Geoarchaeological Investigations in Mesoamerica Move into the 21st Century: A Review

33
Review Article Geoarchaeological Investigations in Mesoamerica Move into the 21st Century: A Review Nicholas P. Dunning, 1, * Carmen McCane, 1 Tyler Swinney, 2 Matthew Purtill, 3 Jani Sparks, 4 Ashley Mann, 1 Jon-Paul McCool, 1 and Chantal Ivenso 1 1 Department of Geography, University of Cincinnati, Cincinnati, Ohio 2 Department of Anthropology, University of Cincinnati, Cincinnati, Ohio 3 Department of Geology and Geography, West Virginia University, Morgantown, West Virginia 4 Department of Geology, University of Cincinnati, Cincinnati, Ohio Correspondence * Corresponding author; E-mail: [email protected] Received 13 November 2013 Revised 7 January 2015 Accepted 8 January 2015 Scientific editing by Gary Huckleberry Published online in Wiley Online Library (wileyonlinelibrary.com). doi 10.1002/gea.21507 Over the past thirty years, geoarchaeology has moved from the fringe to main- stream status within Mesoamerican archaeological investigations. This review focuses on works published since the year 2000. Five themes are identified as central to recent studies: (1) the correlation of environmental change and cultural history; (2) anthropogenic environmental impacts; (3) ancient land cover, land use, and diet; (4) archaeological prospection; and (5) provenance studies. These themes are often interwoven in the application of complex sys- tems approaches that allow scientists to more accurately model the intricacies of ancient human–environment interactions. C 2015 Wiley Periodicals, Inc. INTRODUCTION Geoarchaeology has rapidly emerged as a critical com- ponent in the investigation of the ancient cultures of Mesoamerica. Although environmentally centered stud- ies in Mesoamerican civilization can be found as early as one hundred years ago (e.g., Bennett, 1926; Cook, 1931; Huntington, 1917; Sapper, 1896), more concerted efforts truly began in the 1970s and the pace, depth, and sophis- tication of such investigations has increased substantially in the 21st century. Given the quantity and breadth of geoarchaeological approaches that have emerged in the past two decades, our review cannot pretend to be com- prehensive. Rather, we focus chiefly on work published since 2000. Five principal themes that have received spe- cial attention in these investigations are: (1) the corre- lation of environmental change and cultural history; (2) anthropogenic environmental impacts; (3) ancient land cover, land use, and diet; (4) archaeological prospec- tion; and (5) provenance studies. Although these themes will be treated separately below, it is critical to remem- ber that the first three in particular are, in fact, strongly interrelated and to some degree inseparable. In recogni- tion of this interconnectedness, scientists are increasingly adopting complex systems perspectives to model the dy- namic interaction of natural and cultural processes that have played out in Mesoamerica (e.g., Dunning & Beach, 2004; Turner & Sabloff, 2012). The boundaries of Mesoamerica have not been defini- tively delimited, nor is there universal agreement about the northern and southern extremes of this cultural re- gion. More liberal boundaries extend to the vicinity of Lake Nicaragua in the south and the Yaqui River in Sonora, Mexico in the northwest. We follow a more con- servative placement of the boundaries around the Gulf of Fonseca in El Salvador in the south and the Sinaloa River Valley in Mexico in the north. 1 Even this more constricted region is remarkably environmentally diverse (Figure 1). In the north past plate divergence produced horst and graben terrain and many interior basins that 1 For a summary of recent geoarchaeological work in other parts of upper Central America see Sheets 2009. For work on the northern fringes of Mesoamerica see Metcalfe 2006. Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C 2015 Wiley Periodicals, Inc. 167

Transcript of Geoarchaeological Investigations in Mesoamerica Move into the 21st Century: A Review

Review Article

Geoarchaeological Investigations in Mesoamerica Move into the21st Century: A ReviewNicholas P. Dunning,1,* Carmen McCane,1 Tyler Swinney,2 Matthew Purtill,3 Jani Sparks,4 Ashley Mann,1

Jon-Paul McCool,1 and Chantal Ivenso1

1Department of Geography, University of Cincinnati, Cincinnati, Ohio2Department of Anthropology, University of Cincinnati, Cincinnati, Ohio3Department of Geology and Geography, West Virginia University, Morgantown, West Virginia4Department of Geology, University of Cincinnati, Cincinnati, Ohio

Correspondence*Corresponding author; E-mail:

[email protected]

Received13 November 2013

Revised7 January 2015

Accepted8 January 2015

Scientific editing by Gary Huckleberry

Published online in Wiley Online Library

(wileyonlinelibrary.com).

doi 10.1002/gea.21507

Over the past thirty years, geoarchaeology has moved from the fringe to main-stream status within Mesoamerican archaeological investigations. This reviewfocuses on works published since the year 2000. Five themes are identifiedas central to recent studies: (1) the correlation of environmental change andcultural history; (2) anthropogenic environmental impacts; (3) ancient landcover, land use, and diet; (4) archaeological prospection; and (5) provenancestudies. These themes are often interwoven in the application of complex sys-tems approaches that allow scientists to more accurately model the intricaciesof ancient human–environment interactions. C© 2015 Wiley Periodicals, Inc.

INTRODUCTION

Geoarchaeology has rapidly emerged as a critical com-ponent in the investigation of the ancient cultures ofMesoamerica. Although environmentally centered stud-ies in Mesoamerican civilization can be found as early asone hundred years ago (e.g., Bennett, 1926; Cook, 1931;Huntington, 1917; Sapper, 1896), more concerted effortstruly began in the 1970s and the pace, depth, and sophis-tication of such investigations has increased substantiallyin the 21st century. Given the quantity and breadth ofgeoarchaeological approaches that have emerged in thepast two decades, our review cannot pretend to be com-prehensive. Rather, we focus chiefly on work publishedsince 2000. Five principal themes that have received spe-cial attention in these investigations are: (1) the corre-lation of environmental change and cultural history; (2)anthropogenic environmental impacts; (3) ancient landcover, land use, and diet; (4) archaeological prospec-tion; and (5) provenance studies. Although these themeswill be treated separately below, it is critical to remem-ber that the first three in particular are, in fact, strongly

interrelated and to some degree inseparable. In recogni-tion of this interconnectedness, scientists are increasinglyadopting complex systems perspectives to model the dy-namic interaction of natural and cultural processes thathave played out in Mesoamerica (e.g., Dunning & Beach,2004; Turner & Sabloff, 2012).

The boundaries of Mesoamerica have not been defini-tively delimited, nor is there universal agreement aboutthe northern and southern extremes of this cultural re-gion. More liberal boundaries extend to the vicinity ofLake Nicaragua in the south and the Yaqui River inSonora, Mexico in the northwest. We follow a more con-servative placement of the boundaries around the Gulfof Fonseca in El Salvador in the south and the SinaloaRiver Valley in Mexico in the north.1 Even this moreconstricted region is remarkably environmentally diverse(Figure 1). In the north past plate divergence producedhorst and graben terrain and many interior basins that

1For a summary of recent geoarchaeological work in other partsof upper Central America see Sheets 2009. For work on thenorthern fringes of Mesoamerica see Metcalfe 2006.

Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc. 167

GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA DUNNING ET AL.

Figure 1 Map of Mesoamerica showing general physiography and some of the places highlighted in the text.

have been the foci of much human settlement for mil-lennia. Across much of Mesoamerica the complex con-vergence of the Caribbean, Cocos, North American, andPacific plates has created a mosaic of physiographic zonesthat include merging volcanic arcs, orographic uplift,and subsidence. Adjacent lowlands and basins are therepositories of tens of millions of years of sediment andaccretion. Major physiographic regions include the Mex-ican Plateau and associated ridges and basins, the abut-ting uplifted masses of the Sierra Madres Oriental andOccidental, the Trans-Mexican Volcanic Arc, the CentralAmerican Volcanic Arc and associated folded and faultedhighlands, and the huge, partially emerged carbonateYucatan Platform.

Mesoamerica straddles the Tropic of Cancer, a latitu-dinal position that typically experiences an annual shiftfrom wet to dry seasons as the Intertropical Conver-gence Zone (ITCZ) shifts northward and southward andmutes temperature extremes. The distribution of rain-fall is also affected by steep elevation gradients createdby the region’s mountain systems generating significant

rain shadows in many interior and western coastal ar-eas where Easterly Waves cannot bring wet season rains.On the other hand, windward slopes and eastern coastlowlands experience many months of copious rainfall.Thus, Holocene vegetation zones range from tropical rainforests to true desert. When people first entered theregion in the Pleistocene, conditions across most low-land areas are believed to have been typically cooler anddrier than the present whereas many interior basins werecooler, but wetter. It was within this shifting mosaic ofenvironments that the cultures of ancient Mesoamericadeveloped.

Pre-Hispanic human occupation of Mesoamerica is of-ten divided into the chronological periods as indicated inTable I. Some scholars prefer different terminology forsome time periods, most notably “Formative” in placeof “Preclassic.” Population growth (and decline) and cul-tural developments were not uniform across the entireregion. In each period, some sites and regions experi-enced greater growth or more complete abandonmentthan their neighbors while others experienced much less

168 Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc.

DUNNING ET AL. GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA

Table I Mesoamerican pre-Hispanic chronological periods.

Major Period Sub-period Ages

Paleoindian Pre-7000 B.C.

Archaic 7000 – 2200 B.C.

Preclassic

Early Preclassic 2200 – 1000 B.C.

Middle Preclassic 1000 – 400 B.C.

Late Preclassic 400 B.C. – 100 A.D.

Terminal Preclassic 100 – 250 A.D.

Classic

Early Classic 250 – 600 A.D.

Late Classic 600 – 770 A.D.

Terminal Classic 770 – 900 A.D.

Postclassic

Early Postclassic 900 – 1250 A.D.

Late Postclassic 1250 – 1500 A.D.

fluctuation. Some regions or parts of regions were aban-doned for hundreds of years at a time, while others werenever abandoned.

ENVIRONMENTAL CHANGE AND CULTUREHISTORY

Over the past decade plus, the procurement of deep sedi-ment cores from several Mesoamerica lakes has given sci-entists the ability to reach back into the last portions ofthe Pleistocene and, thus, contextualize Holocene envi-ronmental change in a new light. These older sedimentsand their embedded proxy data also help to distinguishnatural from anthropogenic change in some areas. An 85ka sediment record from Lake Peten Itza in Guatemalaincludes distinct dry climate episodes (marked by gyp-sum precipitation) during stadials and wetter conditionsduring inter-stadials (Hodell et al., 2008). Extreme arid-ity is clearly associated with Heinrich Events and cold seasurface temperatures in the North Atlantic, reduced cir-culation, and the southward displacement of the ITCZ.Oscillations between wetter and drier conditions are alsoevident in the last 10 ky of the Pleistocene before fi-nally giving way to more persistently warmer, moisterconditions in the Holocene. Nevertheless, drier pulsesare evident within the Holocene and the data fromthe Pleistocene raises the question that some of the in-creased sedimentation resulting from accelerated soil ero-sion following Maya occupation may have been triggeredby episodes of aridity and canopy opening as well asby anthropogenic deforestation (Hillesheim et al., 2005;Mueller et al., 2009).

A 52,000 year sediment record in the Zacapu Basinin Michoacan, Mexico shows moist conditions from52 ka until a drying trend initiates around 35 ka

(Ortega et al., 2002). An undefined event occurs at 25ka which the authors suggest as a depositional hiatusdue to tectonic modification of the basin. Further dry-ing continues throughout the Pleistocene–Holocene tran-sition from 14 to 4.8 ka. After 4.8 ka the authors wereunable to separate climatic and anthropogenic environ-mental changes. A 48,000 year sediment record fromnearby Lake Patzcuaro in the volcanic highlands of Mi-choacan in western Mexico exhibits comparatively littlefluctuation in atmospheric moisture. In a 17,000 yearrecord from Lake Zirahuen, Lozano-Garcia et al. (2013)found evidence for cooler and drier conditions in thePleistocene with a transition to moister conditions be-ginning around 13.5 ka. Particularly warm periods werenoted for 9.5–9.0 ka, a notable cold snap around 8.2 ka,and an especially moist period from 7.5–7.1 ka matchingrecords from elsewhere in Mesoamerica and the NorthAtlantic. In contrast to other paleoenvironmental recordsfrom central Mexico, the Lake Patzcuaro study indicatesrelatively moist conditions during the Last Glacial Maxi-mum, whereas the other studies suggest generally greateraridity. After 10 ka, the lake began to experience in-creased eutrophication possibly as the result of losing itsformer connectivity with the Lerma River. Sequences ofpaleosols associated with paleolakes in the Basin of Mex-ico indicate that drying conditions in the early Holocenegave way to greater moisture in the mid-Holocene andbrief lake resurgence, before drying conditions resumedin the late Holocene (Sedov et al., 2001). The gradual re-treat of the basin paleolakes and the development of bet-ter drained soils set the stage for the regional expansionof agriculture in the late Holocene (Sedov et al., 2010).

Perhaps no issue has received more attention in Mayaarchaeology over the past two decades than the roledroughts have played in the course of Maya civilization.Proxy data were originally obtained from a pair of lakesin the northeast Yucatan Peninsula (Chichancanab andPunta Laguna) and focused on changing composition ofsediments (e.g., precipitation of gypsum) and oxygen iso-tope values in ostracods and gastropods (Curtis, Hod-dell & Brenner, 1996; Hodell, Brenner, & Curtis, 1995).These early studies and other syntheses (e.g., Gill, 2000)were greeted with considerable skepticism by many Mayaarchaeologists who viewed them as deterministic andalso questioned the chronological precision or correla-tions between apparent drought and the archaeologi-cal record (e.g., Aimers, 2007; Demarest, Rice, & Rice,2004; McAnany & Gallareta Negron, 2010). However,evidence for episodic or cyclical droughts has mountedover the past decade with additional proxies, additionalstudy sites, and enhanced chronological control. Thesedata have led to an acceptance by most Maya archae-ologists acknowledging the damaging role droughts may

Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc. 169

GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA DUNNING ET AL.

have played in the course of Maya civilization, albeit howcausal that role was versus other cultural factors remainscontentious (cf. the various chapters in Iannone, 2014).

Limnological studies continue to add important data tounderstanding climate change in the Maya Lowlands. Inthe northern Maya Lowlands more detailed studies havebeen conducted at Laguna Chichancanab and LagunaPunta Laguna (Hodell, Brenner, & Curtis, 2005; Hodellet al., 2007), Lake Coba (Leyden, Brenner, & Dahlin,1998), Aguada Xcaamal (a clay-plugged cenote; Hodellet al., 2005), and Lake Tzib (Carrillo-Bastos et al., 2010).In the southern Maya Lowlands, further study of LakePeten Itza, Lake Salpeten, and Lake Quexil has not onlypushed the paleoenvironmental record back tens of thou-sands of years, but has also resulted in a more refinedchronology (Anselmetti et al., 2006; Mueller et al., 2009;Rosenmeier et al., 2002). While earlier studies of the cen-tral Peten lakes were unable to distinguish between an-thropogenic and climate change signals, newer studiesusing multiple proxies and enhanced dating have hadmore success. A brief summary of important findings fol-lows. Since the onset of the Holocene, climate conditionswere considerably variable especially with respect to ef-fective rainfall, that is, the amount of rainfall that fallsduring the typical growing season (Brenner et al., 2002;Hodell, Brenner, & Curtis, 2005). Wetter conditions inthe mid-Holocene gave way to a general, progressive dry-ing trend that began about 4000 years ago (Mueller et al.,2009). For the late Holocene, it has been proposed thatshort-term (centennial scale) fluctuations in rainfall mayhave been driven by the 208-year cycle of solar output(Hodell et al., 2001; Wahl et al., 2006). Statistical analysishas been used to question the validity of the role of solarperiodicity in Maya droughts, but does not question thefact that severe droughts have occurred (Carleton, Camp-bell, & Collard, 2014).

Mounting paleoenvironmental data indicate that pe-riods of particularly intense drought may have afflictedwide areas of the Maya Lowlands in the 4th century B.C.,and the 2nd, 9th, and 11th centuries A.D. Notably, thesewere not single “megadroughts,” but generally a tem-porally concentrated series of droughts (Brenner et al.,2003). Although most attention has been given to theTerminal Classic droughts of the 9th century A.D., ear-lier droughts are beginning to garner more attention,especially those centered on the second century A.D.(Figure 2). These droughts had a devastating effect onLate Preclassic Maya populations, including abandon-ment of many large and small centers, changes that mayhave helped set the stage for the emergence of ClassicMaya civilization (Dunning et al., 2002, 2014a; Hansenet al., 2002; Wahl, Byrne, & Anderson, 2014; Wahl et al.,2006, 2007).

Strong corollary proxy paleoclimate data have beenobtained from outside of Mesoamerica, most notablyfrom the Cariaco Basin offshore Venezuela (Haug et al.2003; Peterson & Haug 2005). Anaerobic conditions inthis deep basin have preserved biannually laminatedwet/dry season sediments that record droughts in thelower Caribbean that seem to correlate well with theMaya Lowlands record; droughts in both places are linkedto the timing and strength of seasonal shifts in the ITCZ.

Detailed records of paleoprecipitation have been ob-tained from speleothems harvested from caves in Yu-catan, Mexico, and Belize. From cave laminae insouthern Belize, Webster et al. (2007) used color, lumi-nescence, δ 13C, and δ 18O as precipitation proxies to sug-gest Preclassic climate flux from drought to pluvial, LatePreclassic (5 B.C. and A.D. 141) severe droughts, ClassicPeriod wetter conditions sandwiching a drought in theMiddle Classic (A.D. 517), and Late Classic through Post-classic (A.D. 780, 910, 1074, and 1139) severe droughts.Analysis of a speleothem from Yok Balum Cave in Belizealso show clear periodic drought in the Terminal Clas-sic, but with the most extreme peak in the period be-tween A.D. 1000 and 1020, a period not flagged by mostother Maya paleoclimate records (Kennett et al., 2012).In northern Yucatan paleoprecipitation records gatheredfrom a speleothem (dubbed “Chaak” for the Maya raingod) found in Tzab Na Cave near Tecoh indicate thatthe northern lowlands experienced a similar chronolog-ical pattern of decreased rainfall including a series ofeight multi-year droughts, 3–18 years in length, betweenA.D. 806 and 935 (Medina-Elizade et al., 2010; Medina-Elizade & Rohling, 2012). These studies show the highpotential of speleothems as paleoclimate proxies in thekarst Maya Lowlands, including the possibility of pro-viding measures of annual precipitation proxies, by farthe finest chronological resolution available. However,it needs to be borne in mind that these records can beseverely affected by localized land use changes that mayaccentuate drought signals for some periods. For exam-ple, land use change, possibly including the constructionof plastered surfaces above a cave, will undoubted effectinfiltration and percolation. Hence, speleothem records,despite their potential for fine-grained chronological res-olution, will be most meaningful when multiple recordscan be obtained for a locality and region. Other at-tempts to analyze speleothems are ongoing, but havebeen limited by dating problems associated with low Thlevels and disruptions found in individual speleothemformation processes (e.g., Dahlin, 2011; Smyth et al.,2011).

The record of recurrent aridity also includes many re-gional soils. The development of Vertisols on alluviumin the lower Usumacinta drainage has been interpreted

170 Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc.

DUNNING ET AL. GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA

Figure 2 Aguada Tintal, San Bartolo, Guatemala: a) Coring aguada floor in 2005; b) sediment composition and pollen frequency of Aguada Tintal core

(adapted from Dunning et al. 2014a, Figure 4). The oxidized sediment in Zone 3 marks a Late/Terminal Preclassic drought above an earlier cultivated

landscape; dredging removed later Classic period sediment, followed by abandonment and reforestation. Pollen analysis by John G. Jones.

to indicate relative aridity during the Preclassic (Solis-Castillo et al., 2013). In northern Quintana Roo, thepunctuated development of Calcisols in depression soilsindicates fluctuating moisture levels that likely reflectclimate change, though changes in wetland hydrologyalso likely contributed to varying pedogenic processes(Wollwage et al., 2012). Cave sediments may also reflectcyclical aridity (Polk, van Beynen, & Reeder, 2007).

