<<

Water and sustainable land use at the ancient tropical city of ,

Vernon L. Scarborougha,1, Nicholas P. Dunningb, Kenneth B. Tankersleya, Christopher Carrb, Eric Weaverb, Liwy Graziosoc, Brian Laned, John G. Jonese, Palma Buttlesf, Fred Valdezg, and David L. Lentzh

Departments of aAnthropology, bGeography, and hBiology, University of Cincinnati, Cincinnati, OH 45221; cUniversidad de San Carlos de Guatemala, Edificio S1, Ciudad Universitaria zona 12, 01012 Ciudad de Guatemala, , Guatemala; dDepartment of Anthropology, University of Hawaii at Manoa, Honolulu, HI 96822; eDepartment of Anthropology, Washington University, Pullman, WA 99164; fSoftware Engineering Institute, Carnegie Mellon University, Pittsburgh, PA 15213; and gDepartment of Anthropology, University of Texas, Austin, TX 78712

Edited by B. L. Turner, Arizona State University, Tempe, AZ, and approved May 28, 2012 (received for review February 22, 2012)

The access to water and the engineered landscapes accommodat- Tikal, previously mapped by the University of Pennsylvania Tikal † ing its collection and allocation are pivotal issues for assessing Project (14) (Figs. S1 and S2). As is typical of many of the largest sustainability. Recent mapping, sediment coring, and formal centers located in the heartland of present day Petén, excavation at Tikal, Guatemala, have markedly expanded our Guatemala, Tikal was positioned some distance from major pe- understanding of ancient Maya water and land use. Among the rennial stream or lake access and dependent on sizable reservoir – landscape and engineering feats identified are the largest ancient construction associated with elevated catchments (5 7). fi The longevity and resilience of the ancient Maya over a 1,500-y dam identi ed in the Maya area of ; the posited fi manner by which reservoir waters were released; construction of period reveal signi cant cultural and environmental adaptations to a seasonally wet–dry tropical ecosystem beset by a sizable a cofferdam for dredging the largest reservoir at Tikal; the pres- population—a condition with contemporary implications. Pop- ence of ancient springs linked to the initial colonization of Tikal; ulation estimates for the southern Maya Lowlands at A.D. 700 the use of sand filtration to cleanse water entering reservoirs; fi suggest as many as 5 million people or a density one full order of a switching station that facilitated seasonal lling and release; magnitude greater than supported in the region today (7, 10); and the deepest rock-cut canal segment in the Maya Lowlands. ironically, this setting is under environmental siege by a much These engineering achievements were integrated into a system reduced contemporary population—one organized very differ- that sustained the urban complex through deep , and they ently than in the past (3, 4). Early colonization of the ridge on have implications for sustainable construction and use of water which the subsequent site of Tikal resides was spurred by springs management systems in tropical forest settings worldwide. at the head end of a natural ravine that was subsequently widened and dammed. Architectural construction and paving archaeology | resilience | intensification | tropics | paleoecology accompanying a population boom in the Classic period required large water storage tanks in the site’s central precinct filled by ow populations have used currently threatened directed seasonal runoff, but the pavements unintentionally Henvironments in a sustainable and managed manner over prevented the normal recharging of the springs that had origi- time can be addressed through archeology and its multidisci- nally attracted colonists. Although the recharge and filtering of plinary collaborations (1). Today, in the geographical core of the pure water source was significantly curtailed, many