You are on page 1of 23

This article was downloaded by: [Universidad de los Andes] On: 23 February 2009 Access details: Access Details:

[subscription number 793898576] Publisher Routledge Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

World Archaeology
Publication details, including instructions for authors and subscription information: http://www.informaworld.com/smpp/title~content=t713699333

Engineered highlands: the social organization of water in the ancient northcentral Andes (AD 1000-1480)
Kevin Lane a a School of Arts, Histories and Cultures, The University of Manchester, Online Publication Date: 01 March 2009

To cite this Article Lane, Kevin(2009)'Engineered highlands: the social organization of water in the ancient north-central Andes (AD

1000-1480)',World Archaeology,41:1,169 190


To link to this Article: DOI: 10.1080/00438240802655245 URL: http://dx.doi.org/10.1080/00438240802655245

PLEASE SCROLL DOWN FOR ARTICLE


Full terms and conditions of use: http://www.informaworld.com/terms-and-conditions-of-access.pdf This article may be used for research, teaching and private study purposes. Any substantial or systematic reproduction, re-distribution, re-selling, loan or sub-licensing, systematic supply or distribution in any form to anyone is expressly forbidden. The publisher does not give any warranty express or implied or make any representation that the contents will be complete or accurate or up to date. The accuracy of any instructions, formulae and drug doses should be independently verified with primary sources. The publisher shall not be liable for any loss, actions, claims, proceedings, demand or costs or damages whatsoever or howsoever caused arising directly or indirectly in connection with or arising out of the use of this material.

Engineered highlands: the social organization of water in the ancient north-central Andes (AD 10001480)
Kevin Lane

Downloaded By: [Universidad de los Andes] At: 15:55 23 February 2009

Abstract
Recent research in the Late Intermediate Period (AD 10001480) north-central Andes has emphasized the segmentary and egalitarian nature of late prehispanic Huaylas communities in the region. This form of devolved power nevertheless conceals investiture of authority in a person or group, albeit temporarily. Given the ubiquitous importance of water in the Andes, I suggest that this type of power could well have rested on special managers closely involved in the organization of water and its concomitant hydrological infrastructure, working for the wider community and its leaders. Centring on archaeological evidence from the Cordillera Negra in the Ancash highlands the area demonstrates the development of a complex suite of hydraulic technology. It is this unique setting, including previously sparsely recorded structures such as silt dams and silt reservoirs, which provides the technological suite necessary to test these preliminary ideas. The approach taken here is an attempt to revisit some of Karl Wittfogels concepts, adapting them to include a community-based bottom-up perspective to the hydraulic hypothesis.

Keywords
Andes; communitized; Huaylas; hydraulic technology; prehispanic; Wittfogel.

Things look bad for great Causes today, in a postmodern era when, although the ideological scene is fragmented into a panoply of positions which struggle for hegemony, there is an underlying consensus: the era of big explanations is over, we need weak thought, opposed to all foundationalisms. (Zizek 2008: 1)

World Archaeology Vol. 41(1): 169190 The Archaeology of Water 2009 Taylor & Francis ISSN 0043-8243 print/1470-1375 online DOI: 10.1080/00438240802655245

170

Kevin Lane

Introduction This article suggests some initial considerations pertaining to the social organization of water in the prehispanic north-central Andes during the phase known as the Late Intermediate Period (AD 10001480, henceforth referred to as the LIP). The area selected for this article lies within the Cordillera Negra in the region above 3000m in the suni/ paramo and puna/jalca ecozones (see Fig. 1). These two ecozones are categorized respectively as alpine and tundra environments (Pulgar Vidal 1946; Tosi 1960). Hydrologically, the area is the setting for a vast array of water complexes and hydraulic features that shaped the landscape (Lane 2005). Politically, the region fell under the Inca province of Huaylas (c. AD 1480) (Parssinen 2003; Rowe 1946), although it is debatable how much of a homogeneous political entity the Huaylas people might actually have been prior to domination by the Incas (Lane 2007). This paper is divided into three main sections. An initial section introduces a new categorization of hydraulic architecture, dividing between wet and dry technologies, followed by a brief description of the more common hydraulic technologies in use in the Andean highlands. The second part details silt dams and reservoirs, as two technologies that have not been described in detail previously for the Andes (a very preliminary introduction to these technologies is available in Lane (2006)), including a brief historical contextualization of these technologies within the area. Finally, I analyse the central role

Downloaded By: [Universidad de los Andes] At: 15:55 23 February 2009

Figure 1 Map of study.

Engineered highlands

171

played by hydraulic technology in structuring social organization in the region from an adapted Wittfogelian perspective (Wittfogel 1957).

Hydraulic architecture in the Andes In the Andes, coastal hydraulic systems have been studied intensely (e.g. Billman 2002; Farrington 1980; Kosok 1965; Ortlo 1993, 1995; Park 1983; Rostworowski 1998) while the highlands (excepting the circum-Titicaca Lake area) have generally escaped close attention (Sherbondy 1969). The reasons for this disparity range from diculty of access through to the low visibility of the specic features within the rugged terrain. This bias has only relatively recently started to be partially addressed in the form of pioneering work into highland terracing, or andenes (Bonavia 19678; Donkin 1979; Treacy 1994), raised eld systems, also known as waru-waru or camellones (Erickson 1986; Erickson and Chandler 1989), pond elds, also called either qochas or cochas (Flores Ochoa and Paz Flores 1986; Flores Ochoa et al. 1996; Rozas 1986; Valdivia et al. 1999) and articially irrigated moors, commonly referred to as bofedales (Palacios Rios 1977, 1981, 1996). Although the technological exigencies of two of the main physical environments of the Andes, the coastal strip and the highlands, dier greatly they do share technology in common, such as the ubiquitous canals and reservoirs. The coastal desert strip is characterized by water ow rather than water storage, as seen in features such as irrigation canals, ltration galleries and diversion embankments, geared towards oodwater and irrigation farming. The central highlands, with their vertical gradients, narrow valleys and highland alpine pampas, favour a series of technologies that emphasize water storage. Coastal hydraulic architecture emphasizes agricultural production, while the highlands display a more varied retinue of hydraulic features that have so far eluded simple categorization into solely agricultural production. In the highlands, all hydraulic technologies have to be viewed with a certain degree of exibility (Lane 2006). By exibility, what is meant is that technology can have a dual productive purpose: use, on the one hand, for farming and, on the other, for herding. This system of dual-productive technology is a pattern that has been detected in the south Andean highland region (see Erickson 2000; Flores Ochoa and Paz Flores 1986; Flores Ochoa et al. 1996), but never really contemplated for the northern region. Nevertheless, the archaeological evidence from the study area conrms a similar pattern of technological use in the northern highlands (Lane 2005). I divide hydraulic architecture into two main types, dry and wet. Dry hydraulic features are those of which water is a periodic component such as canals, terraces and elds, whereas wet features are those for which water is a constant element of the technology, as in the case of dams and reservoirs (see Table 1). Within this last category I include three important features of eld technology: raised elds, sunken gardens and irrigated moorlands. These three types of hydraulic features occur specically in the circum-Titicaca area and singly in other parts of the Andes (e.g. Plazas et al. 1993). In the highlands these features are important in that they have dual productive uses, for both agriculture and herding. For instance, these three features formed an integral part of the

