natural and cultural landscape evolution during the … · natural and cultural landscape evolution...

217
Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands Carlos Enrique Avendaño Mendoza A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Graduate Department of Geography University of Toronto © Copyright by Carlos Enrique Avendaño Mendoza, 2012

Upload: lyhuong

Post on 27-Sep-2018

219 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

Natural and cultural landscape evolution during

the Late Holocene in North Central Guatemalan

Lowlands and Highlands

Carlos Enrique Avendaño Mendoza

A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy

Graduate Department of Geography University of Toronto

© Copyright by Carlos Enrique Avendaño Mendoza, 2012

Page 2: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

ii

Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan

Lowlands and Highlands

Doctor of Philosophy 2012

Carlos Enrique Avendaño Mendoza Graduate Department of Geography

University of Toronto

Abstract  Paleoecology has been only in recent decades applied to Mesoamerica; this thesis

provides new records of paleoenvironmental changes in Guatemala. Paleoecological

reconstructions are developed based mainly on pollen in the Lachuá lowlands and

Purulhá highlands of the Las Verapaces Region. For the first time, quantitative vegetation

and climate analyses are developed, and plant indicator taxa from vegetation belts are

identified. Changes in vegetation are explained partially by elevation and climatic

parameters, topography, drainage divides, and biogeography. Pollen rain and indicator

plant taxa from vegetation belts were linked through a first modern pollen rain analysis

based on bryophyte polsters and surface sediments. The latter contain fewer forest-

interior plant taxa in both locations, and in the highlands, they contain higher local pollen

content than in the lowlands. These calibrations aided vegetation reconstructions based

on fossil pollen in sediment records from the Lachuá and Purulhá regions.

Reconstructions for the last ~2000 years before present (BP) were developed based on

fossil pollen from cores P-4 on a floodplain in Purulhá, and L-3, a wetland in Lachuá.

Page 3: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

iii

Core P-4 suggests that Mayan populations developed a system of agricultural terraces in

a former paleolake-swamp environment, which was abandoned at the time of the Spanish

Conquest (~400 BP). Core L-3 indicates the abandonment of Mayan “Forest Gardens” at

the time of the early Postclassic. These gardens likely prevailed during the Classic period

(~300-1100 yrs BP) at the outskirts of the ancient city of Salinas de los Nueve Cerros.

Following abandonment, forest recovery took place for about 800 yrs. Cultural factors are

found to be more important in determining vegetation dynamics in this region, since no

clear evidence of climate forcing was found. The P-4 and L-3 cores provide likely

evidence that Mayan populations were, contrary to other evidence, innovative landscape

managers. Scenarios in the Las Verapaces Region have been drastically modified in

recent times (e.g. after the European Conquest), as suggested by pollen evidence in the

top of both P-4 and L-3 cores, possibly due mostly to modern large scale natural

resources exploitation, which represent environmental threats greater than any seen in the

last ca. 2000 years.

Page 4: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

iv

Acknowledgments 

I deeply thank my supervisor Sharon Cowling for her sincere and wonderful support

during the development of my Ph.D since the very first day I arrived to Canada. She was

there waiting for me in the Toronto International Pearson Airport with a sign that had my

name on, I can only say “Muchas gracias eternas”. Thanked is my co-supervisor Sarah

Finkelstein for her marvellous support at the Paleoecology Laboratory of the Department

of Geography. I had the honor to be at the start of her Laboratory and see the evolution to

what today is: An excellent place to learn and grow.

I thank Prof. Tenley Conway and Prof. Anthony Davis for their helpful comments as part

of my Ph.D. Academic Committee. Prof. Juan Carlos Berrio is greatly thanked for his

valuable training in tropical paleoecology during field campaign in Guatemala and during

my visits to his laboratory at the Department of Geography, Leicester University,

England. I thank his wife Natalia de Berrio for her support too. I thank too the “Los

Juanetes”, a Latin American rock band in the middle of England, for making my visit a

nice one. I thank the Guatemalan team, “los COMPAI” and more, that supported me

during my field campaign in Guatemala in 2006 and many many more things.

Lachuá National Park and Biotopo del Quetzal Administrations and staff are thanked for

supporting my research. I am greatly thankful to forest guards at Lachuá National Park

for their support in bryophyte polster and core sampling. I thank Santa Lucia Lachuá

Municipality for support in collecting sediments from Salinas de los Nueve Cerros,

especially to Major Pedro Oxom and Family Tun. San Cristobal Verapaz Municipality

Page 5: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

v

Administration and staff are thanked for supporting our research. Fincas Villa Trinidad,

Patal, Chisiguan, and Lesbia Mus and Yolanda Barahona are thanked for their support in

collecting sediments.

I thank CONAP, Franklin Herrera, Escuela de Biología –at the Faculty of CCQQ and

Pharmacy –USAC- for the support in acquiring collection and research licenses. As well

I thank staff and members of Escuela de Biología, Faculty of CCQQ and Pharmacy, and

USAC for their support during my Ph.D.

I thank Dr. Gerald Islebe for his support in pollen identification and feedback during my

thesis development. Enric Aguilar and Melissa Gervais are thanked for obtaining

Guatemalan climatic information. Joan Bunbury is greatly thanked for the support in

creating maps and using CANOCO ©. Dr. Arnoud Boom is thanked for his support

during my visit to Leicester University, England (as well, thanks for introducing me more

into Asian Cinema). Grace Jeon is thanked for her support in developing Loss-on-ignition

measurements for my core samples in the Paleoecology Lab, Department of Geography –

UofT-. The Centre for Global Change Science and their staff, especially Ana Sousa, at

the University of Toronto is greatly thanked for enhancing my Ph.D. experience. I thank

Prof. Jock McAndrews and Charlie Turton for their support during my Ph.D.

I thank everybody at the Department of Geography who supported me during my Ph.D.

years as a student, especially from the main office at Sidney Smith (esp. Marianne

Ishibashi, Marika Maslej, and Jessica Finlayson). I am very grateful to the Physical

Page 6: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

vi

Geography Building (PGB) community who supported and helped me, especially

members of Cowling and Finkelstein’s research laboratories. Members of Chen’s,

Diamond’s and Desloge’s laboratories are greatly thanked for their company and support.

I am greatly grateful to Mircea Pilaf for his support since the first day I arrived to PGB

and for the conversations in Romanian. I thank the “Geography soccer” community

whom I shared many summer, fall, winter, and spring games. Thanked are Maria Johnson

and Family for being my Guatemalan-“Chapina”-Canadian Family.

I am grateful to Claudia Avendaño and Knutt Eissermann for providing help and time in

finding the source vegetation literature for this study. I am grateful to Maria Elena

Hidalgo “mi Ague”, Carlos Avendaño E., Yolanda Mendoza de Avendaño, Gary

Avendaño, and Hector Bol for providing help during field work. I thank my Family in

Guatemala for their spiritual and moral support: Papa, Mama, Clada, Gary, Abue, Hector,

Ti Lili, Dn. Enrique, Kennes ... This thesis is dedicated to my Family, which has

supported me in my entire life in any possible path that I have taken … forever and ever.

Special dedication for Mateo and Belinda, who now have become my triangle of life, joy,

and motivation to become a better being. Mateo:

“No llegó la gota carmín, Llegó en su lugar la noticia de su visita, Certidumbres y rumbos no aleatorios, A esta edad, en este lugar, en esta vida… Semilla liberando indicios de luz, Transformando auras, metamorfosis interna, Milagro de la multiplicación de tu rostro en cada rostro, en el niño de la calle, en el abuelo de la esquina, en el rostro del espejo, Bien leí que en tradiciones ancestrales se entiende como la llegada de un maestro, En silencio quiero aprender de ti… Después de años de ser profecía, la epifanía llego esta mañana: reconocer al prójimo como a mi propio hijo… Traes polvo cósmico celestial, soplas tu aliento en mi oído y me revelas el universo”.