more Classic Maya (A.D. 250–800)—the tropical forest of the amount of water available to the growing population Petén, Guatemala (a subtropical moist forest in the Holdridge was now contained in the formal reservoir system. system) (2)—short-fallow slash-and-burn , logging, Results and cattle ranching have significantly affected portions of the ecosystem and limited access to potable water (3, 4). Neverthe- Our investigations concentrated on the water management fea- less, within this biophysical context, one of the earliest and most tures on the southern side of central Tikal. At the summit of the long-lived tropical flourished. Maya water and land north-to-south–oriented ridge defining the upraised central pre- uses were significantly affected by highly seasonal precipitation cinct of Tikal, three prominent arroyos cut west to east through and karst physiography, which accommodated little perennial the site area (Fig. 1). The southernmost arroyo was deeply incised surface water. In response, the ancient Maya developed a com- and displays significant human alteration, although each of these plex system of water management dependent on water collection drainages was dammed or had their flow redirected. The , and storage devices. The hydraulic system was cleverly tailored to Palace, and Hidden Reservoirs represent a descending chain of the biophysical conditions and adaptively engineered to the artificially dammed tanks that were markedly widened to increase evolving needs of a growing population for more than 1,000 y (5– storage capacity as well as provide necessary construction fill for 7). By identifying how a tropical setting was altered using a Stone Age , methods and techniques associated with long-lived and sustainable landscape engineering are revealed. Establishing Author contributions: V.L.S., N.P.D., K.B.T., C.C., L.G., and D.L.L. designed research; V.L.S., baseline assessments of human impact on an environment before N.P.D., K.B.T., C.C., E.W., L.G., B.L., J.G.J., F.V., and D.L.L. performed research; V.L.S., N.P.D., the extraction and depletion of resources by recent technological K.B.T., C.C., E.W., J.G.J., P.B., F.V., and D.L.L. analyzed data; and V.L.S., N.P.D., K.B.T., C.C., advancements may allow an evaluation of current technology’s E.W., and D.L.L. wrote the paper. effects and the origins of unintended ecological as well as The authors declare no conflict of interest. social consequences. This article is a PNAS Direct Submission. The ancient low-density urban community of Tikal, Guate- 1To whom correspondence should be addressed. E-mail: [email protected]. mala, was recently examined by way of water and landscape – This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10. assessments (8 10).* Our intent was to document the evolution 1073/pnas.1202881109/-/DCSupplemental. of a tropical wet–dry engineered landscape (11) and the manner in which the site was altered from its initial colonization (Middle *Low-density urbanism contrasts with highly nucleated definitions of city identified with ’ many archaic states in semiarid settings elsewhere in the world. Population estimates to Late Preclassic, 600 B.C. to A.D. 250) to the community s range from 60,000 to 80,000 people within a 120-km2 area. apogee (Late Classic, A.D. 650–800) followed by its rapid near † A dry core from Inscription Reservoir yielded four radiocarbon dates, and the most re- abandonment in the late ninth century (12, 13). Six ancient cent was contemporanous with the core from the Tikal Reservoir and likely pre-Maya. reservoirs associated with four separate topographical catchment The small residential Madeira Reservoir was not further explored, although previous 2 areas were examined within the 9-km core, referred to as Central excavators indicate Early Classic to Late Classic usage.