Although the paleoenvironmental proxy data for an-cient droughts in the Maya Lowlands is now compelling,it is also clear that the actual distribution of precipitationwas highly variable across both space and time (as it is to-day, especially during drought years). Such variability isundoubtedly part of the reason that there is an imperfectcorrelation between the drought record and Maya culturehistory, especially from the perspective of individual sitesand sub-regions. Certainly some regions and sites weremore vulnerable to drought, particularly the Elevated In-terior Region of the Yucatan Peninsula (EIR in Figure 1)where access to both surface and subterranean water wasnegligible (Dunning, Beach, & Luzzadder-Beach, 2012).The complex series of historical events that played outacross the Maya Lowlands over time make it clear thatthe ability of different communities and leaders to copeand adapt to drought and other stresses indeed variedgreatly (Aimers & Hodell, 2011; Iannone, 2014; Yeager &Hodell, 2008). Nevertheless, it is also clear that droughtshave had devastating effects on the Maya Lowlands andits inhabitants for millennia (Gill et al., 2007). The Mayathemselves may well have also contributed to the inten-sification of droughts by removing regional forest coverinterfering with water capture and regeneration via tran-spiration (Cook et al., 2012; Griffin et al. 2014; Oglesbyet al., 2010; Shaw, 2003).

Pollen and sediment composition in a core taken fromLake Pompal in the Tuxla Mountains of Veracruz indicatelower water levels and forest canopy opening consistentwith a drying trend after 2600 14C yr B.P. and peakingaround 1300 14C yr B.P., though the record is complicatedby the concurrent intensification of regional agricultureduring this same period (Goman & Byrne, 1998). Farthernorth in Veracruz, Hudson and Heitmuller (2003) docu-mented the evolving system of natural river levees in thelower Panuco River Basin linking these to changing basinhydrology through the Late Pleistocene and Holocene.

A speleothem from Cueva del Diablo north of Aca-pulco in the Mexican state of Guerrero was analyzedby Bernal et al. (2011) for δ18O variation within thecarbonate, and used to infer the environmental condi-tions during the periods of speleothem formation andgrowth. Moist conditions are largely tied with the influxof glacial meltwater to the Atlantic with precipitation de-creasing from the Caribbean as a result of cooler Atlanticsea surface temperatures driving a southern migration ofthe ITCZ. Pacific climate patterns are suggested to inter-fere with Caribbean and Atlantic patterns after � 4.3 ka,specifically an overall decrease in precipitation resultingfrom increased influence of El Nino-Southern Oscillation(ENSO) and marked periods of moisture decrease dur-ing ENSO warm phases. The two largest climatic anoma-lies of the Holocene at 10.3 ka and 8.2 ka, expressedelsewhere by low lake levels at Patzcuaro (the 10.3 kaanomaly) and glacial advances on the Iztaccihuatl Vol-cano, are represented as hiatuses of speleothem growthfollowed by lower δ18O values. Association with theseknown climatic anomalies driving the hiatuses, as op-posed to other processes, is supported by uraniumratios representing increases in the lighter isotope.

Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc. 171

GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA DUNNING ET AL.

Unfortunately, the speleothem record terminates at 1.24ka and could not provide paleoclimate data on Termi-nal Classic and Postclassic droughts that plagued parts ofMesoamerica.

Not unexpectedly, ENSO cycles are fairly apparent inthe sediment records of several west and central Mexi-can lakes (Metcalfe, 2006). Historical records of droughtsafflicting central Mexico between A.D. 1450 and 1900indicate a close correlation between ENSO cycles anddroughts (Mendoza et al., 2005).

Sediment cores recovered from lakes Tixtla and Huizil-tepec, in the West Mexican state of Guerrero documentforest cover change from � 2.70 ka to 1.95 ka indicativeof drier conditions giving way to higher humidity thanthe modern environment, � 1.95 ka to 1.07 ka (Berrıoet al., 2006). In Oaxaca, pollen and geomorphic data re-covered from the lower Rio Verde suggest that periodsof increased aridity and canopy opening in highland wa-tersheds centered around 6.17 ka and 5.34 ka led to in-creased erosion in upper portions of the watershed andaggradation downstream prior to any significant humanimpacts (Joyce & Mueller, 1997). Increased ENSO fre-quencies are also thought to underlie an increase in hur-ricane strikes in coastal Oaxaca in the mid-late Holocene(Goman, Joyce, & Mueller, 2005).

The most detailed record of past precipitation recoveredto date in Mesoamerica comes from Montezuma Bald-cypress (Taxodium mucronatum) growing in Queretaro incentral Mexico (Stahle et al., 2011). Growth rings in thesetrees have provided a 1238 year rainfall record includ-ing droughts at A.D. 810, A.D. 860, and several betweenA.D. 897–922, a record that only partly correlates withthat of the Maya Lowlands but clearly indicates that Ter-minal Classic droughts also affected central Mexico. Im-portantly, periods of droughts are also evident for A.D.1149–1167, during the decline of the Toltec state, andA.D. 1514–1539 during the Spanish conquest of the Aztecstate.

In the Guatemalan Highlands analysis of diatoms in along sediment core from Lake Amatitlan indicate loweredlake levels during two periods (250 B.C. to A.D. 125, andA.D. 875 to 1375), probably indicative of periods of rela-tive aridity (Velez et al., 2011). On the southeastern edgeof Mesoamerica Dull (2004) obtained an 8000-year sedi-ment record from a crater lake in the Sierra de Apaneca,El Salvador. Vegetation and chemical proxies indicate awarm, humid climate during the mid-Holocene optimum(a warm period 9.0 and 5.0 ka), the onset of a cooling,drying trend around 5.5 ka, then stabilization into cli-mate similar to the present day around 3.5 ka. Tephrafrom the Tierra Blanca Joven (TBJ) eruption of IlopangoVolcano is prominent within the core, followed abruptlyby indications of local abandonment and reforestation.

Volcanic eruptions have been linked to climate dis-ruptions and longer term changes around the globe. Thelargest known stratospheric aerosol loading event of thepast 2000 years occurred in A.D. 536 and disrupted globalclimate until around A.D. 550 and creating agriculturalproblems in many parts of the world, including possi-bly within the Maya Lowlands (Dahlin & Chase, 2014).Although the catastrophic TBJ eruption of Ilopango Vol-cano in El Salvador was originally dated to A.D. 260± 114 yrs, a new suite of radiocarbon dates indicatethe eruption occurred sometime between A.D. 440 and550 suggesting a possible link to the A.D. 536 event(Dull et al., 2010; Kutterolf, Freundt, & Perez, 2008). Re-gardless of whether or not the TBJ eruption occurredin A.D. 536, it had catastrophic effects across south-east Mesoamerica leading to widespread abandonmentof settlements and migration to neighboring areas (Dull,Southon, & Sheets, 2001).

Given the location of Mesoamerica astride conver-gent tectonic plate boundaries, it is not surprisingthat volcanism and earthquakes played a significantrole in shaping the region’s cultural history—as wellas the cultures themselves, particularly their cosmolo-gies (Gill & Keating, 2002; Sheets, 1999, 2008). Al-though much less studied, earthquake activity also clearlyinfluenced Mesoamerican cultures and their histories(Kovach, 2004). In some instances, effusive volcaniceruptions have also provided archaeologists with as-tounding opportunities where towns and villages havebeen rapidly entombed by ejecta. The village of Cerenin El Salvador has now been the subject to decades ofproductive investigation (further discussion follows be-low). Another buried site that has revealed a great dealabout daily life in Mesoamerica is Tetimpa on the slopesof Popocatepetl Volcano in Puebla, Mexico (Plunket &Urunuela, 1998). Sheets (2009) astutely points out thatgiven the widespread deposition of tephras dating tothe period of human occupation in Mesoamerica, otherburied sites surely exist and await discovery; he suggeststhat deposits around Ceboruca Volcano (see Sieron &Siebe, 2008) in Nayarit might be an especially promisingplace to look.

The diffusion of volcanic ash is also now known to haveplayed a significant role in the agricultural potential ofnon-volcanic regions of Mesoamerica. Ash is likely theprimary inorganic component in many soils of the south-ern Maya Lowlands (Tankersley et al., 2011, 2015), and isa significant contributor of clay minerals, silica, and nutri-ents to soils in the northern Maya Lowlands in additionto Saharan dust (Das et al., 2011; Cabadas-Baez et al.,2010).

Another natural hazard with a frequent recurrence in-terval in Mesoamerica is hurricanes and tropical storms.

172 Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc.

DUNNING ET AL. GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA

Paleotempestology is still in its infancy in the region,but initial findings are promising. McClosky and Keller(2009) uncovered a 5000-year record of hurricane strikesin two coastal lagoons in southern Belize. Unfortunately,an enormous hurricane in the 15th century A.D. appar-ently scoured large amounts of sediment from both la-goons, creating a 1700-year gap in the record spanningmuch of the course of ancient Maya civilization. A morecomplete sediment record has been recovered in a se-ries of cores taken in the Blue Hole, a submerged Pleis-tocene sinkhole on the Belizean Barrier Reef (Gischleret al., 2008). These cores indicate numerous hurricanesover the past 1500 years (the limit of the cores), includ-ing two major strikes in the second half of the 8th centuryA.D. Further analysis of the cores indicate that hurricanefrequency appears to correlate more strongly with globalclimate than with local conditions (Denommee, Bentley,& Drowxler, 2014).

Dunning and Houston (2011) examined ethnographic,historic, hieroglyphic, and geomorphic evidence for theeffects of hurricanes on Maya history and culture andon the Maya Lowlands environment. Given the highfrequency of hurricane strikes across the Maya Low-lands, it is not surprising that these sometimes devas-tating storms are deeply embedded in Maya cosmologywarranting their own deity, a malevolent “sparking ser-pent.” In the historic and modern era, hurricanes havegenerated major flooding even in interior regions andare a powerful geomorphic agent presenting clear recordsin fluvial sediments. Targeted examination of fluvial se-quences and coastal deposits in other parts of the MayaLowlands (e.g. Beach et al., 2009) and elsewhere inMesoamerica (e.g., Goman, Joyce, & Mueller, 2005; Neffet al., 2006a) will undoubtedly reveal other records ofhurricane strikes. Mud layers within a speleothem fromnorthern Yucatan have been interpreted as proxies fortropical storms (Frappier et al., 2014), but are notablysuppressed during intense drought episodes when lowregional groundwater levels inhibited the filling of caveswith water and sediment during storms.

Another environmental change affecting Mesoamer-ica owes its origin to orbitally driven, long-term climatefluctuations that influence global sea levels. Fluctuatingsea levels are especially important on the low-lying Yu-catan Platform including the submergence of large por-tions with sea-level rise at the end of the Pleistocene(� 17,000–10,000 years ago) and during the earlyHolocene (� 10,000–5000 years ago). Although the rateof sea level rise slowed dramatically by the middleHolocene, it continued to affect coastal regions and res-idents (Van Hengstum et al., 2009).

Geoarchaeological investigations around La Joyancaalong the Rio San Pedro Matir, a major tributary of the

Usumacintla in the northwestern Peten, Guatemala haveuncovered evidence for dynamic environmental changesassociated with the Pleistocene-Holocene transition andwith the ancient Maya occupation of the region (Galopet al., 2004; Metailie et al., 2003). In this region of mutedridge and valley terrain, late Pleistocene stream channelsdraining grabens were aggraded, probably as the result ofbase-level changes in the Holocene. Continued changesin the Preclassic and Classic compelled the Maya to adaptto continued slow sea level rise and spread of wetlandsin the coastal plains. In the Holbox Fracture in northernQuintana Roo, wetlands connected to the coast graduallychanged and likely triggered cycles of site abandonmentand reoccupation and Maya attempts to manipulate wet-land hydrology (Fedick et al., 2000; Sedov et al., 2008). Inthe coastal lowlands of northern Belize and adjacent areasof Mexico, sea level rise triggered changes in groundwa-ter chemistry and aggradation in the region’s low gradientriver systems (Figure 3) (Beach et al., 2009; Luzzadder-Beach & Beach, 2009; Luzzadder-Beach, Beach, & Dun-ning, 2012). Associated wetland field systems were some-times adapted to changing hydrology, but at other timesabandoned. Wetland field systems in coastal Veracruzwere likewise affected by sea level rise (Heimo, Siemens,& Hebda, 2004). Sea level rise is also implicated inchanging fluvial hydrology and associated settlement atthe mouths of the Panuco River in northern Veracruz(Hudson, 2004), the Rio Verde in Oaxaca (Goman, Joyce,& Mueller, 2005) and several drainages on the Pacificcoast of Guatemala (Neff et al., 2006).

A natural disaster of the distant past has also had a last-ing impact on the cultural history of part of the north-ern Maya Lowlands. Though buried under later carbon-ate sediments, the massive 65 million-year-old ChicxulubMeteorite Crater underlying the northwestern YucatanPeninsula continues to profoundly influence groundwa-ter movement and the development of water-bearingcaves and sinkholes (cenotes), and larger structural andsolution depressions in the overlying limestone (Hilde-brand et al., 1995; Perry et al., 2010). Differences inbedrock chemistry and topography associated with theburied crater also influenced soil development (Popeet al., 1996). These hydrologic and pedologic factors have,in turn, influenced settlement distribution and historyacross the region (Winemiller, 2007).

ANTHROPOGENIC ENVIRONMENTALCHANGE

The Lake Patzcuaro Basin in Michoacan, Mexico has beenthe focus of a controversy centering on the timing of, andagency behind accelerated soil erosion. Although the lake

Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc. 173

GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA DUNNING ET AL.

Figure 3 Processes involved in wetland field formation as documented in riparian wetlands in northern Belize: a) ditching through peaty wetland; b)

peat formation over paleosol after sea-level and water table rise followed by ditching; c) aggradation from alluvial deposition over peat or paleosol; d)

“gyp-heave” or subsurface gypsum accumulation or pillowing; e) “raising” or aggradation via anthropogenic deposition (adapted from various figures in

Beach et al., 2009, 2015). Note: in reality multiple of these processes may affect a given location.

has been cored multiple times since the 1940s, much ofthe controversy has been in the last 20 years. A study byO’Hara, Street-Perrott, and Burt (1993) based on corestaken from many different parts of the large lake re-ported a positive correlation between Prehispanic popu-lation growth and sedimentation rates within the basinas a whole indicating erosion peaking in the Classic andPostclassic. A detailed study of a portion of the southernbasin by Fisher et al. (2003) based on deep stratigraphicexcavations refuted this correlation finding that soil ero-sion was at its highest during Preclassic land clearanceand settlement, then stabilized as population increasedin the Classic and Postclassic periods and heavy invest-ment was made in agricultural terracing, and finally ac-celerated again in the early Colonial period when regionalpopulation crashed and soil conservation measures were

abandoned. Detailed analyses of four cores taken fromseveral parts of the basin lend support to the originalmodel of accelerating soil erosion correlating with Pre-hispanic population growth and agricultural expansion(Metcalfe et al., 2007). It is likely that the more detailedstratigraphy and chronological control provided by ex-cavations in a part of the southern basin reveal morenuanced, but perhaps localized effects that may not berepresentative of the lake basin as a whole. Similar strati-graphic excavations in other parts of the basin could helpresolve this controversy.

At lakes Tixtla and Huiziltepec, in the West Mexicanstate of Guerrero increases in watershed cultivation (in-dicated by maize pollen) are documented for the firstmillennium A.D. (Berrıo et al., 2006). Small pulses ofincreased soil erosion are indicated sporadically in both

174 Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc.

DUNNING ET AL. GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA

lakes during this period, but the most notable increasein erosion is associated with the post-Conquest period.At Lake Zirahuen disturbance is first noted around 3ka and continues with some oscillations before peak-ing in the Postclassic around 1350 A.D. (Lozano-Garciaet al., 2013; Ortega et al., 2010). However, the evi-dence for anthropogenic erosion is complicated by indica-tions of episodic aridity that may also have induced ero-sion/sedimentation. Similarly, Park et al.’s (2010) studyof two shallow central Mexican maar lakes produced con-flated signals for aridity and anthropogenic erosion withpeaks in both the Preclassic and Classic times. Anotherstudy of Lake Zirahuen and Lake Juanacatlan in Jaliscodocumented increases in erosion and metal contaminantslinked to Colonial agriculture and mining (Davies et al.,2005). In sum, each drainage basin that has been exam-ined exhibits a unique history of anthropogenic change.

In the Maya Lowlands, scholars have for many yearslinked the decline of Classic Maya civilization with soilerosion and environmental degradation (e.g., Bennett,1926; Cooke, 1931; Morley, 1946). Combined paleoen-vironmental and archaeological studies at Lake Yaxhaand other central Peten, Guatemala lakes were used todevelop a more quantitative version of this model thatposited a direct relationship between population growth,forest clearance, and soil erosion (Deevy et al., 1979; Rice,1993). While this model does indeed appear to be rep-resentative of some investigated catchments, it does notseem to be a universal fit. At Lake Yaxha, high sedimen-tation rates may have been linked as much to urbaniza-tion (the large site of Yaxha sprawled across one side ofthe lake) as with agriculture. In many areas, the relation-ship between population growth and land degradation ismuch less direct. For many studied catchments, soil ero-sion peaks not with maximum population numbers (typ-ically in the Late Classic), but during the first centuries offorest clearance and agricultural expansion (Anselmattiet al., 2006; Dunning et al., 1998; Dunning & Beach,2000, 2010). Evidence suggests that early farmers wereless conservative while land pressure was slight, some-times with catastrophic results when sloping, unprotectedland was subject to erosive tropical rainfall (Beach et al.2006). By sometime in the Late Preclassic, soil conser-vation measures such as terracing began to appear insome areas (Beach et al., 2008; Dunning & Beach, 2000;Carazzo, et al., 2007; Garrison & Dunning, 2009; see alsofurther discussion below).

The relationship between population pressure and soilerosion in the Maya Lowlands was sometimes complex.For example, Carozza et al. (2007) document pulses oferosion/sedimentation around La Joyanca, Guatamala.Light population in the Middle Preclassic initiated littleerosion, rising population in the Late Preclassic gener-

ated large amounts of erosion, but greater populationgrowth in the Late Classic produced notably less erosion,presumably because of the implementation of conserva-tion measures. Following Terminal Classic abandonment,a recolonizing Postclassic population renewed the erosioncycle. There is no question that at times some ancientMaya groups generated catastrophic soil erosion as evi-dent in the utter denudation of hillslopes and the burial ofdownslope/downstream locations including older occu-pation sites (Dunning & Beach, 2000; Holley et al., 2000),but more conservative land management was practicedin other places and times and soilscapes were stabilized(Dunning et al., 2009).

An aspect of urbanization in the Maya Lowlands thatundoubtedly contributed to land degradation was thequarrying of limestone both for stone and for makinglime plaster. This form of degradation may have beenmost acute in the Late Preclassic, when architectural stylefavored massive pyramid complexes coated in extremelythick layers of lime stucco, such as in the Mirador Basinarea (Hansen et al, 2002; Wahl et al., 2007).

Not surprisingly, soil erosion linked to land clearance,especially for agriculture, has been documented through-out Mesoamerica. McAuliffe et al. (2001) documentedsevere Prehispanic soil erosion in the Tehuacan Val-ley of central Mexico, probably during both the Clas-sic and Postclassic periods, and continuing into Colonialand modern times. Prehispanic soil conservation terrac-ing was present in some parts of the valley, but notothers. Where present, the terraces were eventually ne-glected and have largely disintegrated. At Tlaxcala, Bore-jsza et al. (2008) uncovered a complex landscape thatexperienced severe erosion in the Preclassic period, fol-lowed by abandonment, then reclamation with extensiveagricultural terracing in the Postclassic (Figure 4). Someof that erosion was generated by swidden agriculture inhigh elevation areas (tierra fria) that were brought intoproduction by either population pressure of political eco-nomic exclusion (Borejsza, Frederick, & Lesure, 2011).Ancient, Colonial, and modern soil erosion on slopinglands has been widespread in central Mexico (Cordova& Parsons, 1997; Heine, 2003) resulting in the burial ofancient land surfaces within basins and valleys (McClungde Tapia et al., 2003; Sanchez-Perez et al., 2013).

Around Laguna Pompal in the Tuxla Mountains ofVeracruz, the first forest clearance for agriculture beganaround 4830 14C yr B.P., but was followed by a hiatusof nearly 1500 years before cultivation resumed around2600 14C yr B.P. (Goman & Byrne, 1998). Forest clear-ance was then pervasive until about 1300 14C yr B.P. re-flecting the spread and intensification of agriculture inthe Preclassic (Formative) and Classic periods, thoughwithout generating a notable increase in the deposition

Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc. 175

GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA DUNNING ET AL.

Figure 4 Idealized diagram of slope and land use

evolution in Tlaxcala, Mexico (based on figures and

discussion in Borejsza et al., 2008): a)

Preclassic/Formative period cultivation; 2) Classic period

degradation and abandonment; c) Postclassic terracing

and stabilization.

of inorganic sediments indicative of accelerated soil ero-sion. An earlier study of nearby Lake Catemaco founda similar record of forest clearance with little increasein sedimentation (Byrne & Horn, 1989). Whether thelack of indicated erosion is the result of local watershedcharacteristics, natural resistance of the region’s Andis-ols to erosion, or ancient soil conservation practices isunknown.

