pers. comm.).

A distinctive Andean method of camelid slaughter can be accomplished with a small, sharp flake is described by George Miller (1979: 27-36). The ch'illamethod of slaughter consists of laying the animal down, cutting a small incision near the sternum with a small flake of stone, and reaching in and manually breaking the ascending aorta where it leaves the heart. Ethnoarchaeological studies have shown that simple flakes are often used in butchering and shearing, and therefore a prehistoric association between pastoral facilities and lithic flakes, both utilized and unutilized, seems probable.

3.6.1. Variability in Andean obsidian use

In many regions of the world prehistoric artifacts made from obsidian can be generally classified by whether the principal function is for display or for some utility more directly related to subsistence. In Mesoamerica and the ancient Near East, both areas with complex societies and elaborate stone tool production, obsidian was used to make bowls, vases, eccentrics, seal stamps, statuettes, and tables, as well as items of personal decoration such as labrets, ear spools, necklaces, and pendants (Burger, et al. 1994: 246). Craftspeople also developed efficient, high utility obsidian technologies as well, such as prismatic blades, a form that allows archaeologists to quantify cutting edge to edge-length, error rates, and other efficiency measures.

In the south-central Andes the diversity of artifact forms made from obsidian is relatively low, and it is difficult to differentiate display from utilitarian applications. For example, obsidian projectile points are both sharp and highly visible, suggesting that the points had a display function that underscored their utility as a weapon and as a cutting tool.

Hafting in the Andes

Archaeological evidence shows that bifacially flaked obsidian implements were hafted in a variety of ways. The majority of the hafting evidence from the Andes comes from coastal sites due to superior preservation. Projectile points were hafted to spears, spear thrower darts, and arrow shafts. Obsidian bifacial tools were also hafted to wood or bone handles for use as knives. Hafting materials varied regionally, but hafting was often accomplished using gum or resin, and hafts supported with cotton string have been found in some coastal sites (Carmichael, et al. 1998: 79).

Evidence of use of obsidian

The artifact types predominantly manufactured from obsidian include projectile points and tools for cutting and shearing tasks. While simple flakes are wide-spread and were probably used abundantly for butchering, scraping, and shearing purposes, the utility of obsidian flakes is frequently discounted when flakes are relegated to the "debitage" or "debris" class. In Andean studies bifacially-flaked instruments are the most commonly analyzed obsidian artifact class in archaeological reports from sites in the Andes.

Obsidian artifacts are sometimes found in association with iconographic representations of dark colored artifacts that are similar in appearance to the very obsidian artifacts found in that context (Burger and Asaro 1977: 15). Such is the case with black-tipped darts and knives depicted on Ocucaje 8 through Nasca 6 ceramics and textiles, and obsidian artifacts found in tombs from those contexts. Building on the discussion in Burger and Asaro (1977: 13-18), examples of obsidian artifacts from the south-central Andes follow.



Application

Form

Provenience

Description

Reference

Weapon (probable dart point), conflict

Point

Looted tomb at Hacienda Mosojcancha, Huancavelica.

A point made from obsidian was found embedded deeply in a human lumbar. (Figure C-10).

(Ravines 1967)

Weapon, with spear throwers

Point

Grave 16 at Asia,
(central coast of Peru).

Preceramic context.

Found in association with spear throwers

(Engel 1963: 56;Uhle 1909)

Weapon, hafted

Point

Tombs at Hacienda Ocucaje, Epoch 10.
Early Horizon context.

Points hafted with gum and, in one case, cotton thread to wooden foreshafts.

(Burger and Asaro 1977: 14)

Weapon

Point

Carhua (south coast , Peru)

Point penetrating through arm muscle near humerus (Figure C-1).

(Engel 1966: 212)

Weapon,

dart

Point

Paracas Necropolis

Well-preserved harpoon (Figure C-9).

(Engel 1966: 180c)

Weapon,

dart

Hafted projectile

depiction

Nasca Phase B1 and B2 diagnostic attribute

Phase B has "Atlatl darts (arrows) in series as ornaments" (Figure C-6).

(Carmichael, et al. 1998: 151)

Weapon, poison

Point

Eastern Lake Titicaca

Obsidian is among point types dipped in strong poison from herbs, perhaps curare.

(Cobo 1990 [1653]: 216-217)

Weapon, bow and arrow

Point depiction

Tiwanaku pottery

Archers with bow and black

tipped arrows depicted on a Tiwanaku q'ero(Figure C-4).

(Bennett 1934: 426-459;Posnansky 1957: plate XXa)

Weapon, hunting

Point depiction

Nasca B vessel

Depiction of darts sailing towards a group of camelids .

(Burger and Asaro 1977: 16)

Tool, Ritual

Knife depiction

Nasca B pottery

Black knives associated with taking of trophy heads.

(Burger and Asaro 1977: 15)

Tool, Ritual

Knife depiction

Nasca textiles, Epoch 1

of EIP

Black knives associated with taking of trophy heads.

(D'Harcourt 1962: 110, 112)

Tool, Ritual

Knife, hafted

Early Nasca

Bifacial knife hafted to painted dolphin palate (Figure C-5).

(Disselhoff 1972: 277)

Decorative, Ritual

Mirror fragment

Huancayo, Middle

Horizon 2 context

Fragment of obsidian mirror ground and polished to .4 cm thickness.

(Browman 1970: 86)

Decorative, Ritual

Mirror

Huarmey, Wari

Mirror mounded in carved wooden hand (Figure C-8).

(Lavalle and Perú 1990: 185)

Medical

Obsidian knives with blood-stains.

Cerro Colorado, Paracas

Part of medical kit that also contained a chachalote (sperm whale) tooth knife, bandages, balls of cotton, and thread.

(Tello 1929: 55)

Medical

Chillisaa kala, Aymara for "black flint"

Titicaca Basin

Speculation about tools used for trephination.

(Bandelier 1904;Marino and Gonzales-Portillo 2000)

Table 3-10. Examples of obsidian use in the south-centralAndes (part 1).



Application

Form

Provenience

Description

Reference

Medical,
Ritual

Material used in folk cures

Canchis, Cuzco; and elsewhere

Modern use in folk cures, the stone was believed to have curative powers.

(Burger and Asaro 1977: 17;Cobo 1990 [1653])

Medical,
Ritual

"knives of crystalline stone"

Titicaca Basin ( ?)