12408–12413 | PNAS | July 31, 2012 | vol. 109 | no. 31 www.pnas.org/cgi/doi/10.1073/pnas.1202881109 Downloaded by guest on September 28, 2021 Fig. 1. Map showing the main catch- ments and associated reservoirs at Tikal, Guatemala. We focused on six reservoirs in the southern one-half of the area—Tem- ple, Palace, Hidden, Corriental, Perdido, and Terminos Reservoir (not shown) to the east of Central Tikal (Fig. S2B). Three prominent arroyos are shown in blue. From ref. 15. Reprinted with permission from AAAS.

building massive pyramids and palace-like structures in immedi- Reservoir, was engineered from exposed bedrock with added ate proximity (Figs. 2 and 3). Expansive plaster surfaces covering construction fill to prevent the natural movement of water out of plazas and courtyards sealed the otherwise fractured the arroyo head. Four radiocarbon samples (Fig. S3B) were bedrock and were canted to direct runoff into these tanks, an taken from the main tank body and indicate a Late Classic adaptation that prevented the natural percolation of surface (eighth century) remodeling. Another three samples were re- moisture into the porous substratum and concentrated seasonal trieved from the silting tank (Fig. S3C), and both contexts in- precipitation for ameliorating annual (15). Two other dicate a Late Classic use (Fig. 3 and Table S1). One sample from low-lying bajo- or swamp-margin reservoirs—Corriental and the main tank excavation unit (OP 7C) (Fig. S3A) was associated Perdido—separated by two independent but adjacent catchment with a Late Preclassic date; sediments were trapped between the areas were examined through a program of extensive coring and floor of the tank and subsequent matrices, and the former was formal excavation (Figs. 1 and 3), although Perdido Reservoir introduced from earlier activities flanking a late enlargement of datasets will not be developed here. The Inscription and Tikal the Temple Reservoir (Fig. S3B). Reservoirs, also part of the southern catchment system, received A substantial berm separated the silting tank from the main limited attention in this study (14). Temple Reservoir, with a constricted 2-m-wide plastered spill- way sealing an underlying Late Classic dedicatory burial and an Temple Reservoir. Temple Reservoir is the most elevated but overall vertical drop of 3 m before cascading into the lower-lying smallest public tank, with a conservative volume of 27,140 m3 basin 8 m below the base of the silting tank (Fig. S3A). Of special (16). Surrounded by some of the most impressive pyramidal interest was the amount of water that percolated from one of our structures at the site, together with the feature’s circular sym- test units during the dry season (recharge rate of 7.5 liters/h in metry and associated silting tank, it was a ritual place (Figs. 2 and April of 2009, the driest month of the year; workmen routinely 3). Fourteen excavation units were systematically positioned, filled their canteens from this source while it was exposed, pre- principally in the upper silting tank, with two dry cores taken ferring it to water available in their village). At 1.6 m below the from the basal reaches of the silting tank and the low-lying main surface and all cultural deposition, potable water seeped through ANTHROPOLOGY tank body (Fig. S3A) —the latter featuring a poorly preserved the profile wall of fine silty sediment—a condition not repeated clay lining (Fig. S3B). Excavation results indicate that the ancient from our excavations in other reservoirs at Tikal at significantly Maya expanded the head end of the natural arroyo by creating lower elevations and associated with much larger catchments and an elevated silting tank to the south and west of the original tanks (Fig. S3C). The Temple Reservoir area positioned at the shallow headwater drainage before quarrying the deeper main head end of the ravine may tap a fault spring (17), a source of tank body to the north and east off the original course of the freshwater available all year and a focused initial attraction for

increasingly incised eastward-flowing arroyo. Based on present- Preclassic settlers. A date of 521–216 B.C. was obtained imme- SCIENCE

day contours and subsurface testing, the Temple Causeway, diately above the seep interface (all dates recorded by way of SUSTAINABILITY which separated the Temple Reservoir from the larger Palace a2-σ calibrated range). Earlier excavations along the exterior

Scarborough et al. PNAS | July 31, 2012 | vol. 109 | no. 31 | 12409 Downloaded by guest on September 28, 2021 Fig. 2. Temple, Palace, and Hidden Reservoir chain with location of excavations and ancient arroyo drainage. The spring in the silting tank feeding Temple Reservoir and the cofferdam separating the Temple and Palace Reservoirs are indicated. The release sluice gates run through the Palace dam into Hidden Reservoir. A patch of the Palace Reservoir map has been locally georeferenced based on project Total Station and global positioning system measurements. The base map is courtesy of the Penn Museum (16).

western margins of the reservoir revealed an elaborate ramp and kind of filtered groundwater issuing from the elevated dating to the Late Preclassic (18), contemporaneous with the margins of the arroyo. North Acropolis (19) and the Mundo Perdido complex (20) that wraps around the head end of the arroyo and spring (Fig. S1). Palace Reservoir. The Palace Reservoir extended from the lower Ironically, with the continued growth in population and the built margins of Temple Reservoir to the well-defined Palace Cause- environment, water demands increased markedly, resulting in way or Dam, and it once contained an estimated water volume of extensively paved surfaces designed to direct and retain seasonal 74,631 m3 (21). Thirteen discrete excavations and one dry core runoff (14). This condition, however, capped the recharge po- (Fig. S4A) revealed a diminutive basal channel ∼1-m deep as- tential of the elevated spring and significantly altered the amount sociated with the original arroyo drainage and filled with thin

Fig. 3. Cross-sectional schematic profile of southern one-half of the Tikal hillock, with reservoirs posi- tioned by elevation and associ- ated with dated stratigraphic sed- imentational histories (Table S1).