The Rio Verde drainage basin includes the NochixtlanValley, Ejutla Valley, and the Valley of Oaxaca. Manyyears of research have demonstrated that episodes of soilerosion were triggered by land clearance for agriculturein the upper watersheds during the Preclassic as early as1500 B.C. and continued episodically through the Clas-sic period, punctuated by periods of stability and soil for-mation in individual tributaries (Joyce & Goman, 2012;Joyce & Mueller, 1992, 1997; Perez Rodrıguez, Anderson,

& Neff, 2011). Erosion in the upper watershed initiatedsedimentation in the lower Rio Verde drainage, creatingalluvial soils that became a new focus for agricultural in-tensification and population concentration, though thisaggradation also created new flooding problems. Depo-sition at the coastal outlet of the Rio Verde created adynamic lagoon environment with new, but shifting re-sources (Goman, Joyce, & Mueller, 2005).

A core taken from Aguada Petapilla in a small valleynear Copan, Honduras showed two peaks in deforesta-tion: around 900 B.C. when agriculture began in the area,and around A.D. 400 when a Maya kingdom was estab-lished at Copan (McNeil, Burney, & Burney, 2010). Thesefindings are contrary to earlier studies of Petapilla that in-dicated accelerating deforestation in the region throughthe Classic period (Abrams & Rue, 1988; Rue, 1987). Thenew finding, however, fails to account for the evidence

176 Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc.

DUNNING ET AL. GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA

of significant soil erosion in the Copan Valley during theClassic period (Abrams et al., 1996).

In the Guatemalan Highlands, Velez et al. (2011) corre-late increased inorganic sedimentation with forest clear-ance and soil erosion brought on by population increasesin the Middle Preclassic, Early Classic, and Late Postclas-sic. On the nearby Pacific coast, periods of populationgrowth also correlate with increased fluvial sedimenta-tion (Neff et al., 2006a, 2006b). Population levels, forestclearance, and agriculture are also linked in Laguna Cuz-cachapa in El Salvaror (Dull, 2007).

In reality, human–environment interactions are oftenhighly intertwined and modeling these interactions re-quires a complex systems perspective. A good exampleof the complexity of such interrelationships can be il-lustrated by the effects of forest removal in the MayaLowlands (Dunning, Beach, & Luzzadder-Beach, 2012;Turner & Sabloff, 2012), a set of interactions that beginswith the nature of bedrock on the Yucatan Peninsula.Much of this limestone bedrock is nearly pure calcite. Inthe southern Lowlands the inorganic fraction comprisingregional soils is largely derived from windblown volcanicash; in the north the parent material is a compound ofvolcanic ash, Saharan and North American eolian dust,plus local calcite. While these soils are fairly fertile, theyare notoriously low in phosphorus. Over the long term,P is recharged by new inputs of volcanic ash. On an an-nual basis, P is recharged by inputs of eolian dust (Daset al., 2011; Lawrence et al., 2007). However, as for-est is progressively removed from a location, the captureof dust is proportionally diminished leading to decreas-ing P levels in the soil. Thus, as agriculture was intensi-fied across the Maya Lowlands, P levels in regional soilswould have progressively declined. Reductions in forestfallowing would also likely decrease nitrogen levels byreducing inputs from leguminous Leucaena species, whichcomprise a significant portion of regional forests. Declin-ing soil fertility would have had cascading effects. Yieldswould have declined necessitating clearance of more for-est if available. Plant health would have declined makingcrops more vulnerable to diseases such as maize mosaicvirus. Declining soil fertility would also increase vulnera-bility to field invasion by weedy species, most notably ofbracken fern (Pteridium aquilinium), an aggressive speciesthat thrives in nutrient poor soil and which is noted forinvading over-cropped fields in the Southern Maya Low-lands and for being difficult to remove once established(Dunning & Beach, 2010; Perez-Salicrup, 2004). Thus,ironically, while the high base status soils and tropical cli-mate of the Maya Lowlands were favorable for cultivatingmaize and supporting high population densities, popula-tion growth led to the incremental reduction of forest forconstruction material, fuel, and farm land would have

created a risk spiral within the region, especially whencoupled with other environmental and cultural risk fac-tors (Dunning, Beach, & Luzzadder-Beach, 2012; Turner& Sabloff, 2012).

The Maya are a forest people who have accumulateda cultural storehouse of the economic properties associ-ated with trees and other plants including food, timber,fuel, fiber, and medicine among others. Hence, it is likelythat for millennia the Maya selectively removed somespecies and encouraged others. For many years, it hasbeen suspected that even in areas abandoned for cen-turies, the species composition of the tropical forest inthe Maya Lowlands partly reflects ancient forest manage-ment (e.g., Gomez-Pompa et al., 1987). Surveys of forestmantling ancient Maya sites have been used to documentthe species that were likely incorporated into gardens, or-chards, or managed forest (Ford, 2008; Ross, 2011; Ross &Rangle, 2011; Thompson et al., 2015). Analysis of woodand charcoal recovered from archaeological excavationsalso sheds light on the changing composition of ancientforests (Hageman & Goldstein, 2009; Lentz et al., 2014,2015). At Tikal, Lentz and Hockaday (2009) found thatManilkara zapata, a tree favored for lintels in monumen-tal architecture, was likely carefully managed but even-tually became scarce enough to be replaced by anotherless desirable species. Charcoal recovered from many sitesindicate that pine may have become intentionally culti-vated in areas outside its natural range because it wasfavored in ritual contexts (Lentz et al., 2005; Morehart,Lentz, & Prufer, 2005). Genetic analysis is beginning toadd a new dimension to understanding the managementof tree species through domestication or “wild” manage-ment (e.g., Galindo-Tovar et al., 2008; Petersen, Parker,& Potter, 2012; Thompson et al., 2015).

Anthropogenic land use changes also had lasting im-pacts on hydrology. Forest clearance, agriculture, andquarrying triggered erosion on sloping land surroundingclosed depressions (bajos) in the southern Maya Low-lands and the resulting sedimentation often resulted inthe blockage of groundwater recharge of once peren-nial wetlands (Beach et al., 2003; Dunning et al., 2002;Dunning, Beach, & Luzzadder-Beach 2006; Hansen et al.,2002). Sedimentation, eutrophication, and climate dryingall likely contributed to the long-term transformation ofmany perennial wetlands and shallow lakes into seasonalswamps in which deep clay Vertisols developed. How-ever, although many wetlands or parts of wetlands were“degraded”, other parts were actually enhanced from aMaya point of view. Wide aprons of calcium-rich collu-vial and alluvial sediment derived from upslope erosionaccumulated along the margins of bajos. These aprons ofdeep, high-base-status soils developed into some of thebest agricultural land available in the southern lowlands

Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc. 177

GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA DUNNING ET AL.

and became the focus of intensive cultivation (Dunninget al., 2002; Dunning, Beach, & Luzzadder-Beach 2006;Gunn et al., 2002). In some perennial wetlands in ripar-ian settings in the coastal lowlands, canals were dug tofacilitate drainage and agriculture (see further discussionbelow). More than a thousand years later, these canalsystems often continue to influence wetland hydrology,ecology, and pedogenesis (Figure 3) (Beach et al., 2009).In the Valley of Guatemala, a canal system associatedwith former Lake Miraflores may have abetted the per-manent desiccation of the lake in conjunction with dry-ing climate (Popenoe de Hatch et al., 2002).

LAND COVER, LAND USE, PALEODIET

By 1000 B.C., much of Mesoamerica was becoming anever-changing patchwork of forest and fields (Dunning& Beach, 2010). Charting how this increasingly human-ized landscape changed over time and space has becomean ever more sophisticated scientific endeavor. As dis-cussed above, investigations of ancient forests have uti-lized palynology, macrobotanical analysis of wood andcharcoal, surveys of contemporary forests, and studies ofgenetic variation. As will be discussed in the next sec-tion, analysis of carbon isotopes in soil organic matteralso offers information about changing vegetation com-position. The nature of forest cover can also be partlyreconstructed by analysis of fauna recovered in archaeo-logical investigations. The diversity of animal species ex-ploited by residents of investigated Maya sites suggeststhat a wide range of habitats were visited by hunters(Emery & Thornton, 2008a; Freiwald, 2010). Carbonisotope ratios in the bones of species such as white-tailed deer and peccary indicate that these animals con-sumed a significant quantity of maize and were proba-bly procured from farmer’s fields (Emery, 2003), thoughthere are indications at some sites that deer were cap-tured and fed maize before slaughter, a practice also usedwith dogs (White, Longstaffe, & Law, 2001). Variation instrontium isotopes in animal skeletons indicates that atsome larger Maya sites animals were being procured fromconsiderable distances, perhaps through trade or tribute(Freiwald, 2010; Thornton, 2011). In the northern MayaLowlands, species composition of animals recovered fromarchaeological contexts indicate that most animals werelikely procured locally, though higher status house-holds consumed more exotic species as well (Gotz, 2008;Masson & Lope, 2008). In sum, these studies indicate thatmost Maya sites were surrounded by a mosaic of fieldsand forest, some fairly mature, some secondary scrub.However, the specific composition of this field/forest mixvaried greatly across space and time as reflected in vary-

ing mixes of exploited animals (Emery, 2008; Emery& Thornton, 2008b; Somerville, Fauvelle, & Froehle,2013).

Estimates of the amount of forest cover that existedduring regional population zeniths (most typically in theLate Classic) range from as high as 75% to as low as15%. This range likely reflects both actual place-to-placevariation as well as differences in the methods used toreconstruct vegetative cover. Pollen percentages are themost commonly used method, but can lead to underrep-resentation of many tree species (e.g., the many insect-pollinated species predominant in tropical forest; Ford,2008). Macrobotanical remains recovered from archae-ological contexts can provide a partial picture of actualforest use, though one biased by both cultural and preser-vation filters – as well as recovery methods (Hageman &Goldstein, 2009). At Teotihuacan and neighboring sites,the relatively consistent use of certain woody species forfuel and for construction from Preclassic through Post-classic times suggests that some type of forest manage-ment was in play within at least this part of the Basin ofMexico to ensure the continued availability of these treesand shrubs (Adriana-Moran & McClung de Tapia, 2008).Nevertheless, widespread sedimentation over long peri-ods in the Teotihuacan Valley indicate that substantialforest clearance occurred resulting in persistent erosionon surrounding highlands (Sanchez-Perez et al., 2013).Similarly, soil erosion in many watersheds in the MayaLowlands and other regions of Mesoamerica indicates pe-riods of widespread land clearance, though sedimentationrates also indicate that denudation varied greatly acrossspace and time (see above).

The buried village of Ceren, El Salvador and sur-rounding land provide the best evidence on the na-ture of agriculture thus far recovered in Mesoamerica(Sheets, 2002, 2008). At Ceren, rapidly deposited vol-canic ash entombed not only the small town, but thegardens and fields worked by its inhabitants (Figure 5).Excavations for many years were concentrated withinthe village site and its immediate environs. These ex-cavations revealed areas of intensive garden produc-tion that included small plots of maize and manioc, aswell as beans, squashes, hog plum (Spondias), avocado,guava, calabash, and cacao, among others species (Lentz& Ramırez-Sosa, 2002; Sheets, 2008). These gardens wereclearly carefully planned and well tended. More recently,investigations have been extended into the buried land-scape beyond the margins of the village site and haverevealed extensive fields of maize and manioc (Sheetset al., 2011, 2012). While manioc has been posited asa major crop in Mesoamerica, data had been previouslylacking to support this hypothesis. Since the cultiva-tion of manioc and other root crops does not involve

178 Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc.

DUNNING ET AL. GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA

Figure 5 Household 1, Ceren, El Salvador. a) ridges in the foreground are the remains of ancient garden plots rapidly buried by volcanic ash along with

the house structures; ridges are about 15 cm high and 30 cm apart, crest to crest; b) map of the crops and trees revealed around Household 1. Photo and

image courtesy of David Lentz, Dept. of Biological Sciences, University of Cincinnati.

pollination and the plants naturally produce little pollen,its absence in most archaeological pollen studies is notsurprising. Nevertheless, examples of manioc and otherroot crops are increasingly being found in pollen andmacrobotanical records (e.g., Akpinar-Ferrand et al.,2012; Dunning et al., 2003; Lentz et al., 2014, 2015). Al-though lacking the advantageous burial by ash, investiga-tions at the Maya village site of Chan in Belize have pro-duced paleobotanical and archaeological data that closelyparallels that found at Ceren with a diversified and inten-sified agricultural system that additionally included agri-cultural terracing (Robin, 2012).

Agricultural terracing was used widely in Mesoamer-ica occurring to greater or lesser extents in any regionwith sloping land. In Oaxaca, terracing began to appearin the Middle Preclassic likely in response to increasingsoil erosion as land was brought into cultivation (Joyce &Goman, 2012; Kowalewski et al., 2009; Perez Rodrıguez& Anderson, 2013). Similar land degradation processesprobably prompted other early investments in terracingin other parts of Mesoamerica including Michoacan andthe Basin of Mexico, the highlands of Veracruz, and theMaya Lowlands (Beach et al., 2009; Fisher, 2005; Gar-rison & Dunning, 2009; Heine, 2003). Terraces systemsshow great variation in scale and degree of organization.In most cases, accretionary growth appears evident, typ-ically in close association with rural household groups;this association and the lack of apparent central plan-ning indicate that the terraces systems evolved infor-mally as farmers brought more sloping land into produc-tion (Beach et al., 2002; Dunning, 2004; Dunning et al.,2003; Dunning & Beach, 2010; Lemonnier & Vanniere,2013; Wyatt, 2012). In other instances, centralized orga-nization is evident in large-scale, carefully orchestrated

systems. Most dramatically, the Maya city of Caracol, Be-lize grew amidst a system of tens of square kilometers ofterracing in a hilly landscape (Figure 6) that would oth-erwise not have allowed such a large urban populationconcentration (Chase & Chase, 1998; Dunning & Beach,2010; Chase et al., 2011). Concentrated terrace systemsalso were a significant part of urbanization at the majorcenters of Monte Alban in Oaxaca and Anagamuco nearLake Patzcuaro, Michoacan (Fisher, 2005; Fisher, Leisz, &Outlaw, 2011; Perez Rodrıguez & Anderson, 2013; Joyce& Goman, 2012).

One puzzling aspect of land use in Mesoamerica is thelack of agricultural terracing in some places where suchan investment would seem to have been highly benefi-cial. The ancient Maya city of Tikal is a case in point.Much of urban Tikal lies on gently to steeply sloping landin a region where terracing has been found at nearbysites (e.g., San Bartolo and Xultun; Garrison & Dunning,2009), only a handful of probable agricultural terraceshave been identified at Tikal despite numerous surveysand detailed mapping (Dunning et al., 2015). Carbon iso-topes recovered from organic matter within sedimentsin two of Tikal’s reservoirs, as well as a slow rate ofsedimentation in those reservoirs, suggest that the ur-ban residential catchment areas feeding these reservoirswere protected by perennial tree cover (e.g. orchardsor tree-dominated gardens) (Scarborough et al., 2012;Tankersley et al., 2011, 2015). This “garden city” modelseems to apply to many other Maya cities in the Clas-sic periods as well (Dunning & Beach, 2010; Lentz et al.,2015) and elsewhere, such as the Gulf Lowlands of Ve-racruz (Stark & Ossa, 2007). Investigations at two sitesin the Tuxla Mountains of Veracruz found archaeobotan-ical evidence for increasing use of tree crops over time,

Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc. 179

GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA DUNNING ET AL.

Figure 6 LiDAR 2-D image of the Pachituk Terminus, Caracol Belize showing settlement groups and extensive agricultural terracing. Courtesy of Arlen

and Diane Chase, Caracol Archaeological Project.

which would have helped conserve sloping lands (Van-Derwarker, 2005).

Much of Tikal’s staple food production may have takenplace on colluvial aprons around localized depressions(“pocket bajos”) and the edges of the sprawling Bajode Santa Fe on the eastern flank of the city (Balzottiet al., 2013; Dunning et al., 2015; Lentz et al. 2014).An unexpected outcome of Preclassic erosion near manysouthern Maya Lowlands population centers was thedumping of huge quantities of base-rich eroded soil intothe region’s many karst depressions (bajos) which hadpreviously been largely unsuitable for maize agriculture(Beach et al., 2003; Dunning, Beach, & Luzzadder-Beach,2006; Dunning et al., 2002; Gunn et al., 2002). These ar-eas of new, deep colluvial and alluvial soils became thefoci of intensified agricultural production in the ensuingClassic period; at some sites this cultivation is evident inextensive field wall systems, though these are not evidentat Tikal.

Another focus of intensive and specialized cultivationin both the southern and northern Maya Lowlands werelimestone sinkholes or rejolladas. Deeper, more humidsoils in rejolladas allowed these places to be used for dry

season cultivation and production of crops with high wa-ter demand, most notably cacao (Dunning, Beach, & Rue,1997; Lentz et al. 2014, 2015; Munro-Strasiuk, Man-ahan, & Stockton, 2014; Munro-Strasiuk, Manahan, &Stockton, 2014).

Wetlands were a widespread locus of agriculturein Mesoamerica. Wetland agriculture developed earlywithin the Gulf Lowlands of Veracruz, where it mayhave evolved from cultivation along natural alluvial lev-ees (Arnold III, 2009). Macrobotanical, phytolith, andpollen evidence from this area indicates maize cultivationby 7100 14C yr B.P., the earliest known in Mesoamerica(Pohl et al., 2007). Surprisingly, this area has also pro-duced evidence for the Mesoamerican domestication ofsunflowers (Lentz et al., 2008). Investigations in an ex-tensive wetland field complex at Mandinga in central Ve-racruz indicate that development of the island fields be-gan in the Preclassic and intensive cultivation of the fieldscontinued until the 7th century A.D., when several factorsincluding sea-level rise, aggradation, a volcanic eruption,and regional population decline contributed to abandon-ment of the system (Heimo, Siemens, & Hebda, 2004).Early agriculture also appears to have been established

180 Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc.

DUNNING ET AL. GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA

along alluvial levees along the Pacific Coast as well,though this cultivation did not evolve into a system ofwetland cultivation (Neff et al., 2006b; Rosenswig, 2006).

Large and small patches of island fields are found inriparian and connected perennial wetlands across north-ern Belize and contiguous areas of Mexico. Many of thesefield systems originated in the Preclassic, though use andformation history varies from place to place reflectingboth differing and changing environmental and politi-cal economic conditions (Figure 3) (Beach et al., 2009;Pyburn, 1998). One notable finding has been the im-portant though variable role played by the gradual cap-illary precipitation of gypsum and calcium carbonate inthe formation of these field systems (Luzzadder-Beach &Beach, 2009). These systems appear to have given depen-dent populations greater resilience during the troublingtimes of the Terminal Classic, but ultimately they toosuccumbed and the fields were abandoned (Luzzadder-Beach, Beach, & Dunning, 2012). Similar field systemsare found along the Candelaria River (Siemens et al.,2002) and wetlands around the Laguna de Terminos insouthwestern Campeche, Mexico, but have been sub-ject to relatively little investigation. The true extent ofthe Campeche fields has only recently become apparent(Figure 7).

The wetlands of the Yalahau Fracture zone in north-ern Quintana Roo, Mexico were also manipulated bythe ancient Maya with a system of dikes and dams, buthow this system functioned has not been firmly estab-lished (Fedick et al., 2000; Fedick & Morrison, 2004; Se-dov et al., 2008). One important use of these wetlandswas the harvesting of periphyton that was used to fer-tilize nearby dryland agricultural fields. Analysis of theagricultural fields in the Yalahau was aided greatly by mi-cromorphological study of soil thin sections (Sedov et al.,2008), a technique that has been little used to date inMesoamerica (but see Beach et al. 2009; Sanchez-Perezet al., 2013). Periphyton may also have been harvestedfrom wetlands in northwestern Yucatan to enrich urbangarden plots at Chunchucmil, a large urban center locatedin an agriculturally inhospitable area, though with readyaccess to groundwater (Beach, 1998; Dahlin et al., 2005;Luzzadder-Beach, 2000).

The best known wetland agriculture system inMesoamerica is the chinampa or island field system oncefound over large areas around the margins of lakes withinthe Basin of Mexico. Unlike other areas of Mesoamer-ica where wetland agriculture began in Preclassic times,the chinampa system appears to have been developedlargely in the Postclassic (Parsons & Morett, 2004). Whilethe chinampa system is known to have been extensivein the Basin of Mexico in the Postclassic, its full ex-tent is not known; analysis of historical documents and

remotely sensed imagery is helping to map the formerdistribution of wetland fields (Morehart, 2012). Chi-nampa systems are also now known to have extendedinto lake basins in western Mexico (Fisher, 2005).