Downloaded By: [Universidad de los Andes] At: 15:55 23 February 2009

172

Kevin Lane

Table 1 Dry and wet hydraulic architecture Category Hydraulic architecture Dry Type Description

Terraces Irrigation canals Fodder chakras Camellones Qochas Water dams Water reservoirs Silt dam Silt reservoir

Wet

Built steps or benches creating cultivation platforms on sloped terrain Canals distributing water to elds and land features Fields for the cultivation of animal fodder Raised eld cultivation elds Sunken pond elds Articial lake Articial pond Large articial silt trap Small articial silt trap or silt terrace

Downloaded By: [Universidad de los Andes] At: 15:55 23 February 2009

LIP agro-pastoralist Lupaca confederacy of the Titicaca basin (Erickson 2000; Murra 2002 [1968]). One type of dry hydraulic system requires further explanation, namely fodder chakras. Fodder chakras are an ethnographically recorded example of the human cultivation of fodder for animals (Lane 2005). Although previously unknown for the Andes, the cultivation of fodder is a longstanding practice in Europe where during the harsh winter months, especially in the north, natural pasture is scarce. In a similar fashion, it is possible that prehispanic Andean agro-pastoralists cultivated plants for use as fodder in areas above 3500m during the months in the Antipodean winter (August to September) when almost all natural pasture has been either exhausted or sun- and wind-burnt through lack of rains and exposure to the areas high rate of solar radiation. Circumstantial evidence, on the basis of isotope analysis, for the feeding of maize stalks to prehispanic camelids for the region of Chavin would seem to support this hypothesis (Burger and van der Merwe 1990). Ethnographic evidence from the study area attests to the rehabilitation of ancient rain-fed sloping elds (chakras) solely for the purpose of planting of ryegrass (Lolium perenne) and alfalfa (Medicago sativa) for use as fodder at c. 3950m. As part of the dual productive system, suggested above, these fodder chakras would be used to cultivate bitter potatoes (Solanum x juzepczukii and Solanum x curtilobum) in the odd years when the soil matrix has been enriched suciently by animal defecation and plant-matter decomposition. Indeed, this system of herding/bitter potato rotation has analogues in the southern Andes and is an important means of replenishing spent soils (Flannery et al. 1989). In a similar sense, articial bofedales store water geologically (see Fairley 2003) for use in agricultural irrigation and provide important niche grassland habitats for animal pasture (Lane 2005, 2006; Palacios Rios 1977, 1981, 1996). Irrigated bofedales prevail across a large geographical scope and are also obliquely referenced in an important early ethnohistoric source by Guaman Poma de Ayala (1993 [1615]: 780). Archaeologically, the prehistoric existence of large-scale articial bofedales is corroborated by Erickson (2000: 343). In the north-central Andean region geological water storage has emphasized the construction of two types of previously sparsely documented hydraulic architecture,

Engineered highlands

173

namely silt reservoirs (reservorios de limo) and silt dams (represas de limo). These silt dams and reservoirs have been found in close association with water dams and reservoirs.

Wet hydraulic technology in the Cordillera Negra The four types of dams and reservoirs share certain features in common. For instance all construction is of double-faced stone walling inlled with rubble (see Plate 1) with a lower layer of larger foundation stones (see Plate 2); all the stone is rough cut and locally sourced. Extant archaeology and literature attests to the existence of some of these hydraulic features: water reservoirs are known from Guaman Poma de Ayala (1993 [1615]: 937) and more recently recorded by others (Donkin 1979; Farrington 1980; Gambini 19834; Kosok 1965; Netherly 1984; Venturi and Villanueva 2002); although water dams have received scant mention in the archaeological literature (see Denevan 2001: 157; Donkin 1979: 105; Scarborough 2003: 59), they are mentioned ethnohistorically (de Avila 1999 [1598?]: 419) Also, new evidence emerging from current investigations in the Cordillera Negra demonstrates a marked reliance on water from large highland dams. This strategy of reservoir and dam construction would seem to be a logical reection of the naturally occurring, wet weather conditions that exist at this ecological juncture between

Downloaded By: [Universidad de los Andes] At: 15:55 23 February 2009

Plate 1 Section cut of the water dam wall of Estanque (Ti 1) in the Quebrada Tinko.

174

Kevin Lane

Downloaded By: [Universidad de los Andes] At: 15:55 23 February 2009

Plate 2 Detail of dam wall of Rakacocha (Pa 4) in the Quebrada Uchupacancha.

the puna and paramo habitats. It should be noted that these dams and reservoirs form one feature in a complex hydraulic system, which include silt dams, silt reservoirs, terraces and canals that can encompass whole side valleys (Lane 2006). These large dams are gravity structures, in that the volume of their own weight holds back the water. Our survey revealed twenty-nine such sites, from major dams to reservoirs in series (see Table 2). Eighteen of the lakes identied were the result of damming1 and only one, Itchicocha, was wholly natural. The dams and lakes are all located above 4100m. Meanwhile, water reservoirs are located in close association with cultivated elds and therefore occur at a much lower altitude, usually between 3200 and 4000m, hugging the limits of viable cultivation. The reservoirs are generally round, ovular or roughly rectangular and vary between 10 and 15m in relative diameter or length. Construction is of a rough coursed-stone internal wall and either a similar external wall or an earthen embankment. Unlike the other hydraulic structures considered here, they are closed constructions, meaning they enclose a specically bounded area on all sides. Aside from water dams and reservoirs two new hydraulic features, silt dams and reservoirs, have been identied within the study area. These features are local technological adaptations that enhance the available moorland (bofedales) available for pastoralist use. They are closer in type to the large moorlands created and nurtured by the Chichillapi herders of Puno (Palacios Rios 1977, 1981, 1996) than to the household-scale bofedales identied in various parts of the Andes (see Aldenderfer 1998; Carhuallanqui 1998; Kuznar 1995). Through the construction of silt dams and reservoirs, large areas of the puna were turned into a rich plant biota and mineral trap for use by camelids, especially alpacas (Lama pacos). Five silt dams have been identied for the study area with enough combined acreage to feed 1200 animals adequately in a total area of 372,027m2 of articial bofedal (calculation based on formula elaborated by Browman (1990)). Outwardly, these dam-like structures bear striking similarities to the normal water dams yet, unlike the water dams, the silt