12410 | www.pnas.org/cgi/doi/10.1073/pnas.1202881109 Scarborough et al. Downloaded by guest on September 28, 2021 alternating lenses of granular sand and silty clays, interpreted as episodes of seasonal storm flows followed by less energetic de- position (Fig. S4 B and C). Stratigraphically dating to the initial colonization of the Tikal hillock or sculpted plateau in the Middle to Late Preclassic, the bedrock channel was modified by way of a carved limestone bench or landing (Fig. S4 B and C)on the immediate northern bank below the precipitous 10-m climb to the upper crest of the ancient ravine. The modified channel bottom dated to 358–55 B.C., although a more elevated stratum yielded a 1739–1535 B.C. date—the latter reversal interpreted as an earlier matrix eroding from the adjacent slope and associated with the original unaltered arroyo (see below). Because banded sediments filled the ancient channel and covered the carved landing without signs of dredging, seasonal access to the flow of water through the ancient ravine was neglected at some point. By the Classic Period, the Palace Dam was constructed to contain the waters that were now directed from the many sealed plaster surfaces in the central precinct. With sloping basal dimensions of 80 × 60 × 10-m high and a conservative volume approximated at 14,000 m3, this gravity dam represents the largest hydraulic architectural feature known in the Maya area, and it is second only to the huge Late Preclassic/Early Classic Purrón Dam in ’s Tehuacán Valley for greater Meso- america (22). Given the narrow course of the incised prequarried arroyo (Fig. S4C), it is likely that portions of the original es- Fig. 4. Palace Reservoir depictions. (A) Schematic of dam function showing carpment were left in place to form part of the dam. Trenching vertically stacked sluice gates (prepared by R. Weaver). (B) Microstratigraphy efforts into the feature suggest an early cut-stone building effort beneath dam collapse debris (OP 6Q) (Fig. S5). (C) Flagstone pavement lining followed by a massive wall of rubble and earthen construction, reservoir floor (northernmost flagstone) (Fig. S4A). the latter sealed with cut veneer stone (Fig. S5). Sluice gates through the massive feature were poorly preserved, but there are evidential suggestions of vertically stacked openings no more maintenance, the collected waters from Temple and Palace than 30 cm in diameter defined by slab stones secured with Reservoirs were subject to a controlled release into the low-lying plaster and imported dark viscous bajo clays (Figs. S5 and S6). Hidden Reservoir before dropping a total of 45 m into and These clays were dated to 15,360 ± 50 C14 y BP (Fig. 3 and around the bajo margin Tikal Reservoir (15) (Fig. 1 and Fig. Table S1), and they were likely sourced to Pleistocene surfaces S2B). One wet core was taken from the lowest lying Tikal Res- found deeply buried in the nearby Bajo de Santa Fe. ervoir (Fig. S2B), yielding only an Early Preclassic date (1430– Water discharge through the dam did not entail massive 1260 B.C.) (Fig. 3 and Table S1). The apparent integration of openings—gates were subject to intense erosion pressures but this tank with the above chain of reservoirs, however, suggests gradually released near the ponded surfaces associated with a contemporaneous date, with the overall system well-defined by ‡ dropping water levels (Fig. 4A). Dating of the dam comes from the Late Classic (15, 16, 21). matrices sealed by elongated cut-stone collapse debris