Compared to many regions of early civilization,Mesoamerica is not incised with any great river systems.Nevertheless, shorter, less voluminous rivers abound insome areas, but are notably absent in others such as theYucatan Peninsula. Water management was of critical im-portance both in areas of abundant surface water as wellas those without. The Elevated Interior Region of theMaya Lowlands is an example of the latter, a karst regionessentially lacking in perennial surface water and largelyinaccessible groundwater (Dunning, Beach, & Luzzadder-Beach, 2012). Yet this region formed the heartland ofMaya civilization. Urbanization was achieved by harvest-ing rainwater during the rainy season, storing it, andmeting it out through the dry. Large urban centers re-quired sophisticated hydraulic systems. At Tikal, a multi-tiered system of reservoirs was developed to impound andstore rainwater (Scarborough et al., 2012). The major-ity of harvested water at Tikal was likely used to meetdomestic needs, including filtering water entering somereservoirs. A few reservoirs appear to have functionedto protect downstream fields during heavy rains, andmay have also provided irrigation water to nearby fields.Given the heavy rains and tropical storms that period-ically affect the Maya Lowlands, flood control was animportant part of water management systems (Davis-Salazar, 2006). Not surprisingly, sites in the Elevated In-terior Region that invested heavily in water managementoften fared better during times of increased drought fre-quency (Akpinar-Ferrand et al., 2012; Dunning, Beach,& Luzzadder-Beach 2012; Dunning et al., 2015).

At the large Puuc region site of Xcoch in the north-ern Maya Lowlands a system of large reservoirs was de-vised in the Preclassic period (Dunning et al., 2014b).This reservoir system was functionally integrated with thesite’s monumental architecture (which funneled waterinto the reservoirs), systemically linked to the networkof household cisterns, symbolically linked to a deep caveunderlying the site center (and home to rain god wor-ship), and possibly linked to nearby fields for irrigation.Similar relationships are evident in the hydraulic systemsat other regional sites (Isendahl, 2011).

Although irrigation in ancient Mesoamerica wasnowhere practiced at the same scale as in parts of the OldWorld, it was nevertheless of critical importance in someareas. Localized irrigation was a vital part of the agricul-tural system that supported the huge Classic period city ofTeotihuacan (Nichols, Spence, & Berland, 1991; Sanchez-Perez et al., 2013). With the fall of Teotihuacan, settle-ment shifted onto the lake plains of the Basin of Mexico

Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc. 181

GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA DUNNING ET AL.

Figure 7 GoogleEarth image showing extensive wetland field systems near the Laguna de Terminos north of themouth of the Rıo Candelaria, Campeche,

Mexico. The large variety of shapes and sizes of the fields suggest that they date to multiple cultural time periods. These fields were first noted by

Dr. Hugo Lopez Rosas, an ecologist with Universidad Autonoma de Mexico.

and irrigation became less important (Parsons & Morett,2004). Kaminaljuyu, a close political and economic ally ofTeotihuacan in highland Guatemala, also adopted an irri-gation system (Popenoe de Hatch et al., 2002). Irrigationwas also important in other areas including the Pacificpiedmont of Guatemala and Chiapas, where it was usedto facilitate cacao cultivation (Kaplan, 2008).

PROSPECTION

A long-standing problem for archaeologists seeking tounderstand ancient Maya settlement and land use pat-terns is the pervasive cover of tropical vegetation overlarge areas. Aerial surveys have been used since a se-ries of regional overflights by Charles Lindbergh andAlfred Kidder in 1929, but with limited effectiveness ex-cept in areas where the forest has been cleared; other-wise only the largest of features are typically detectableunder the forest mantle. Recent developments in the useof both satellite and airborne remote sensing techniquesoffer the promise of revealing the previously invisible an-cient landscape at multiple scales. The combined use of

several types of multi-spectral satellite imagery and air-borne sensors is proving to be of considerable use forlocating sites and large landscape features obscured bytropical forest canopy in the southern Maya Lowlands(Sever & Irwin, 2003). Saturno et al. (2007) successfullydeveloped a spectral signature based on IKONOS imagesto identify Maya sites taking advantage of the differen-tial drainage and nutrient status of vegetation growingon the ruins. However, this signature may be region-specific and needs to be refined, or alternate signaturesdeveloped to facilitate site detection in other parts of thesouthern lowlands (Garrison et al., 2008). Airborne radarhas been employed in archaeological survey in the MayaLowlands since the late 1970s, but system resolution wasinitially too coarse to be useful—despite errant claims ofsuccess. Newer systems such as AIRSAR (Airborne Syn-thetic Aperture Radar) offer much better resolution andinitial experiments have proven successful in site detec-tion (Garrison et al., 2011). Combinations of differentimagery can also be used to provide invaluable land-scape context data such as topography, physiographic fea-tures, drainage, and vegetation (Estrada-Belli & Kock,2007). However, none of these satellite sensor systems or

182 Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc.

DUNNING ET AL. GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA

airborne radar is capable of mapping features down to thescale of individual house mounds.

Airborne LiDAR (light detection and ranging) wasflown over the ancient Maya city of Caracol, Belize in2009; the results after processing were spectacular in theamount of land surface detail revealed including thou-sands of house mounds and agricultural terraces (Figure6) (Chase et al., 2011). While these features were al-ready known to exist at Caracol (existing detailed sur-face maps allowed for ready ground truthing), the LiDARsurvey revealed sprawling, previously unmapped terracesystems and residential areas. Previously unknown phys-iographic features such as cave openings were also de-tected (Weishampel et al., 2011).

LiDAR has also been tested on the large Tarascan(Purepucha) site of Anagamuco near Lake Patzcuaro, Mi-choacan, Mexico (Fisher, Leisz & Outlaw, 2011). Thoughnot tropical, heavy forest cover on much of the site madetraditional mapping a labor-intensive process, whereasthe LiDAR survey revealed many square kilometers ofurbanized landscape. High initial costs make LiDAR cur-rently too expensive for most archaeological projects, butits pay off in detailed, digital maps readily incorporatedinto a Geographic Information System (GIS) makes ita technology of choice for thorough site survey (Carteret al., 2012; Chase et al., 2012). LiDAR has also now beenflown over the sites of Uxbenka in Belize and Yaxnohcah,Campeche and other Maya sites in Mexico with promis-ing results still forthcoming. LiDAR has also been used togood effect at Izapa in Chiapas, Mexico (Rosenswig et al.,2012). The effectiveness of LiDAR suggests it would be ahighly useful tool for archaeological survey for other ar-eas in Mesoamerica where ground visibility is very poor(e.g., the tropical lowlands of Veracruz; Stark & Garraty,2008).

Satellite remote sensing has also been used to goodeffect in other parts of Mesoamerica, perhaps provingmost useful in documenting ancient wetland field sys-tems. Morehart (2012) used analysis of Landsat 7 ETM+(Enhanced Thematic Mapper Plus) imagery in conjuc-tion with historical maps to reconstruct the extent ofchinampa agriculture in the northern Basin of Mexico.Heimo, Siemens, & Hebda (2004) document the environ-mental parameters of wetland field systems in coastal Ve-racruz. Including digital imagery as part of a research de-sign has the advantage of being readily incorporated intoa study GIS (e.g., Estrada-Belli & Kock, 2007; Morehart,2012).

Subsurface remote sensing techniques are also beingeffectively employed in some Mesoamerican archaeolog-ical projects. Ground Penetrating Radar (GPR) has beenused successfully in aggrading landscapes where buriedfeatures were constructed at least partially with hard

stone like basalt, such as Cotzumalhuapa in the PacificPiedmot of Guatemala (Safi et al., 2012) and Texcocoin the Basin of Mexico (Arciniega-Ceballos et al., 2009).Magnetic Susceptibility (MS) has also been used success-fully in naturally aggrading environments including theTuxla Mountains of Veracruz (Venter et al., 2006). GPRand MS have also been highly useful in urbanized or ur-banizing landscapes, both ancient and modern – essen-tially anthropogenically aggrading contexts (Arciniega-Ceballos et al., 2009; Chavez et al., 2001, 2005). Inenvironments such as alluvial wetlands of Veracruzwhere hard stone was infrequently used, gridded auger-ing programs have proven necessary to map subsurfacesites (Wendt, 2003). In humid tropical environments itis important to note that low elevation structures canbe rendered essentially invisible by bioturbation and or-ganic accumulation, and site surveys should systemati-cally sample open areas (Johnston, 2002, 2004).

Site prospection must also be sensitive to geomorpho-logically dynamic settings such as within fluvial systems.Given the attraction of riparian environments for settle-ment, archaeologists should actively look for sites in suchsettings even though early sites especially might be deeplyburied (e.g., Holley et al., 2000). This is especially neededin the search for early, more ephemeral, pre-ceramic sitesin fluvial environments that have undergone multiple cutand fill episodes (Figure 8) (Borejsza et al., 2014).

The study of chemical signatures remaining on resid-ual lime plaster floors in order to identify ancientactivity areas has been pioneered in central Mexico, par-ticularly at Teotihuacan, then extended to other sites inMexico, including an ethnoarchaeological study of tradi-tional homes in Tlaxcala which helped developed activ-ity “signatures” (e.g., Barba & Ortız, 1992; Manzanilla &Barba, 1990). Over time, the sophistication of these stud-ies has increased from the analysis of a few elements toover 20 elements and compounds such as fatty acids andproteins (Barba, 2007).

Chemical prospection for ancient activity areas hasbeen extended to soil floors. Studies have included asignature-establishing ethnoarchaeological study of con-temporary earth-floored homes in Las Pozas, Guatemala(Fernandez et al., 2002). At the rapidly buried site ofCeren, investigators had the advantage of chemicallytesting for activities areas inside and outside houseswith in situ architecture, artifacts, and plant remains(Parnell, Sheets, & Terry, 2002). As expected highphosphorus levels were found in food processing, con-sumption, and disposal areas; iron with areas of animalbutchery; and heavy metals with pigment processing.At the rapidly abandoned site of Aguateca, Guatemalaphosphorus was similarly found to closely correlate withartifact assemblages indicative of food processing and

Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc. 183

GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA DUNNING ET AL.

Figure 8 Cross section of a hypothetical entrenched stream valley (barranca) in Oaxaca showing the effects of multiple alluvial cut-and-fill episodes over

time resulting in the creation of multiple occupation surfaces, many of which are buried or partially removed. Many pre-ceramic sites are likely located in

such settings, but may be difficult to detect (after Borejsza et al., 2014).

consumption (Terry et al., 2004). At the ancient site ofCancuen, Guatemala, Cook et al. (2006) used ICP-MS(Inductively Coupled Plasma Mass Spectrometry) analy-sis to test for 60 major, minor, and rare-earth elementscollected from two Late Classic house floors, with 30 el-ements showing enrichment over controls. The most no-table concentrations were mercury (likely from process-ing Cinnabar as a pigment) and gold (since there is noevidence that the Maya at Cancuen processed gold, thisenrichment might reflect micro-debitage from jade pro-cessing). Like Ceren, at Piedras Negras, Guatemala, Par-nell, Terry and Nelson (2002b) found phosphorus to bethe most useful indicator of human activity, especiallyfood processessing and disposal, iron as an indicator ofanimal slaughter, and various metals as indicators of min-eral pigments. Testing of newly available portable X-rayfluorescence scanner (pXRF) at the Colonial era HaciendaTelchaquillo in Yucatan indicates that this unit was use-ful for detecting trace metals, but not for phosphorus(Coronel et al., 2014).

Beyond the immediate vicinity of houses, phospho-rus continues to be the most useful chemical prospec-tion toolfor identifying middens, however, Eberl et al.(2012) found that not all middens show up as phosphorushotspots, probably because not all trash disposal involved

high concentrations of organic wastes. Phosphorus canalso be used to identify agricultural gardens and fieldswhere organic amendments were repeatedly applied aspart of intensification of cultivation (Dunning, Beach, &Rue, 1997). Once soil use patterns have been identified,soil mapping can be used to extrapolate probable agricul-tural potentials around and between sites (e.g., Fernan-dez et al., 2005). Where available, the folk soil classifi-cations used by local traditional farmers can be used toanalyze probable land use patterns in the soilscape (Dun-ning & Beach, 2003; Jensen et al., 2007).

Analysis of residual carbon isotopes in the humin frac-tion of soil organic matter (SOM) has proven to be an ef-fective tool for understanding ancient rural land use inparts of Mesoamerica. Research in several parts of theworld has shown that δ13C enrichment occurs when ahigh percentage of vegetation contributing to the buildupof SOM utilizes the C4 photosynthetic pathway such asmany tropical grasses, including maize. Changes downthe soil profile in carbon isotope ratios can provide animportant line of evidence for vegetation change and ero-sion over time, especially in well dated aggrading profiles.In the Maya Lowlands, paleosols from known agriculturalcontexts such as terraces and wetland fields provide themost secure evidence for intensive cultivation of maize

184 Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc.

DUNNING ET AL. GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA

resulting in significant δ13C enrichment (Beach et al.,2011; Webb, Schwarcz, & Healy, 2004). Studies aroundseveral ancient Maya sites including Aguateca (Johnson,Wright, & Terry, 2007; Wright, Terry, & Eberl, 2009),Motul de San Jose (Webb et al., 2007), Piedras Negras(Johnson et al., 2007b), and Ramonal (a satellite centerof Tikal; Burnett et al., 2012), have documented distinc-tive spatial patterns for δ13C enrichment (greater than the2.5–3‰ typical from bacterial fractionation) suggestive offields kept in maize production for extended periods. Al-together, these studies indicate that C4 plants (presum-ably mainly maize) made up � 25% of the vegetationat these sites in the Maya Classic period, but only a fewpercent today. At Tikal GIS-based soil maps have beenused to extrapolate probable zones of maize productionbeyond sampled areas (Balzotti et al., 2013). However,such extrapolations must remain tenuous for the timebeing because data points are quite limited. The identi-fication of δ13C enrichment with in situ maize productionmust also remain provisional at present because of de-ranging factors such as slope transport, burial, and thepresence of other vegetation that can cause δ13C enrich-ment (other C4 grasses and CAM pathway vegetation,both of which can be expected to increase during peri-ods of drier climate; cf. Solıs-Castillo et al., 2013). Ulti-mately, DNA analysis of soil organic matter may providemore definitive information on ancient cropping practices(Speller, 2013).

PROVENANCE

Many provenance studies in Mesoamerican archaeologyhave focused on obsidian (Braswell, 2004). Many of thesestudies have been used to elicit trade relations and re-source procurement over time (e.g., Brown, Dreiss, &Hughes, 2004; Knight, 2003; Knight & Glascock, 2009).These assessments have become refined enough thatchanging obsidian sources evident in the archaeologi-cal record can be used to help refine site-specific or re-gional chronologies (Braswell, 2011). Laboratory tech-niques such as NAA (neutron activation analysis), XRF(X-ray florescence), and ICP-MS have partially replacedvisual inspection by refining the elemental “signatures” ofobsidian sources (e.g., Ponomarenko, 2004), differentiatesub-sources (e.g., Argote-Espino et al., 2012), and in theprospection for, and identification of more obscure obsid-ian sources such as that at San Luis, Honduras (Aoyama,Tashiro, & Glascock, 1999). Nevertheless, because visualinspection is capable of distinguishing many Mesoameri-can obsidians at almost no cost this method is still widelyemployed, often as a compliment to machine-based tests

and allowing for much greater numbers of samples to beassessed (e.g., Pool, Knight, & Glascock, 2014).

A variety of laboratory techniques have also been ap-plied to other material in Mesoamerica to identify sourcesand movement. Material for the manufacture of groundstone tools is a scarce resource in the Maya Lowlandsand manos and metates made from hard igneous andmetamorphic rocks from the Maya Mountains were avaluable item. Abramiuk and Meurer (2006) used thinsection analysis to pinpoint sources for ground stone arti-facts in the Bladen District of Belize. Jade is also a scarceresource, with the only known source area in Mesoamer-ica occurring in the Motagua Valley of Guatemala. How-ever, ICP-MS analysis has been employed to begin to dis-tinguish the homogenous jadeites on the north side ofthe valley from more chemically heterogeneous sourceson the south side (Neff, Kovacevich, & Bishop, 2010).Chert, perhaps the most widely stone used for tools inMesoamerica, has largely defied attempts to develop ele-mental signatures tied to specific sources (Bishop, 2014).

Other provenance studies have sought to identify ordifferentiate important clay sources, largely in an effortto develop and refine trade models. This endeavor hasproven most successful in areas with greater geologic di-versity which produce distinctive elemental signatureswithin clay, temper, and pigments used in ceramic man-ufacture such as in the Mexican and Guatemalan high-lands (Alex, Nichols & Glascock, 2012; Charlton et al.,2008; Inanez et al. 2010; Neff et al., 2000; Reents-Budetet al., 2006). Although in use in archaeology since the1930s, petrographic analysis continues to be a useful toolfor identifying distinctive minerals within ceramics aswell as microstructure (e.g., Stoltman, 2011), though of-ten coupled with other techniques such as NAA and ICP-MS (e.g., Cecil, 2007). Despite the increasing precision ofdata being produced by the application of instrumentaltechniques to Mesoamerican ceramics, interpretation ofthese data can still be contentious as seen in the livelydebate over directionality of trade among the Olmec andOlmec-linked cultures (cf., Blomster, Neff, & Glascock,2005; Neff et al. 2006c; Neff et al. 2006d; Sharer et al.,2006; Stoltman et al., 2005).

Attempts to establish clay/ceramic provenance in theMaya Lowlands have met with much more limited suc-cess because of more limited geologic diversity (e.g.,Bartlett et al., 2000), and seemingly require a multiplicityof techniques to be successful (e.g. Cecil, 2007). However,exceptions occur, including the rare palygorskite that wasmixed with indigo to make the unique Maya Blue pig-ment (Arnold et al., 2007).

One localized provenance study sought the sources forbitumen found in Olmec archaeological sites (Wendt &Lu, 2006). Provenance studies have even helped bring

Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc. 185

GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA DUNNING ET AL.

scientific scrutiny to items of pseudo-scientific interest: acombination of microscopy and mineralogy was used todemonstrate that at least two of the “mysterious” crys-tal skulls could not be of pre-Columbian manufacture orof Mesoamerican origin as had been purported by some(Sax et al., 2008).

One of the many effects of the Chicxulub meteor im-pact on the Yucatan Platform was the creation of a dis-tinctive generally north-south differential distribution ofstrontium isotopes across the landmass (Gilli et al., 2009;Hodell et al., 2004). Strontium isotopes, of course, varyacross Mesoamerica along with geologic substrates. Hu-man bones absorb small amounts of Sr during child-hood growth years. Hence, both places and the humanswho grew up there have distinctive Sr signatures, a factthat has led to what are essentially human provenancestudies (Price et al., 2008). One notable study of a largeskeleton population from Tikal, Guatemala was able toidentify immigrants to the city (Wright, 2005). Skeletaldetective work has been used to check hieroglyphic in-scriptions, iconography, and artifacts that suggested inter-relationhips between Tikal, Teotihuacan, Kaminaljuyu,and Copan (Price et al., 2010; Wright, 2005; Wright,2012; Wright et al., 2010), and movement of popula-tion through the port town of Xcambo in northern Yu-catan (Cucina et al., 2011). Oxygen isotope ratios havealso proven useful for differentiating populations andindividuals of different geographic origins, particularlybetween Teotihuacan and the Maya Lowlands and High-lands (White et al., 2000, 2001).

In a review of the use of instrumental technologies toaddress provenance and other economic issues in the an-cient Mesoamerican economy, Bishop (2014) extols themany advances in precision and variety of techniquesavailable to archaeologists today. However, he also de-cries the apparent declining use of these techniques tohigh costs and the closure of research laboratories by uni-versities and other entities as part of fiscal cutbacks. Onthe other hand, the incorporation of the growing body ofprovenance data into centralized geographic informationsystems is helping create more robust spatial-temporalmodels of ancient exchange patterns, such as that be-ing developed for Central Mexican ceramics (Nicholas,Stoner, & Crider, 2014).

CONCLUSIONS

Forty years ago, a review of geoarchaeological researchin Mesoamerica would have been short, but not sweet.Studies in which paleoenvironmental concerns weregiven more than passing consideration were just emerg-ing. Today, a review of geoarchaeological research of

even a few years finds a large and expanding array ofstudies that either employ geoarchaeological techniquesor are driven by concerns central to geoarchaeoloogysuch as those discussed in the preceding pages.