Abdominal surgery by "sorcerers"

(Cobo 1990 [1653])

Animal castration

Flakes, unmodified or retouched

Colca

"We use sharp pieces of obsidian or glass to castrate herd animals it doesn't cause infection like rusted metal knives."

T. Valdevia 2003, Pers. Comm. (my translation).

Shearing

Flakes, unmodified or retouched

Andes

"Aboriginal shearing required special implements, perhaps obsidian knives." Some modern pastoralists use broken glass and tin lids for shearing.

(Dransart 2002: 108-109;Gilmore 1950: 446)

Table 3-11. Examples of obsidian use in the south-centralAndes (part 2).

Many of these examples are shown in Appendix C. of this volume. While the diversity of artifact forms was relatively low, it is evident that the visual and fracture properties of obsidian were relevant to the tools that were used from this material. Further north in the Andes, in Ecuador, a greater percentage of obsidian artifacts seems to have filled a primarily decorative role, including abead, an ear spool, and three polished mirrors(Burger, et al. 1994: 246).

The evidence of use lithics may also take the form of cutting and scraping marks on faunal remains, but the lithic material type can rarely be established by these means. The continued use of glass and obsidian in modern contexts for shearing, butchery, and castration suggests that the prehispanic metals such as copper and bronze did not displace obsidian and other lithic materials for utilitarian tasks. Prehispanic metals were used largely for display, although metals were used in some weaponry, such as mace-heads (Lechtman 1984).

Obsidian in warfare

The evidence for obsidian use in conflict in the Andes comes from a variety of archaeological and ethnohistorical sources, but it is primarily in the form of indirect evidence. The data reveal a great expansion in the production of small obsidian projectile points after the onset of a food producing economy. The majority of Late Horizon weapons that the Spanish faced during their invasion appear to have been bola stones and percussion weapons like maces and slings(Cahlander, et al. 1980;Korfmann 1973), as well as padded armor.

Archaeologists working in the south-central Andes vary historically in their assessment of the use of projectiles in the highland region(Giesso 2000: 43). Bennett(1946: 23)asserts that the use of bow and arrow were not important on the altiplano, Metraux(1946: 244-245)says that spear throwers were in use, while Kidder(1956: 138)indicates that evidence from projectiles show that arrows were widely used at Tiwanaku. Of the projectile points analyzed by Giesso from Tiwanaku, 19% were made from obsidian, with the highest concentration of points coming from excavations in the civic/ceremonial core at the Akapana East, K'karaña, and Mollo Kontu mounds(Giesso 2000: 228-238).

In a dramatic example of obsidian use as a weapon, one archaeological findError! Reference source not found.shows a probable spear or dart point penetrated the victim's abdomen from the left anterior abdomen and lodged on the anterior side of the lumbar vertebra.

Evidence of use of the bow and arrow in warfare

One of the strongest patterns in obsidian distributions in the south-central Andes is the sudden onset of obsidian use with series 5 projectile points in the Terminal Archaic, after 3300 cal BCE In defining the small type 5D projectile points Klink and Aldenderfer(2005: 54)suggest that the widespread adoption of these point types may reflect the use of bow and arrow technology, as these points fallwithin the size range described by Shott (1997: Table 2) as associated with arrow points (although with substantial overlap with the size of the smaller dart points).

/misc/image011.jpg

Figure 3-14. Type 5d projectile points from a Terminal Archaic level at Asana and Early Formative levels at Qillqatani, from illustration in Klink and Aldenderfer (2005: 49).

Additional metrics for differentiating arrowheads from dart and spear points have been suggested by Thomas (1978: 470) and Patterson (1985). With Mesoamerican projectile points, Aoyama (2005: 297) has found a very significant correlation between evidence from microwear analysis and projectile dimensions in his effort to differentiate arrowheads from dart and spear points. The relationship between the mass of the item and the velocity and distance of the projectile are discussed in detail by Hughes (1998). She notes that the innovation of fletching allowed for balanced projectiles that had smaller projectile tips and lighter shaft materials, which in turn permitted greater distance and velocity in weapons systems (Hughes 1998). On a related point, the notched base in all type 5 projectile styles reflects a change in hafting technology. Greater impact loads can be absorbed by mounting projectiles tips with notched bases into slotted hafts (Hughes 1998: 367;Van Buren 1974).

Obsidian has a number of attributes that make it a particularly effective material for projectile point tips, most notably are the predictableconchoidal fracture qualities of obsidian that allow it to be pressure flaked into very small projectile points. The sharpness of freshly knapped obsidian margins is unrivalled, andethnographicallyobsidian is renowned as a brittle material that can fracture on impact, causing the fragments to produce greater bleeding in the victim of an obsidian projectile (Ellis 1997: 47-53). In materials science, obsidian has extremely low compression strength because it has no crystalline structure (Hughes 1998: 372). Due to the brittleness or lack of compression strength, missed shots can result in many broken obsidian projectile tips, and perhaps for this reason obsidian was not the predominant material for projectile point production until after the domestication of camelids when hunting was decreasingly the primary source of meat (as inferred from the ratio of camelid to deer bone in excavated assemblages).

Increased velocity and distance is possible with the use of bow and arrow, and obsidian has greater penetrating power, although lower durability, than other material types. Do these changes suggest a greater use of projectiles for warfare? Greater use of small, fletched projectiles might be expected in contexts where the individual needs (1) increased distance from the target (2) increased velocity, perhaps due to change in prey or to the use of padded armor. The use of obsidian may increase in contexts where one needs (1) a material that can be knapped into small, light projectiles as discussed above, (2) increased penetration with sharper material, (3) decreased concern for durability because in warfare the weapon will perhaps be retrieved by the opponent.

Although empirical evidence for warfare is scarce in the Terminal Archaic in the highlands, a study of 144 Chinchorro individuals from Terminal Archaic contexts in extreme northern Chile found that one third of the adults suffered from anterior cranial fractures that probably resulted from interpersonal violence, and men were three times more likely than women to have these wounds (Standen and Arriaza 2000). These wounds appear to have been caused by percussion weapons like slings, but the coastal evidence for interpersonal violence in this time period is strong. Additional support for the introduction of the bow and arrow in the Terminal Archaic come from the region among Chinchorro burials in coastal Northern Chile. Bows among the Chinchorro grave goods date to circa 3700 - 1100 cal. BCE, or the Terminal Archaic and Early Formative (Bittmann and Munizaga 1979;Rivera 1991;Standen 2003).