both to- ward the end and immediately after the use life of the feature Corriental Reservoir. Another water catchment and reservoir sys- (the three contexts sampled providing the same A.D. 670–880 tem was positioned south of the chain of reservoirs detailed dates; OP 6L and Op 6Q) (Figs. S4A and S5). Excavations above (23) (Fig. 5A). The Corriental Reservoir is one of four revealed a 10-cm-thick sediment deposit with 11 distinct micro- bajo-margin tanks within Central Tikal, situated just above iso- strata sealed beneath dam collapse (Fig. 4B); darker strata were lated, seasonally inundated depressions (pocket bajos) that were deposited during the annual dry season when water levels likely the foci of intensive farming (21). Corriental was circum- dropped, concentrating organic matter, whereas lighter strata scribed by a 4- to 7-m-high berm with two ingresses or gates and were calcite clay bands laid down by runoff moving across plaster a complex egress. Thirty dry cores and nine trenches (Fig. S7A) in the rainy season. Predictably, partial dam collapse took place revealed the feature’s shape, an estimated maximum capacity of during a rainy season inundation. A 1-m-thick apron of fine silty 57,559 m3 (21), and the tank’s sedimentation history (Figs. 3 and clay was heaped over the collapse debris, suggesting a sub- 5 A and B). A basal date of 8,960 ± 60 C14 y B.P. underlies the sequent crude Late or Terminal Classic period repair attempt. reservoir at 4-m below surface datum (BSD) and was associated The Early Preclassic (1870–1850 B.C.) arroyo channel (OP 6L) with a buried soil identified by a very low δ13C signature in- (Fig. S5) was likely buried by sediment backed up by a small dicative of a natural drainage having coursed through this for- interior dam built in the Early Classic given ceramic associations ested area before Maya colonization (Tables S1 and S2). Wetter and evidence for an initial diversion feature (OP 6U) (Fig. S5). conditions in the Holocene generated subsequent erosion within Sometime thereafter, the present Palace Dam was constructed, the natural catchment and an initial aggradation of sediment by falling into disrepair and partial collapse by the end of the Late the Middle Preclassic (760–400 B.C.). Evidence from the in- Classic period. A flagstone pavement lining was found immedi- terface between the built berm and bedrock indicates a date of ately above the reservoir floor on the southern slope of the tank A.D. 400–570 (core 17) (Fig. S7A and Table S1) for this con- to aid in seepage control (Fig. 4C and Fig. S4A) and is likely struction, a feature at this northern upstream margin of the associated with Late Classic expansion of the reservoir. tank’s circumference established as much for diverting and The isolation of the elevated Temple Reservoir from the more directing flow as containment. Excavation results and X-ray ANTHROPOLOGY voluminous Palace Reservoir suggests that the dam dividing diffraction (XRD) analysis of both the berm and reservoir these two tanks functioned as a coffer when dredging—or repairs sediments indicate limited dredging and a source other than the were needed in the lower reservoir—as indicated by the overall infilling reservoir matrices for the berm’s construction (23). paucity of sediment in the huge Palace tank. After routine The sediments in the reservoir show stratified layers of quartz sands interbedded with laminated organic clays (Fig. 5B and Fig. S7B), the latter indicating slow-moving recharge. Given the lack