Page 7: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

vii

Table of Contents Abstract ............................................................................................................................... ii 

Acknowledgments...................................................................................................... iv List of Tables ............................................................................................................. ix List of Figures ............................................................................................................. x List of Appendices .................................................................................................... xii 

Chapter 1: Background Information ................................................................................... 1 1.1 Pollen as a Paleoecological Proxy ........................................................................ 1 1.2 Climate Variability Over the Holocene ................................................................ 4 1.3 Reconstructing vegetation and landscapes............................................................ 8 1.4 Reconstructing Cultural Landscapes................................................................... 10 1.5 Thesis Objectives and Research Questions......................................................... 13 1.6 Geomorphological and Vegetational Setting of Study Region........................... 15 1.7 Cultural History of Study Region ....................................................................... 20 

Chapter 2 Vegetation Distribution along the Las Verapaces region in North Central Guatemala.... 27 

2.1 Introduction......................................................................................................... 27 2.2 Methods............................................................................................................... 30 2.3 Results................................................................................................................. 35 2.4 Discussion ........................................................................................................... 47 2.5. Chapter summary ............................................................................................... 54 

Chapter 3 Modern pollen rain in the north-central Guatemalan lowlands and highlands................. 56 

3.1 Introduction......................................................................................................... 56 3.2 Methods............................................................................................................... 59 3.3 Results................................................................................................................. 64 3.4 Discussion ........................................................................................................... 85 3.5. Chapter summary ............................................................................................... 95 

Chapter 4 Late-Holocene History of a Highland Floodplain in Las Verapaces, Guatemala............. 98 

4.1 Introduction......................................................................................................... 98 4.2 Methods............................................................................................................... 99 4.3 Results............................................................................................................... 103 4.4 Discussion ......................................................................................................... 115 

Page 8: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

viii

4.5. Chapter Summary ............................................................................................ 132 Chapter 5 The Lachuá Lowlands Rain Forest in Guatemala: 2,000 yrs of forested landscape? ..... 134 

5.1 Introduction....................................................................................................... 134 5.2 Methods............................................................................................................. 136 5.3 Results............................................................................................................... 139 5.4 Discussion ......................................................................................................... 149 5.5 Chapter summary .............................................................................................. 161 

Chapter 6 Conclusions..................................................................................................................... 163 

6.1 What are the factors that explain vegetation distribution along the Las Verapaces environmental gradient and what taxa can be used as "indicator species"? ........... 164 6.2 Can paleoecological calibrations for fossil pollen be constructed from a comparison of modern pollen rain from surface sediments and bryophyte polsters?................................................................................................................................. 166 6.3 What are the major vegetation changes recorded in the highland core from the Las Verapaces region? ............................................................................................ 168 6.4 What are the major vegetation changes recorded in the lowland core from the Las Verapaces region? ............................................................................................ 170 6.5 What is the role of natural variability and cultural factors related to the Maya Civilization in the evolution of landscapes in the Las Verapaces Region? ............ 172 

References....................................................................................................................... 174 Appendices...................................................................................................................... 196 

Page 9: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

ix

List of Tables 

Table 2.1. Sites included in the Las Verapaces Gradient, providing a total of 23 sampling units (SU) from 9 sites spanning an elevation gradient of 170 to 2532 m asl. Table 2.2. Indicator plant taxa for the three vegetation belts along the Las Verapaces Gradient, selected from DCA axis scores for species (see text for details). Table 2.3. Generalist plant taxa for the Las Verapaces Gradient, as determined by DCA axis scores for species (see text for details). Table 2.4. Disjunctive plant taxa distributed in Lowland Rain Forest and Montane Cloud Forest in the Las Verapaces Gradient generated from DCA axis scores for species (see text for details). Table 3.1. Pollen types and their % range for bryophyte polsters and surface sediments. Information about vegetation belt, plant habit, and pollen dispersal syndrome is provided. Table 3.2. Lachuá bryophyte polsters and surface sediments samples. Table 3.3. Purulhá bryophyte polsters and surface sediments samples. Table 3.4. Factor Analysis scores for pollen types with highest amount of variance. Table 4.1. P-4 core stratigraphic sequence. Table 4.2. Radiocarbon dates, calibrated age ranges from dated sediment from P-4 core of the Cahabón Floodplain, Purulhá Table 5.1. Radiocarbon dates, calibrated age ranges from dated sediment from P-4 core of the Cahabón Floodplain, Purulhá.

Page 10: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

x

List of Figures 

Figure 1.1. Temperature changes in the late Pleistocene and the Holocene inferred from oxygen-18 measurements in Greenland ice core (taken from Dergachev and van Geel, 2004). Figure 1.2. Location of Guatemala in Central America. Numbers indicate location of meteorological stations. Figure 1.3. Topographic map of the Las Verapaces Region in Guatemala. Figure 1.4. Topographic map of study sites in the Las Verapaces Region in Guatemala. Map I= Lachuá lowlands, Map II= Purulhá highlands. Figure 2.1. Detrended Correspondence Analysis diagram of Las Verapaces sites along the first two DCA axes. Figure 2.2. Linear regression of Detrended Correspondence Analysis (DCA) Axis 1 scores (raw scores) of indicator plant taxa (A) and sites (B) against elevation values per site of the Las Verapaces Gradient. Figure 2.3. Linear regression curves for temperature (°C) from meteorological stations from Central and Northern Guatemala. Figure 2.4. Detrended Correspondence Analysis diagram for the Las Verapaces Gradient sites and climatic variables. Figure 2.5. A) Indicator (in one vegetation belt) and B) generalist (across two vegetation belts) taxa separated according to their biogeographic origin along the Las Verapaces gradient vegetation belts Figure 3.1. Location of Guatemala in Central America. Las Verapaces Region is enclosed in rectangle. Figure 3.2. Pollen diagram from Lachuá and Purulhá based on bryophyte polsters and surface sediment samples. Figure 3.3. Lachuá pollen diagram based on bryophyte polsters and surface sediment samples. Figure 3.4. Lachuá DCA Q-mode diagrams of arboreal pollen data with Pinus removal. Figure 3.5. Purulhá pollen diagram based on bryophyte polster and surface sediment samples.

Page 11: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

xi

Figure 3.6. Purulhá DCA Q-mode diagrams based on arboreal pollen and non-arboreal pollen data. Figure 3.7. Las Verapaces DCA Q-mode diagram AP shared data. Figure 4.1. P-4 core paleoecological diagram taken from Cahabón River Flooplain. Figure 4.2. Graph showing depth (cm) vs. calendar age (cal yrs BP) of sediments from core P-4 taken from the Cahabón River floodplain. Figure 4.3. Principal Component Analysis (PCA) of sampled levels from core P-4. Figure 4.4. Pollen percentage diagram of P-4 core from the Cahabón River floodplain. Figure 4.5. Location of the headwaters of the Cahabón River and the floodplain. Figure 4.6. Principal Component Analysis (PCA) of modern pollen rain samples from Purulhá highlands and fossil samples from core P-4. Figure 4.7. Principal Component Analysis (PCA) of modern pollen rain samples from Purulhá highlands and sampled levels from core P-4. Figure 4.8. Cahabón River and its floodplain. Series of 3 arrows indicate river flow direction. N= North, masl=meters above sea level. Taken by J.C. Berrio © 2006. Figure 5.1. L-3 core paleoecological diagram taken from a wetland next to Lake Lachuá. Figure 5.2. Principal Component Analysis (PCA) of sampled levels from core L-3. Figure 5.3. Pollen percentage diagram of L-3 core from a wetland next to Lake Lachuá. Figure 5.4. Location of the ancient Mayan city of Salinas de los Nueve Cerros on the banks of the Chixoy River, Alta Verapaz, Guatemala. Figure 5.5. Principal Component Analysis (PCA) of modern pollen rain samples from Lachuá lowlands and fossil samples from core L-3.

Page 12: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

xii

List of Appendices 

Appendix 2.1. Indicator, generalist, and disjunctive plant checklist. Appendix 3.1. Pollen types found in modern pollen calibrations and fossil pollen spectra from cores P-4 and L-3. Associated plant and uses by ancient Mayan populations are shown. Appendix 4.1. Pollen counts (raw) from P-4 core. Appendix 5.1. Pollen counts (raw) from L-3 core.