Possible use of poisons on projectile tips

A potential explanation of widespread adoption of the small type 5D projectile points is the greater availability of poisons that reduced the need for heavy, destructive projectiles.South American poison arrows are usually quite small and such arrows are frequently tipped only with a sharpened wooden point. According to Ellis(1997: 55),virtually all ethnographic examples of arrow use include some kind of poison applied to the arrow in order to have either a toxic or a septic effect on the victim. Ellis observes that due to the great variety of substances used to create such toxins, in many regions of study these substances would probably contaminate any chemical attempt to use residue analysis to differentiate the types of poisons used, or even the prey that was hunted, with a particular used projectile point.

A variety ofhighly effective poisons are applied to the tips of projectiles by hunters in the Amazon Basin today(Heath and Chiara 1977).In the prehispanic Andean highlands, trade contacts with the Amazonian lowlands to the east may have made available poison concoctionsfor application to projectiles, most notoriously the fast acting paralysis alkaloid curareprepared from the vine Chondrodendron tomentosum(Casarett, et al. 1996).

Bernabé Cobo (1990 [1653]: 216-217) discusses the use of bow and arrow with poisons by "expert marksmen" in his chapter on warfare. Hedescribes the widespread and expert use of the sling for warfare, which is consistent with reports elsewhere on the use of slings, but then he states that bow and arrow were more significant in warfare.

The most widespread weapon of all the Indies, not only in war but also in the hunt, was the bow and arrow. Their bows were made as tall and even taller than a man, and some of them were eight or ten palms long, of a certain black palm called chontawhose wood is very heavy and tough; the cord was made of animal tendons, cabuya, or some other strong material; the arrows, of a light material such as rushes, reeds, or cane, or other sticks just as light, with the tip and point of chontaor some other tough, barbed wood, bone, or animal tooth, obsidian point, or fish spine.
Many used poisoned arrows, their points anointed with a strong poison; but, among the nations of this realm, only the Chunchosused this poisonous herb on their arrows, and it was not a simple herb, but a mixture of various poisonous herbs and vermin; and it was so effective and deadly that anyone hit by one of those poisoned arrows who shed blood, even though it might be no more than the blood resulting from the prick of a needle, died raving and making frightful grimaces(Cobo 1990 [1653]: 216-217).

The Chunchos ethnic group is described as living in the "forests east of Lake Titicaca on the border of the Inca Empire".It is possible that Amazonian poisons became available in altiplano during the Terminal Archaic due to expanding exchange networks, together with Amazonian hallucinogenics and other lowland products, dramatically altering the efficacy of arrows. As a sharp but lightweight weapon, obsidian tipped arrows would have represented an effective poison delivery system to animal and human victims alike.

While there is no direct evidence for the use of Amazonian poisons on the altiplano, the dramatic change in projectile technology with the type 5D type, co-eval with expanding exchange networks in the region, suggest that a new technology for weapons systems, such as poisons and new shaft and fletching materials, may have influenced the design of projectiles in this time.The transitional economic context of the Terminal Archaic involved many changes, including shifts in both food production and interregional exchange, and the technology of lithic production show significant alterations that correspond to this period.

Multiethnic access to geological sources in the Andes

Salt procurement in the Amazon basin on the eastern flanks of the Andes provides an example of multiethnic access to a raw material source. During annual voyages to the Chanchamayo salt quarry ten days outside of their territory, the Asháninka of the Gran Pajonal (Ucayali, Peru) combine salt procurement with exchange with neighboring groups who share access to the source(Varese 2002: 33-35). Arturo Wertheman, a missionary who traveled through the region in 1876, explains

Throughout the year small bands of Asháninka traders traveled [the Gran Pajonal] paths to obtain salt, carrying with them tunics or ceramics to exchange for other items and for hospitality. With them traveled their traditions, their hopes, and the information of interest to their society. The Pajonal, the vast center of Campa territory not yet invaded by whites, appears to have been the center of culture and tradition through which the Indians journeyed, like a constant flow of life through their very society(Varese 2002: 120)

Other ethnographic examples of multiethnic access to salt quarries are mentioned in the ethnographic literature. Oberem(1985 [1974]: 353-354)describes access by both the Quijo and the Canelo people to a large rock salt quarry located on the Huallaga River, a tributary to the Amazon lying further north on the border of Peru and Ecuador.

In the Cotahuasi valley of Arequipa, to the north of the Colca valley, the rock salt mine of Warwa [Huarhua] has been exploited since Archaic times(Jennings 2002: 217-218, 247-251, 564-566). Access to the Warwa quarry, which lies near the border of the departments of Arequipa, Ayacucho, and Apurímac, is described by Concha Contreras(1975: 74-76)as including caravan drivers from all three of those neighboring departments.

El primer viaje lo hacen, mayormente, en el mes de abril. En esta época cientos de pastores se concentran en esta mina. El camino es estrecho y accidentado hasta llegar a la misma bocamina…Desde el fondo de la mina los pastores cargan a la espalda la cantidad de sal que necesitan llevar, de tal manera que hacen muchos viajes al socavón de la mina. En todo este tramo tardan 4 días, porque después de la mina, siguen cargando en la espalda hasta una distancia de aproximadamente cinco kilómetros, donde quedaron las llamas pastando, puesto que hasta la misma mina no pueden entrar juntamente con sus llamas(Concha Contreras 1975: 74).

There appear to be a number of protocols associated with salt acquisition at Warwa. Notably, the multiethnic visits by caravanners from various departments coincide in April despite the tight working quarters at the salt mine. Furthermore, there seems to be concern for impacts in the much-visited the mine area itself. The cargo animals are actually grazing a distance from the quarry and humans are obliged to carry the loads to these areas instead of attempting to load the animals close to the mine. These cases of multiethnic raw material access provide examples of the social and institutional nature of access to unique geological sources, and the focalized attention that these source receive from surrounding ethnic groups. The analogy with regional acquisition and diffusive transport of the raw material is perhaps the closest modern analogy to the nature of prehispanic obsidian quarrying that remains in the Andes.