‡ SCIENCE Logs used as grand bottle brushes are posited to have allowed the removal of silt of natural quartz sand sources within the greater Tikal region

accumulation with possible vertical inspection by way of limited access manholes at and their frequent identification in strata within several of the SUSTAINABILITY the summit of the dam. Causeway surface excavations by our teams were not possible. reservoirs (Fig. 5B and Figs. S3B and S4 B and C), deliberately

Scarborough et al. PNAS | July 31, 2012 | vol. 109 | no. 31 | 12411 Downloaded by guest on September 28, 2021 Water recharge into the reservoir was closed from this source by A way of an expedient though substantial earth-moving effort in damming the East Gate, only to resume infilling with runoff from the other major catchments entering from the Northwest and South Gates (Fig. S7A).§ During the dry season, waters are posited to have been incrementally released by dismantling the earthen dam and switching the flow out of the tank at this East Gate location. Sometime after the Late Preclassic, a period of greater precipitation (27), the deep V-shaped ingress canal was carefully infilled and sealed with a dense composite of crushed limestone cement, not unlike the matrices used in plaza floors or portions of the most elaborate weight-bearing wall sections at uptown Tikal, thus preventing the redirection of waters into the tank (the decommissioning of the canal segment into the reser- voir corresponds sequentially with the water diversion function of the northern berm construction). A postulated or dam (Fig. S7A) was added to complete the circumscribing berm at the former East Gate switching station, although it has since eroded out as a result of the area’s natural contours and the present-day B C corriental or rushing movement of rainy season waters from this quadrant of Central Tikal [the work by Carr and Hazard (16) noted the incidence of construction stone found below the East Gate drainage and suggest a post-Maya collapse of the plug]. Discussion Ancient climatic instability is recorded in the region, with several researchers suggesting drought peaks during the Terminal Pre- classic period in the second century (27, 28) followed by con- ditions more like today or until the 9th and 10th centuries, when a series of (29–31) aided in severely fragmenting the established political and social order (32). Our work suggests that the system of reservoirs and early water diversion features were established at the onset of a Terminal Preclassic drying trend, an adaptation that likely helped Tikal and some other centers survive, whereas many others were abandoned (27). Subsequent periods of water redirection indicate the decom- missioning of the Corriental Reservoir switching station and the Fig. 5. Corriental Reservoir depictions. (A) Plan map. The principle catch- infilling of the carved bedrock canal leading into it, a reflection ment is to the northwest, with the northeastern catchment diverted away of the greater abundance of water during the Classic period. To from the reservoir during the Classic Period. (B)Profile of the sediment accommodate the need for potable water, sand filtration beds control unit for correlating coring operation; dates in the brackets are from were built at the inflow gates of several of the reservoirs. The core 8 (Fig. S7A). (C) Canal exposure (OP 1G) (Fig. S7A). The base map is water management system was highly resilient and adaptable in courtesy of the Penn Museum (16). the face of both natural and human deranging forces until about A.D. 900. Coring operations have revealed an elevated presence of positioned sand boxes are suggested to have been located above minerals derived from volcanic ash and ejecta in the reservoirs several of the ingresses, directing catchment runoff into a tank (23). Originating hundreds of kilometers away, ash enriched the and acting as a form of filtering to create degrees of potable region’s soils and likely provided the temper for a huge in- water (24). Low δ15N isotopic signatures suggest an absence of vestment in ceramic production throughout the Maya region algal blooms related to pollutants in reservoir waters (23) (Table (33). The frequency of these volcanic events and their short-term S2). Such features may have been replaced frequently, given the negative effects suggest the need for a social mechanism to caprice of storm flow from hurricanes and related tropical mobilize labor to remove as much of this material as possible depressions as well as seasonal variations in rainfall. A terminal from the surficial water catchment system. Nevertheless, the date of A.D. 1010–1170 (Fig. 5B) for reservoir use is recorded at most persistent water quality concern was quotidian excreta and 65 cm BSD, indicating some Early Postclassic occupation of the the human-affected, organic waste that influenced water pota- catchment area (12, 13). Given the steepness of the catchment bility within the engineered microwatershed (24). Although gradient with its associated erosional rates and a reservoir use silting tanks and sand filters helped in controlling both inorganic life of over 1,000 y, the amount of sedimentation into the de- and organic pollutants, respectively, water may have been boiled pression was minimal, with little evidence of dredging—a condi- or more likely, combined with carbohydrates, such as tion unlike that apparent in the summit arroyo tanks. Although gruel, and fermented to accommodate thirst (34). only the Temple Reservoir manifests a well-defined silting tank, Isotopic assessments of bulk carbon in reservoir sediment re- negative evidence suggests that shallow silting tanks were located veal significantly lower δ13C values in Central Tikal and higher above the inflow gates of all reservoirs (25, 26). δ13C values on its periphery during most periods of occupation Of special note were two unusual features identifying an East (Tables S1 and S2). These findings suggest less maize agriculture Gate ingress and egress posited as a seasonally adjusted or presence of other C4 plants immediately within the site cen- switching station or control point for one of the inflowing ter, with potential garden plots farther away. Although the use, catchments recharging the Corriental Reservoir (Fig. 5A and Fig. variety, and abundance of tree species through time are currently S7A). A 2-m-deep canal was sectioned immediately above the ingress, revealing a V-shaped channel excavated through the

indurated bedrock and believed to date to the Late Preclassic, § fl Excavations at the East Gate revealed an elevational drop from the bottom of the with in owing catchment waters directed into a shallow and early V-shaped canal into the upper edge of the Late Preclassic tank of 30 cm. The drop from version of the tank (Fig. 5C). The end of the Late Preclassic has this East Gate lip or elevated margin to the basal Late Preclassic reservoir depths was at been frequently associated with an intensified drought cycle (27). least 1 m.