Page 13: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

1

Chapter 1: Background Information

1.1 Pollen as a Paleoecological Proxy  

One of the main objectives of paleoecological research is to reconstruct environmental

changes occuring at different scales of resolution, from global to local scales (Bennington

et al., 2009; Birks, 2005; Hunter, 1998; Willis and Birks, 2006). Many Holocene

examples can be cited that demonstrate how natural and cultural factors influence the

evolution of landscapes and regions (Berrio et al., 2001; Lorimer, 2001; Muñoz and

Gajewski, 2010; Ye et al., 2010). The likely reason for the emphasis on separating natural

from cultural factors relates to our understanding of whether current global

environmental trends are due to natural variability, cultural factors, or some combination

(Harris, 2003; Cao et al., 2010).

Vegetation is a fundamental component of ecosystems, landscapes and regions, and has

been used widely as a paleoecological indicator (Markgraf et al., 2009; Valsecchi et al.,

2010; Cheng, 2011). Vegetation was chosen as a proxy for landscape evolution because

of its intimate relationship with climatic and topographic variability (Clark, 2007;

Davidar et al., 2005; Simona et al., 2009). Vegetation reflects the environmental and/or

cultural regimes that control landscapes and regions at different spatio-temporal ranges.

The chosen proxy for vegetation reconstruction is pollen because of its taxonomic

specificity and because it reflects processes related to vegetation dynamics (i.e.

pollination), in addition to the fact that it has been studied thoroughly and used often for

Page 14: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

2

different applications in biogeography, climate change, biome reconstructions, and

archaeology (Berrio et al., 2001; Birks and Birks, 2003.; Graham, 2006; Marchant et al.,

2009). The relationship between vegetation and pollen found in depository records, either

superficial or sedimentary, is not 1:1 because of the multiple factors that are involved in

pollen release, transportation, deposition and preservation (Brown et al., 2007; Bunting et

al., 2004; Campbell, 1999; Fægri and Iversen, 1989). It is necessary to understand the

relationship between vegetation and pollen collected from depositories, in order to

understand pollen representation at modern or past times for a determined landscape and

region. Therefore the concept of uniformitarism underlies palynological research: it is

assumed that the chosen proxy has had a response in the past similar to its responses to

present-day natural and cultural changes (Bradley, 1999).

Vegetation has been closely linked to human history and activities (e.g. agriculture and

forestry) because vegetation provides a resource source for multiple needs: timber,

fuelwood, medicine, food, and resins (Fuller et al., 2010; Innes et al., 2009; Rokaya et al.,

2010; Weiser and Lepofsky, 2009). Complementary use of archaeological methods helps

to broaden our ability to understand human impact on landscapes (Li et al., 2010; McKey

et al., 2010; Weiss and Brunner, 2010). Pollen grains (i.e. as micro-botanical remains or

microfossils) have been widely used in paleoecology and have become relevant proxies

to reveal natural and cultural factors in landscape evolution (Lozano-García et al., 2010;

Scharf, 2010). Changes in pollen composition, pollen abundance, and information related

to the presence or absence of specific taxa provide the foundation for paleoecological

reconstructions of past environmental change.

Page 15: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

3

Modern pollen spectra have been studied from bryophyte polsters (e.g. mosses and

liverworts) and are collected mostly from the interior of non-disturbed forests

(Domínguez-Vázquez et al., 2004). These studies are important in understanding the

conditions under which the pollen is deposited; these studies are also necessary for

comparisons (i.e. in presence, absence and abundance) between observed pollen signal

and surrounding plant taxa. This modern-day calibration process is necessary for the best

possible interpretation of the fossil pollen record.

Topography affects energy distribution in landscapes, such as water and wind flows

where pollen transportation occurs (Schueler and Schluenzen, 2006; Vogler et al., 2009).

The role of topography in affecting pollen transport, however, is not entirely understood

(Higgins et al., 2003). These processes have mostly studied with respect to maize pollen

in terms of cross-pollination in agricultural fields (Klein et al., 2003). On its own,

elevation above sea level has an influence on pollen dispersal and deposition because of

orographic effects related to patterns of wind circulation (Fægri and Iversen, 1989).

Regionally-dispersed pollen is sensitive to atmospheric conditions, for example, because

surface convection (i.e. air turbulence from heating) can raise pollen above the canopy-

level, causing long-distance, horizontal transfer until the air parcel eventually cools and

descends (Murray et al., 2007), or it encounters a “disturbance” in flow such as a lake

basin causing pollen to fall out of the atmosphere (Sugita 1993). Provenance of pollen

may also be a source of bias in interpreting paleoecological signals because where

sediment is deposited (i.e. lakes, rivers, oceans) and how it is transported (i.e. by wind,

water, terrestrial and aquatic animals) is important (Traverse, 1994; Nielsen, 2005). Once

Page 16: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

4

pollen grains land on a surface, they will respond to the physical, chemical, and

biological processes occurring on the surface, that in turn determine sedimentation and

preservation of pollen samples. The understanding of taphonomy and pollen-environment

relationship is determinant in pollen analysis, since an important assumption is that the

pollen assemblage recorded from a sediment sample is the same as the originally

deposited (Twiddle and Bunting, 2010).

The wide ranging applications of pollen analysis in paleoecology have increased the

research scope to conservation biology and biogeography. For example, conservation

efforts have been directed where plant communities in riparian environments have been

identified as relicts (i.e. early Holocene), after studying pollen spectra in sedimentary

records found in floodplains (Southgate, 2010). At the geological scale, pollen records

have been the basis to explain the evolution of biomes coupled to tectonic processes (e.g.

orogeny) based on pollen spectra collected in lakes sediments in the Andes

(Hooghiemstra et al., 2006).

1.2 Climate Variability Over the Holocene 

Reconstucting past climate changes is important for explaining roles of external and

internal forcings on the climate system and for predicting future trends. External forcings

on the climate system include changes in orbital parameters of the Earth, and solar

variability; internal forcings, by contrast, are related to processes that occur within the

Earth system (e.g. volcanic activity) (Beniston, 2005). The Milankovitch Cycles are

important variations in Earth’s orbit, known mostly for their role in promoting the

Page 17: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

5

Pleistocene Ice Ages (Lisiecki, 2010). The parameter of eccentricity (the measure of the

shape of the Earth's orbit around the sun) varies on a timescale of ~ 100,000 yrs and

contributes to glacial-interglacial cycling (Berger, 1989). The other two Milankovitch

parameters are: obliquity (measure of the Earth’s rotation tilt from 22 to 24.8° every 41

ky) that is responsible for the definition of tropical and circum-polar latitudes, and

precession of the equinoxes (which cycles on a scale of 19 to 26 ky), which is related to

solar insolation variability as a function of the Earth-Sun distance at the moment of the

vernal equinox. The interaction of the three Milankovitch parameters is consistent with

recorded climatic variability at the multi-millenial timescale, by producing a complex

pattern of solar radiation reception on Earth’s atmosphere (Mendoza, 2005). Large-scale

biotic processes such as migration and colonization have been affected by these cycles

and modern day biogeography has been greatly influenced by the glacial – interglacial

cycling of the Quaternary (Erwin, 2009; Kerhoulas and Arbogast, 2010).

At a much smaller time scale, solar variability as evidenced through the sunspot cycles of

11, 22 and 240 years, result in changes in the amount of short wave radiation reaching the

Earth (Rapp, 2010). Decreased occurrence of sunspots is believed to be one of the factors

explaining reductions of global temperature (see Little Ice Age below) (Haase-Schramm

et al., 2005).

Internal forcing of climate is related to volcanic activity (i.e. tectonics), ocean circulation,

and critical changes in the biosphere (marine and terrestrial) and cryosphere (Beniston,

2005). Volcanic activity cools the climate because particulate matter emitted from the

Page 18: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

6

eruption changes the Earth's albedo, increasing solar energy reflectance. Ocean

circulation patterns affect the climate, such as the associated drop in global temperatures

due to a weakening of the thermohaline circulation during the Younger Dryas (ca. 11,000

yrs BP) and Little Ice Age (ca. 300 yrs BP) (Bradley and England, 2008; Helama et al.,

2009). Variability in other circulations could have more regional effects at decadal time

scales such as El Niño Southern Oscillation (ENSO) and the North Atlantic Oscillation

(NAO) (Seager et al., 2010). The former is associated to the contraction and expansion of

warm waters in the west Pacific, and the latter is believed to account for ca. 50% of

variability in sea level pressure on both sides of the Atlantic Ocean. The internal forcing

factors have in common that they operate at a sub-millenial time scale.