A common element in each of the areas reviewedabove is the growing recognition of the importance, andconsequent use of convergent approaches in order to cre-ate more robust models of ancient environments andeconomies. These approaches include multi-proxy studiesof paleoenvironments, land use, and ancient diet to takeadvantage the growing array of instrumental techniquesavailable to researchers (e.g., Lentz et al., 2014; Morell-Hart, Joyce, & Henderson, 2014; Wahl, Byrne, & Ander-son, 2014). A highly promising new approach involvesthe analysis of hydrogen and carbon isotopes in plant-wax lipids preserved in sediments, because it potentiallyprovides a direct measurement of climate signals alongwith land cover/use data (Douglas et al., 2012), thoughtime and costs currently constrain its application. In thearea of prospection, the introduction of airborne LiDARto the archaeological study of heavily forested areas ofMesoamerica has revolutionized the ability of perform-ing complete site surveys, a formerly cost and time pro-hibitive possibility. In provenance studies, the convergentuse of multiple instrumental techniques has allowed fortremendous gains in the accuracy of sourcing ancient ma-terials. However, as Bishop (2014) has cautioned, theseadvances are threatened by their very complexity and ex-pense in a time of fiscal constraint in many academic andother research institutions.

As information about the ancient cultures and envi-ronment of Mesoamerica accumulates at a rapid rate, ourmodels need to reflect the complexity of interactions thattook place in the past. Complex systems models can pro-vide a useful means of intertwining cultural and envi-ronmental factors and processes (e.g., Turner & Sabloff,2012). The power and success of these models dependsin great part on their ability to synthesize and correlatelarge quantities of data in effective and meaningful ways.In pursuit of this goal it is imperative that the rapidlygrowing body of data being produced by geoarchaologicaland other types of archaeological investigations be incor-porated into databases capable of supporting integratedanalyses. Geospatial databases offer one effective way tosynthesize and compare data from multiple sources insupport of temporally dynamic models (e.g., Chase et al.,2012; Heckbert, 2013; Kennett & Beach, 2013; Nichols,Stoner, & Crider, 2014).

As Sheets (2009) has pointed out, the large majorityof investigations in which geoarchaeology plays an im-portant role are based in processual theory. Althoughthis focus has proven highly productive, it has tendedto exclude more humanistic perspectives. In issues such

186 Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc.

DUNNING ET AL. GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA

as the role of drought in the Terminal Classic col-lapse of Maya Civilization, the dominance of processualstudies has led to charges of determinism (Aimers, 2007;McAnany & Gallareta Negron, 2010; Webster, 2002). Asnoted by Sheets, cave investigations are one area whereprocessessual and humanistic approaches are being wed(e.g., Brady & Prufer, 2010). Given the symbolic impor-tance attached to caves in Mesoamerican cultures, theconvergence of investigations seeking the “meaning” ofcaves and their uses and those seeking paleoenvironmen-tal proxies is to be expected. For example, speleothemsare an archive of paleoclimate information and are alsoimportant artifacts used in ancient rain-related rituals(Brady et al., 1997; Moyes et al., 2009).

Resilience theory (Holling & Gunderson, 2002) offersone useful avenue for synthesizing human-environmentinteractions (Redman, 2005; Scarborough, 2009). Thisperspective sees human-environment systems as an openset of relationships and interdependencies that eitheradapt to change and survive, or fragment and collapse;human beliefs and actions are an integral part of theoutcomes. An important aspect of coming to under-stand human-environment systems from this perspectiveis that not only does it help us understand the successand failure of past societies, but can help predict thesustainability and resilience of these systems and theircomponents in the future (Fisher et al., 2009). The per-spective of using information gathered from archaeo-logical investigations of all kinds, especially those thatmost directly involve human-environment interactions,underlies a growing body of research (for example, thatsupported by the IHOPE initiative: Costanza, Graumlich,& Steffen 2007; Chase & Scarborough 2014). Ultimatelywhat we can learn about the capacity to adapt to envi-ronmental change, including that induced by human ac-tivities, may be the most enduring product of geoarchae-ological studies of the ancient world.

This paper is partly the result of a graduate seminar held inthe Department of Geography, University of Cincinnati duringthe Spring Semester of 2013. All of the co-authors were par-ticipants in that seminar. David Lentz is to be thanked for pro-viding the photograph and map in Figure 5. Arlen Chase pro-vided the image in Figure 6. Jon-Paul McCool drafted the mapin Figure 1. The authors thank two anonymous reviewers fortheir many helpful comments that led to the improvement of thearticle.

REFERENCES

Abramiuk, M.A., & Meurer, W.P. (2006). A preliminary

geoarchaeological Investigation of Ground Stone Tools in

and around the Maya Mountains, Toledo District, Belize.

Latin American Antiquity, 17, 335–354.

Abrams, E.M., & Rue, D.J. (1988). The causes and

consequences of deforestation among the prehistoric Maya.

Human Ecology, 16, 377–395.

Abrams, E.M., Freter M., Rue, D.J., & Wingard, J. (1996). The

role of deforestation in the collapse of the Late Classic

Copan Maya state. In M.C. Pearl (Ed.), Tropical

deforestation: The human dimension (pp. 55–75). New

York: Columbia University Press.

Adriana-Moran, C.C., & McClung de Tapia, E. (2008). Trees

and shrubs: The use of wood in Pre-Hispanic Teotihuacan.

Journal of Archaeological Science, 35, 2927–2936.

Aimers, J.J. (2007). What Maya collapse? Terminal Classic

variation in the Maya Lowlands. Journal of Archaeological

Research, 15, 329–377.

Aimers, J.J., & Hodell, D. (2011). Societal collapse: Drought

and the Maya. Nature, 479, 44–45.

Alex, B. A., Nichols, D. L. & Glascock, M. D. (2012).

Compositional analysis of Formative Period ceramics in the

Teotihuacan Valley, Mexico. Archaeometry, 54, 821–834.

Akpinar-Ferrand E., Dunning, N.P., Jones, J.G., & Lentz, D.L.

(2012) Aguadas and ancient Maya water management in

the southern Maya Lowlands. Ancient Mesoamerica, 23,

85–101.

Anselmetti, F.S., Ariztegui, D., Hodell, D.A., Hillesheim, M.B,

Brenner, M., Gilli, A., and McKenzie, J.A., 2006. Late

Quaternary climate-induced lake level variations in Lake

Peten-Itza, Guatemala, inferred from seismic stratigraphic

analysis. Palaeogeography. Palaeoclimatoloy.

Palaeoecology, 230, 52–69.

Anselmetti, F.S., Ariztegui, D., Brenner, M., Hodell, D., &

Rosenmeier, M.F. (2007). Quantification of soil erosion

rates related to ancient Maya deforestation. Geology, 35,

915–918.

Aoyama, K., Tashiro, T., & Glascock, M. (1999). A

Pre-Columbian obsidian source in San Luis, Honduras.

Ancient Mesoamerica, 10, 237–250.

Arciniega-Ceballos, A., Hernandez-Quintera, E., Cabral, Cano,

E., & Morett-Alatorre, L. (2009). Shallow geophysical

survey at the archaeological site of San Miguel Pocuila

Basin of Mexico. Journal of Archaeological Science, 36,

1199–1205.

Argote-Espino, D., Sole, J., Lopez-Garcia, P., & Sterpone, O.

(2012). Obsidian subsource identification in the Sierra de

Pachuco and Otumba Volcanic Regions, central Mexico, by

ICP-MS and DBSCAN statistical analysis. Geoarchaeology,

27, 48–62.

Arnold III, P.J. (2009). Settlement and subsistence among the

Early Formative Gulf Olmec. Journal of Anthropological

Archaeology, 28, 397–411.

Arnold, D. E., Neff, H., Glascock, M.D., & Speakman, R.J.

(2007). Sourcing the palygorskite used in Maya Blue: A

pilot study comparing the results of INAA and LA-ICP-MS.

Latin American Antiquity, 18, 44–58.

Balzotti, C.S., Webster, D.L., Murtha, T.M, Petersen, S.L.,

Burnett, R.L., & Terry, R.E. (2013). Modeling the ancient

Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc. 187

GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA DUNNING ET AL.

amize agriculture potential of landforms in Tikal national

Park, Guatemala. International Journal of Remote Sensing,

34, 5868–5891.

Barba, L. (2007). Chemical residues in lime-plastered

archaeological floors. Geoarchaeology, 22, 439–452.

Barba, L. and Ortız, A. (1992). Analisis quımico de pozos de

ocupacıon: Un caso etnografico en Tlaxcala, Mexico. Latin

American Antiquity, 3, 63–82.

Bartlett, M.L., Neff, H., & McAnany, P.A. (2000).

Differentiation of clay resources on a limestone plain: The

analysis of clay utilization during the Maya Formative at

K’axob, Belize. Geoarchaeology, 15, 95–133.

Beach, T. (1998). Soil constraints on northwest Yucatan:

Pedo-archaeology and Maya subsistence at Chunchucmil.

Geoarchaeology 13, 759–791

Beach, T., Luzzadder-Beach, S., Dunning, N.P., Hageman, J.,

& Lohse, J. (2002). Upland agriculture in the Maya

Lowlands: Ancient Maya soil conservation in northwestern

Belize. Geographical Review, 92, 372–397.

Beach, T., Luzzadder-Beach, S., Dunning, N.P. &

Scarborough, V. (2003). Depression soils in the lowland

tropics of northwestern Belize. In A. Gomez-Pompa, M.

Allen, S. Fedick, & J. J. Jimenez-Orsonio (Eds.), Lowland

Maya area: Three millennia at the human-wildland

interface (pp. 139–174). Binghamton, NY: Haworth Press.

Beach, T., Dunning, N.P., Luzzadder-Beach, S., Cook, D. &

Lohse, J., (2006). Ancient Maya. impacts on soils and soil

erosion. Catena, 65, 166–178.

Beach, T., Luzzadder-Beach, S., Dunning N.P., & Cook, D.

(2008). Human and natural impacts on fluvial and karst

systems in the Maya Lowlands. Geomorphology, 101,

301–331

Beach, T., Luzzadder-Beach S., Dunning, N., Jones, J, Lohse,

J., Guderjan, T., Bozarth, S., Millspaugh, & Bhattacharya,

T. (2009) A review of human and natural changes in Maya

Lowlands wetlands over the Holocene. Quaternary Science

Reviews, 28, 1710–1724.

Beach, T., Luzzadder-Beach, S., Terry, R., Dunning, N.P.,

Houston, S., & Garrison, T. (2011). Carbon isotopic ratios

of wetland and terrace soil sequences in the Maya

Lowlands of Belize and Guatemala. Catena, 85,

109–119.

Beach, T., Luzzadder-Beach, S., Guderjan, T & Krause, S.

(2015). The floating gardens of Chan Cahal: Soils, water,

and human interaction. Catena. In press.

Bennett, H.H. (1926). Agriculture in Central America. Annals

of the Association of American Geographers, 16, 63–84.

Bernal, J.P., Lachniet, M., McCulloch, M., Mortimer, G.,

Morales, P., & Cienfuegos, E. (2011). A speleothem record

of Holocene climate variability from southwestern Mexico.

Quaternary Research 75, 104–113

Berrıo, J.C., Hooghiemstra, H., vanGeel, B., &

Ludlow-Wiechers, B. (2006). Environmental history of the

dry forest biome of Guerrero, Mexico, and human impact

during the last c. 2700 years. The Holocene, 16, 63–80.

Bishop, R.L. (2014). Instrumental approaches of

understanding Mesoamerican economy: Elusive promises.

Ancient Mesoamerica, 25, 251–269.

Blomster, J.P., Neff, H. & Glascock, M.D. (2005). Olmec

pottery production and export in ancient Mexico

determined through elemental analysis. Science, 307,

1068–1072.

Borejsza, A., Rodrıguez-Lopez, I., Frederick, C.D., & Bateman,

M.D. (2008). Agricultural slope management at La Laguna,

Tlaxcala, Mexico. Journal of Archaeological Science, 35,

1854–1866.

Borejsza, A., Frederick, C.D., & Lesure, R.G. (2011). Swidden

agriculture in the tierra fria? Evidence from sedimentary

records in Tlaxcala. Ancient Mesoamerica, 22, 91–106.

Borejsza, A., Frederick, C.D., Morett Alatorre, L. & Joyce,

A.A. (2014). Alluvial stratigraphy and the search for

preceramic open-air sites in Highland Mesoamerica. Latin

American Antiquity, 25, 278–299.

Brady, J.E., & Prufer, K.M. (Eds.) (2010). In the Maw of the

Earth Monster: Studies of Mesoamerican ritual cave use.

Austin: University of Texas Press.

Brady, J.E., Scott, A., Neff, H., & Glascock, M.D. (1997).

Speleothem breakage, movement, removal, and caching:

An aspect of ancient Maya cave modification.

Geoarchaeology, 12, 725–750.

Braswell, G.H. (2004). Lithic analysis in the Maya area. In G.

Borgstede & C. Golden (Eds.), Maya archaeology at the

millenium (pp. 187–199). London: Routledge.

Braswell, G.H. (2011). The obsidian and ceramics of the Puuc

region: Chronology, procurement, and production at

Xkipche, Yucatan, Mexico. Ancient Mesoamerica, 22,

135–154.

Brenner, M., Rosenmeier, M.F., Hodell, D.A., & Curtis, J.H.

(2002). Paleolimnology of the Maya Lowlands: Long-term

perspectives on interactions among climate, environment,

and humans. Ancient Mesoamerica, 13, 141–157.

Brenner, M., Hodell, D.A., Curtis, J., Rosenmeier, M.F.,

Anselmetti, F.S. & Ariztegui, D. (2003). Paleolimnological

approaches for inferring past climate change in the Maya

region: Recent advances and methodological limitations. In

A. Gomez-Pompa, M.F. Allen, S.L. Fedick, & J.J.

Jimenez-Osornio (Eds.), The Lowland Maya area: Three

millennia at the human-wildland interface (pp. 45–75).

New York: Haworth Press.

Brown, D.O., Dreiss, M.L., & Hughes, R.E. (2004). Preclassic

obsidian procurement and utilization at the Maya site of

Colha, Belize. Latin American Antiquity, 15, 222–240.

Burnett, R.L., Terry, R.E., Alvarez, M., Balzotti, C., Murtha,

T., Webster, D., & Silverstein, J. (2012). The ancient

agricultural landscape of the satellite settlement of

Ramonal near Tikal, Guatemala. Quaternary International,

265, 101–115.

Byrne, R. & Horn, S.B. (1989). Prehistoric agriculture and

forest clearance in the Sierra de los Tuxlas, Mexico.

Palynology, 13, 181–193.

188 Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc.

DUNNING ET AL. GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA

Cabadas-Baez, H., Solleiro-Rebolledo, E., Sedov, S., Pi, T., &

Gama-Castro J. (2010). Pedosediments of karstic sinkholes

in the eolianites of NE Yucatan: A record of Late

Quaternary soil development, geomorphic processes, and

landscape stability. Geomorphology, 122, 323–337.

Carillio-Bastos, A., Islebe, G.A., Torrescano-Valle, N., &

Gonzalez, N.E. (2010). Holocene vegetation and climate

history of central Quintana Roo, Yucatan Peninsula,

Mexico. Review of Paleobotany and Palynology, 160,

189–196.

Carleton, W.C., Campbell, D., & Collard, M. (2014). A

reassessment of the impact of drought cycles on the Classic

Maya. Quaternary Science Reviews, 105, 151–161.

Carozza, J.-M., Galop, D., Metaile, J.-P., Vanniere, B.,

Bossuet, G., Monna, F., Lopez-Saez, J.A., Arnauld, M.-C.,

Breuil, V., Forne, M., & Lemonnier, E. (2007). Landuse

and soil degradation in the southern Maya Lowlands, from

Pre-Classic to Post-Classic times: The case of La Joyanca

(Peten, Guatemala). Geodinamica Acta, 20, 195–207.

Carter, W.E., Shrestha, R.L., Fisher, C.T., & Leisz, S. (2012)

Geodetic imaging: A new tool for Mesoamerican

archaeology. EOS, 93, 413–415.

Cecil, L.G. (2007). Postclassic Maya ceramic advances:

Conjoining stylistic, technological, and chemical

compositional data. In D. Rosslere (Ed.), Developments in

ceramic materials research (pp. 1–34). Hauppauge, NY:

Nova Science Publishers.

Charlton, T.H., Otis-Charlton, C.L., Nichols, D.L. & Neff, H.

(2008). Aztec Otumba, AD 1200–1600: Patterns of the

production, distribution, and consumption of ceramic

products. In C.A. Pool & G.J. Bey III (Eds.), Pottery

economics in Mesoamerica (pp. 237–266). Tucson:

University of Arizona Press.

Chase, A., & Chase, D. (1998). Scale and intensity in Classic

Maya agriculture: Terracing and agriculture in the ‘‘garden

city” of Caracol, Belize. Culture and Agriculture, 20, 60–77.

Chase, A.F., Chase, D.Z., Wishampel, J.F., Drake, J.B.,

Shrestha, R.L., Slatton, K.C., Awe, J.J., & Carter, W.E.

(2011). Airborne LiDAR archaeology, and the ancient

Maya landscape at Caracol, Belize. Journal of

Archaeological Science, 38, 387–398.

Chase, A.F., Chase, D., Fisher, C., Leisz, S., & Weishampel, J.

(2012). The geospatial revolution in Mesoamerican

archaeology. Proceedings of the National Academy of

Sciences, 109, 12916–12921.

Chase, A.F. & Scarborough, V.L. (Eds.) (2014). The resilience

and vulnerability of ancient landscapes: Transforming

Maya archaeology through IHOPE. Archaeological Papers

of the American Anthropological Association 24 (special

issue).

Chavez, R.E., Camara, M.E., Tejero, A., Barba, L. &

Manzanilla, L. (2001). Site characterization by geophysical

methods in the archaeological zone of Teotihuacan,

Mexico. Journal of Archaeological Science, 28,

1265–1276.

Chavez, R.E., Camara, M.E., Ponce, R., & Argote, D. (2005).

Use of geophysical methods in urban archaeological

prospection: The Basilica de Nuestra Senor de Salud,

Patzcuaro, Mexico. Geoarchaeology, 20,

505–519.

Cook, B.I., Anchukaitis, K.J., Kaplan J.O., Puma, M.J., Kelley,

M., & Gueyffier, D. (2012). Pre-Columbian deforestation as

an amplifier of drought in Mesoamerica. Geophysical

Research Letters, 39, DOI: 10.1029/2012GL052565.

Cooke, C.W. (1931). Why the Mayan cities of the Peten

district, Guatemala, were abandoned. Journal of the

Washington Academy of Sciences, 21, 283–287.

Cook, D.E., Kovacevich, B., Beach, T., & Bishop, R. (2006).

Deciphering the inorganic chemical record of ancient

human activity using ICP-MS: A reconnaissance study of

late Classic soil floors at Cancuen, Guatemala. Journal of

Archaeological Science, 33, 628–640.

Cordova, C.E., & Parsons, J.R. (1997). Geoarchaeology of an

Aztex dispersed village on the Texcoco Piedmont of central

Mexico. Geoarchaeology, 12, 177–210.

Coronel, E.G., Bair, D.A., Brown, C.T. & Terry, R.E. (2014).

Utility and limitations of portable X-ray fluorescence and

field laboratory conditions on the geochemical analysis of

soils and floors at areas of known human activities. Soil

Science, 179, 258–270.

Costanza, R., Graumlich, L.J. & Steffen, W.L. (2007).

Sustainability or collapse: Lessons from integrating the

history of humans and the rest of the world. In R.

Costanza, L. Graumlich, & W. Steffen (Eds.), Sustainability

or collapse? An integrated history and future of people on

earth (pp. 3–17). Cambridge: MIT Press.

Cucina, A., Tiesler, V., Sierra Sosa, T., & Neff, H. (2011).

Trace-element evidence for foreigners at a Maya port in

northern Yucatan. Journal of Archaeological Science, 38,

1878–1885.

Curtis, J.H., Hodell, D.A, & Brenner, M. (1996). Climate

variability on the Yucatan Peninsula (Mexico) during the

last 3500 years, and implications for Maya cultural

evolution. Quaternary Research. 46, 37–47.

Dahlin, B.H. (2011). Final Report 0716015 Collaborative

Research: Speleothem Proxies for Maya Culture – Climate

Interactions. Washington, DC, National Science

Foundation.

Dahlin, B.H., Beach, T., Luzzadder-Beach, S., Hixon, D.,

Hutson, S., Magnoli, A., Masell, E., & Mazeau, D.E. (2005).

Reconstructing agricultural self-sufficiency at

Chunchucmil, Yucatan, Mexico. Ancient Mesoamerica, 16,

229–247.

Dahlin, B.H., & Chase, A. (2014). A tale of three cities: Effects

of AD 536 event on the lowland Maya heartland. In G.

Iannone (Ed.), The Great Maya Droughts in cultural

context. Boulder: University of Colorado Press.

Das, R., Lawrence, D., D’Odorico, D., & DeLonge, M. (2011).