3.6.2. Symbolic significance of obsidian

Inferring the symbolic significance of obsidian in the prehispanic Andes involves approaching the topic from a several lines of evidence because there are few direct indicators of valuation and symbolic meaning of obsidian in the Andes. Below, these issues are raised as a series of questions that may, or may not, be answered with archaeological evidence from the region:

(1) How did the symbolic importance of obsidian change through prehistory?

(2) Do changes in use of obsidian reflect local geographic availability?

(3) Did the visual properties of obsidian resonate with Andean aesthetic traditions?

(4) Was there social significance in the visible differences between major obsidian types?

Part of the difficulty in inferring the symbolic properties of obsidian is the result of a relative reduction in the importance of obsidian during the Late Horizon and Colonial period, to judge from ethnohistoric accounts. Ethnohistoric sources are a prime source of information, although through Spanish eyes, of cultural significance and symbolic value in the prehispanic Andean world. Despite the apparent decline of obsidian use in the late Prehispanic and Colonial periods, archaeological distributions suggest that obsidian had substantial cultural and ceremonial associations in particular contexts during in Prehispanic times.

How did the social and symbolic importance of obsidian change through time and space?

Obsidian flakes and bifacial tools were sometimes included as grave goods from the Terminal Archaic (3300 cal BCE) and onwards. Ethnohistoric accounts and archaeological evidence indicate that obsidian "knives" were employed in ritual practice as well as in medical procedures throughout the region, although it is not evident if these were bifacially-flaked tools or simply freshly struck flakes. In addition, concentrations of flakes of quartz as well as obsidian from a variety of non-local sources were found in the Tiwanaku ceremonial mound of Mollo Kontu in a regular pattern that was interpreted by Couture (2003) to indicate deliberate inclusion in clay fill rather than accidental redeposition.

Based on analogy from other goods, it is possible that in areas where obsidian is used in ceremonial contexts a finely-made obsidian implement may have been less likely to have been exchanged as an alienable product traded in barter. In some historical contexts, a ritual item may become a singular item and acquire a distinctive 'genealogy' related to its social history that precludes it from being bartered and circulated in equivalence for something ordinary such as a sack of potatoes (Hodder 1982;Kopytoff 1986).

However, on the whole, obsidian appears to be one of a number of products that escapes easy classification. On one hand, obsidian circulated in established networks controlled by llama caravan drivers who were responsible for the distribution of a variety of goods on the altiplano. These caravans appear to have been organized by households and derived communities of relatively humble means, to judge from other aspects of their material lives excavated at sites like Qillqatani. On the other hand, as a luminescent material with irregular spatial availability, obsidian had properties that qualify it as a "prestige good" in many contexts (Hayden 1998). In terms of rituals organized and performed on a local level, obsidian was probably one of a number of materials that were traditional but non-the-less slightly rare like ochre and shell. Craig (2005: 683-693) explores the evidence for the symbolic importance of discrete color groups in ritual items like ochre and obsidian in the Titicaca Basin Archaic and Formative site of Jiskairumoko by examining the patterning of obsidian and ochre in comparison to what were common-place, functional applications of these goods.

From the perspective of early aggrandizers in the Terminal Archaic and Formative Lake Titicaca Basin, obsidian probably represented a somewhat elusive product for labor investment and therefore was not a principal substance used in elite strategies by the time of the Middle and Late Formative. From the perspective of prestige technologies serving the agenda of aggrandizers by locking up surplus labor (Hayden 1998), the issue might be considered in terms of the following factors:

(1) Obsidian was relatively easy to acquire, perhaps too easily acquired, for the people living in the highlands of Arequipa; and it was not concentrated in specific, controllable points accessed only through mine shafts.

(2) Obsidian has functional properties that assured its continued circulation for the production of cutting and piercing implements, making it insufficiently rare to have served as a "preciosity" (Clark and Blake 1994;Goldstein 2000).

(3) While craft specialization was fostered in more closely-controllable contexts in regional centers during the Formative, obsidian sources were perhaps sufficiently distant to have escaped this kind of specialized craft production. In contrast, consider the various non-utilitarian obsidian products made by craftspeople in Mesoamerica, discussed above.

(4) As obsidian was primarily made into projectile point tips, it was perhaps associated with hunting, or the threat of violence. In some regional traditions, such as in Nasca iconography, obsidian appears to have been associated with trophy head taking. However, in the south-central Andean highlands there is no evidence of a direct association between violence, emerging leadership, and control of surpluses along the lines of Hayden's (1995) despotic leader model.

Obsidian products, particularly finely-made bifacial tools, were perhaps one of a series of items that served to differentiate status-seeking individuals in early transegalitarian contexts, but with later crafts investment obsidian appears to have been assigned a relatively specific role for projectile point production and for cutting implements.

Do changes in use of obsidian reflect local geographic availability?

One way to examine the social and symbolic significance of obsidian is to examine the changing use of the material as availability declines with distance from the geological source. To judge from the distant consumption zone, obsidian was used in both mundane contexts and also in ritual or ceremonial contexts. At Tiwanaku, obsidian was found dispersed into the fill of ceremonial mounds (Couture 2003). It is also sometimes found in Titicaca Basin burials as of the Terminal Archaic. As reported by Craig (2005: 570-574, 679-682), at Jiskairumoko and at other sites in the Ilave valley, obsidian has been found in burials dating to 3300 cal BCE and later along with other non-local goods including lapis lazuli (sodalite) beads, gold discs, and gold beads, as well as ritual items like a camelid effigy made from bone. If the associations of obsidian with ritual power were related to its non-local origins, one might expect this pattern to have been weaker in closer proximity to the Chivay source. Items from ritual contexts, such as grave goods from the Colca Valley, and in close proximity to the Chivay obsidian source, may provide data to test this hypothesis.

Do the unusual visual qualities of obsidian resonate with Andean traditions?

Andean traditions place a priority on visual attributes, and links between the visual purity of a material and its essence have been widely noted in some segments of the Andean literature. As a natural glass, the aqueous properties of obsidian cause the material to reflect light which may display the workmanship of obsidian artifacts, as well as the potential sharpness of obsidian tools. Obsidian used for prehistoric tool production was often a homogeneous glass that was visibly consistent, pure in color, and sometimes transparent or banded. The importance of visual qualities of metals in the Andes has received greater attention:

The social arena in which metallurgy received its greatest stimulus in the Andes was the arena dominated by status and political display. An underlying cultural value system that appears to have strongly influenced the visual manifestation of status and power was a color symbolism oriented around the colors of silver and of gold. The most innovative and interesting aspects of Andean metallurgy arose from attempts by Andean metalsmiths to produce metallic gold and metallic silver surfaces on metal objects that were made of neither metal (Lechtman 1984: 15).