Downloaded By: [Universidad de los Andes] At: 15:55 23 February 2009

Table 2 Table of hydraulic architecture within the study area Volume capacity (m3) Source

Site code

Sector

Site name

Altitude (m)

Surface area (m2)

Breque-Chorrillos-Rico area Chorrillos Cho 1 Cho 2 Cho 6 Cho 7 Cho 8 Yanacocha Oleron Cocharuri Orconcocha Putacayoc/Kaukayoc Llanapaccha 4550 4200 4660 39004050 36003900 55,468 53,125 35,000 150,000 170,000 554,680 Water dam Silt Dam Water dam Silt reservoirs/estancia Necropolis/water reservoirs and terraces

Freisem 1998 Survey 20023 Survey 20023 Survey 20023 Survey 20023

Rico-Breque Cj 1 Cj 4 Co 1 Co 2 Ra 1 3,000,000 97,600 C A Nununga Represa Decision Copalococha Intiauran Ricococha Baja 3800 3890 3825 3985 4485 340 480 284,375 75 20,625

Water reservoirs 26 Silt Reservoirs Silt dam Silt reservoir Water dam

Ra 2

Ricocochoa Alta

4560

18,750

92,500

Water dam

Quebrada Rico Togllakita Alalakmachay Estanque Pacarinancocha Tayapucro 4450 4140

40504400 4310

150,000 6,000 10,000 6,000 20,535

Silt reservoirs Water dam

Survey 2000 Survey 2000 Freisem 1998 Survey 20003 Programa Cordillera Negra 1999 Programa Cordillera Negra 1999 Survey 2002 Survey 2003

Engineered highlands

600,000 4250 286

Water dam Water dam Water dam Estancia/lake

N/A Pc 7 Paucarmas-Cruzcatac area Pc 11 Ti 1 N/A Other Pa 3

Survey 2003 Survey 1999 Freisem 1998 Survey 2002 (continued)

175

Downloaded By: [Universidad de los Andes] At: 15:55 23 February 2009

176 Kevin Lane

Table 2 (Continued) Sector Site name Altitude (m) Surface area (m2) Volume capacity (m3) Source

Site code

383,920

Pa 4 Pa 5 Pa 6 Rac 1 Rac 2 Rac 3 Rac 4 Rac 5 Rac 6 Rac 7 Uc 2 225,000 133,920 687,500

Rakacocha Tsaquicocha Carhuacocha Huancacocha Sacracocha Huaytacocha Iscaycocha Yanacocha Macho Yanacocha Hembra Alichococha Agococha/ Negrahuacanan Tsaquicocha 4300 16,000

4350 4625 4550 4425 4590 4500 4575 4725 4725 4325 4525

7,850 1,027 38,392 17,500 35,000 22,500 13,392 32,291 21,875 17,500 68,750

Water dam Silt dam Water dam Silt dam Water dam Water dam Water dam Water dam Water dam Water dam Water dam

Uc 3

Silt dam

Survey 2002 Survey 2002 Freisem 1998 Survey 2002 Survey 2002 Freisem 1998 Freisem 1998 Freisem 1998 Survey 2002 Survey 2002 Ministerio de Agricultura 1996/ Freisem 1998 Survey 2002

Engineered highlands

177

dams are located lower down the altitudinal ladder of the Andes, occurring between 3825 and 4425m. Silt dams also tend to be free-standing rather than anchored onto rock (as many of the water dams are) with usually only one discernible outtake sluice located along the base of the structure, normally at its centre. The basic principle governing the silt dam is that of geologic water storage (Fairley 2003); the accumulated soil basin acts as an aquifer in which water is stored and puried. More sieve than dam, these structures siphon excess water out of the basin while retaining enough saturated moisture for the growth of bofedal-type conditions ideal for camelids. Silt dams are the result of a process of construction and years of careful nurture. In the main, the initial silt basin would probably have been small and would have grown slowly through silt accretion during the annual rains, a system of accretion known as varve formation (Boreham and Moxey 1997; Leet 1982). It is probable that periodic de-silting of these structures occurred with the excess being used for cultivation in terraces and elds; this has been observed among similar structures in South Asia, the gabarband silt-traps (Possehl 1975). The example of the 28.5ha silt dam of Collpacocha (Co 1), located at the conuence of the Rico and Huinchos Rivers, is a classic of this type of technology (see Plates 3 and 4). Collpacocha (Co 1) is 100m long, 11m wide, constructed of three major stone steps to a height of 5.4m. The bofedal at Collpacocha (Co 1) would have allowed a massive concentration of animal wealth at this juncture, which might well have been a major consideration in the placement of the Inca administrative site of Intiauran (Co 2) in close proximity. This develops from the concept of storage on the hoof suggested by Bokonyi (1989). As a counterpart to water reservoirs, the area has yielded evidence for silt reservoirs. These occur in clusters along three of the side valleys in the Breque-Chorrillos-Rico area and there is evidence to suggest that they occur in the side valley that leads from Pacarinancocha Lake, located at the headwaters of the Loco River. Over-silting makes these structures hard to detect, yet they cover an approximate area of 300,415m2, or 300ha of bofedal, enough to sustain 975 animals adequately. They thus constitute an important technology for the provision of pasture at a local level. Unlike water reservoirs they are open, usually semi-circular or horseshoe shaped, ranging between 7 and 20m in length, with an average height that ranges between 60 and 200cm. Within the broad sloping elds in which they occur silt reservoirs are positioned so as to distribute water and silt across the widest extent possible (see Fig. 2 for an example of how this works in practice). Another notable dierence is that these features have a drainage sluice, usually along the centre of the structure, which drains excess water from the reservoir in much the same way as occurs with the silt dams (see Plate 5 for examples of silt reservoirs from the study area). Similar to the silt dams, they occur at a height (39004400m) low enough for them to benet from natural erosion episodes that occur annually in the highlands and high enough so as not to impede generalized farming at lower altitudes. Their range is signicant, as these reservoirs represent a natural progression from the slightly higherplaced silt dams, which they succeed, and crucially at 3900m they edge, by over 200m, into the upper suni area, which is traditionally viewed as the upper limits of primary agriculture. This returns us to the ongoing debate about the nature of the agricultural frontier during the prehispanic period. The fact that bofedal construction extends well into the suni zone makes it pertinent to question the use of the suni area solely for agriculture

Downloaded By: [Universidad de los Andes] At: 15:55 23 February 2009

178

Kevin Lane

Downloaded By: [Universidad de los Andes] At: 15:55 23 February 2009

Figure 2 Schematic of part of a silt reservoir system (Scale 1:500).

and suggests a fully integrated agro-pastoralist economy utilizing this liminal zone between agriculture and herding for both types of livelihood (Lane 2006).