Impact of land use change on atmospheric P inputs in a

tropical dry forest. Journal of Geophysical Research,

Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc. 189

GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA DUNNING ET AL.

Biogeosciences, 116, G01027,

doi:10.1029/2010JG0014032011

Davies, S.J., Metcalfe, S.E., Bernal-Brooks, F.,

Chacon-Torres, A., Farmer, J.G., MacKenzie, A.B. &

Newton, A.J. (2005). Lake sediments record sensitivity of

two hydrologically closed upland lakes in Mexico to

human impact. AMBIO: A Journal of the Human

Environment, 34, 470–475.

Davis-Salazar, K.L. (2006). Late Classic Maya drainage and

flood control at Copan, Honduras. Ancient Mesoamerica,

17, 125–138.

Deevey, E.S., Rice, D.S., Rice, P.M., Vaughan, H.H., Brenner,

M. & Flannery, M. (1979) Mayan urbanism: Impact on a

tropical karst environment. Science, 206,

298–306.

Demarest, A.A., Rice, P.M. & Rice, D.S. (Eds.) (2004). The

Terminal Classic in the Maya Lowlands: collapse, transition,

and transformation. Boulder: University Press of Colorado.

Denommee, K.C., Bentley, S.J. & Droxler, A.W. (2014).

Climate controls on hurricane patterns: A 12000-y

near-annual record from Lighthouse Reef, Belize. Nature

Research Reports, 3876. doi: 10.1038/srep03876.

Douglas, P.M.J., Pagani, M., Brenner, M., Hodell, D.A., &

Curtis, J.H. (2012) Aridity and vegetation composition are

important determinants of leaf-wax δD values in

southeastern Mexico and Central America. Geochimica et

Cosmochimica Acta, 97, 24–45.

Dull, R.A. (2004). An 8000-year record of vegetation, climate,

and human disturbance from the Sierra de Apeneca, El

Salvador. Quaternary Research, 61,

159–187.

Dull, R.A. (2007). Evidence for forest clearance, agriculture,

and human-induced erosion in Precolumbian El Salvador.

Annals of the Association of American Geographers, 97,

127–141.

Dull, R.A., Southon, J.R., & Sheets, P. (2001). Volcanism,

ecology, and culture: A reassessment of the Volcan

Ilopango TBJ eruption in the southern Maya realm. Latin

American Antiquity, 12, 25–44.

Dull, R.A., Southon, J.R., Kutterolf, S., Freundt, A., Wahl, D.,

& Sheets, P. (2010). Did the TBJ Ilopango eruption cause

the AD 536 event? American Geophysical Union, Fall

Meeting 2010, abstract #V13C-2370.

Dunning, N.P. (2004). Down on the farm: Classic Maya

houselots as farmsteads. In J. Lohse & F. Valdez (Eds)

Ancient Maya commoners (pp. 96–116). Austin: University

of Texas Press.

Dunning, N.P., & Beach, T. (2000). Stability and instability in

prehispanic Maya landscapes. In D. Lentz (Ed.), An

imperfect balance: Landscape transformations in the

pre-Columbian Americas (pp. 179–202). New York:

Columbia University Press.

Dunning, N.P., & Beach, T. (2003). Fruit of the luum: Lowland

Maya soil knowledge and agricultural practices. Mono y

Conejo, 2, 1–25.

Dunning, N.P., & Beach, T. (2004). Noxious or nurturing

nature? Maya civilization in environmental context. In

C.W. Golden & G. Borgstede (Eds.), Continuities and

changes in Maya Archaeology: Perspectives at the

millenium (pp. 125–141). New York: Routledge.

Dunning, N. & Beach, T. (2010). Farms and forests: spatial

and temporal perspectives on ancient Maya landscapes. In

I.P. Martini & W. Chesworth (Eds.), Landscapes and

societies (pp. 369–389). Berlin: Springer-Verlag.

Dunning, N.P., & Houston, S. (2011). Hurricanes as a

disruptive force in the Maya Lowlands. In C. Isendahl &

B.L. Persson (Eds.), Ecology, power, and religion in Maya

landscapes (pp. 49–59). Mockmuhl: Verlag Anton

Sauerwien.

Dunning, N., Beach, T., & Rue, D. (1997). The paleoecology

and ancient settlement of the Petexbatun Region,

Guatemala. Ancient Mesoamerica, 8, 255–266.

Dunning, N.P., Beach, T., Farrell, P., & Luzzadder-Beach, S.

(1998). Prehispanic agrosystems and adaptive regions in

the Maya Lowlands. Culture and Agriculture, 20, 87–100.

Dunning, N.P., Rue, D., Beach, T., Covich, A., & Traverse, A.

(1998) Human–environment interactions in a tropical

watershed: The paleoecology of Laguna Tamarindito,

Guatemala. Journal of Field Archaeology, 25, 139–151.

Dunning, N.P., Luzzadder-Beach, S., Beach, T., Jones, J.G.,

Scarborough, V.L., & Culbert, T.P. (2002) Arising from the

bajos: The evolution of a Neotropical landscape and the rise

of Maya civilization. Annals of the Association of American

Geographers, 92, 267–282.

Dunning, N., Jones, J.G., Beach, T., & Luzzadder-Beach, S.

(2003). Physiography, habitats, and landscapes of the

Three Rivers Region. In V. Scarborough, V. F. Valdez, &

N.P. Dunning (Eds.), Heterarchy, political economy and

the ancient Maya: The Three Rivers Region of the. east-

central Yucatan Peninsula (pp. 14–24). Tempe: University

of Arizona.

Dunning, N.P., Beach, T., & Luzzadder-Beach, S. (2006)

Environmental variability among bajos in the Southern

Maya Lowlands and its implications for ancient Maya

civilization and archaeology. In L. Lucero & B. Fash (Eds),

Precolumbian water management (pp. 81–99). Tempe:

University of Arizona Press.

Dunning, N.P., Beach, T., Luzzadder-Beach, S., & Jones, J.G.

(2009). Creating a stable landscape: Soil conservation

among the ancient Maya. In C. Fisher, B. Hill, & G.

Feinman (Eds), The archaeology of environmental change:

Socionatural legacies of degradation and resilience (pp.

85–105). Tempe: University of Arizona Press.

Dunning N.P, Beach T, & Luzzadder-Beach S. (2012) Kax and

kol: Collapse and resilience in lowland Maya civilization.

Proceedings of the National Academy of Sciences, 109,

3652–3657.

Dunning, N., D. Wahl, T. Beach, J. Jones, S.

Luzzadder-Beach, and C. McCormick (2014a). The end of

the beginning: Drought, environmental change and the

190 Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc.

DUNNING ET AL. GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA

Preclassic to Classic transition in the east-central Yucatan

Peninsula. In G. Iannone (Ed.), The Great Maya Droughts

in cultural context (pp. 107–129). Boulder: University of

Colorado Press.

Dunning, N., Weaver, E., Smyth, M.P., & Ortegon Zapata, D.

(2014b). Xcoch: Home of ancient Maya rain gods and

water managers. In T.W. Stanton (Ed.). The archaeology of

Yucatan: New directions and data (pp. 65–80). Oxford, UK:

BAR International Series.

Dunning, N.P., Griffin, R.E., Jones, J.G., Terry, R.E., Larsen,

Z., & Carr, C. (2015). Life on the edge: Tikal in a Bajo

landscape. In D.L. Lentz, N. P. Dunning, & V.L.

Scarborough (Eds.), Tikal: Paleoecology of an ancient Maya

city (pp. 95–123). New York: Cambridge University Press.

Eberl, M., Alvarez, M., & Terry, R.E. (2012). Chemical

signatures of middens at a Late Classic Maya residential

complex, Guatemala. Geoarchaeology, 27, 426–440.

Emery, K.F. (2003). Natural resource use and Classic Maya

economics: Environmental archaeology at Motul de San

Jose, Guatemala. Mayab, 16, 33–48.

Emery, K.F. (2007). Assessing the impact of ancient Maya

animal use. Journal for Nature Conservation, 15, 184–195.

Emery, K.F. (2008). A zooarchaeological test for dietary

resource depression at the end of the Classic Period in the

Petexbatun, Guatemala. Human Ecology, 36, 617–634.

Emery, K.F., & Thornton, E.K. (2008a). Zooarchaeological

habitat analysis of ancient Maya landscape changes.

Journal of Ethnobiology, 28, 154–178.

Emery, K.F., & Thornton, E.K. (2008b). A regional

perspective on biotic change during the Classic Maya

occupation using zooarchaeological isotopic chemistry.

Quaternary International, 191, 131–143.

Estrada-Belli, F., & Kock, M. (2007). Remote sensing and GIS

analysis of a Maya city and its landscape, Holmul,

Guatemala. In J. Wiseman & F. El-Baz (Eds.), Remote

sensing in archaeology (pp. 263–281). New York: Springer.

Fedick, S., & Morrison, B. (2004). Ancient use and

manipulation of landscape in the Yalahau region of the

northern Maya lowlands. Agriculture and Human Values,

21, 207–219.

Fedick, S., Morrison, B., Andersen B.J., Boucher, S., Acosta, J.

C., & Mathews, J. P. (2000). Wetland manipulation in the

Yalahau Region of the northern Maya Lowlands. Journal

of Field Archaeology, 27. 131–152.

Fernandez, F.G., Terry, R.E., Inomata, T., & Eberl, M. (2002).

An ethnoarchaeological study of chemical residues in the

floors and soils of Q’eqchi’ Maya houses at Las Pozas,

Guatemala. Geoarchaeology, 17, 487–519.

Fernandez, F.G., Johnson, K.D., Terry, R.E., Nelson, S., &

Webster, D. (2005). Soil resources of the ancient Maya at

Piedras Negras, Guatemala. Soil Science Society of America

Journal, 69, 2020–2032.

Fisher, C.T. (2005). Abandoning the garden: demographic and

landscape change in the Lake Patzcuaro Basin, Mexico.

American Anthropologist, 107, 87–95.

Fischer, C.T., Pollard, H.P., Israde-Alcantara, I.,

Garduno-Monroy, V.H., & Banerjee, S.K. (2003). A

reexamination of human-induced environmental change

within the Lake Patzcuaro Basin, Michoacan, Mexico.

Proceedings of the National Academy of Sciences, 100,

4957–4962.

Fisher, C.T., Hill, J.B., & Feinman, G.M. (Eds.) (2009). The

archaeology of environmental change: Socionatural

legacies of degradation and resilience. Tempe: University of

Arizona Press.

Fisher, C.T., Leisz, S., & Outlaw, G. (2011) LiDAR: A valuable

tool uncovers an ancient city in Mexico. Photogrammetric

Engineering and Remote Sensing, 77, 962–967.

Ford, A. (2008). Dominant plants of the Maya forest and

gardens of El Pilar: Implications for paleoenvironmental

reconstructions. Journal of Ethnobiology, 28,

179–199.

Ford, A., & Nigh, R. (2009). Origins of the Maya forest

garden: Maya resource management. Journal of

Ethnobiology, 29, 213–236.

Frappier, A.B., J. Pyburn, A.D. Pinkey-Drobnis, X. Wang, D.R.

Corbett, & B.H. Dahlin (2014), Two millennia of tropical

cyclone-induced mud layers in a northern Yucatan

stalagmite reveal: Multiple overlapping climatic hazards

during the Maya Terminal Classic megadroughts.

Geophysical Research Letters, 41, 5148–5157.

Freiwald, C.R. (2010). Dietary diversity in the upper Belize

River Valley: A zooarchaeological and isotopic perspective,

In J.E. Staller & M.D. Carrasco (Eds.), Pre-Columbian

foodways: Interdiscipilanary approaches to food, culture,

and markets in ancient Mesoamerica, (pp.399-420). New

York: Springer.

Galindo-Tovar, M.E., Ogata-Aguilar, N., & Arzate-Fernandez,

A.M. (2008). Some aspects of avocado (Persea Americana

Mill.) diversity and domestication in Mesoamerica. Genetic

Resources and Crop Evolution, 55, 441–450.

Galop, D., Lemonnier, E., Carozza, J.-M., & Metailie, J.-P.

(2004). Bosques, milpas, casas, y aguadas de antano. In C.

Arnauld, V. Breuil-Martınez, & E. Ponciano Alvarado

(Eds.), La Joyanca (La Libertad, Guatemala): Antigua

Ciudad Maya del Noreste del Peten (pp. 55–71). Guatema

la City: Asociacion Tikal.

Garrison T.G., & Dunning N.P. (2009). Settlement,

environment, and politics in the San Bartolo – Xultun

territory, El Peten, Guatemala. Latin American Antiquity,

20, 525–552.

Garrison, T.G., Houston, S.D., Golden, C., Inomata, T.,

Nelson, Z., & Munson, J. (2008). Evaluating the use of

IKONOS satellite imagery in the lowland Maya settlement

archaeology. Journal of Archaeological Science, 35,

2770–2777.

Garrison, T.G., Chapman, B., Houston, S., Roman, E., &

Lopez, J.L.G. (2011). Discovering ancient Maya settlements

using airborne radar elevation data. Journal of

Archaeological Science, 38, 1655–1662.

Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc. 191

GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA DUNNING ET AL.

Gill, R.B. (2000) The Great Maya Droughts. University of New

Mexico Press, Albuquerque.

Gill, R.B., & Keating, J.P. (2002). Volcanism and

Mesoamerican archaeology. Ancient Mesoamerica, 13,

125–140.

Gill, R.B., Mayewski, P.A., Nyberg, J., Haug, G.H. & Peterson,

L.C. (2007). Drought and the Maya Collapse. Ancient

Mesoamerica, 18, 283–302.

Gilli, A., Hodell, D.A., Kamenov, J.D., & Brenner, M. (2009).

Geological and archaeological implications of strontium

isotope analysis of exposed bedrock in the Chicxulub crater

basin, northwestern Yucatan, Mexico. Geology, 37,

723–726.

Gischler, E., Shinn, E.A., Oschmann, W., Fiebig, & Buster,

N.A. (2008). A 1500-year Holocene Caribbean climate

archive from Blue Hole, Lighthouse, Belize. Journal of

Coastal Research, 24, 1495–1505.

Goman, M., & Byrne, R. (1998). A 5000-year record of

agriculture and tropical forest clearance in the Tuxlas,

Veracruz, Mexico. The Holocene, 8, 83–89.

Goman, M., Joyce, A., & Mueller, R. (2005). Stratigraphic

evidence for anthropogenically induced coastal

environmental change from Oaxaca, Mexico. Quaternary

Research 63, 250–260.

Gomez-Pompa, A., Flores, S.J., & Sosa, V. (1987). The

“pet-kot”: A man-made tropical forest of the Maya.

Interciencia, 12, 10–15.

Gotz, C.M. (2008). Coastal and inland patterns of faunal

exploitation in the prehispanic northern Maya lowlands.

Quaternary International, 191, 154–169.

Griffin R., R. Oglesby, Sever, T. & Nair, U. (2014). Agricultural

landscapes, deforestation, and drought severity. In G.

Iannone (Ed.), The Great Maya Droughts in cultural

context (pp. 71–86). Boulder: University of Colorado Press.

Gunn, J.D., Foss, J.E., Folan, W.J., del Rosario Dominguez

Carrasco, M., & Faust, B.B. (2002). Bajo sediments and the

hydraulic system of Calakmul, Campeche, Mexico. Ancient

Mesoamerica, 13, 297–316.

Hageman, J.B., & Goldstein, D.J. (2009). An integrated

assessment of archaeobotanical recovery methods in the

neotropical rainforest of northern Belize: Flotation and dry

screening. Journal of Archaeological Science, 36,

2841–2852.

Hansen, R.D., Bozarth, S., Jacob, J., Wahl, D., & Schreiner, T.

(2002). Climatic and environmental variability in the rise

of Maya civilization. Ancient Mesoamerica, 13, 273–295.

Haug, G.H., Gunther, D., Peterson, L.C., Sigman, D.M.,

Hughen, K.A., & Aeschliman, B. (2003). Climate and the

collapse of Maya civilization. Science, 299, 1731–1735.

Heckbert, S. (2013). MayaSim: An agent-based model of the

ancient Maya socioecological system. Journal of Artificial

Societies and Social Simulation, 4, 11–25.

Heimo, M., Siemens, A.H., & Hebda, R. (2004). Prehispanic

changes in wetland topography and their implications to

past and future wetland agriculture at Laguna Mandinga,

Veracruz, Mexico. Agriculture and Human Values, 21,

313–327.

Heine, K. (2003). Paleo-pedological evidence of

human-induced environmental change in the

Puebla-Tlaxcala area (Mexico) during the last 3500 years.

Revista Mexicana de Ciencias Geologicas, 20, 235–244.

Hildebrand, A.R., Pilkington, M., Connors, M., Ortiz-Aleman,

C., & Chavez, R. E. (1995). Size and structure of the

Chicxulub Crater revealed by horizontal gravity gradients

and cenotes. Nature, 376, 415–417.

Hillesheim, M.B., Hodell, D.A., Leyden, B.W., Brenner, M.,

Curtis, J.H., Anselmetti, F.S., Ariztegui, D., Buck, D.G.,

Guilderson, T.P., Rosenmeier, M.F., & Schnurrenberger,

D.W. (2005). Lowland neotropical climate change during

the late deglacial and early Holocene. Journal of

Quaternary Science, 20, 363–376.

Hodell, D.A., Brenner, M. & Curtis, J. (1995). Possible role of

climate in the collapse of Classic Maya Civilization. Nature,

375, 391–394.

Hodell, D.A., Brenner, M., Curtis, J.H., & Guilderson, T.

(2001) Solar forcing of drought frequency in the Maya

Lowlands. Science, 292, 1367–1370.

Hodell, D.A, Quinn, R.L., Brenner, M., & Kamenov, G.

(2004). Spatial variation of strontium isotopes (87Sr/86Sr)

in the Maya region: A tool for tracking ancient human

migration. Journal of Archaeological Science, 31,

585–601.

Hodell, D.A., Brenner, M., & Curtis, J.H. (2005). Terminal

Classic drought in the northern Maya Lowlands inferred

from multiple sediment cores in Lake Chichancanab

(Mexico). Quaternary Science Review, 24, 1413–1427.

Hodell, D.A., Brenner, M., Curtis, J.H., Medina-Gonzalez, R.,

Rosenmeier, M.F., Guilderson, T.P., Chan-Can, E.I., &

Albornaz-Pat, A. (2005). Climate change on the Yucatan

Peninsula during the Little Ice Age. Quaternary Research,

63, 109–121.

Hodell, D.A., Brenner, M., & Curtis, J.H. (2007). Climate and

cultural history of the northeastern Yucatan Peninsula,

Quintana Roo, Mexico, Climatic Change, DOI

10.1007/s10584-006-9177-4

Hodell D.A., Anselmetti, F.S., Ariztegui, D., Brenner, M.,

Curtis, J.H., Grilli, A., Grzesnik, D.A., Guilderson, T.J.,

Mueller, A.D., Bush, M.B., Correa-Metrio, Escobar, J., &

Kutteroff, F. (2008). An 85-ka record of climate change in

lowland Central America. Quaternary Science Reviews, 27,

1152–1165.

Holley, G.R., Dalan, R., Woods, W., & Watters, H. (2000).

Implications of a buried Preclassic site in western Belize. In

S.R. Ahler (Ed.), Mounds, Modoc, and Mesoamerica:

Papers in Honor of Melvin L. Fowler (pp. 111–124). Illinois

State: Museum Scientific Papers vol. 28, Springfield, IL.

Holling, C.S., & Gunderson, L.H. (2002). Resilience and

adaptive cycles. In L.H. Gunderson and C.H. Holling (Eds.),

Panarchy: Understanding transformations in human and

natural systems (pp. 25–62). Washington, DC: Island Press.

192 Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc.

DUNNING ET AL. GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA

Hudson, P.F. (2004). Geomorphic context of the prehistoric

Huastec floodplain environments: Lower Panuco Basin,

Mexico. Journal of Archaeological Science, 31,

653–668.

Hudson, P.F., & Heitmuller, F.T. (2003). Local- and

watershed-scale controls on the spatial variability of natural

levee deposits in a large fine-grained floodplain: Lower

Panuco Basin, Mexico. Geomorphology, 56, 255–269.

Huntington, E. (1917). Maya civilization and climate changes.

In Proceedings of the XIX International Congress of

Americanists (pp. 150–164), 5–10 October 1914,

Washington, DC.

Iannone, G. (Ed.) (2014). The Great Maya Droughts in

cultural context. Boulder: University Press of Colorado.

Inanez, J.G., Bellucci, J.J., Rodriguez-Alegria, E., Ash, R.,

McDonough, W. & Speakman, R.J. (2010). Romita Ware

revisited: A reassessment of the provenance of Colonial

Mexican ceramics by LA-MC-ICP-MS and INAA. Journal of

Archaeological Science, 37, 2698–2704.