In Andean metallurgy, the appearance of consistency in color, reflectivity and material was prioritized because visual characteristics conveyed information about the inherent essence and animation of the object (Lechtman 1984: 33-36).

There is little direct support in Andean archaeological or ethnohistorical sources for inference regarding how obsidian was perceived, but it would be consistent for obsidian, a stone with the appearance of watery luster, to be associated with ceremonial power and ritual sacrifice given the well-demonstrated importance of stone and water in Andean cosmology.

Were the visual differences between obsidian types in the Andes important?

Visible differences between obsidian chemical types in the region are principally in terms of nodule size, fracture characteristics, glass color, and cortex. While nodule size and fracture characteristics are believed to have been important determining factors in explaining which obsidian types were circulated widely in prehistory (Burger, et al. 2000: 348), glass color is more of visual aesthetic issue. The obsidian sources in the south-central Andes are predominantly black or grey. Obsidian from the Chivay source is often a transparent grey and banded, and Quispisisa obsidian sometimes has a red coloration, although Burger et al. (2000: 314) state that the Quispisisa type is visually indistinguishable from Alca obsidian. As with many goods with discrete places of origin, these visual differences communicate information about the spatial origin of the stone that would have visually linked the material with regions and socio-political groups to knowledgeable viewers. To Brooks (1998: 452;1997) the transparency of Chivay obsidian was a reason for its wide circulation in prehistory. Others have commented on the transparency of the material including Giesso (2000;2003), and Burger et al (2000: 296). Giesso (2003: 368) observes that archaeological and ethnographic evidence from the Andes indicate that "transparent elements were viewed as mediators between different cosmological worlds". Further study may permit evaluation of observed patterns in the contexts of obsidian use that are linked to color.

Discussion

The social and ritual significance of obsidian in particular prehispanic Andean contexts appears to have varied across time and space, and with further research in the region into evidence of production and consumption these differences will be better understood. Archaeologists have established that the visual attributes of particular materials like metal were important in the late Prehispanic Andes. If this mode of interpretation may be extended, high quality obsidian shares some visual characteristics with metal such as shininess and an appearance of material purity. Furthermore, obsidian from particular areas was often visually distinct and this may have conveyed information in regions such as Moquegua where a variety of obsidian types have been encountered. Obsidian was irregularly available across the landscape, and the mere possession of this highly visible material in obsidian-poor regions had possible social significance because it suggested that the holder participated in long distance exchange networks or had alliances with groups in obsidian-rich areas.

3.7. Models for the Procurement and Circulation of Chivay Obsidian in Prehistory

A number of models have been presented for regional interaction and exchange in the south-central Andes (Bandy 2005;Browman 1974;Browman 1980;Browman 1981;Burger and Asaro 1978: 68-70;Dillehay and Nuñez 1988;Nuñez and Dillehay 1995 [1979];Stanish 2003). This study evaluates a selection of these models at the Chivay obsidian source in the highlands of Arequipa. Renfrew explored various configurations for interaction and his "exchange modes" are reviewed above in Chapter 2 (Figure 2-2). These models will be discussed with respect to activity at the Chivay source area and material expectations for what may result from each model in the vicinity of the obsidian source. It should be noted that due to the extremely thin cortex on many Chivay obsidian nodules, decortication is not a consistently useful measure of reduction level or labor investment, but none-the-less the Upper Colca lab analysis sought to measure percentage of remaining cortex on flaked stone artifacts. When the geological cortex is of the extremely thin variety, it is sometimes left on the face of tools and it does not pose an obstacle to knapping. As cortical flakes from obsidian with a thin cortex are often smooth, and can be equally sharp, one should therefore not assume that nodules will be decorticated in the quarry area.

A number of modes of procurement are explored here, but these acquisition and exchange modes are not mutually exclusive either in time or space. That is, a variety of processes were likely to have been occurring simultaneously. For example, a independent caravan have transported an obsidian nodule to a site in the Ilave river valley, and then obsidian nodule may have been transmitted through down-the-line trade from the Ilave area to the Tiwanaku area. These models, therefore, will focus specifically on procurement and initial transport from the Chivay source because that is where more direct material correlates for these different models can be expected.

3.7.1. Direct acquisition Model

Description

Direct acquisition by the end user entails high mobility and multiethnic access to the Chivay source. In this model, those traveling to the source would procure only sufficient obsidian for their household or community needs and no more. The evidence of procurement at the source would be the direct impacts of communities of consumers where obsidian was perhaps circulated in a context of generalized reciprocity but specifically not exchanged against goods in truequebarter, as that is a different type of procurement. Thus, this category consists of direct, personal visits by the immediate consuming household throughout prehistory.

Model

Exchange Mode

Description

Material Correlates

Direct Access

None

Personal (household) procurement through visits to the source. No exchange.

High variability in procurement, advanced reduction, low density production. Possible presence of discarded non-local low-value materials, and non-local temporary architecture.

Down-the-line

Reciprocity, including barter

Local procurement supplying regional demand through exchange. Barter relationships, delayed reciprocity, and other arrangements between neighbors may have been reciprocated with obsidian.

Low variability in procurement, medium reduction, low to medium density of procurement and reduction. Local debris and architecture at procurement area. Non-local portable objects brought into region possible from reciprocation.

Caravans, Independent

Freelance or non-market central places, reciprocity and barter.

Household organized caravans transporting a variety of goods near the Colca. Procurement and transport of obsidian as one of these goods.

Production associated with pastoral facilities, dense processing activity that is moderately systematic. Small triangular proj. point production, possible evidence from variety of non-local goods and architecture.

Caravans, Administered

Emissary, colonial enclave, entrepôt. Redistribution, barter.

Elite commissioned and delegated caravans transporting goods near the Colca. Obsidian procurement for elite use at regional center.

Production with pastoral facilities. Systematic reduction by part time specialists. Possibly non-local elite-related material. Possible evidence of control of source.