Towards a prehistory of use The use of all these varied hydraulic features represents a signicant breakthrough in our conception of these prehispanic highland peoples economic organization, in which we have herders at ease with the use of technology that aided in the maximization of animal yields, as well as beneting the farming economy. To understand this interrelated relationship between water, crops and animals it is necessary to understand that silt and water reservoirs and dams were particular elements in a system that implemented an integrated vision of water use for both agriculture and pastoralism across valleys. In the high Andes, the pursuit of a successful economy has always necessitated the interlinking of two types of production agriculture and herding as a symbiotic whole that aimed to control risk through economic diversication (Browman 1990). In the Cordillera Negra, by the LIP, we see the development of a highly specialized highland mixed-farmer and mountain agro-pastoralist economy where herding in the puna (4000 and above) was integrated with tuber, legume and semi-cereal cultivation in the suni (30004000m) (Lane 2005, 2006). This farmer-herder economy was based on small dispersed settlements located near to primary production zones (from Mayer 1985, 2002) along the tightly compressed, tiered ecological niches of the high Andes, zones made all the more viable through the construction of a exible hydraulic management system that seamlessly accounted for the exigencies of these two types of production. Ethnohistoric sources for the period in question allude to these two forms of production being exclusively practised by two separate ethnic groups Huari farmers and Llacuaz herders (Duviols 1973, 1986, 2003; Rostworowski 1988). I disagree. In a recent article I argue that by the LIP the Huari and Llacuaz were not separate ethnic groups, but rather

Engineered highlands

179

Downloaded By: [Universidad de los Andes] At: 15:55 23 February 2009

Plate 3 Detail of aerial photograph showing silt dam of Collpacocha (Co 1) and silt basin.

Plate 4 Silt dam of Collpacocha (Co 1).

180

Kevin Lane

complementary, though with unequal moiety divisions within the ayllu (ayllu is an Andean term generically signifying community) that existed side by side exploiting the resources of the highlands (Lane 2009). A similar form of organization has been observed among the farmers and herders of Sardinia (Angioni 1996). As in the Sardinian case, this does not preclude dierence between these economic groups. Rostworowski (1988) and Gose (1992) have shown how tension existed between these herders and farmers for control of resources, and especially water. As custodians of the puna ecozone, Rostworowski (1988) in particular has demonstrated that in many cases it was the Llacuaz element of the community that primarily controlled water access within the community. Yet, suggesting that we have a set hierarchical system of elites or castes based around a form of economic production simplies matters in what was a uid and dynamic relationship between high-altitude farmers and herders that shifted according to geography, ecology and local politics across the north-central Andean highlands. For instance, in the study area we know from the domestic settlement pattern and material evidence (such as ceramic) that there is little apparent dierentiation between houses or indeed settlements other than that of size (Lane 2007). In the mortuary evidence the emphasis is rather on communal family sepulchre-type tombs commonly found throughout the region (Lane 2005); see also Isbell (1997) for a regional perspective. Again the sepulchres, also known locally as chulpas, machays or pukullos (Herrera and Lane 2004), demonstrate little to dierentiate them in the manner of function or scale that might indicate a more important family clan vis-a`-vis their peers within the community. In eect the only anomaly seems to be the truly monumental scale of the hydraulic structures in the region. Chronologically within the study area the settlement and mortuary evidence demonstrates a steady colonization of the puna highlands commencing in the Early Intermediate Period (EIP) (100 BCAD 600). Currently there is very little evidence to suggest that prior to the EIP the area was used for more than transit or ephemeral and non-permanent hunting and subsequently small-scale herding (Lynch 1980). Intensive colonization of the suni and puna eco-zones was followed by the adoption of the system of agro-pastoralism described above during the Middle Horizon (MH) (AD 6001000) and the LIP (AD 10001480). The construction of hydraulic infrastructure to facilitate these strategies is clearly reected by the establishment of permanent settlements and cemeteries close to these features within the study area during this period. The archaeological evidence points to a substantial increase in pastoralist production during this period culminating with the Inca occupation of the region (AD 14801532). During this long period of highland occupation, structural construction and accretion over time are crucial in understanding the chronology of these hydraulic systems. Given the complexity and scale of these hydraulic features, it is necessary to consider their construction as part of a long process of harnessing the regions scarce water resources. This process took a long time, with settlement and mortuary evidence demonstrating extensive occupation of the area from perhaps AD 100, and denitely from AD 600 onwards. This timescale would allow for the periodic innovations and re-modelling that these hydraulic systems would have necessitated, taking into consideration important cultural and human uctuations through time. Abandonment of these structures occurred during the Spanish colony (AD 15321821).

Downloaded By: [Universidad de los Andes] At: 15:55 23 February 2009

Engineered highlands

181

The wealth of ancient hydraulic architecture correlates directly with the lack of a constant, reliable source of water in the region. As opposed to the Cordillera Blanca, the Cordillera Negra has no permanent snow-caps to provide glacial melt; therefore water is restricted to rainfall, highland lakes and spring. This lack of water coupled with the severe natural gradient of the area means that modern communities usually complain of water scarcity two months after the rainy season (November to April). Essentially, the vertical landscape renders a region with rainfall between 600 and 800mm per year into a semiarid region. The archaeological evidence posits that this might not always have been so, and that past indigenous responses to variable yearly water supply were both innovative and technologically complex. This technological variability in turn begs the question of what type of organization underwrote the maintenance of these hydraulic systems.

Downloaded By: [Universidad de los Andes] At: 15:55 23 February 2009

Water and social authority Given the postmodernist reaction to grand theories it is perhaps with a degree of trepidation that one might re-approach Wittfogels original hydraulic hypothesis. Yet, taking on board Zizeks (2008) recent clarion call in defence of unfashionable grand ideas, this is precisely what I intend to do. Nevertheless, my approach to Wittfogelian theory is tempered by consideration of the large body of scholarly criticism that has emerged around his original ideas (e.g. Erickson 2006; Lansing 1987; Mitchell 1973; Stanish 1994, 2006). I suggest a more critical approach to Wittfogel that shies away from broad crosscultural commonalities and considers rather questions of scale and context. Nevertheless, I subscribe to the basic Wittfogelian tenet that water was a crucial agent in the negotiation of power in prehistoric societies, especially those in arid and semi-arid regions, and created hydraulic societies (from Steward 1955b). I disagree with Stanishs (1994, 2006) neo-Wittfogelian position on the primacy of the state in creating the necessary conditions for the intensication of hydraulic systems. I do not dispute the fact that this can happen, I argue, though, against its determinism. Instead, I subscribe to Mitchells (1973) concept that, in pre-industrial societies, centralized control of irrigation systems is rather the exception than the norm. This concept has been popular with Clark Erickson (2000, 2006) and his investigations in the circum-Titicaca area. Erickson is vehemently against many of the undertones of Wittfogels up-down position to power and water control, opting rather for a bottom-up approach of community water organization and power. Yet I believe that there is room within Wittfogelian theory to accommodate much of Ericksons criticism. A major problem of Ericksons work has been that his studies have been conducted in the Lake Titicaca region, an area for which there has been abundant evidence of complex socio-cultural formations such as Tiahuanaco and the Inca (Stanish 2001). This is not the case in the study area considered here: other than for the coming of the Inca (c. AD 1480) there is a lack of evidence suggesting state-like formations in the ancient province of Huaylas from at least the end of the EIP (100 BCAD 600) onwards, and probably longer (Lane 2007). It is this condition which, I believe, favours the application of Ericksons bottom-up community-centred model of water control and social authority within a Wittfogelian reinterpretation.