Isendahl, C. (2011). The weight of water: a new look at

pre-hispanic Puuc Maya reservoirs. Ancient Mesoamerica,

22, 185–197.

Jensen, C.T., Moriarty, M.D., Johnson, K.D., Terry, R.E., &

Emery, K.F. (2007). Soil resources of the Motul De San

Jose Maya: Correlating soil taxonomy and modern Itza

Maya soil classification within a Classic Maya

archaeological zone. Geoarchaeology, 22, 337–357.

Johnson, K.D., Terry, R.E., Jackson, M.W., & Golden, C.

(2007). Ancient soil resources of the Usumacinta River

region, Guatemala. Journal of Archaeological Science, 34,

1117–1129.

Johnson, K.D., Wright, D.R., & Terry, R.E. (2007).

Application of carbon isotope analysis to ancient maize

agriculture in the Petexbatun region of Guatemala.

Geoarchaeology, 22, 313–336.

Johnston, K.J. (2002). Protrusion, bioturbation, and

settlement detection during surface survey: The lowland

Maya case. Journal of Archaeological Method and Theory,

9, 1–67.

Johnston, K.J. (2004). The “invisible” Maya: Minimally

mounded residential settlement at Itzan, Peten, Guatemala.

Latin American Antiquity, 15, 145–175.

Joyce, A.A., & Goman, M. (2012). Bridging the theoretical

divide in Holocene landscape studies: Social and ecological

approaches to the ancient Oaxacan landscapes. Quaternary

Science Reviews, 55, 1–22.

Joyce, A.A., Mueller, R. (1992). The social impact of

anthropogenic landscape modification in the Rıo Verde,

Oaxaca, Mexico. Geoarchaeology, 7, 503–526.

Joyce, A.A., Mueller, R. (1997). Prehispanic human ecology

of the Rıo Verde drainage basin, Oaxaca, Mexico. World

Archaeology, 29, 75–94.

Kaplan, J. (2008). Hydraulics, cacao, and complex

developments at Preclassic Chocola, Guatemala: Evidence

and implications. Latin American Antiquity, 19, 399–413.

Kennett, D.J. & Beach, T.P. (2013). Archaeological and

environmental lessons for the Anthropocene from the

Classic Maya Collapse. Anthropocene, 4, 88–110.

Kennett, D.J., Breienbach, S.F.M., Aquino, V.V., Asmerom,

Y., Awe, J., Baldini, J.U.L., Bartlein, P., Culleton, B.J.,

Ebert, C., Jazwa, C., Macri, M.J., Marwan, N., Polyak, V.,

Prufer, K.M., Ridley, H.E., Sodemann, H., Winterhalder,

B., & Haug, G.H. (2012). Development and disintegration

of Maya political systems in response to climate change.

Science, 338, 788–791.

Knight, C.L.F. (2003). Obsidian production, consumption,

and distribution at Tres Zapotes: Piecing together political

economy. In C.A. Pool (Ed.), Settlement archaeology and

political economy at Tres Zapotes, Veracruz, Mexico (pp.

69–89). Los Angeles: The Cotsen Institute of

Archaeology.

Knight, C.L.F., & Glascock, M.D. (2009). The terminal

formative to Classic period obsidian assemblage at Palo

Errado, Veracruz, Mexico. Latin American Antiquity, 20,

507–524.

Kovach, R.L. (2004). Early earthquakes of the Americas. New

York: Cambridge University Press.

Kowalewski, S.A., Balkansky, A.K., Stiver Walsh, L.R.,

Pluckhahn, T.J., Chamblee, J.F., Perez Rodrıguez, V.,

Heredia Espinoza, V.Y., & Smith, C.A., (2009). Origins of

the Nuu: Archaeology in the Mixteca Alta, Mexico.

University Press of Colorado, Boulder.

Kutterolf, S., Freundt, A., & Perez, W. (2008). Pacific offshore

record of plinian arc volcanism in Central America: 2.

Tephra volumes and erupted masses. Geochemistry,

Geophysics, Geosystems 9, Q02S02,

doi:10.1029/2007GC001791.

Lawrence, D., D’Odorico, P., Diekmann, L., DeLonge, M., Das,

R., & Eaton, J. (2007). Ecological feedbacks following

deforestation create the potential for catastrophic

ecosystem shift in dry tropical forest. Proceedings of the

National Academy of Sciences, 104, 20696–20701.

Lemonnier, E., & Vanniere, B. (2013). Agrarian features,

farmsteads and homesteads in the Rıo Bec nuclear zone

(Mexico). Ancient Mesoamerica, 24, 397–414.

Lentz, D.L., & Hockaday, B. (2009). Tikal timbers and

temples: Ancient Maya agroforestry and the end of time.

Journal of Archaeological Science, 36, 1342–1353.

Lentz, D.L. & Ramırez-Sosa, C.R. (2002). Ceren plant

resources: Abundance and diversity. In P. Sheets (Ed.),

Before the volcano erupted: The Ceren village in central

America (pp. 33–42). Austin: University of Texas

Press.

Lentz, D.L., Yaeger, J., Robin, C., & Ashmore, W. (2005).

Pine, prestige and politics of the Late Classic Maya at

Xunantunich, Belize. Antiquity, 79, 573–585.

Lentz, D.L., Pohl, M.D., Alvarado, J.L., Tarighat, S., & Bye, R.

(2008). Sunflower (Heliathus annuus L.) as a

pre-Columbian domesticate in Mexico. Proceedings of the

National Academy of Sciences, 105, 6232–6237.

Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc. 193

GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA DUNNING ET AL.

Lentz, D.L., Dunning, N.P., Scarborough, V.L., Magee, K.,

Thompson, K.M., Weaver, E., Carr, C., Terry, R.E., Islebe,

G., Tankersley, K.B., Graziosa Sierra, L., Jones, J.G.,

Buttles, P., Valdez, F., & Ramos Hernandez, C. (2014).

Farms, forests and the edge of sustainability at the ancient

Maya city of Tikal. Proceedings of the National Academy of

Sciences, 111, 18513–18518.

Lentz, D.L., Magee, K., Weaver, E., Jones, J.G., Tankersley,

K.B., Hood, A., Islebe, G., Ramos, C., & Dunning, N.P.

(2015). Agroforestry and agricultural practices of the

ancient Maya of Tikal: Resilience and management of an

essential resource. In D.L. Lentz, N.P. Dunning, & V.L.

Scarborough (Eds.), Tikal: Paleoecology of an ancient Maya

city (pp. 152–185). New York: Cambridge University

Press.

Leyden, B.W. (2002). Pollen evidence for climatic variability

and cultural disturbance in the Maya Lowlands. Ancient

Mesoamerica, 13, 85–101.

Leyden, B.W., Brenner, M., & Dahlin, B. (1998) Cultural and

climatic history of Coba: A lowland Maya city in Quintana

Roo, Mexico. Quaternary Research, 49, 111–122.

Lozano-Garcia, S., Torres-Rodrıguez, E., Ortega, B., Vazquez,

G., & Caballero, M. (2013). Ecosystem responses to climate

and disturbance in western central Mexico during the late

Pleistocene and Holocene. Palaeogeography,

Palaeoclimatology, Palaeoecology, 370, 184–195.

Lucero, L.J. (2002). The collapse of the Classic Maya: A case

for the role of water control. American Anthropologist,

104, 814–826.

Luzzadder-Beach, S. (2000). Water resources of the

Chunchucmil Maya. Geographical Review, 90,

493–510.

Luzzadder-Beach, S., & Beach, T. (2009). Arising from the

wetlands: Mechanisms and chronology of landscape

aggradation in the northern coastal plain of Belize. Annals

of the Association of American Geographers, 99, 1–26.

Luzzadder-Beach, S., Beach, T., & Dunning, N.P. (2012)

Wetland fields as mirrors of drought and the Maya

abandonment. Proceedings of the National Academy of

Science, 109, 3646–3651.

McAnany, P.A., & Gallareta Negron, T. (2010). Bellicose rulers

and climatological peril? Retrofitting twenty-first-century

woes on eighth-century Maya society. In P.A. MCAnany &

N. Yoffee (Eds.), Questioning collapse: Human resilience,

ecological vulnerability, and the aftermath of empire (pp.

142–175). New York: Cambridge University Press.

McAuliffe, J.C., Sundt, P.C., Valiente-Banuet, A., Casas, A., &

Viveros, J.L. (2001). Pre-Columbian soil erosion, persistent

ecological changes, and collapse of a subsistence

agricultural economy in the semi-arid Tehuacan Valley,

Mexico’s ‘Cradle of Maize’. Journal of Arid Environments,

47, 47–75.

McClosky, T.A., & Keller, G. (2009). 5000 year sedimentary

record of hurricane strikes on the central coast of Belize.

Quaternary International, 195, 53–68.

McClung de Tapia, E., Solleiro-Rebolledo, E., Gama, J.,

Villalpando, J.L., & Sedov, S. (2003). Paleosols in the

Teotihuacan Valley, Mexico: Evidence for

paleoenvironment and human impacts. Revista Mexicana

de Ciencias Geologicas 20, 270–282.

McNeil, C.L., Burney, D.A., & Burney, L.P. (2010). Evidence

disputing deforestation as the cause for the collapse of the

ancient Maya polity of Copan, Honduras. Proceedings of

the National Academy of Sciences, 107, 1017–1022.

Manzanilla, L., & Barba, L. (1990). The study of activities in

Classic households: Two case studies from Coba and

Teotihuacan. Ancient Mesoamerica, 1, 41–49.

Masson, M.A. & Peraza Lope, C. (2008). Animal use at

Mayapan. Quaternary International, 191, 170–181.

Medina-Elizade M., Burns, S., Lea, D.W., Asmerom, S.,

vanGunten, L., Polyak, V., Vuille, M., & Karmalkar, A.

(2010) High resolution stalagmite climate record from the

Yucatan Peninsula spanning the Maya Terminal Classic

period. Earth & Planetary Science Letters, 298, 255–262.

Medina-Elizade, M., & Rohling, E.J. (2012). Collapse of

Classic Maya civilization related to modest reduction in

precipitation. Science 335, 956–959.

Mendoza, B., Jauregui, E., Diaz-Sandoval, R., Garcia-Acosta,

V., Velasco, V., & Cordero, G. (2005). Historical droughts in

central Mexico and their relation with El Nino. Journal of

Applied Meteorology and Climatology, 44, 709–716.

Metcalfe, S.E. (2006). Late Quaternary environments of the

northern deserts and Central Transvolcanic Belt of Mexico.

Annals of the Missouri Botanical Garden, 28, 258–273.

Metailie, J.P., Corozza, J.-M., Galup, D., & Arnauld, M.-C.

(2003). Lagos, bajos y paleo-paisajes en el Peten

noroccidental: El inicio de una investigacion geografica y

arqueologica (La Joyanca). In A. Breton, A.

Monod-Becquelin, & M.H. Ruz (Eds.), Espacios Mayas:

Representaciones, Usos, Creencias (pp. 23–48). Mexico

City: Universidad Autonoma de Mexico.

Metcalfe, S.E., Davies, S.J., Braisby, J.D., Leng, M.J., Newton,

A.J., Terrett, N.L., & O’Hara, S.L. (2007). Long- and

short-term change in the Patzcuaro Basin, central Mexico.

Palaeogeography, Palaeoclimatology, Palaeoecology, 247,

272–295.

Morehart, C.T. (2012). Mapping ancient chinampa landscapes

in the Basin of Mexico: a GIS and remote sensing approach.

Journal of Archaeological Science, 39, 2541–2551.

Morehart, C.T., Lentz, D.L., & Prufer, K.M. (2005). Wood of

the gods: The ritual use of pine (Pinus spp) by the ancient

lowland Maya. Latin American Antiquity, 16, 255–274.

Morell-Hart, S., Joyce, R.A. & Henderson, J.S. (2014).

Multi-proxy analysis of plant use at Formative Period Los

Naranjos, Honduras. Latin American Antiquity, 25, 65–81.

Morley, S.G. (1946). The ancient Maya. Stanford, CA:

Stanford University Press.

Moyes, H., Awe, J.J., Brook, G.A., & Webster, J.W. (2009).

The ancient Maya drought cult: Late Classic cave use in

Belize. Latin American Antiquity, 20, 175–206.

194 Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc.

DUNNING ET AL. GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA

Mueller, A.D., Islebe, G.A., Hillsheim, M.B., Grzesik, D.A.,

Anselmetti, F.S., Ariztegui, D., Brenner, M., Curtis, J.H.,

Hodell, D.A. & Venz, K.A. (2009). Climate drying and

associated forest decline in the lowlands of northern

Guatemala during the late Holocene. Quaternary Research,

71, 133–141.

Mueller, A.D., Islebe, G.A., Anselmetti, F.S., Ariztegui, D.,

Brenner, M., Hodell, D.A., Hajdas, I., Hamann, Y., Haug,

G.H., & Kennett, D.J. (2010). Recovery of the forest

ecosystem in the tropical lowlands of northern Guatemala

after disintegration of Classic Maya polities. Geology, 38,

523–526.

Munro-Strasiuk, M., Manahan, T., & Stockton, T. (2014).

Spatial and physical characteristics of Rejolladas in

NorthernYucatan Peninsula, Mexico: Implications for

understanding ancient Maya agriculture and settlement

patterns. Geoarchaeology, 29, 156–172.

Neff, H., Glascock, M.D., Charlton, T.H., Otis-Charlton, C. &

Nichols, D.L. (2000). Provenience investigation of ceramics

and obsidian from Otumba. Ancient Mesoamerica, 11,

307–321.

Neff, H, Pearsall, D.M., Jones, J.G., Arroyo de Pieters, B. &

Freidel, D.E. (2006a). Climate change and population

history in the Pacific Lowlands of southern Mesoamerica.

Quaternary Research, 65, 390–400.

Neff, H., Pearsall, D.M., Jones, J.G., Arroyo, B., Collins, S.K. &

Freidel, D.E. (2006b). Early Maya adaptive patterns:

Mid-Late Holocene paleoenvironmental evidence from

Pacific Guatemala. Latin American Antiquity, 17, 287–315.

Neff, H., Blomster, J.P., Glascock, M.D, Bishop, R.L.,

Blackman, M.J., Coe, M.D., Cowgill, G.L., Cyphers, A.,

Diehl, R.A., Houston, S., Joyce, A.A., Lipo, C.P. & Winter

M.C. (2006c). Smokescreens in the provenance

investigations of Early Formative Mesoamerican ceramics.

Latin American Antiquity, 17, 104–118.

Neff, H., Blomster, J.P., Glascock, M.D, Bishop, R.L.,

Blackman, M.J., Coe, M.D., Cowgill, G.L., Diehl, R.A.,

Houston, S., Joyce, A.A., Lipo, C.P., Stark, B.L. & Winter

M.C. (2006d). Methodological issues in the provenance

investigations of Early Formative Mesoamerican ceramics.

Latin American Antiquity, 17, 54–76.

Neff, H., Kovacevich, B. & Bishop, R.L. (2010).

Caracterizacion de los compuestos de la jadeite

Mesoamercana: Breve revision a partir de low resultados

obtenidos durante el studio de la mascara de K’inich

Janaab’ Pakal. In L. Filloy Nadal (Ed.), Jabaab’ Pakal de

Palenque: Vida y Muerte de un Gobernante Maya (pp.

131–138. Mexico City: Instituto Nacional de Antropologıa

y Historia.

Nichols, D.L, Spence, M., & Borland, M. (1991). Watering the

fields of Teotihuacan: Early irrigation at the ancient city.

Ancient Mesoamerica, 2, 119–129.

Nicholas, D.L., Stoner, W. & Crider, D. (2014). A geospatial

approach to ceramic chemical composition data in the

Basin of Mexico. Paper presented at the 79th Annual

Meeting of the Society for American Archaeology, Austin,

TX, April 23–27.

Oglesby, R.J., Sever, T.L., Saturno, W., Erickson, D.J. III, &

Srikishen, J. (2010). Collapse of the Maya: Could

deforestation have contributed? Journal of Geophysical

Research 115, D12106, doi:10.1029/2009JD01 1942.

O’Hara, S.L., Street-Perrott, F.A. & Burt, T.P. (1993).

Accelerated soil erosion around a Mexican highland lake

caused by prehispanic agriculture. Nature, 362, 48–51

Ortega, B., Caballero, C., Lozano, S., Israde, I., & Vilaclara, G.

(2002). 52000 years of environmental history in Zacapu

basin, Michoacan, Mexico: The magnetic record. Earth and

Planetary Science Letters, 202, 663–675.

Ortega, B., Vazquez, G, Caballero, M., Israde, I.,

Lozano-Garcıa, S., Schaaf, P., & Torres, E. (2010). Late

Pleistocene-Holocene record of environmental changes in

Lake Zirahuen, central Mexico. Journal of Paleolimnology,

44, 745–760.

Park, J., Byrne, R., Bohnel, H., Molina Garza, R., & Conserva,

M. (2010). Holocene climate change and human impact,

central Mexico: A record based on maar lake pollen and

sediment chemistry. Quaternary Science Reviews, 29,

618–632.

Parnell, J.J., Jacob, J., Terry, R.E., & Golden, C. (2001). Using

in-field soil phosphate testing to rapidly identify middens at

Piedras Negras, Guatemala. Geoarchaeology, 16,

855–863.

Parnell, J.J., Sheet, P., & Terry, R. (2002a). Soil chemical

analysis of ancient activities in Ceren, El Salvador: A case

study of a rapidly abandoned site. Latin American

Antiquity, 13, 331–345.

Parnell, J.J., Terry, R.E., & Nelson, Z. (2002b). Soil chemical

analysis applied as an interpretive tool for ancient human

activities at Piedras Negras, Guatemala. Journal of

Archaeological Science, 29, 379–404.

Parsons, J.R., & Morett, L. (2004). Recursos acuaticos en la

subsistencia Azteca. Arqueologıa Mexicana, 12, 38–53.

Perez Rodrıguez, V., Anderson, K.C., & Neff, M.K. (2011).

The Cerro Jazmın Archaeological Project: Investigating

prehispanic urbanism and its environmental impact in the

Mixteca Alta, Oaxaca, Mexico. Journal of Field

Archaeology, 36, 83–99.

Perez Rodrıguez, V. & Anderson, K.C. (2013). Terracing in the

Mixteca Alta, Mexico: Cycles of resilience of an ancient

land-use strategy. Human Ecology, 41, 335–349.

Perez-Salicrup D. (2004) Forest types and their implications.

In B.L. Turner II, J. Geoghagen, & D.R. Foster (Eds.),

Integrated land-change science and tropical deforestation

in the southern Yucatan (pp. 63–80). New York: Oxford

University Press.

Perry, E.C., Velazquez-Oliman, G., Wagner, N., Paytan, A., &

Street, J. (2010). Sr ions and isotopes in groundwater,

Campeche, Mexico: Insight into the development of Edzna

Valley. In P. Birkle & I.S. Torres-Alvardo (Eds.),

Water-rock interaction (pp. 349–352). London: CRC Press.

Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc. 195

GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA DUNNING ET AL.

Petersen, J.J., Parker, I.M., & Potter, D. (2012). Origins and

close relatives of a semi-domesticated Neotropical fruit tree:

Chrysophyllum cainito (Sapotaceae). American Journal of

Botany, 99, 585–604.

Peterson, L.C., & Haug, G.H. (2005). Climate and the collapse

of Maya civilization: A series of multi-year droughts helped

to doom an ancient culture. American Scientist, 93,

322–329.

Plunkett, P., & Urunuela, G. (1998). Preclassic household

patterns preserved under volcanic ash at Tetimpa, Puebla,

Mexico. Latin American Antiquity 9,

287–309.

Pohl, M.E.D., Piperno, D.R., Pope, K.O., & Jones, J.G. (2007).

Microfossil evidence for Pre-Columbian maize dispersals in

the Neotropics from San Andres, Tabasco, Mexico.

Proceedings of the National Academy of Sciences, 104,

6870–6875.

Polk, J.S., vanBeynen, P.E., & Reeder, P.P. (2007). Late

Holocene environmental reconstruction using cave

sediments from Belize. Quaternary International, 68,

53–63.

Ponomarenko, A.L. (2004). The Pachuca obsidian source,

Hidalgo, Mexico: A geoarchaeological perspective.

Geoarchaeology, 19, 71–91.

Pool, C.A., Knight, C.L.F., & Glascok, M.D. (2014). Formative

obsidian procurement at Tres Zapotes, Veracruz, Mexico:

Implications for Olmec and Epi-Olmec political economy.

Ancient Mesoamerica, 25, 271–293.