The explanation of the Holocene climatic variability requires understanding the coupled

effects of external and internal forcings. During the last 10,000-12,000 years the

Holocene stands as an epoch of warmth and steady climate, characterized by centennial

and millennial-scale alternating of cold and warm periods, superimposed over a long-

term trend of first warming and then cooling (Bjune et al., 2004). The onset of the

Holocene climate has been shaped by the cyclical transition from a glacial to an

interglacial where the maximum insolation was experienced (~10 ky BP) (Solanki et al.,

2004) (Figure 1.1.). Thereafter four warming maxima, alternated by cold stages, have

been deducted from paleoecological data during the intervals: 6700-5700, 4500-3200,

2300-1600, and during 1150-900 yrs BP (the Medieval Climatic Optimum) (Dergachev

and van Geel, 2004). Cold Heinrich events (stadials) and Dansgaard-Oeschger warm

stages (interstadials) are important factors that are believed to play a role determining

Page 19: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

7

climatic shifts at the millennial scale (Merkel et al., 2010). Millennial to centennial

natural variability modifies macro-regional climatic regimes and therefore more localized

dynamics such as forest humidity and temperature (Jouzel et al., 2007; Popescu et al.,

2010).

Although global climatic synchronicities have been recognized, regional variations play a

critical role in understanding biogeographical patterns found at smaller spatio-temporal

scales (Viau and Gajewski, 2009). Variability in the location of the Intertropical

Convergence Zone (Chiang and Bitz, 2005; Holbourn et al., 2010) and cyclicity of ENSO

(Merkel et al., 2010) are of major importance to understanding climatic variability at

more regional scales in Mesoamerica. Evidence of climatic variability in the Yucatán

Peninsula, is derived from the 206-year period oscillations of oxygen isotopes and

gypsum precipitation from Lake Chicancanab, and possibly related to variation in solar

radiation (Hodell et al., 2001). Similar paleoclimatic patterns have been gathered from

Figure 1.1 Temperature changes along the late Pleistocene and the Holocene inferred from oxygen-18 measurements in Greenland ice core (taken from Dergachev and van Geel, 2004).

Page 20: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

8

the Circum-Caribbean, Lake Valencia and the Cariaco Basin in Venezuela, and when

combined with the Chicancanab data, aligns with critical processes along major cultural

periods (Alley et al., 2003; Hodell et al., 1991; Peterson et al., 1991). Arid events have

been associated with cyclic events and include observed droughts between 150 and 250

AD (Pre-Classic abandonment), 750-1050 AD (Terminal Classic Collapse) and 1450 AD

(Post-Classic) (Hodell et al., 2007).

The climate system is currently understood as the product of the coupled interactions

between the atmosphere, hydrosphere, lithosphere, cryosphere, and biosphere.

Information provided by paleoclimatic studies provide scientific basis for hypothesis

testing of climatic variability in determined locations under different temporal scales of

resolution.

 

1.3 Reconstructing vegetation and landscapes  

A large number of vegetation reconstructions based on pollen have been conducted

around the world, spanning time periods from hundred to millions of years ago, and have

provided important information to determine the roles of natural factors in landscape

evolution. Based on changes in pollen composition, it has been possible to identify a high

correlation between tectonic processes of the Andean orogeny of the last 3 million years

(Mya) with altitudinal changes in North Andean biomes (Hooghiemstra and Van der

Hammen, 2004; Torres et al., 2005). In coastal environments, sea level changes at the

multi-millenial scale have been analyzed based on regressive and transgressive phases

Page 21: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

9

reconstructed from sedimentary sequences, and have been used in conjunction with

pollen information to show an inland-to-coast migration of vegetation (Torrescano and

Islebe, 2006; Gabriel et al., 2009). Milankovitch cycles affect the retreat and advance of

glacial ice caps, events that can be recognized in pollen diagrams showing latitudinal tree

line oscillations (Kramer et al., 2010). Other periods of natural climatic variability such

as the Younger Dryas stadial (cold event) (Kokorowski et al., 2008) and solar cycles are

evident in pollen diagrams (Morner, 2010). Pollen from the Arctic specialist Dryas

octopetala is used as an indicator of the Younger Dryas because of the increase in

distribution and abundance of D. octopetala at this stadial (Joosten, 1995).

In places such as the Mexican Central Highlands and the Lacandon rain forest in Chiapas,

evidence of the Maya Terminal Classic (800-900 century AD) drought event has been

interpreted based on the increase of Pinus pollen (Almeida et al., 2005; Domínguez-

Vázquez and Islebe, 2008). In contrast, reconstructions from neighboring regions such as

the Mexican Sierra Madre Oriental (East-Central Mexico) (Conserva and Byrne, 2002)

and Sierra de Los Tuxtlas (Lozano-García et al., 2010) show no evidence of drought and

actually indicate slightly moister conditions. Geographical variability in precipitation

may be because of orographic effects in topographically complex regions, which creates

climatic envelopes at the regional scale. For example, Wendt (1989) proposed the

existence of a wet belt across the Gulf of Mexico, Southeast of Mexico, Central

Guatemala, and the Izabal province (Caribbean Guatemalan Coast), which possibly

allowed the permanence of hypothesized tropical rain forests pleistocenic refuges.

Page 22: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

10

Data from Los Tuxtlas show evidence of the Little Ice Age (LIA) in the Gulf of Mexico;

multi-proxy records indicate wetter conditions around 1500 to 1700 AD (e.g. increased

lake levels and increased accumulation rates of pollen of lowland or highland forest taxa)

(Lozano-García et al., 2007; Lozano-García et al., 2010). In contrast, reconstructions

based on oxygen isotopes and titanium content from the Yucatán peninsula (Aguada

X’caamal) and the Cariaco basin (respectively) show lower precipitation between 1500

and 1800 AD (Hodell et al., 2005). Climate proxies and the presence of Zea pollen from

Lake Tzib at Quintana Roo, Mexico (Carrillo-Bastos et al., 2010) likely indicate higher

precipitation around 1200 AD contrary to what would be expected during the time of the

Medieval Warm Period (MWP) (just before the LIA).

1.4 Reconstructing Cultural Landscapes  

Pollen can also be used to reconstruct anthropogenic impacts on landscapes during

different cultural periods, for example, during the early Holocene phase of hunters-and-

gatherers (Kunes et al., 2008). Cultural impacts on the environment are of greater interest

for more recent times, including the transition from nomadic human populations to fully

sedentary communities (Rowley-Conwy, 2009). It is when human groups started to

remain in one area for longer periods of time that we can see a clear anthropogenic signal

in the paleo-record, reflecting the evolution of agriculture as an important modifier of

landscapes.

In different culturally important regions around the globe (the Near East, Ganges Delta,

Yellow River watershed), the origins and development of agriculture have been

Page 23: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

11

reconstructed based on the first appearance of cereal pollen cultigens. In the case of

Mesoamerica, corn pollen (Zea mays L.) is traced (Zizumbo-Villarreal and Colunga-

García, 2010) whereas in China, the initial presence of rice pollen (Oryza sativa) is used

to signify the beginning of agriculture (Fuller et al., 2009). The reconstruction of

landscapes histories based on architectural and ceramic remains (i.e. archaeological

methods) is complemented with the use of pollen because it can tell a more complete

story about an area, including information on landscape management, levels of

disturbance, and conservation efforts (Bettis III et al., 2008; Dambrine et al., 2007;

Delhon et al., 2009; Mercuri, 2008).

The magnitude to which anthropogenic activities influence landscapes is a topic of much

discussion between researchers (Horrocks et al., 2007; Williams et al., 2010; Yu et al.,

2010; Zhao et al., 2010). Numerous scholarly theories have been derived depending on

the type of evidence collected (i.e. paleoecological versus archaeological) and the cultural

context in which that evidence is found. On the one hand, ancient cultures have been

considered responsible for major modifications to landscapes; involving activities that

generally bring upon detrimental societal consequences as a result of natural factors such

as soil erosion and resource depletion (Diamond, 2009; McWethy et al., 2009). It has

been suggested that anthropogenic activities (particularly changes in land-use) can alter

regional climate, such as precipitation (Shaw, 2003; Gill et al., 2007) and therefore could

play an interactive role in prolonging periods of drought and/or deepening the magnitude

of water stress. From this perspective, anthropogenic activities are considered the critical

trigger in the collapse of past societies (Diamond, 2005).