182

Kevin Lane

In essence, what I am advocating is a non-state, community-managed complex hydraulic agro-pastoralism operating in the north-central Andean highlands during the late prehispanic period. As can be seen, in the area, the archaeological evidence posits an eschewing of the centralization of water and power around a permanent dominant clique or elite; yet the environmental conditions a naturally semi-arid region whose scarcity of water is circumvented by the construction of hydraulic technology would suggest that water is the key raison detre around which society is organized. When characterizing a hydraulic society, Wittfogel (1957: 22) considered three separate elements as integral to its constitution: intensication of production, a substantial cooperative arrangement and a specic kind of labour division. If the concept of an intensication of production and a cooperative factor are self-explanatory and understood as given within the construction of a hydraulic society, then I believe that it is the nal element that needs further detailing, especially within the social context of the northcentral Andes. Originally, Wittfogel (1957) envisaged this labour division as one between elites and commoners. Here I posit that another relationship sequenced between this hierarchical extreme and an opposite stance centred on unitary family-type subsistence farming can exist, a relationship centred on the community. The Cordillera Negra has had to contend with water scarcity as an essential environmental element for at least the last thousand years. Given this fact it is highly probable that in the past water availability was indeed the factor around which social organization pivoted, much as it does nowadays. This would in some way explain the juxtaposition between poorly stratied settlements and large complex hydraulic systems. Given these conditions a communitized system of water control would emerge, in which the unifying social unit for this hydraulic society would be the community. The implication is that a group elected from within the community, either hereditary or more likely for a particular period of time, oversees access to water and necessary work within the hydraulic system. Under this remit the central role of the Wittfogelian despot would be taken up by a sub-segment of the community. Therefore, while espousing certain key concepts of the hydraulic hypothesis (Wittfogel 1957), I see power in the study area embedded around the community in a manner similar to that advocated by Erickson (2006): technological and organizational complexity at a rural level without obvious elites a bottom-up approach to authority and power. The lack of distinguishable elites suggests that power and authority were distributed according to segmentary or acephalous leadership organized around the ayllu or community. In this case the elites would be the heads of households or individual ayllus who would comprise the acephalous directorship from which the individual water managers would be chosen. The commoners would be comprised, as nowadays, of non-heads of households (such as younger sons or grandsons) and widowed women. These, especially the former, would comprise the bulk of the community. To summarize, within the context of segmentary communities we would witness a chosen group or set of individuals working as adjunct bureaucracy, at the behest of heads of households, regulating water for the greater group a veritable non-state hydraulic society. Incidentally, it would be from within these elites that the chosen water managers would emerge, to an independent though transient position whose power was circumscribed, by the nature of their election, by the dictates of the community, especially the elite group of household heads. Nevertheless, within their transitory roles their regulatory powers did

Downloaded By: [Universidad de los Andes] At: 15:55 23 February 2009

Engineered highlands

183

imbue the holders with authority. Paerregaard (1994) makes this point when he states that water in the past was both ritual and politics, such that water judges and distributors were relatively independent of the rest of the community. This meant that a particularly foresighted group of water judges could, with community cooperation, initiate or continue the construction of valley-wide systems of water management such as the ones evidenced here. This last point is crucial, as we can see that the hydraulic systems described here were most probably the end eect of centuries of careful construction and nurturing of the water table and concomitant landscape. This would have been possible only through concentrated community action united around a worldview that understood and planned around the limitations of the land and water towards the fullment of their own needs. That the system seems to have survived in use for at least the better part of 500 years is as much a tribute to the people of this region as to their organizational capacity. What we would be evidencing is a fully communitized managerial form of organization. This form of communitized organization could well come under the blanket term of heterarchy (Crumley 1995), although I prefer the more particularist denition used here as it implies the organizational supremacy of the community unit. These empowered communities would be heavily involved in non-subsistence surplus production for trade and exchange with nearby intermontane communities and even the coast, evidenced in the importation of exotic goods such as ceramics, textiles and shell (Lane 2005). Ethnohistorical evidence highlights the importance of the community within this form of social organization, for instance the position of the water judge or cilquiua named by 2 Guaman Poma de Ayala (1993 [1615]: 7801) or the water camayuc or cochacamayuc mentioned in the Huarochir Manuscript (de Avila 1999 [1598?]: 935). Similarly, the documents state that these people were in charge of water distribution and importantly of the rites and oerings to lakes, canals and rivers. In the past, control of water always incorporated the physical with the cosmological, and hence the political. Yet, whereas in the case of the cilquiua the position was chosen, that of the position of cochacamayuc of the Huarochir district seems to have been vested in two lineages, those of Llacsamisa and of Rapacha clans, from which water priests or yancas were chosen (Salomon 1998). These two ancestors represented two lineages and it is possible that they alluded to moiety halves of a given ayllu, or community, and thus in fact comprised the majority of the community. Therefore there is the subtle implication that people might have been in a manner of speaking elected from within these moiety halves. This last example would seem to bear similarities to the Llacuaz and Huari situation described for the study area. The scale of hydraulic architecture documented in the Huarochir district accords well with that found in the study area, and suggests that water was perhaps managed in a similar communal fashion in the Huaylas highlands (contra sensu stricto Steward 1955a; Wittfogel 1957). The low population of the study area, estimated at between 2250 and 4000 people for the whole study area (Lane 2005: 205), supports this hypothesis as it seems to preclude signicant social stratication beyond that of the local groups or clans (Johnson and Earle 2000). These low population gures also support the concept of slow accretion of hydraulic construction over time within a communitized setting. The ethnographic present also suggests the communitized organization of water seen in the position of the alcalde de agua (water mayor) or varayocs (sta-holders) recorded for

Downloaded By: [Universidad de los Andes] At: 15:55 23 February 2009

184

Kevin Lane

Downloaded By: [Universidad de los Andes] At: 15:55 23 February 2009

Plate 5 Examples of silt reservoirs from study area.