12412 | www.pnas.org/cgi/doi/10.1073/pnas.1202881109 Scarborough et al. Downloaded by guest on September 28, 2021 being studied, previous research indicates that the ancient Maya over deep time and based on their reduced impact on the bio- of Tikal—at least leading up to the Late Classic period—were physical environs. Today, these simple systems are consistent conservative in their forest management practices (35). Given with conservation efforts (36) and warrant attention in situations the steepness of Tikal’s topography and the near absence of less accommodating of Western technological breakthroughs or terracing, whatever vegetation covered the drainages surround- access to limited—even unobtainable—energy sources. Perhaps ing the reservoirs seems to have anchored soils and slowed by assessing the adaptive behaviors of our distant human erosion—a condition apparent in the relatively slight amount ancestors and their engineered landscapes, we can better inform of sediment infilling into the bajo-margin reservoirs for over our own suite of environmental options and degrees of resilience a millennium. for the future. Hydraulic and potability studies are crucial for understanding urban centers, especially areas where little precipitation is Materials and Methods – available during 3 5 mo of the year (2, 15). The low-density Our work reveals a representative sample of the water catchment and res- urban conditions at Tikal (8, 9) were accommodated by the ad- ervoir system through time at Central Tikal (15, 21). Other tests were con- aptation sustained for more than a millennium. The hydraulic ducted beyond the core, principally along an existing eastward transect (37), system accretionally developed dependent on the interplay be- but they are not discussed here (Fig. S2). This research involved several tween the evolving built environment and the biogenic demands methods and approaches inclusive of both dry and wet coring of ancient of an increasingly anthropogenic forest and soilscape. These reservoirs and natural depressions, conventional excavation and surveying relationships were complicated by climate change, the distribu- techniques, and a suite of paleoenvironmental analyses. Both previous fi tion of reservoirs throughout the site, and the clever modi cation studies and ongoing botanical assessments of past and present ecosystems of the drainage system; a highly resilient urban landscape may complement the water management investigations described (35). To date, have hastened reservoir construction at the end of the Late 45 accelerator mass spectrometry radiocarbon dates from key strata within Preclassic during extended drought conditions, changes that reservoirs and related contexts provide our principle chronological control, subsequently resulted in a different set of water diversion prac- although redeposited ceramics of known ages provide terminus postquem tices associated with greater rainfall. The system was resilient assessments of human use. Because of the extensive and exclusive use of and sustainable based on a Stone Age technology, which pre- wood for cooking fires, field clearance, kiln firing, and lime plaster pro- vented an overexploitation of the biophysical resources under- duction during the construction and maintenance of Tikal, slight but con- girding water access. tinuous charcoal rain on the surfaces of the often undredged tanks provides Although the current intent for creating a more productive use useful dating contexts. of natural resources is driven by improved for extracting and distributing water, other pathways to enduring ACKNOWLEDGMENTS. We thank Pat Culbert and Peter Harrison for their longevity exist. Returning to past life ways is not a sought or pioneering efforts at Tikal and Denise Knisely and Janine Sparks for their viable option, but a deeper understanding of simple technologies editing skills. The two reports submitted to the Guatemalan authorities and associated labor allocations on an engineered landscape further acknowledge the efforts of the Guatemalan–United States field sup- could curb some of the unintended consequences affecting port team. Joyce Marcus assessed an earlier version of this article. We ac- ’ knowledge our colleagues associated with the Integrated History of the today s environs. Archeology is in a position to recognize low- People on Earth (IHOPE) initiative. Alphawood, Wenner-Gren Grant #7799, technology adaptations—as assessed by current measures of National Science Foundation Grant BCS-0810118, and the University of Cin- technology—systems noted for their resilience and sustainability cinnati funded this research.