Table 3-12. Models of procurement and exchange for Chivay obsidian. Compare terms with those used in Figure 2-2 and Figure 3-3.

Based on low population densities during the Early Holocene it is assumed that the earliest regional consumers of Chivay obsidian in the south-central Andes, the residents of Asana circa 9400 BP, acquired obsidian directly. Direct household acquisition of resources, such as salt procurement, persists to this day in a few places in the Andean highlands. It is possible that the multi-ethnic nature of access to Andean salt mines may serve as a model for procurement that occurred at Chivay during much of prehistory.

As it is generally the altiplano pastoralists that possess the major means of transport, llama pack animals, and direct household procurement by such groups was perhaps common for many types of goods. For example, herders from the community of Paratía (Flores Ochoa 1968: 87-109), to the north-west of Juliaca in the department of Puno, made regular trips to the Colca valley to acquire agricultural goods. If Paratíans traveled directly, it is likely that they used the Quebrada Escalera route passing to the north of Nevado Huarancante which would pass only a few kilometers from the Maymeja area of the Chivay source. On the return voyage from the Colca, if llamas were not overly burdened, a special stop could have been made to acquire a few nodules of obsidian. Similarly, the household-organized llama caravans described by Nielsen (2001) could well have obtained lithic raw material for household consumption if their travel route passed adjacent to a scarce raw material source on a return journey.

This mode of transport requires that consuming households had the social and physical means to travel to the obsidian source. The ability to partake in regional transport, even in a logistical fashion described by Flores Ochoa and by Nielsen, required strong animals, the food security to undergo a risky journey, peaceful conditions and personal security on the isolated travel routes, probable social relationships or contracts with communities encountered along the way, and knowledge of the extraction source area. Conditions of multiethnic access were likely variable in prehistory and knowledge of travel routes and sources of raw material were best obtained through cordial relations with local groups. In short, this means of obsidian distribution requires relatively cosmopolitan travelers with the resources to personally visit the sources of goods needed by the household. This mode of transport is relatively inefficient because individual households or communities sharing resources, by definition, have to personally acquire and produce obsidian in this model of obsidian procurement.

Material Expectations

In the immediate vicinity of the obsidian source archaeological evidence of the direct acquisition mode would produce high variability in procurement, but relatively small quantities in production because consumption would be limited to the households of those visiting the source. Greater amounts of advanced reduction evidence, including bifacial thinning flakes and tools, broken and discarded during manufacture, will be in evidence at the source area. This type of procurement would have the greatest chance of resulting in diagnostic projectile points in the area as advanced reduction and potential discard in the area of the source is expected. When temporal control is available, primarily from datable organic material in excavated contexts, direct acquisition should result in irregular visits to the source area based on household need. In excavation units cultural material will likely be low-density, as soil and perhaps ash from adjacent active volcanoes will have time to accumulate as visitation rates are low.

There is some chance that undecorated, non-local pottery might be found in association with quarrying or with adjacent rest area bofedal zones. For example, Formative Period ceramics from the southern Lake Titicaca are typically fiber tempered, but in the north basin and in the Colca they are grit tempered. Evidence of fiber tempered plainware may have resulted from discard during direct access procurement by southern Titicaca Basin visitors. Other temper and paste characteristics may serve to identify non-local pottery. Alternately, non-local pottery may result from reciprocation activities (i.e., the Down-the-line model), but the notion here is that non-decorated, utilitarian pottery may have been too crude to have served as barter goods.

3.7.2. Multiple Reciprocal Exchanges (Down the line) Model

Description

The multiple reciprocal exchanges model involves the direct acquisition of obsidian by local people for the express purpose of transferring the obsidian to neighboring communities in exchange for other items. In this model, people residing in Chivay area, perhaps no more than one day's travel from the source, procure material and transport it to a location where the obsidian is then exchanged with neighbors.

This kind of procurement resembles Renfrew's (1975: 520) "Reciprocity" and the "Down-the-line" modes (see Figure 2-2), where goods are exchanged with neighboring groups of roughly equal status through a variety of configurations that are essentially reciprocal in some form. While a synchronic barter of obsidian for, say, a chunk of ochre from the neighboring region is easiest to describe conceptually, reciprocity relationships are a manifestation of a wide range of social mechanisms (Section 2.2.2). In the context of the Chivay area, reciprocal exchange arrangements such as barter for other products, or for grazing rights, for labor, or for social functions such as bridewealth, may be evident. Furthermore, delayed reciprocal arrangements between neighbors are extremely common. Evidence of Down-the-line exchange may be encountered in a wide variety of socio-political contexts from Archaic Foragers to agro-pastoralists living on the periphery of states during the Late Prehispanic. If demand for a product is sufficiently high, archaeological evidence may be encountered of individuals devoting themselves to procurement in order to satisfy regional demand, but substantial quantities of goods would have to be reciprocated because the households that are sponsoring this increased procurement would have fewer provisioners working to bring subsistence to the household. Thus, if down-the-line demand is sufficiently great then the households devoting themselves to elevated rates of procurement would need to barter products for subsistence goods. In the Andes, the pattern observed from items like salt (Concha Contreras 1975: 74-76;Nielsen 2000) suggests that in modern circumstances when demand, and barter values, are sufficiently high, then the down-the-line network may be simply by-passed in favor of procurement through personal or caravan acquisition. Caravans from the consuming zone or adjacent highland areas will make the journey to procure the material and transport it for household use and for barter (a combination of the Direct Access model and the Independent Caravans model).

In a functioning Down-the-line system the flow of information is also important. The changes in the regional demand for a product like obsidian can return to the source area procurers through direct requests, or it might be reflected in increased barter value in a market context. Information exchange may also return specific demands from consumers as to the size, form, or quality of the source material. The temporal regularity of down-the-line reciprocal trade may also be quite variable, as down-the-line networks may dwindle and then be revived during a seasonal gathering or ceremonial occasion. Reciprocal relationships can take the form of mutualism and buffering, they may result from a need to complement the resources on a neighbor's territory, and they often present opportunities for ambitious individuals to advance their interests through differential access to non-local goods.