many communities in the south (see Boelens et al. 2006; Gelles 2000; Mitchell and Guillet 1994; Paerregaard 1994; Trawick 2003). In the study area the present-day system of water organization is similarly embedded in the community through its Comite de Regantes (Irrigators Committee) which comprises all the heads of households within a given community. This committee in turn elects either tomeros (sluicers) or aguaceros (waterers) from among its number, depending on the community, for a single year, with the possibility of the position being ratied for another term. The tomero or aguacero is nominally independent and in charge of distributing water and, crucially, regulating the cleaning and maintenance of the hydraulic systems, mainly canals, in the area. This canal cleaning, involving the whole of the community, usually takes place twice a year. Canal cleaning is undertaken through cajoling and the actual threat of water restrictions on recalcitrant parties imposed by the tomero or aguacero, the underlying reason being that negligence by members of a community in their water duties can have larger repercussions across the larger community. There is usually more than one tomero or aguacero per community as this reduces the possibility of water access bias by these individuals on behalf of their own family (Gonzalez Rosales and Lane 2007). Nowadays, these positions are seen as particularly onerous and the recipients, usually upstanding members of the communities, often resent their election. This is due to the

Engineered highlands

185

conicts and tensions that often occur during the allocation of access to canal, and hence water, rights by the various members of the communities in respect to their elds. Tensions often erupt both at community and at intra-community level; it is the duty of the tomero or aguacero to negotiate through these disputes as the designated representative of the Comite de Regantes. This last point is an important one, for, although the tomero or aguacero is seen as independent, his position and power are derived from and ultimately dependent on the Comite de Regantes on behalf of the community which sets the rules and future direction of water management for the given hydraulic society, much in the same way as has been suggested for the prehispanic past.

Conclusion Although more than 500 years separate these modern communities from the prehispanic ones, I believe that it has been protable for our purposes to evaluate the communitized element of water organization inherent in these modern systems and what it might elucidate about the Andean past. These modern communities similarly do not possess a large degree of internal social dierentiation, although this is changing as individual families engage to a greater and lesser degree with the modern nation-state, and particularly the inherent inequities of access to wage labour. It is therefore tempting to suggest that it is this lack of social dierentiation that underwrites the heart of community organization in the region in the present as it did in the past (Lane 2007). Similarly, the thin veneer of modern state control of water distribution is a possible model for reecting on Inca control of the prehispanic system post-AD 1480. Also, this study, hopefully, serves to demonstrate the viability of adapting Wittfogelian theory to a community-centred non-state managerial system within a given context and time. Indeed, the fact that we still argue the viability of Wittfogels hydraulic hypothesis is a sure indication of its allure and the appreciable applicability of the underlying structure of many of these grand theories. It is tempting to suggest that in many cases these grand old thinkers may have given us the tools and basic structure to important currents of thought; it is for us to criticize, but also to revisit and were possible recycle. In a small way this is what I have attempted here. A nal point: in the past, this communitized managerial system of control might well have been further complicated by the existence of Llacuaz agro-pastoralists and Huari mixed farmers. The data recovered to date from the study area do not elucidate further the intricacies of their internal relationship regarding water rights. While circumstantial evidence from the Canta province would seem to indicate Huari negotiation of water rights with the dominant Llacuaz (Rostworowski 1988), we also know from other sources that this was not always the case, with Huari dominance registered for other areas further to the north (Duviols 1973). Indeed, within the study area the archaeology would seem to indicate that the southern Loco Valley was more agricultural and hence possibly Huari dominated, while Llacuaz herders seemed to have favoured the northern Chaclancayo Valley (Lane 2005). The matter of the relationship between these complementary halves within this system of communitized managed water authority is an important subject that requires further reection and study.
Downloaded By: [Universidad de los Andes] At: 15:55 23 February 2009

186

Kevin Lane

Acknowledgements The eldwork for this article was made possible by generous sponsorship from the Leverhulme Foundation, the British Academy, the Wingate Trust, the Cambridge European Trust, Trinity Hall Cambridge. I would also like personally to thank Alexander Herrera and all the people who have worked with me on these various eld projects and especially the communities of the Cordillera Negra. Thanks also go to the anonymous reviewers. As always any errors remain my own.

School of Arts, Histories and Cultures, The University of Manchester Kevin.Lane@manchester.ac.uk


Downloaded By: [Universidad de los Andes] At: 15:55 23 February 2009

Notes 1 The dammed water lakes are Cho 1, 6; Ra 1, 2; Pc 7, 11; Ti 1; Pacarinancocha; Pa 4, 6; Rac 2-7, Uc 2, 3. 2 There is no translation for the word cilquiua and it probably represents a lost Quechua word; cochacamayuc translates as keeper of the waters.

References
Aldenderfer, M. S. 1998. Montane Foragers: Asana and the South-Central Andean Archaic. Iowa City: University of Iowa Press. Angioni, G. 1996. On agro-pastoral space in Sardinia. In The Anthropology of Tribal and Peasant Pastoral Societies: The Dialectics of Social Cohesion and Fragmentation (ed. U. Fabietti and P. C. Salzman). Como, Italy: IBIS Publishers. Billman, B. R. 2002. Irrigation and the origins of the southern Moche State on the north coast of Peru. Latin American Antiquity, 13: 371400. Boelens, R., Getches, D. and Guevara Gil, A. (eds) 2006. Agua y Derecho: Polticas Hdricas, Derechos Consuetudinarios e Identidad Locales. Lima, Peru: WALIR/IEP. Bokonyi, S. 1989. Denitions of animal domestication. In The Walking Larder: Patterns of Domestication, Pastoralism and Predation (ed. J. Clutton-Brock). London: Unwin Hyman, pp. 227. Bonavia, D. 19671968. Investigaciones arqueologicas en el Mantaro Medio. Revista del Museo Nacional, 35: 21194. Boreham, S. and Moxey, P. A. 1997. A century of vegetation change in Epping Forest determined from pollen analysis of pond sediments. The London Naturalist, 76: 2135. Browman, D. L. 1990. High altitude camelid pastoralism of the Andes. In The World of Pastoralism: Herding Systems in Comparative Perspective (eds J. G. Galaty and D. L. Johnson). New York: Guilford Press, pp. 32352. Burger, R. L. and van der Merwe, N. 1990. Maize and the origin of highland chav n civilization: an isotopic perspective. American Anthropologist, 92: 8595. Carhuallanqui, R. M. 1998. Pastores de Altura: Magia, Ritos y Danzas. Lima: REDES/Red de Solidaridad.