Page 24: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

12

In contrast, past societies can be viewed from the perspective of practicing sustainability

of resources in their everyday activities, such as the planting of trees to prevent large-

scale erosion of highlands (Smith and Demarest, 2001; Aimers, 2007). From this second

point of view, the “collapse” of an ancient civilization has foundations in multi-factorial

processes, both anthropogenic and natural (e.g. biotic, abiotic) (Demarest et al., 2004;

Demarest, 2009).

Regardless of how human activity is viewed within ecosystem dynamics, evidence shows

an increasing effect of changes in greenhouse gas concentrations in the atmosphere, since

the onset of agricultural activities, the introduction of large-scale herding of grazers, and

most recently due to the burning of fossil fuels (Olofsson and Hickler, 2008; Brook,

2009). The "Anthropocene", a controversial naming of the latter period of the Holocene,

has been defined by the period of over-arching effects of humans on climatic, hydrologic

and edaphic cycles (Ruddiman, 2003; Crutzen, 2006).

Page 25: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

13

1.5 Thesis Objectives and Research Questions 

1.5.1. Rationale.

The role of natural and cultural factors in the evolution of landscapes within the Las

Verapaces region of north-central Guatemala is the focus of this thesis. Natural and

cultural factors can intermingle or act independently at different spatio-temporal scales

(Clark, 2007; Díaz and Stahle, 2007; Partel et al., 2007; Sarmiento et al., 2008;

Wainwright, 2008). The separation of past cultural and natural processes by using

paleoecological methodology is needed to help provide a solid scientific basis to assess

modern-day impacts of human activities at the global, regional and landscape scales. This

thesis is developed in the Lachuá lowlands and the Purulhá highlands of the Las

Verapaces region, an important location in the Mesoamerican context due to its high

biological and cultural diversity, which nevertheless lacks exploration in paleoecological

terms.

1.5.2. Approach.

My approach involves paleovegetation reconstructions of the Lachuá lowlands and the

Purulhá highlands in the Las Verapaces Region from the Preclassic to modern-day times,

covering the past two millennia. To develop paleo-vegetation reconstructions for the Las

Verapaces Region, it was necessary to first determine the taxonomic composition of

vegetation communities and the altitudinal distribution of vegetation types, including

explanations for their geographical variation (Chapter 2). Since the relationship between

the abundance of pollen grains and the abundance of corresponding vegetation is not 1:1,

Page 26: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

14

it was necessary to develop the first calibration study of the region by comparing pollen

sources such as lake sediments and bryophyte polsters and analyzing the modern pollen

rain (Chapter 3). Paleoecological reconstructions were developed based on fossil pollen

spectra collected from a core (P-4) from the Cahabón River floodplain at the Purulhá

highlands spanning the last ~2390 years (Chapter 4) and a wetland core (L-3), taken

adjacent to Lachuá Lake, within the Lachuá lowlands (Chapter 5) spanning the last

~2000 years.

Research Questions. The main research questions addressed in this thesis include:

a) What are the factors that explain vegetation distribution along the Las Verapaces

environmental gradient and what taxa can be used as "indicators"?

b) Can paleoecological calibrations for fossil pollen be constructed from a

comparison of modern pollen rain from surface sediments and bryophyte polsters?

c) What are the major vegetation changes recorded in the two (lowland, highland)

cores from the Las Verapaces region?

d) What is the role of natural variability and cultural factors related to the Maya

Civilization in the evolution of landscapes in the Las Verapaces Region?

Page 27: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

15

1.6 Geomorphological and Vegetational Setting of Study Region 

The Las Verapaces region is located in north central Guatemala, encompassing sharp

environmental gradients from the Lachuá lowlands (~170 masl) to the Purulhá highlands

(~2500 masl) (Figure 1.2). In addition to being characterized by environmental gradients,

I also selected the region because of the absence of paleoecological research (Islebe and

Leyden, 2006) despite its importance in both natural and historical cultural diversity. Las

Verapaces is distributed across two Guatemalan provinces: Alta Verapaz and Baja

Verapaz (Figure 1.2 and 1.3). The geological structure of the area is primarily karstic

terrain of Cretaceous and Tertiary origin (Alta Verapaz), with metamorphic regions

dating from the Lower Paleozoic (Baja Verapaz and Alta Verapaz) (Ortega-Gutiérrez et

al., 2007).

1.6.1 Lachuá Lowlands

The Lachuá lowlands are located in a transitional zone between the Petén Lowlands and

the Cordilleran central highlands (Weyl, 1980) and contain one of the last remnants of

Lowland Rain Forest remaining in Guatemala (Figure 1.3) (for vegetation belt

description see results Chapter 2). The site has a protected area, the Lachuá Lake

National Park, which covers approximately 14,500 ha in addition to a surrounding buffer

zone of approximately 28,000 ha (Monzón, 1999). An inventory of Lachuá’s forest

species (as well as other vegetation types) was undertaken within the past 10 years

(García, 2001; Ávila, 2004; Cajas, 2009; Castañeda, 1997), and more recently, a modern

pollen reference collection of the thirty most abundant plant species has been collated

(Barrientos, 2006). There is a Lowland Rain Forest remnant (~300 ha) northeast of

Page 28: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

16

Lachuá Lake National Park located in the top of a hill 285 masl in elevation with a series

of small ponds known as Tortugas (Tun personal communication, 2006). The remnant is

known as Salinas de los Nueve Cerros Regional Park, where an archaeological site of the

same name is located.

Geomorphologically, the area contains undulated karstic hills and varied landforms

ranging from low- to mid-elevations (170-600 masl) (Avendaño et al. 2007). The Lachuá

Lake is found at the Lachuá Lake National Park; a circular depression (400 hectares) with

a depth of 200 m, draining into the lower sedimentary basin of the Chixoy River

(Granados, 2001). Moisture-laden winds from the northwest and east originate from

within the Caribbean Sea, creating a mean annual precipitation of approximately 2000-

2499 mm. The rainy season occurs between May and October, with mean annual

temperatures between 25.5–28°C (Monzón, 1999).

1.6.2 Purulhá highlands

The Purulhá highlands cover the Cahabón River headwaters, and the Polochic and

Chixoy upper basins, ranging in elevation from 1560-2300 masl (Figure 1.3). Purulhá

contains a main remnant of cloud forest (1044 ha) that is protected under the jurisdiction

of “Biotopo Universitario para la Conservación del Quetzal” (BUCQ) (CONAP, 2000).

This site is underlain by the metamorphic and karstic system of the Sierra Chuacús

mountain range (Weyl, 1980). Moisture-laden Caribbean winds from the east, northeast,

and northwest result in mean annual precipitation around 2092 mm and mean annual

Page 29: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

17

Figure 1.2. Location of Guatemala in Central America. Circle encloses location of the Las Verapaces Region. Numbers indicate locations of meteorological stations. 1= Flores, 2= Puerto Barrios, 3= Las Vegas, 4= Panzos, 5= Cahabón, 6=Papalhá, 7= Cobán, 8= Suiza Continental.

1

2346

57

8

MéxicoBelize

Honduras

NicaraguaEl Salvador

Costa Rica

Panama

Guatemala

Pacific Ocean

Caribbean Sea

Gulf of Mexico

1

2346

57

8

1

2346

57

8

MéxicoBelize

Honduras

NicaraguaEl Salvador

Costa Rica

Panama

Guatemala

Pacific Ocean

Caribbean Sea

Gulf of Mexico

Page 30: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

18

Figure 1.3. Topographic map of the Las Verapaces Region in Guatemala. Encircled numbers indicate study sites. 1= Lachuá lowlands, 2=Sierra Chinajá, 3= Rio Tinajas, 4= Chelemhá, 5=Tucurú, 6= Tamahú, 7=Purulhá (BUCQ)-, 8= Tactic, 9= Santa Cruz Verapaz. Locations #6 to #9 are part of the Purulhá highlands. Watershed names are indicated in italics.