Material Expectations

Down-the-line procurement involves local people visiting the Chivay source and acquiring goods to supply the reciprocal exchange network, however large that it may be. In the quarry area one should expect local visitors, and therefore local styles in both discarded materials and local architecture. Procurement may take place in the context of embedded economic activities, such as hunting forays into the high country or pasturing of camelids in the rich bofedal adjacent to the source.

If reciprocation for obsidian takes the form of portable objects, such as non-local ceramics, one may encounter diagnostic, non-local goods in the communities adjacent to the obsidian source. These may be in the form of styles belonging to neighboring communities, or more exotic styles may be found on non-local goods that could have arrived through the exchange network from even more distant areas. There is a high likelihood that reciprocation for obsidian would have taken place in other forms as well: goods that are perishable, labor, or other assets that are otherwise less easy to detect.

If locals are involved obsidian procurement at the Chivay source one may also find that the large nodules available at the Chivay source are used in the local economy as well. That is, if nodules in the Chivay source are up to 30cm in length then large flakes, either cortical or non-cortical, may be expected to have been discarded in residential contexts in local communities. If large nodules are available then those that are not exchanged with reciprocal partners are put to use for local needs. Thus when large cores and flakes are procured in the Chivay source area, then appropriately large flakes should be discarded in the middens of communities in the adjacent consumption zone in the upper Colca.

Procurement and initial reduction at the source will have relatively low variability because it is conducted by the same local methods. Local people will have better knowledge of high quality extraction loci and perhaps there is lower variability in procurement locations as a result. As reciprocity networks, and particularly barter arrangements, are contingent on visual attributes of bartered items one should expect medium reduction of material at the source or in the adjacent communities. At the very least, nodules will be partially decorticated and an initial strike that provides entry into the core should be expected, as this serves to expose the quality of material on the interior to barter partners. Furthermore, if transport does not involve camelid cargo animals (because reciprocity is either taking place pre-domestication during the Archaic Forager period, or otherwise does not involve camelids) one might expect a greater concern for the weight of the nodules and therefore further reduction in the vicinity of the obsidian source.

More advanced reduction may also be expected as it minimizes risk and waste by producing blanks, preforms, and prepared cores in the vicinity of the source where obsidian is abundant. However, according to the Down-the-line model producers have the greatest social distance, and therefore the least information, about their consumers. Advanced reduction limits the possible forms that artifacts may take, and therefore producers would need to know what kinds of tools consumers were planning to produce in order to move beyond initial stages of reduction. Thus medium level reduction might be expected, but not an abundance of advanced reduction at the Chivay obsidian source.

3.7.3. Independent Caravans Model

Description

Long distance transport of goods by way of camelid caravans was well established in the prehispanic Andes. The strongest evidence for the importance of caravan transport comes in the form of ethnohistoric and ethnoarchaeological studies described earlier in this chapter, however archaeological evidence of caravan traffic is usually very light and it often requires inference from indirect evidence. The "Independent Caravans Model" described here consists of caravans organized on the household level, although ethnographic studies show that, in practice, the members of several households will often band together for company and for safety while participating in long distance caravans. It is worth pointing out that independent long distance transport does not necessarily involve cargo animals. It is possible that small quantities of obsidian were carried bytraveling peddlers. As a variant to this model, one should consider that peddlers carrying small portable items, mostly cultural goods like herbs, shell, feathers, but potentially small obsidian tools or cores, could have circulated objects widely without the assistance of llamas.

According to this model, a household with a sufficient number of cargo animals, usually castrated male llamas, will initiate a trade caravan by transporting goods that they expect will be in demand, to regions that they anticipate will have complementary goods to offer them. According to some descriptions, caravans are pursuing a directed acquisition of specific goods and then they return directly home, while other models describe entire circuits where herders acquire goods, travel, barter for other goods, travel some more, perhaps re-trade their new goods and so on; finally returning to their place of origin several months later.

The mere presence of products distributed over larger distances is not proof of caravan transport, either household organized or administered, because other modes such as direct acquisition and down-the-line models actually result in widely dispersed goods as well. Furthermore, many of the distinctive objects that archaeologists recognize as non-local are often small enough to have been transported without cargo animals. Establishing the beginnings of caravan transport is not a simple task because there is no one signature for long distance caravan organization that is distinctive from other modes of transport. Furthermore, many of the goods are believed to have been perishable, complicating efforts to interpret prehispanic trade caravan patterns. Finally, studies of contemporary caravans emphasize that diversified strategies characterize caravan driving, whether in making daily decisions while on the trail, or in the larger context of economy and exchange. It is thus difficult to define a consistent indicator for caravan activity.

Portable diagnostic artifacts, whether decorated ceramics or other exotic goods, are often relatively small and therefore the artifact weight and total quantity frequently cannot be used to differentiate between caravan transport, traveling peddlers, and down-the-line exchange. The temporal regularity of exchange, however, is a consistent measure that archaeologists can recover from stratified deposits. When regular caravan transport routes developed then the scheduling of such transport may have been linked to the timing of annual events such as harvests and annual ceremonies, and if so these cyclical patterns would result a steady accumulation of non-local goods through time. In contrast, down-the-line exchange depends upon the articulation of many individual exchanges and it is not linked to the acquisition of scheduled harvest products in the same manner as caravan transport and therefore the presence non-local goods would have been irregular.

In terms of the network configurations discussed earlier (Figure 2-3), the configuration that describes the diffusion of obsidian in the region is distinct from the configuration of the regular articulation between herders and farmers that involved the barter of pastoral products for agricultural products. However the regular conveyance of some agricultural goods adjacent to a raw material source creates a context for conveying larger quantities of obsidian regionally. Ethnographic studies indicate that caravans will opportunistically embed exchange into other activities. For example, Nielsen (2000: 488) explains that caravans primarily organized around salt transport would carry a variety of other trade items, and they would occasionally stop to procure raw materials, such as lithics, when the caravan route travels past a known source. Similarly, there is caravans that visited the Colca valley in prehistory from the Titicaca Basin were passing with 3 km of the Chivay source if they used one of the popular routes into the Colca from the south-east direction. Thus, procurement of obsidian was likely to have been associated with long distance exchange opportunities.