Engineered highlands

187

Crumley, C. L. 1995. Heterarchy and the analysis of complex societies. In Heterarchy and the Analysis of Complex Societies (eds R. Ehrenreich, C. L. Crumley, and J. E. Levy). Washington, DC: American Anthropological Association, pp. 15. de Avila,, F. 1999 [1598?]. Ritos y Tradiciones de Huarochir, 2nd edn. Lima: IFEA, Banco Central de Reserva Del Peru, Universidad Particular Ricardo Palma. Denevan, W. M. 2001. Cultivated Landscapes of Native Amazonia and the Andes. Oxford: Oxford University Press. Donkin, R. A. 1979. Agricultural Terracing in the Aboriginal New World. Tucson, AZ: University of Arizona Press. Duviols, P. 1973. Huari y Llacuaz: agricultores y pastores: un dualismo prehispanico de oposicion y complementaridad. Revista del Museo Nacional, 39: 15391. Duviols, P. 1986. Cultura Andina y Represion: Procesos y Visitas de Idolatrias y Hechicerias, Vol. 17. Cuzco: Centro de Estudios Rurales Andinos, Bartoleme de las Casas.
Downloaded By: [Universidad de los Andes] At: 15:55 23 February 2009

Duviols, P. 2003. Procesos y Visitas de Idolatras: Cajatambo, Vol. 17. Lima: IFEA/PUCP. Erickson, C. L. 1986. Agricultura en camellones en la cuenca del Lago Titicaca. In Andenes y Camellones en el Peru Andino: Historia, Presente y Futuro (eds C. de la Torre and M. Burga). Lima: CONCITEC, pp. 33150. Erickson, C. L. 2000. The Lake Titicaca Basin: a Precolumbian built landscape. In Imperfect Balance: Landscape Transformations in the Precolumbian Americas (ed. D. L. Lentz). New York: Columbian University Press, pp. 31156. Erickson, C. L. 2006. Intensication, political economy, and the farming community: in defense of a bottom-up perspective of the past. In Agricultural Strategies (eds J. Marcus and C. Stanish). Los Angeles, CA: Cotsen Institute of Archaeology, University of California, pp. 33463. Erickson, C. L. and Chandler, K. L. 1989. Raised elds and sustainable agriculture in the Lake Titicaca Basin of Peru. In Fragile Lands of Latin America: Strategies for Sustainable Development (ed. J. O. Bowder). Boulder, CO: Westview Press, pp. 23048. Fairley, J. P. 2003. Geologic water storage in Precolumbian Peru. Latin American Antiquity, 14: 193 206. Farrington, I. S. 1980. The archaeology of irrigation canals, with special reference to Peru. World Archaeology, 11: 287305. Flannery, K. V., Marcus, J. and Reynolds, R. G. 1989. The Flocks of the Wamani: A Study of Llama Herders on the Puna of Ayacucho, Peru. San Diego: Academic Press. Flores Ochoa, J. and Paz Flores, M. P. 1986. La agricultura en lagunas (qochas). In Andenes y Camellones en el Peru Andino: Historia, Presente y Futuro (eds C. de la Torre and M. Burga). Lima: CONCITEC, pp. 85105. Flores Ochoa, J., Paz Flores, M. P. and Rozas, W. 1996. Un (re-)descubrimeinto reciente: la agricultura en lagunas temporales (qocha) en el Altiplano. In Comprender la Agricultura Campesina en los Andes Centrales: Peru-Bolivia (ed. P. Morlon). Lima: IFEA/ CBC, pp. 24755. Gambini, W. 19834. Santa y Nepena: Dos Valles, Dos Culturas. Lima: Imprenta M. Castillo R. Gelles, P. H. 2000. Water and Power in Highland Peru: The Cultural Politics of Irrigation and Development. New Brunswick, NJ, and London: Rutgers University Press. Gonzalez Rosales, D. and Lane, K. 2007. Un Acercamiento Antropologico al Distrito de Pamparomas: Informe Preliminar. Proyecto Arqueologico Regional Ancash Cochayoq (ParaCo). Gose, P. 1992. Segmentary state formation and the ritual control of water under the Incas. Comparative Studies in Society and History, 35: 480514.

188

Kevin Lane

Guaman Poma de Ayala, F. 1993 [1615]. Nueva Cronica y Buen Gobierno, Vols 13 Lima: Fondo de Cultura Economica. Herrera, A. and Lane, K. 2004. Project gallery: issues in Andean highland archaeology: the Cambridge Round Table on Ancash Sierra Archaeology. Antiquity, 78: Available at: http:// antiquity.ac.uk/ProjGall/herrera/index.html. Isbell, W. H. 1997. Mummies and Mortuary Monuments: A Processual Prehistory of Central Andean Social Organization. Austin, TX: University of Texas Press. Johnson, A. W. and Earle, T. 2000. The Evolution of Human Societies: From Foraging Group to Agrarian State, 2nd edn. Stanford, CA: Stanford University Press. Kosok, P. 1965. Life, Land and Water in Ancient Peru. New York: Long Island University Press. Kuznar, L. A. 1995. Awatimarka: The Ethnology of an Andean Herding Community. Fort Worth, TX: Harcourt Brace College. Lane, K. 2005. Engineering the Puna: the hydraulics of agro-pastoral communities in a north-central Peruvian valley. PhD thesis, University of Cambridge. Lane, K. 2006. Through the looking glass: re-assessing the role of agro-pastoralism in the northcentral Andean highlands. World Archaeology, 38: 493510. Lane, K. 2007. The state they were in: community, continuity and change in the north-central Andes, 1000AD1608AD. In Socialising Complexity: Structure, Integration and Power (eds S. Kohring and S. Wynne-Jones). Oxford: Oxbow. Lane, K. 2009. Earth, water and ethnogenesis: the imprinting of myth on landscape in the Cordillera Negra, north-central Peru (A.D.9001480). Antiquity. Lansing, J. S. 1987. Balinese water temples and the management of irrigation. American Anthropologist, 89: 32641. Leet, L. D. 1982. Physical Geology. Englewood Clis, NJ: Prentice Hall. Lynch, T. F. 1980. Guitarrero Cave: Early Man in the Andes. New York: Academic Press. Mayer, E. 1985. Production zones. In Andean Ecology and Civilization: An Interdisciplinary Perspective on Andean Ecological Complementarity (eds S. Masuda, I. Shimada and C. Morris). Tokyo: University of Tokyo Press, pp. 4584. Mayer, E. (ed.) 2002. The Articulated Peasant: Household Economies in the Andes. Boulder, CO: Westview Press. Mitchell, W. P. 1973. The hydraulic hypothesis: a reappraisal. Current Anthropology, 14: 5324. Mitchell, W. P. and Guillet, D. (eds) 1994. Irrigation at High Altitudes: The Social Organization of Water Control Systems in the Andes. Washington, DC: American Anthropological Association. Murra, J. V. 2002 [1968]. Un reino Aymara en 1567. In El Mundo Andino: Poblacion, Medio Ambiente y Economa. Lima: IEP & PUCP, pp. 183207. Netherly, P. J. 1984. The management of late Andean irrigation systems on the north coast of Peru. American Antiquity, 49: 22754. Ortlo, C. R. 1993. Chimu hydraulic technology and statecraft on the north coast of Peru, AD 1000 1470. In Economic Aspects of Water Management in the Prehispanic New World (eds V. Scarborough and B. Isaacs). Greenwich, CT: JAI Press, pp. 32767. Ortlo, C. R. 1995. Surveying and hydraulic engineering of the pre-Columbian Chimu state, 9001450. Cambridge Archaeological Journal, 5: 5574.
AD