1

2

3

4

56

7

89

México

La Pasión

Chixoy

Cahabón

Polochic

1800 m

1000 m

200 m

Chinaja

Watershed boundary

1

2

3

4

56

7

89

México

1

2

3

4

56

7

89

1

2

3

4

56

7

89

México

La Pasión

Chixoy

Cahabón

Polochic

1800 m

1000 m

200 m

Chinaja

Watershed boundary

1800 m

1000 m

200 m

Chinaja

Watershed boundary

Page 31: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

19

temperatures between 13.9–20.4°C (García, 1998). The rainy and dry seasons occur

between June and September and January and April, respectively. The Cahabón River

headwaters are located in the municipality of Purulhá town, province of Baja Verapaz, at

an elevation of approximately 1570 masl. The Cahabón River floodplain is characterized

by the presence of entisols and inceptisols in the low valley sections, surrounded by

andisols and ultisols in the surrounding mountains (MAGA, 2001). The floodplain is

located close to the upper limit of the Lower Montane Rain Forest (1000-1800 masl),

surrounded by valley slopes covered by Montane Cloud Forest (1800-2500 masl in my

study region) (for vegetation belts description see results Chapter 2). Local inhabitants

from Purulhá town have mentioned of the possible existence in the past of a lake in the

environs of the town (Vázquez C. personal communication 2011).

1.6.3. Geographical setting and study design

Vegetation sampling (Chapter 2) of lowland sites took place in separate watersheds: (1)

the Chixoy watershed which is composed of mainly Cretaceous-Tertiary marine

sediments and Quaternary alluvium, and (2) the Polochic watershed located over a pull-

apart type basin containing Quaternary alluvium (Fourcade et al., 1999). Highland sites

are located in the upland portions of the Cahabón and Polochic watersheds, which are

underlain by Pennsylvanian to Permian eclogitic rocks and gneisses (Ortega-Gutiérrez et

al., 2007). Rio Tinajas vegetation sampling sites are located in a sub-watershed that

drains into the Polochic Watershed (Tot, 2000).

Page 32: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

20

Modern pollen samples for palynological calibration (Chapter 3) and core samples

(Chapter 4 and 5) were collected in two sites located at both ends of the Las Verapaces

elevational gradient (Figure 1.3): (1) Lowland Rain Forests at the Lachuá lowlands in

Alta Verapaz (~ 170 masl), and (2) the Montane Cloud Forest and the transitional

vegetation belt at the lower limit at the Purulhá highlands and its environs in Alta and

Baja Verapaz (~ 1400-2000 masl). The Purulhá highlands in our study region represent

the highest geographical point.

1.7 Cultural History of Study Region 

According to the cultural succession and temporal differentiation for Mesoamerican

civilizations such as Olmec, Maya and Aztec (Chase et al., 2009), standardized periods

have been defined as the following: 1) Pre-Classic (3000 BC-300 AD), 2) Classic (300 -

900 AD), and 3) Post-Classic (900~1500 AD). These periods are delineated based on

critical changes to the political, economic and ceremonial development of Mesoamerican

civilizations. The most studied transition includes the end of the Classic Period of the

Maya Lowlands, known as the Terminal Classic Period (Demarest et al., 2004; Demarest,

2006).

Paleoecological studies in the Guatemalan Northern Petén Lowlands (Figure 1.2) have

reconstructed environmental changes dating back to the Last Glacial before any human

settlement took place in the region (Leyden, 2002), but emphasis has been placed on

Mayan cities that flourished mostly during the Classic Cultural period (300-900 AD)

(Islebe and Leyden, 2006). The heightened interest in this time period occurs mostly

Page 33: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

21

because the majority of Classical cities underwent a regional transformation process at

the time of the Terminal Classic, largely known as Classic Mayan collapse (Aimers,

2007). Conclusions from some authors indicate that environmental anomalies, such as

droughts (Diamond, 2005; Gill et al., 2007), have played a critical role in determining the

fate of human societies, sometimes enhanced by human disturbances, which brought

together social instability and revolts due to natural resource demise. Contrasting research

approaches have concluded that environmental variability could have played more of a

secondary role on the transformation of societies, and that intrinsic societal characteristics

have a more relevant role in societal collapse (Demarest et al., 2004). This latter approach

emphasizes the idea that societies like the Mayan are able to cope with extrinsic

disturbances such as environmental extreme events, even when facing intrinsic

instabilities that requires substantial societal transformations.

Mesoamerican paleoecological research has provided explanations regarding the role of

environmental and societal factors on the shaping of landscapes along both highlands and

lowlands. Based on different fossil proxy evidence found in sedimentary records, some

lowland locations indicate the occurrence of drastic droughts, which are believed to have

had a dramatic impact on the transition between the Classic and Postclassic (900-1000

AD) . On the other hand, at some other locations experiencing possible arid events, there

were relatively few cultural changes or negative anthropogenic environmental impacts

even when human populations were highest. The Classic-Postclassic transition is

delineated mostly as a socio-political and religious transformation, that in some locations

promoted total or temporary abandonment of cities, semi-destruction due to warfare,

Page 34: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

22

while in other locations, cultural flourishment took place (Demarest, 2009). Evidence

indicates that the most dramatic changes to all aspects of the Mayan Culture and the

environment occurred during the Spanish Conquest and Colonization (Elliot et al. 2010).

The Spanish settlers brought new diseases that contributed in part to the Mayan

population demise, and ultimately the introduction of new economic, political,

sociological, and religious systems (Van Buren 2010).

There is an obvious void in the Mesoamerican paleoecological record that must be filled

due to the contextual importance of the Las Verapaces region. The Las Verapaces

lowlands represent an important geographical transition from the Northern Petén region

to the Las Verapaces Highlands, and Southern Maya Area (i.e. Kaminal Juyu, Copán, and

Takalik Abaj) (Rice et al., 1985; Fowler et al., 1989). The lack of paleoecological

information for the Las Verapaces Region places this thesis as critical for providing

information about the landscape evolution of the last two millennia. Natural and cultural

factors have been explored in this thesis to provide a baseline for continuing

paleoecological research in this region as well as in neighboring regions in Mesoamerica.

The Lachuá lowlands are located east of the neighboring Petexbatún cultural region

where important cities were developed along the Pasión and Chixoy rivers banks

(Demarest, 2006). The Petexbatún region had different political elites that established a

succession of Kingdoms, where military control was critical to maintain privileged

economic riverine routes. Cancuen, located approximately 60 km east of Lachuá, was an

important city since the late Pre-Classic until its abandonment during the Late Classic

Page 35: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

23

(Aimers, 2007). Paleoecological and paleoagronomic evidence from the Petexbatún

lowlands indicate that sustainable agriculture and forestry were practiced in succession

(Demarest, 1997). Sustainable management practices likely involved soil conservation to

mitigate environmental deterioration with time (Beach and Dunning, 1995; Beach et al.,

2008; Dunning et al., 1997).

At the Lachuá lowlands, the ancient city of Salinas de los Nueve Cerros was established

as an important salt producing center along the Chixoy river banks (Figure 1.4) (Dillon,

1977; 1990; Garrido, 2009). There is no direct evidence describing landscape

management practices, but it is possible that similar soil ammendment measures observed

in the Petexbatún region were also occuring in Salinas de los Nueve Cerros.

Archaeological studies indicate that the economic importance of the Mayan site of

Salinas de los Nueve Cerros was salt production practiced from the Preclassic to Post-

Classic times. At present, population pressure in the Lachuá Region is beginning to

encroach on the Lachuá Lake National Park and Salinas de los Nueve Cerros; over the

past 50 years, 50% of the forest has been lost to anthropogenic land-use change

(Avendaño et al, 2007). The population generally consists of people from the q’eqchi’

ethnic group who mostly ended up in the region as a result of territorial displacement and

colonization projects following the Civil War (Hurtado, 2008).

In order to better understand the cultural processes that were taking place in the Maya

Lowlands, it is critical to concomitantly address the environmental and cultural history of

the Maya Highlands. The scarcity of lakes on the one hand explains why highlands (i.e.