Pope, K.C., Ocampo, S.C., Kinsland, G.L., & Smith, R. (1996)

Surface expression of the Chicxulub Crater. Geology, 24,

527–530.

Popenoe de Hatch, M., Ponciano, E., Brenner, M. & Ortloff, C.

(2002). Climate and technological innovation at

Kaminaljuyu, Guatemala. Ancient Mesoamerica, 13,

103–114.

Price, T.D., Burton, J.H., Fullagar, P.D., Wright, L.E., Bulkstra,

J.E., & Tiesler, V. (2008). Strontium isotopes and the study

of human mobility in ancient Mesoamerica. Latin

American Antiquity, 19, 167–180.

Price, T.D., Burton, J.H., Sharer, R., Bulkstra, J.E., Wright,

L.E., Traxler, L.P., & Miller, K.A. (2010). Kings and

commoners at Copan: Isotopic evidence for origins and

movement in the Classic Maya Period. Journal of

Anthropological Archaeology, 29, 15–32.

Pyburn, K.A., Dixon, B., Cook, P. & McNair, A. (1998). The

Albion Island settlement pattern project: Domination and

resistance in Early Classic northern Belize. Journal of Field

Archaeology 25, 37–62

Redman, C.L. (2005). Resilience in archaeology. American

Anthropologist, 107, 70–77.

Reents-Budet, D., Bishop, R.L., Valdes, J.A., Blackman, J.,

2006. La ceramica de Kaminaljuyu: Nuevos datos quımicos.

In Laporte, J.P., Arroyo, B., & Mejia, H. (Eds.), XIX

Simposio de Investigaciones Arqueologicas en Guatemala,

2005 (pp. 183–188). Guatemal a City: Asociacion Tikal.

Rice, D.S. (1993). Eighth-century physical geography,

environment, and natural resources in the Maya Lowlands.

In J.A. Sabloff & J.S. Henderson (Eds.), Lowland Maya

civilization in the eighth century A.D. (pp. 11–63).

Washington, DC: Dumbarton Oaks.

Robin, C. (Ed.) (2012). Chan: An ancient Maya farming

community. Gainsville: University of Florida Press.

Rosenmeier, M.F., Hodell, D.A., Curtis, J.H., & Guilderson,

T.P. (2002). A 4000-year lacustrine record of

environmental change in the southern Maya Lowlands,

Peten, Guatemala. Quaternary Research, 57, 183–190.

Rosenswig, R.M. (2006). Sedentism and food production in

early complex societies of the Soconusco, Mexico. World

Archaeology, 38, 330–355.

Rosenswig, R.M., Lopez, R., Antonelli, C. & Mendelsohn, R.

(2012). LiDAR mapping and surface survey of the Izapa

state on the tropical piedmont of Chiapas, Mexico. Journal

of Archaeological Science, 40, 1493–1507.

Ross, N.J. (2011). Modern tree species composition reflects

ancient Maya “forest gardens” in northwest Belize.

Ecological Applications, 21, 75–84.

Ross, N.J., & Rangel, T.F. (2011). Ancient Maya agroforestry

echoing through spatial relationships in the extant forest of

NW Belize. Biotropica, 43, 141–148.

Rue, D.J. (1987). Early agriculture and early Postclassic Maya

occupation in western Honduras. Nature, 326,

285–286.

Safi, K.N., Chinchilla Mazariegos, O., Lipo, C.P., & Neff, H.

(2012). Using ground-penetrating radar to examine spatial

organization at the Late Classic Maya site of El Baul,

Cotzumalhuapa, Guatemala. Geoarchaeology, 27,

410–425.

Sanchez-Perez, S., Solleiro-Rebollado, E., Sedov, S., McClung

de Tapia, E., Golyeva, A., Prado, B. & Ibarra-Morales, E.

(2013). The Black San Pablo Paleosol of the Teotihuacan

Valley, Mexico: Pedogenesis, fertility, and use in ancient

agricultural systems. Geoarchaeology, 28, 249–267.

Sapper, K. (1896). La geografıa fisica y la geologıa de la

peninsula de Yucatan. Boletin del Instituto de Geologıa de

Mexico, 3, 1–57.

Saturno, W., Sever, T.L., Irwin, D.E., Howell, B.F., &

Garrison, T.G. (2007). Putting us on the map: Remote

sensing investigation of the ancient Maya landscape.In J.

Wiseman & F. El Baz (Eds.), Remote sensing in

archaeology (pp. 137–160). New York: Springer.

Sax, M., Walsh, J.M., Freestone, I.C., Rankin, A.H., & Meeks,

N.D. (2008). The origins of two purportedly pre-Columbian

crystal skulls. Journal of Archaeological Science, 35,

2751–2760.

Scarborough, V.L. (2009). The archaeology of sustainability:

Mesoamerica. Ancient Mesoamerica, 20, 197–203.

Scarborough, V.L., Dunning, N.P., Tankersley, K.B., Carr, C.,

Weaver, E., Grazioso, L., Lane, B., Jones, J.G., Buttles, P.,

Valdez, F., & Lentz, D.L. (2012). Water and sustainable

land use at the ancient tropical city of Tikal, Guatemala.

196 Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc.

DUNNING ET AL. GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA

Proceedings of the National Academy of Sciences, 109,

12408–12413.

Sedov S, Solleiro-Rebolledo E, Gama-Castro, J.E.,

Vallejo-Gomez, E., & Gonzalez-Velazquez, A. (2001).

Buried paleosols of Navado de Toluco: An alternative recod

of Late Quaternary environmental change in central

Mexico. Journal of Quaternary Science, 16,

375–389.

Sedov S, Solleiro-Rebolledo E, Fedick S, Pi-Puig T,

Vallejo-Gomez, E., & del Lourdes Flores-Degadillo, M.

(2008) Micromorphology of a soil catena in Yucatan:

pedogenesis and geomorphological processes in a tropical

karst landscape. In S. Kapur, A. Mermut, & G. Stoops

(Eds.), New trends in soil micromorphology (pp. 19–37).

Berlin: Springer.

Sedov, S., Lazano-Garcıa, S., Solleiro-Rebolledo E, McClung

de Tapia, E., Ortega-Guerrero, B., & Sosa-Najera, S. (2010).

Tepexpan revisited: A multi-proxy of local environmental

changes in relation to human occupation from a paleolake

shore section in Central Mexico. Geomorphology, 122,

309–322.

Sever, T.L., & Irwin, D.E. (2003). Landscape archaeology:

Remote-sensing investigation of the ancient Maya in the

Peten rainforest of northern Guatemala. Ancient

Mesoamerica, 14, 113–122.

Sharer, R.J., Balkansky, A.K., Burton, J., Feinman, G.M.,

Flannery, K.V., Grove, D.C., Marcus, J., Moyle, R.G., Price,

T.D., Redmond, E.M., Reynolds, R.G., Rice, P.M., Spencer,

C.S., Stoltman, J.B. & Yeager, J. (2006). On the logic of

archaeological inference: Early Formative pottery and the

evolution of Mesoamerican societies. Latin American

Antiquity, 17, 90–103.

Shaw, J.M. (2003). Climate change and deforestation:

Implications for the Maya collapse. Ancient Mesoamerica,

14, 157–167.

Sheets, P. (1999). The effects of explosive volcanism on

ancient egalitarian, ranked, and stratified societies in

Middle America. In A. Oliver-Smith & S. Hoffman (Eds.),

Angry earth: Disasters in anthropological perspectives (pp.

36–58). New York, Routledge.

Sheets, P. (Ed.) (2002). Before the volcano erupted: The

ancient Ceren Village in central America. Austin:

University of Texas Press.

Sheets, P. (2006). The Ceren site: An ancient village buried by

volcanic ash in central America. Belmont, CA: Thomson

Higher Education.

Sheets, P. (2008). Armageddon to the Garden of Eden:

Explosive volcanic eruptions and societal resilience in

ancient Middle America. In D.H. Sandweiss, & J. Quilter

(Eds.), El Nino, catastrophism, and culture change in

ancient America (pp. 167–186). Cambridge, MA: Harvard

University Press.

Sheets, P. (2009). Contributions of geoarchaeoloogy to

Mesoamerican studies. Ancient Mesoamerica, 20,

205–2010.

Siemens, A.H., Soler Graham, J.A., Hebda, R. & Heimo, M.

(2002). “Dams” on the Candelaria. Ancient Mesoamerica,

13, 115–123.

Sheets, P., Dixon, C., Guerra, M. & Blanford, A. (2011).

Manioc cultivation at Ceren, El Salvador: occasional

kitchen garden plant or staple crop? Ancient Mesoamerica,

22, 1–11.

Sheets, P., Lentz, D., Piperno, D., Jones, J., Dixon, C., Maloof,

G. & Hood, A. (2012). Ancient manioc agriculture south of

the Ceren Village, El Salvador. Latin American Antiquity,

23, 259–281.

Sieron, K., & Siebe, C. (2008). Revised stratigraphy and

eruption rates of Ceboruco Stratovolcano and

surrounding monogenetic vents (Nayarit, Mexico) from

historical documents and new radiocarbon dates.

Journal of Volcanology and Geothermal Research, 176,

241–264.

Smyth, M.P., Zubrow, E., Ortegon Zapata, D., Dunning, N.P.,

Weaver, E., & van Beynan, P. (2011). Paleoclimatic

reconstruction and archaeological investigations at Xcoch

and the Puuc region of Yucatan, Mexico. Washington, DC,

National Science Foundation.

Solar V.L. (Ed.) (2010). El sistema fluvial Lerma-Santiago

durante el Formativo y el Clasico Temprano. Mexico City:

Instituto Nacional de Antropologıa e Historıa.

Solis-Castillo, B., Solleiro-Rebolledo, E., Sedov, S., Liendo, R.,

Ortiz-Perez, M., & Lopez-Rivera, S. (2013).

Paleoenvironments and human occupation of the Maya

Lowlands of the Usumacinta River, southern Mexico.

Geoarchaeology, 28, 268—288.

Somerville, A.D., Fauvelle, M., & Froehle, A.W. (2013).

Applying new approaches to modeling diet and status:

Isotopic evidence for commoner resiliency and elite

variability in the Classic Maya lowlands. Journal of

Archaeological Science, 40, 1539–1553.

Speller, C. (2013). Analisis de los suelos de Yaxnohcah, sur de

Campeche, Mexico: Aplicando Tecnicas de AND. In K.

Reese-Taylor & A.A. Hernandez (Eds.), Proyecto

Arqueologico Yaxnocah 2011: Informe de la Primara

Temporada de Investigaciones (pp. 82–86).

http://www.mesoweb.com/resources/informes/Yaxnohcah

2011.pdf (accessed 23 September 2013)

Stable, D.W., Villanueava-Diaz, J., Burnette, D.J., Cerano

Paredes, J., Heim, R.R. Jr., Fye, F.K., Acuna Soto, R.,

Therrell, M.D., Cleaveland, M.K., & Stahle, D.K. (2011).

Major Mesoamerican droughts of the past millennium.

Geophysical Research Letters, 38, DOI:

10.1029/2010GL046472.

Stahle, D.W., Villanueva Diaz, J., Burnette, D.J., Cerano

Paredes, J., Heim Jr., J.J., Fye, F.K., Acuna Soto, R.,

Therell, M.D., Cleaveland, M.K. & Stahle, D.K. (2011).

Major Mesoamerican droughts of the past millennium.

Geophysical Research Letters, 38, 1–4.

Stark, B.L., & Garraty, C.P. (2008). Parallel archaeological and

visibility survey in the western lower Papaloapan basin,

Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc. 197

GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA DUNNING ET AL.

Veracruz, Mexico. Journal of Field Archaeology, 38,

177–196.

Stark, B.L., & Ossa, A. (2007). Ancient settlement, urban

gardening, and environment in the Gulf Lowlands of

Mexico. Latin American Antiquity, 18, 385–406.

Stoltman, J.B. (2011). New petrographic evidence pertaining

to ceramic production and importation at the Olmec site of

San Lorenzo. Archaeometry, 53, 510–527.

Stoltman, J.B., Marcus, J., Flannery, K.V., Burton, J.H. &

Moyle, R. (2005). Petrographic evidence shows that

pottery exchange between the Olmecs and their neighbors

was two way. Proceedings of the National Academy of

Sciences, 102, 11213–11218.

Tankersley K.B., Scarborough, V., Dunning, N.P., Huff, W.,

Maynard, B., & Gerke, T. (2011). Evidence for volcanic ash

fall in the Maya lowlands from a reservoir at Tikal,

Guatemala. Journal of Archaeological Science, 38,

2925–2938.

Tankersley, K.B., Dunning, N.P., Scarborough, V.L., Jones,

J.G., Carr, C., & Lentz, D.L. (2015). Fire and water: The

archaeological significance of Tikal’s Quaternary sediments.

In D.L. Lentz, N.P. Dunning, & V.L. Scarborough (Eds.),

Tikal: Paleoecology of an ancient Maya city (pp. 186–211).

New York: Cambridge University Press.

Terry, R.E., Fernandez, F.G., Parnell, J.J., & Inomata, T.

(2004). The story in the floors: Chemical signatures of

ancient and modern Maya activities at Aguateca,

Guatemala. Journal of Archaeological Science, 31,

1237–1250.

Thompson, K., Hood, A., Cavellero, D., & Lentz, D.L. (2015).

Connecting contemporary ecology and ethnobotany to

ancient plant use practices of the Maya at Tikal. In D.L.

Lentz, N.P. Dunning, & V.L. Scarborough (Eds.), Tikal:

Paleoecology of an ancient Maya city (pp. 124–151). New

York: Cambridge University Press.

Thornton, E.K. (2011). Reconstructing ancient Maya animal

trade through strontium isotope (87Sr/86Sr) analysis.

Journal of Archaeological Science, 38, 3254–3263.

Turner, B.L., & Sabloff, J.A. (2012). Classic Period collapse of

the Central Maya Lowlands: Insights about

human-environment relationships for sustainability.

Proceedings of the National Academy of Science, 109,

13908–13914.

Van Derwarker, A.M. (2005). Field cultivation and tree

management in tropical agriculture: A view from Gulf

Coastal Mexico. World Archaeology, 37, 275–289.

Van Hengstum, P.J., Reinhardt, E.G., Beddows, P.A., &

Gabriel, J.J. (2009). Linkages between Holocene

paleoclimate and paleohydrogeology preserved in a

Yucatan underwater cave. Quaternary Science Reviews 29,

2788–2798.

Velez, M.L., Curtis, J.H., Brenner, M., Escobar, J., Leyden,

B.W., & Popenoe de Hatch, M. (2011). Environmental and

cultural changes in highland Guatemala inferred from Lake

Amatitlan sediments. Geoarchaeology, 26, 346–364.

Venter, M.K., Thompson, V.D., Reynolds, M.D., & Waggoner,

J.C. (2006). Integrating shallow geophysical survey:

Archaeological investigations at Totogal in the Sierra de los

Tuxtlas, Veracruz, Mexico. Journal of Archaeological

Science, 33, 767–777.

Wahl, D., Byrne, R., Schreiner, T., & Hansen, R. (2006).

Holocene vegetation change in the northern Peten and its

implications for Maya prehistory. Quaternary Research, 65,

380–289.

Wahl, D., Byrne, R., Schreiner, T., & Hansen, R. (2007).

Palaeolimnological evidence of late-Holocene settlement

and abandonment in the Mirador Basin, Peten, Guatemala.

The Holocene, 17, 813–820.

Wahl, D., Byrne, R. & Anderson, L. (2014). An 8700 year

paleoclimate reconstruction from the southern Maya

lowlands. Quaternary Science Reviews, 103,

19–25.

Webb, E.A., Schwarcz, H.P., & Healy, P.F. (2004). Detection

of ancient maize in lowland Maya soils using stable carbon

isotopes: Evidence from Caracol, Belize. Journal of

Archaeological Science, 31, 1039–1052.

Webb, E.A., Schwarcz, H.P., Jensen, C.T., Terry, R.E.,

Moriarty, M.D., & Emery, K.F. (2007). Stable carbon

isotope signature of ancient maize agriculture in the soils of

Motul de San Jose, Guatemala. Geoarchaeology, 22,

291–213.

Webster, D.L. (2002) The fall of the ancient Maya: Solving the

mystery of the Maya collapse. London and New York:

Thames and Hudson.

Webster, J.W., Brook, G.A., Railsback, L.B., Cheng, H.,

Lawrence, R, Alexander, C., & Reeder, P. (2007).

Stalagmite evidence from Belize indicating significant

droughts at the time of Preclassic Abandonment, the Maya

Hiatus, and the Classic Maya collapse. Palaeogeography,

Palaeoclimatology, Palaeoecology, 250, 1–17.

Wendt, C.J. (2003). Buried occupational deposits at Tres

Zapotes: The results of an auger testing program. In C.A.

Pool (Ed.), Settlement archaeology and political economy

at Tres Zapotes, Veracruz, Mexico (pp. 32–46). Los Angeles:

The Cotsen Institute of Archaeology.

Wendt, C.J., & Lu, S.-T. (2006). Sourcing archaeological

bitumen in the Olmec Region. Journal of Archaeological

Science, 33, 89–97.

Weishampel, J.F., Hightower, J.N., Chase, A.F., Chase, D.Z., &

Patrick, R.A. (2011) Detection and morphologic analysis of

potential below-canopy cave openings in the karst

landscape around the Maya polity of Caracol using

airborne LiDAR. Journal of Cave and Karst Studies, 73,

187–196.

White, C.D., F.J. Longstaffe, M.W. Spence, & K. Law. (2000).

Testing the nature of Teotihuacan imperialism at

Kaminaljuyu using phosphate oxygen-isotope ratios.

Journal of Anthropological Research, 56, 535–558.

White, C.D., F.J. Longstaffe, & K.R. Law. (2001). Revisiting

the Teotihuacan connection at Altun Ha: Oxygen-isotope

198 Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc.

DUNNING ET AL. GEOARCHAEOLOGICAL INVESTIGATIONS IN MESOAMERICA

analysis of Tomb F-8/1. Ancient Mesoamerica, 12,

65–72.

White, C.D., Schwarcz, H.P., Pohl, M.D., & Longstaffe, F.J.

(2004). Feast, field, and forest: Deer and dog diets at

Lagartero, Tikal, and, Copan. In K.F. Emery (Ed.), Maya

zooarchaeology: New directions in method and theory

(pp. 141–158). Los Angeles, Costen Institute of

Archaeology.

Winemiller, T.L. (2007). The Chicxulum Meteor impact and

ancient locational decisions on the Yucatan Peninsula,

Mexico: The application of remote sensing, GIS, and GPS in

settlement pattern studies. ASPRS Proceedings.

http://www.asprs.org/a/publications/proceedings/tampa

2007/0080.pdf (accessed 7 October 2013).

Wollwage, L., Fedick, S., Sedov, S., & Solleiro-Rebolledo, E.

(2012). The deposition and chronology of Cenote T’isil: A

multiproxy study of human/environment interaction in

the northern Maya Lowlands of southeast Mexico.

Geoarchaeology, 27, 441–446.

Wright, D.R., Terry, R.E., & Eberl, M. (2009). Soil properties

and stable carbon isotope analysis of landscape features in

the Petexbatun region of Guatemala. Geoarchaeology, 24,

466–491.

Wright, L.E. (2005). Identifying immigrants to Tikal,

Guatemala: Defining local variability in strontium isotope

ratios of human tooth enamel. Journal of Archaeological

Science, 32, 555–566.

Wright, L.E. (2006). Diet, health, and status among the Pasion

Maya: A reappraisal of the collapse. Nashville: Vanderbilt

University Press.

Wright, L.E. (2012). Immigration to Tikal: Evidence from

stable strontium and oxygen isotopes. Journal of

Anthropological Archaeology, 31, 334–352.

Wright, L.E., Valdes, J.A., Burton, J.H., Price, T.D., &

Schwarz, H.P. (2010). The children of Kaminaljuyu:

Isotopic insights into diet and long distance interaction in

Mesoamerica. Journal of Anthropological Archaeology, 29,

155–178.

Wyatt, A.R. (2012). Agricultural practices at Chan: Farming

and political economy in an ancient Maya community. In

C. Robin (Ed.), Chan and ancient Maya farming

community (pp. 71–88). Gainsville: University of Florida

Press.

Yeager J., & Hodell, D. (2009). The collapse of Maya

civilization: Assessing the interaction of culture, climate,

and environment. In D.H. Sandweiss & J. Quilter (Eds), El

Nino, catastrophism, and culture change in ancient

America (pp. 187–242). Cambridge, MA: Harvard

University Press.

Geoarchaeology: An International Journal 30 (2015) 167–199 Copyright C© 2015 Wiley Periodicals, Inc. 199