Downloaded By: [Universidad de los Andes] At: 15:55 23 February 2009

Paerregaard, K. 1994. Why ght over water? Power, conict, and irrigation in an Andean village. In Irrigation at High Altitudes: The Social Organization of Water Control Systems in the Andes (eds W. P. Mitchell and P. Guillet). Washington, DC: American Anthropological Association.

Engineered highlands

189

Palacios Rios, F. 1977. Pastizales de regad o para Alpacas. In Pastores de Puna: Wywamichiq Punarunakuna (ed. J. Flores Ochoa). Lima: Instituto de Estudios Peruanos. Palacios Rios, F. 1981. Tecnolog a del pastoreo. In In La Tecnologa en el Mundo Andino: Runakunap Kawsayninkupaq Rurasqankunaqa (eds H. Letchman and M. Soldi). Mexico: Universidad Nacional Autonoma de Mexico, pp. 21732. Palacios Rios, F. 1996. Pastizales de regad o para alpacas en la puna alta (el ejemplo de Chichillapi). In Comprender la Agricultura Campesina en los Andes Centrales: Peru-Bolivia (ed. P. Morlon). Lima: IFEA/CBC, pp. 20712. Park, C. C. 1983. Water resources and irrigation agriculture in pre-Hispanic Peru. The Geographical Journal, 149: 15366. Parssinen, M. 2003. Tawantinsuyu: El Estado Inca y su Organizacion Poltica. IFEA/Fondo Editorial de la Ponticia Universidad Catolica del Peru/ Embajada de Finlandia. Plazas, C., Falchetti, A. M., Saenz Samper, J. and Archila, S. 1993. La Sociedad Hidraulica Zenu: Estudio Arqueologico de 2,000 Anos de Historia en las Llanuras del Caribe Colombiano. Santa Fe de Bogota, Colombia: Banco de la Republica. Possehl, G. L. 1975. The chronology of Gabarbands and Palas in western South Asia. Expedition, 17: 337. Pulgar Vidal, J. 1946. Historia y Geografa del Peru. Lima: Universidad Nacional de San Marcos. Rostworowski, M. 1988. Conicts over Coca Fields in XVIth-Century Peru. Ann Arbor, MI: University of Michigan Press. Rostworowski, M. 1998. Sistemas hidraulicos de los Senor os Costenos Prehispanicos. In Ensayos de Historia Andina II: Pampas de Nasca, Genero, Hechicera (ed. M. Rostworowski). Lima: IEP/BCRP, pp. 12549. Rowe, J. H. 1946. Inca culture at the time of the Spanish conquest. In Handbook of South American Indians, Vol. 2, The Andean Civilization (ed. J. H. Steward). Washington, DC: Bureau of American Ethnography, Smithsonian Institution, pp. 183330. Rozas, W. 1986. Es Sistema de cultivo en qocha. In Andenes y Camellones en el Peru Andino: Historia, Presente y Futuro (eds C. de la Torre and M. Burga). Lima: CONCYTEC, pp. 10726. Salomon, F. 1998. Collquiris dam: the colonial re-voicing of an appeal to the Archaic. In Native Traditions in the Postconquest World: A Symposium at Dumbarton Oaks, 2nd through 4th October 1992 (eds E. Hill Boone and T. Cummins). Washington, DC: Dumbarton Oaks Research Library and Collection, pp. 26593. Scarborough, V. L. 2003. The Flow of Power: Ancient Water Systems and Landscapes. Santa Fe, NM: SAR Press. Sherbondy, J. E. 1969. El regad o en el area Andina central: ensayo de distribucion geograca. Revista Espanola de Antropologa Americana, 4: 187223. Stanish, C. 1994. The hydraulic hypothesis revisited: Lake Titicaca Basin raised elds in theoretical perspective. Latin American Antiquity, 5: 31232. Stanish, C. 2001. The origin of state societies in South America. Annual Review of Anthropology, 30: 4164. Stanish, C. 2006. Prehispanic agricultural strategies of intensication in the Titicaca Basin of Peru and Bolivia. In Agricultural Strategies (eds J. Marcus and C. Stanish). Los Angeles, CA: Cotsen Institute of Archaeology, University of California, pp. 36497. Steward, J. H. (ed.) 1955a. Irrigation Civilizations: A Comparative Study. Washington, DC: Pan American Union.

Downloaded By: [Universidad de los Andes] At: 15:55 23 February 2009

190

Kevin Lane

Steward, J. H. 1955b. Some implications of the symposium. In Irrigation Civilizations: A Comparative Study (ed. J. H. Steward). Washington, DC: Pan American Union, pp. 5878. Tosi, J. A. 1960. Zonas de Vida Natural en el Peru. Lima: Instituto Interamericano de Ciencias Agricolas de la OEA, Zona Andina. Trawick, P. B. 2003. The Struggle for Water in Peru: Comedy and Tragedy in the Andean Commons. Stanford, CA: Stanford University Press. Treacy, J. M. 1994. Las Chacras de Coporaque. Lima: Instituto de Estudios Peruanos. Valdivia, R., Reinoso, R. and Mujica, E. 1999. Qochas en la cuenca Norte del Titicaca. Gaceta Arqueologica Andina, 25: 14766. Venturi, F. and Villanueva, S. 2002. El Manejo del Agua en la Cordillera Negra: Estudio Elaborado por Encargo del Programa Cordillera Negra en los Distritos de Huaylas y Huata. Programa Cordillera Negra/Programa de Lucha Contra la Pobreza en Zonas Rurales de la Region Chavin. Wittfogel, K. A. 1957. Oriental Despotism: A Comparative Study of Total Power. New Haven, CT: Yale University Press. Zizek, S. 2008. In Defense of Lost Causes. London and New York: Verso.

Downloaded By: [Universidad de los Andes] At: 15:55 23 February 2009

Kevin Lane completed an MA and PhD in Andean archaeology at the University of Cambridge in 2005. He has undertaken extensive eldwork in Spain, Italy, UK and Germany. He is currently a Leverhulme Fellow at the University of Manchester.

You might also like