Material Expectations

Caravan procurement would have consisted of pastoralists traveling to the Chivay source, acquiring obsidian that they believed to have exchange value, perhaps processing the nodules to some extent, and then transporting the material to consumers or to other traders in areas far from Colca. As temporality is a significant part of caravan organization, regular and dense procurement activities are expected when a caravan undertakes a detour from the principal travel route. Some preparation or maintenance of the trail from the principal travel route to the obsidian source is expected if the animals are heavily laden. The amount of processing and initial reduction in the source area probably reflected the number of days that a caravan would have been willing stop, and conversely the transport of whole nodules is conceivable with the assistance of cargo animals such that some extra grams of weight were less of a limiting factor. As mentioned, cortex is often very thin on Chivay obsidian, therefore it would not be surprising if some percentage of the material transported away from the source area included unreduced, whole nodules.

Ethnoarchaeological studies report that caravan drivers look for camps that include high quality pasture, water, and corrals if possible. When these features occur close to an extraction site, a relatively dense obsidian processing area may be found nearby because herders can nourish their livestock while simultaneously working stone. Because of the temporality of caravan activity, procurement was probably intensive, though episodic, through time. When caravans or animals were present, processing would occur and then the material would be conveyed away and the quarry area would see little use until the next episode of intense production. Evidence of production may be variable, however, because caravans could also transport whole nodules without very much difficulty.

Many of the artifact types that were produced from obsidian transported by caravans can deduced based on archaeological evidence acquired from consumption contexts dating to the Terminal Archaic and later. The principal artifact form made from obsidian during the time that camelid caravans conceivably operated (subsequent to the Late Archaic), is the small triangular projectile point diagnostic to the Terminal Archaic and later. This is not to say that other tool forms were not being made, as other non-diagnostic biface forms could also date to the Terminal Archaic or later, however the vast majority of bifacially-flaked obsidian artifacts are projectile points. Furthermore, cores were likely transported away from the obsidian source because simple flakes serve as valuable cutting tools. Therefore, the majority of obsidian production would have prioritized the production of cores and flakes that serve as blanks for triangular point production. If advanced reduction occurred in the quarry area it is likely that small, triangular point forms were the objective.

Reduction strategies that target the production of small triangular points were probably relatively flexible because these point forms are not especially long or delicate, and therefore it would have been possible to produce the appropriate blanks from a variety of core forms. None-the-less, the overall variability in formal tools forms produced from obsidian is exceptionally low because series 5 points look quite similar. Therefore in terms of intensified production from the Terminal Archaic and onward, relatively consistent reduction strategies in the procurement workshop zone may be encountered.

Because these caravans were independent, one might expect slightly greater variability than in the other models of obsidian distribution because various individual households were participating in this procurement, and methodological variation by region might be reflected in the reduction methods. Finally, independent caravans may behave in other, relatively variable ways such as in architecture, in divergent ceramic styles, and in the types of obsidian pieces that were being exported.

3.7.4. Elite-Sponsored Caravans Model

Description

Ethnohistoric evidence suggests that elites commissioned long distance trade caravans to procure materials that were used in a variety of elite strategies at the regional center. As discussed above, Stanish argues that Titicaca Basin elite-administered long distance trade involved acquiring goods for favorable barter rates in distant valleys and then acquiring prestige from the redistribution of these goods. Late Prehispanic elites were probably in a good position to initiate large caravans: they had immense camelid herds and their followers owned them tax payments in the form of labor. In addition, elites would have had the surpluses necessary to initiate a large scale trading venture.

Would such caravans have visited the Chivay source and extracted obsidian for elite consumption or redistribution? Stanish (2003: 69) argues that all Prehispanic trade was administered trade, as opposed to market-driven trade, but he specifically excludes trade in obsidian as "small and light" and capable of being transported through down-the-line exchange. Obsidian does not appear to have been a high prestige item along the lines of precious metals in the prehispanic Andes. Under the Inka there is evidence of control of access to tunnels leading to rich gold mines in the Andes (Burger and Glascock 2002: 364). With obsidian, however, there is no evidence of elite control either in obsidian consumption patterns in regional centers, or in the Arequipa obsidian quarry areas that are generally dispersed and it would have proved difficult to limit access to them (Jennings and Glascock 2002: 115-116). Thus extensive elite-administered acquisition or redistribution of obsidian should not be expected. Nevertheless, given the importance of exchange and non-local goods in issues relating to the origins of social complexity, any evidence of elite-organized raw material procurement should be studied closely.

Material Expectations

Elite-administered caravans would be relatively difficult to differentiate from independent caravans in their source activities. Elite related diagnostic pottery may be encountered in the vicinity of the source area. There may have been some degree of greater standardization if these were part time or full-time specialists working for the elites. Elite-sponsored procurement may involve greater intensification than would be expected from independent caravan procurement because these task groups likely have been organized and dedicated to the procurement objective. Finally, the export of large nodules may have increased as elite-sponsored caravan trains were reportedly large and capable, and the weight of nodules would not have overly interfered with the progress of the caravan following this model. In addition, if the object of elites was prestige building, larger nodules would probably have been more impressive in the distant consumption zone. In short, differentiating elite sponsored caravans at the source may be relatively difficult unless pottery or some other diagnostic material is found to have been associated with procurement.

3.8. Summary

This chapter establishes the background for further research into the production and circulation of Chivay obsidian in the prehispanic Andes. The chapter began with a review of different forms of exchange based on Polanyi's framework but discussed in light of Andean prehistory. The specific geographical and cultural conditions that distinguish ancient Andean economies were placed in a context that is comparable with other mountain culture regions.

Theories of culture change might suggest that regional exchange requires the managerial or coercive power of elites to organize caravans and benefit differentially from the trade goods. However, archaeological evidence shows that the persistent obsidian exchange that moved small quantities of obsidian since the Early Archaic saw a significant increase during the Terminal Archaic. This increased circulation of obsidian is an early symptom of the dramatic economic and social changes that would occur in the south-central Andean highlands in Formative times.

The evidence of this circulation was perhaps not apparent at early centers because these centers were not yet dominating regional exchange patterns through the control of labor and large herds of camelids that are believed to have been a feature of Late Formative primate centers. The mechanism responsible for disseminating obsidian to the region beginning in the Terminal Archaic is perhaps related to a number of interrelated phenomena. The principal factors governing this change may include: the lowered cost of interaction and transport of weight facilitated by caravan animals, the expansion of regional-scale social networks, and the impetus provided by the needs of some to differentiate themselves in this time of emergent social ranking through the ownership of non-local goods such as obsidian.