1. Fisher CT, Hill JB, Feinman GM (2009) The Archaeology of Environmental Change: 20. Laporte JP, Fialko V (1985) Reporte Arqueologico (1979–1984): Mundo Perdido y Socionatural Legacies of Degradation and Resilience (University of Arizona Press, Zonas de Habitacion, Tikal, Peten (Ministerios de Educacion y Communicaciones, Tucson, AZ). Transporte y Obras Publicas, Guatemala City, Guatemala). 2. Hartshorn GS, Robert N (1984) : Country Environmental Profile, a Field Study (R. 21. Gallopin GG (1990) Water Storage Technology at Tikal, Guatemala (University of Nicolait & Associates, Belize City, Belize). Cincinnati, Cincinnati). 3. Barrientos C, Fernandez VH (1998) Water and Population Dynamics: Case Studies and 22. Aiuvalasit MJ, Neely JA, Bateman MD (2010) New radiometric dating of water man- Policy Implications (American Association for the Advancement of Science, Wash- agement features at the prehistoric Purrón Dam Complex, Tehuacán Valley, , – ington, DC). México. J Archaeol Sci 37:1207 1213. 4. Schwartz NB (1990) Forest Society: A Social History of Petén, Guatemala (University of 23. Tankersley KB, et al. (2011) Evidence for volcanic ash fall in the Maya Lowlands from – Pennsylvania Press, Philadelphia). a reservoir at Tikal, Guatemala. J Archaeol Sci 38:2925 2938. 5. Scarborough VL (1994) Maya water management. Res Explor 10:184–199. 24. Luzzaddeer-Beach S (2000) Water resources of the Maya. Geogr Rev 90: – 6. Scarborough VL (2003) The Flow of Power: Ancient Water Systems and Landscapes 493 510. (SAR Press, Santa Fe, NM). 25. Scarborough VL, Connolly RP, Ross SP (1994) The pre-Hispanic Maya reservoir system – 7. Scarborough VL, Burnside WR (2010) Complexity and sustainability: Perspectives from at , Peten, Guatemala. Ancient Mesoam 5:97 106. 26. Akpinar-Ferrand E, Dunning NP, Jones JG, Lentz DL (2011) Aguadas as ancient Maya the ancient Maya and the modern Balinese. Am Antiq 75:327–363. water sources. Ancient Mesoam 23:85–101. 8. Fletcher R (2009) Low-density, agrarian based urbanism: A comparative view. Insights. 27. Dunning NP, Beach TP, Luzzadder-Beach S (2012) Kax and kol: Collapse and resilience Institute of Advanced Study. Durham Univ J 2:1–19. in lowland . Proc Natl Acad Sci USA 109:3652–3657. 9. Scarborough VL, Lucero LJ (2010) The non-hierarchical development of complexity in 28. Mueller AD, et al. (2009) Climate drying and associated forest decline in the lowlands the semitropics: Water and cooperation. Water Hist 2:185–205. of northern Guatemala during the late Holocene. Quat Res 71:133–141. 10. Culbert TP, Rice DS (1990) Precolumbian Population History in the Maya Lowlands 29. Hodell DA, Brenner M, Curtis JH, Guilderson T (2001) Solar forcing of drought fre- (University of New Mexico Press, Albuquerque, NM). quency in the Maya lowlands. Science 292:1367–1370. 11. Scarborough VL (2003) How to interpret an ancient landscape. Proc Natl Acad Sci USA 30. Haug GH, et al. (2003) Climate and the collapse of Maya civilization. Science 299: – 100:4366 4368. 1731–1735. 12. Harrison PD (1999) The Lords of Tikal: Rulers of an Ancient (Thames and 31. Medina-Elizalde M, Rohling EJ (2012) Collapse of Classic Maya civilization related to Hudson, New York). modest reduction in precipitation. Science 335:956–959. 13. Culbert TP (2003) Tikal: Dynasties, Foreigners, and Affairs of State: Advancing Maya 32. Demarest AA, Rice PM, Rice DS (2004) The Terminal Classic in the Maya Lowlands: Archaeology (School of Advanced Research Press, Santa Fe, NM). Collapse, Transition, and Transformation (University Press of Colorado, Boulder, CO). 14. Inglesiz MJ (1987) Ponce de León. PhD thesis (Universidad Complutense de Madrid, 33. Ford A, Rose WI (1995) Volcanic ash in ancient of the limestone low- ANTHROPOLOGY Madrid). lands: Implications for prehistoric volcanic activity in the Guatemala highlands. J 15. Scarborough VL, Gallopin GG (1991) A water storage adaptation in the Maya low- Volcanol Geotherm Res 66:149–162. lands. Science 251:658–662. 34. Dahlin BH, Litzinger WJ (1986) Old bottle, new wine—the function of in the 16. Carr RF, Hazard JE (1961) Map of the Ruins of Tikal, El Peten, Guatemala (University Maya lowlands. Am Antiq 51:721–736. Museum, University of Pennsylvania, Philadelphia). 35. Lentz DL, Hockaday B (2009) Tikal timbers and : Ancient Maya agroforestry 17. Johnstone KJ (2004) Lowland Maya water management practices. Geoarchaeology and the end of time. J Archaeol Sci 36:1342–1353. 19:265–292. 36. Smith EA, Wishnie M (2000) Conservation and subsistence in small-scale societies. 18. Fialko V (1988) Mundo Perdido, Tikal. Mayan 4:13–24. Annu Rev Anthropol 29:493–524. SCIENCE 19. Coe WR (1990) Excavations in the Great Plaza, North , and North Acropolis of 37. Puleston DE (1983) The Settlement Survey of Tikal (University Museum, University

Tikal (University Museum, Philadelphia). of Pennsylvania, Philadelphia). SUSTAINABILITY

Scarborough et al. PNAS | July 31, 2012 | vol. 109 | no. 31 | 12413 Downloaded by guest on September 28, 2021