Page 36: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

24

Figure 1.4. Topographic map of study sites in the Las Verapaces Region in Guatemala. Map I= Lachuá lowlands, Map II= Purulhá highlands. Bryophyte polsters are indicated in letters and surface sediments in numbers. 1= L1, 2= L2, 3= L3, 4= Sa2 (Salinas de los Nueve Cerros archaeological site and natural reserve), 5=J1, 6= T1, 7= P4, 8= P1, A=samples Ca-Ce, B= samples Ra-Re. Both A and B are located at the Lachuá Lake National Park. C= Samples N1-N10 (“Biotopo Universitario para la Conservación del Quetzal”). Highland archaeological sites: VP=Valpraiso, CH= Chican, CX= Cerro Xucaneb, S= Sulin. National parks are represented as dark grey polygons. Rivers are irregular black thick lines. Lachuá Lake is represented as light gray polygon in map I. Chichoj Lake is represented as light gray polygon in map II. Samples 6-8 are located in the Cahabón River Floodplains. Isolines every 50 m in Map I (lowest point 150 masl, highest point 700 masl). Isolines every 100 m in Map II (lowest point 1400 masl, highest point 2300 masl).

Page 37: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

25

Las Verapaces) paleoecological research lags behind its lowlands counterpart. But on the

other hand, this scarcity is related to the main interest of researchers in wanting analyze

paleoenvironmental records related to archaeological findings from major Classic Maya

centers which were distributed mostly in Mesoamerican lowlands (Anselmetti et al.,

2006; Hillesheim et al., 2005; Wahl et al., 2007). This trend has dominated despite of the

importance that the multi-factorial interaction (i.e. political, economical, ceremonial, etc.)

lowlands-highlands had for the development of the Maya Civilization during the last

3000 years (Freidel et al., 1993). Nevertheless, recently there has been an increase in

addressing paleoecological questions related to highlands environments in Mesoamerica

(Almeida et al., 2005).

There is scarce paleoecological information about highlands landscape management

practices, but archaeological investigations indicate that relatively high gradient

environmental (e.g. topographic) boundaries promoted the evolution of relatively small

(regional) and well-bounded cultural systems (Sharer and Sedat, 1987). In the Purulhá

highlands, there are many minor archaeological sites that range from the Pre-Classic to

the Post-Classic, including such sites as Cerro Xucaneb, Chican, Sulin, and Valparaiso

(Figure 1.4) (Arnauld, 1978, 1987; Ichon et al., 1996). In contrast to the lowlands,

expansion and alliances of these highland cultural entities was limited in part to

constrained communication over mountainous landscapes, and not strictly to economic,

social, political and ideological factors (Ichon et al., 1996). Natural trade routes have

been traced between lowland and highland archaeological sites that cross mountain ridges

and valleys, therefore indicating that commerce and cultural interregional exchange were

Page 38: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

26

occurring at this time (Andrews, 1984). It is precisely the connection between disparate

regions that was important for the development of the Maya Civilization (Arnauld, 1997).

It is the exchange of socio-political, cosmological and ceremonial knowledge, in addition

to landscape management practices, that unifies the Mayan cultural region. Little has

been discussed about the Mayan Highlands terminal Classic and the occurrence of city-

center collapse (Demarest, 2009). There is need for further investigation about what

causes some cities to be abandoned while others to be founded and flourished.

Land-use at the Purulhá highlands during the late-19th century was dominated by coffee

plantations, whereas today the area is dominated by a complex mosaic of cattle fields,

agricultural crops (mainly corn), and ornamental species. Population density in this

highland area (primarily comprising achi, poqomchi’, q’eqchi’, and ladino ethnic groups)

is steadily increasing, and has created an ever heightening demand for land for agriculture

and urbanization (CONAP, 2000). Following European conquest (ca. 500 yrs BP) socio-

economic and political pressures led to dramatic changes in (1) land-use patterns (i.e.

introduction of cash crops and plantations), (2) foreign investment, and (3) displacement

of indigenous populations (Van Buren, 2010). More recently, anthropogenic disturbances

associated with civil war, strong military rule, colonization, deforestation and pollution

related to natural resource extraction (i.e. mining) have contributed to the character of the

landscape in the Las Verapaces Region (Hurtado, 2008).

Page 39: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

27

Chapter 2:

Vegetation Distribution along the Las Verapaces region in North Central Guatemala

2.1 Introduction  

Understanding the controls on vegetation distribution in the tropics will improve

predictions of responses to future climate change (Freycon et al., 2010) and help to better

determine factors behind centers of high biological diversity ("biodiversity"). Climate is

usually considered a first-order control on vegetation type and distribution (Tietjen et al.,

2010); however, other factors such as watershed topography (Bertoldi et al., 2010) and

evolutionary history (Vanderpoorten et al., 2010) can also play critical roles in shaping

biogeography. Guatemala currently does not have a formal protocol for describing

vegetation types or belts based on floristic and environmental criteria, however, the

following approaches have been used in the past: (1) qualitative integrations of flora with

physiographic and geomorphologic factors (Villar, 1998), (2) quantitative local

adaptations of Holdridge Life Zones (De La Cruz, 1982), or (3) qualitative adaptations of

classifications from neighboring regions like Mexico (Rzedowski, 2006). More

formalized vegetation identification surveys are needed, particularly in light of the fact

that Guatemala is located in Nuclear Central America and is home to the Mesoamerican

Tropical Forest Hotspot (Harvey et al., 2008). The Mesoamerican hotspot is renowned for

its high vegetation diversity (Knapp and Davidse, 2006), despite being located in an area

influenced by humans for over the past 7,000 years (Chinchilla, 1984). Guatemala’s rich

Page 40: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

28

biological and cultural complexity highlight the necessity for better understanding the

roles of natural and cultural factors in vegetation distribution.

The first research objective is to identify changes in vegetation communities and to

delineate boundaries between vegetation belts along an elevational gradient located in the

Purulhá highlands and the Lachuá lowlands of the Las Verapaces region in north central

Guatemala. The names of the vegetation belts applied for Las Verapaces region were

adapted and integrated from different vegetation regional studies (Breedlove, 1981; de la

Cruz, 1982; Kappelle et al., 1995; Kappelle, 1996; Domínguez-Vázquez et al., 2004)

(Table 2.1): (a) Lowland Rain Forest below 1000 masl, (b) Lower Montane Rain Forest

between 1000 and 1800 masl and (c) Montane Cloud Forest above 1800 masl up to

approximately 2500-3000 masl. Other vegetation belts found in neighbouring regions

include (de la Cruz, 1982; Islebe and Kappelle, 1994; Islebe and Velázquez, 1994; Islebe

et al., 1995) (Table 2.1): (d) Lowland Humid Forest, with less precipitation than the

Lowland Rain Forest, such as in the northern Petén region; (e) Montane Mixed Forest,

where the endemic tree Abies guatemalensis is found; (f) Sub-Alpine Forest, being the

tree line limit in Guatemalan forests; and (g) Páramo (Alpine bunchgrassland), in the

Sierra de los Cuchumatanes and in the Western Volcanic Chain.

In order to achieve the first research objective a meta-data analysis of different local

literature sources has been created, where the distribution of plant taxa within one site or

among different sites in the elevation gradient is included. The existence of plant taxa

with discrete elevational distributions is responsible for the delineation of vegetation

Page 41: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

29

Table 2.1 Description of vegetation belts found in the Las Verapaces region* and neighbouring regions in Guatemala. Vegetation Belt

Elevation range (masl)

Mean annual precipitation (mm)

Associated plant taxa

Lowland Humid Forest

~ 0 to <600

~1100-1700

Alseis yucatanensis, Aspidoderma megalocarpon, Manilkara zapota, Sabal morisiana.

Lowland Rain Forest*

~ 0 to 1000

~2100-4300

Sapium, Terminalia amazonia, Trema, Ulmus.

Lower Montane Rain Forest*

~ 1000 to 1800

~2000-2500

Alchornea, Croton draco, Persea schiediana, Rapanea, Myrica.

Montane Cloud Forest*

~1800 to 2500-3000

~ >4100

Hedyosmum mexicanum, Quercus, Podocarpus oleifolius.

Mixed Montane Forest

~ 2500 to 3000-3100

~2500

Abies guatemalensis, Alnus, Pinus ayacahuite, P. montezumae, Quercus.

Sub-Alpine Forest

~ 3100 to 3800

~1100-1800

Alnus, Buddleja, Juniperus, Pinus hartwegii.

Páramo (Alpine bunchgrassland)

~ >3800

~1275

Cardamine, Poa venosa, Senecio, (Sierra de los Cuchumatanes); Calamagrostris, Luzula, Halencia, Oxylobus, Poa tacanae (Western Volcanic Chain).

Page 42: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

30

belts; alternatively, plant taxa that have more continuous distribution create landscape

continuums (Kessler, 2000; Hemp, 2006).

The second objective is to evaluate the factors responsible for vegetation distribution and

turnover of plant communities along the Las Verapaces region. Three key deriving

factors will be examined: (1) elevation and associated changes in climate (i.e.

environmental lapse rate), (2) landscape position and topography in drainage divides, and

(3) biogeographic origin (i.e. over geological timescales). The findings from this analysis

will also provide a critical baseline from which to conduct palaeoecological research

because we can relate fossil pollen spectra to indicator taxa from modern-day vegetation

belts. Ultimately, by studying the natural (biotic, abiotic) factors influencing vegetation I

can begin to tease apart complex interactions between the natural environment and

anthropogenic processes.

 

2.2 Methods   

2.2.1 Compilation of the vegetation database

For areas with few published records, forest inventory databases and unpublished

academic theses provide a rich source from which to better understand the biotic and

abiotic factors influencing vegetation trends observed across modern-day landscapes

(Kitahara et al., 2009; Veen et al., 2010). Data on vegetation community composition,

plant species identification and abundance were collected from multiple sources

including silvicultural, ecological and landscape research reports (Table 2.2). Five out of

Page 43: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

31

ten of the sources report ecological data using traditional experimental design, including

large sample sizes and multiple replicates. Dissertation research conducted by University

students in Guatemala was invaluable to the collation of the database. These sources

included: (1) four undergraduate theses from Lachuá (Ávila, 2004; Cajas, 2009),

Purulhá (García, 1998), and Chelemhá (López, 2009.), (2) one Master of Science thesis

from Sierra Chinajá (Bonham, 2006), and (3) forestry inventories extracted from

undergraduate theses for Tucurú (Paz, 2001), Tamahú (Alonso, 1999), Santa Cruz

(Palala, 2000), Tactic (Mollinedo, 2002), and Rio Tinajas (Tot, 2000).

Because six studies only presented qualitative data (presence/absence), sources that had

quantitative data (abundances) were transformed to presence/absence to standardize my

database. Taxonomic nomenclature was also standardized when necessary and updated

(Gentry, 1982; Smith et al., 2004). In some cases, standardization required retention of

genus-level information only, correction of spelling, and revision of taxonomic

synonymies. The end result is a matrix showing distributions (presence/absence) of 794

angiosperm taxa across 23 sampling units.

Although I recognize the ecological, biogeographic and economic importance of

gymnosperms, I am not incorporating them in my study because in Guatemala little

information on their distributions is available outside of a plantation/forestry context.

Therefore, my analysis focuses on exclusively angiosperms.

Page 44: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

32

2.2.2 Creation of climate databases

To create a regional climate database for sites along our selected gradient, I used

information from eight meteorological stations located in Central and Northern

Guatemala (Figure 1.1). Data from seven stations at different elevations were collected

directly from INSIVUMEH (Volcanology, Meteorology, and Hydrology National

Institute) in Guatemala City, each having temporal coverage from the years 1990-2005

inclusive. From a longer climate database (42 years; 1961-2003) (Aguilar et al., 2005),

data from Flores (123 masl) was used for my analysis. Of all meteorological variables

available, I selected three temperature variables that best represent both extremes and

average indicators of regional climate. The chosen variables include: (1) maximum

absolute temperature (TXx) defined as the recorded annual maximum value of daily

maximum temperature, (2) minimum absolute temperature (TNn) defined as the recorded

annual minimum value of daily minimum temperature, and (3) mean annual temperature

(MAT). Temperature parameters such as MAT have been used to estimate upper limits of

low-elevation taxa (Latorre et al., 2006), and TNn and TXx are useful to estimate

physiological barriers for plants survival (e.g. drop of temperatures close to overnight

frosting and dessication stress related high temperatures, respectively).

2.2.3 Statistical Analysis

A multivariate analysis was run on a total of 794 angiosperm plant taxa from nine sites

with a total of 23 sampling units (presence/absence data), to determine the degree of

similarity between sites and the relationships between taxonomic assemblages and

climate variables. Through a detrended correspondence analysis (DCA) sites were

Page 45: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

33

arranged in a diagram along ordination axes to indirectly identify possible underlying

environmental gradients (Jongman et al., 1995). The presented DCA diagram presents

axis scores transformed into percentages to help visualize the data variability (McCune

and Mefford, 2006). I created dendrograms through hierarchical cluster analysis (HCA)

(relative Euclidean distance and Unweighted Pair Group Method Algorithm; UPGMA) of

sites with similar taxonomic composition, which were of aid to establish groups of sites

in the DCA diagram (Jongman et al., 1995). Where consistent agglomeration of sites was

observed through ordination axes and cluster analysis, a vegetation belt was delineated as

a correlation of elevation and species composition (Axis scores). The software PC-Ord

was used to conduct all statistical analyses (McCune and Mefford, 2006). Plant taxon that

presented a unique DCA Axis 1 score were chosen as representative of a particular

distribution pattern along the altitudinal gradient, instead of utilizing a group of taxa with

the same Axis 1 score. Species scores are known to represent a particular site or groups

of sites, as DCA Mode-Q analysis indicates that sites are “centroids” for an assemblage

or array of determined species (Jongman et al., 1995). Species and sites scores are known

to be illustrative of each other (Chase et al. 2000).

As mentioned earlier, the names of the vegetation belts and their elevation limits were

established a priori (Table 2.1): (a) Lowland Rain Forest below 1000 masl, (b) Lower

Montane Rain Forest between 1000 and 1800 masl and (c) Montane Cloud Forest

between above 1800 and 2500 masl. Indicator taxa were defined as those found

exclusively inside a vegetation belt (i.e. in a discrete elevation range like 1000-1800

masl) whereas generalist taxa are those with a wide distribution that span across one or

Page 46: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

34

Table 2.2. Sites included in the vegetation database of the Las Verapaces region, providing a total of 23 sampling units (SU) from 10 studies spanning an elevation gradient of 170 to 2532 masl. Where indicated, elevation ranges used to calculate average elevations are shown in parentheses. If researchers pooled sampling units (SU) of a site into one vegetation data set, the average elevation was calculated for the site. When sampling units of one site were not pooled, their elevations and corresponding vegetation data were entered directly into our database. If elevation ranges for sampling units were given, the average elevation was calculated. For data type, Q indicates studies that used abundance as measurement and C indicates studies that used presence/absence as measurement. Source Ávila (2004)

and Cajas (2009)

Bonham (2006)

Paz (2001) Alonso (1999)

Palala (2000)

Mollinedo (2002)

Tot (2000) García (1998)

López (2009)

Sites Lachuá (n=1)

Sierra Chinajá (n=1)

Tucurú (n=3)

Tamahú (n=1)

Santa Cruz (n=1)

Tactic (n=1)

Rio Tinajas (n=6)

Purulhá (n=5)

Chelemhá (n=4)

SU codes (in bold)

Lach Chin

Buena Vista (Bvta) Cumbre de Florida, (Flo) Chelemá (Che)

Tam Scruz

Tac

Tin1 Tin2 Tin3 Tin4 Tin5 Tin6

Pur1 Pur2 Pur3 Pur4 Pur5

Che1 Che2 Che3 Che4

Elevation (m asl)

170 400 (200-600)

1200 1100 1260

1048 1500 1650 200 (0-400) 600 (400-800) 1000 (800-1200) 1400 (1200-1600) 1800 (1600-2000) 2200 (2000-2400)

1800 1900 2000 2100 2200

1900 (1800-2000) 2100 (2000-2200) 2300 (2200-2400) 2466 (2400-2532)

Data Q C C C C C C Q Q Watershed (see Fig. 1.3)

Chixoy Chixoy Polochic Polochic Cahabón Cahabón Tinajas/ Polochic

Cahabón /

Polochic

Cahabón / Polochic

Page 47: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

35

two neighboring vegetation belts (i.e. from 400 to 1800 masl). I defined disjunctive taxa

as those found at two discrete vegetation belts but not in three (i.e. 400 and 1800 masl).

Disjunctive taxa were considered when they were distributed in two non-neighboring

elevation belts. I created my plant checklist based on taxa from these three different

categories (indicator, generalist, and disjunctive). DCA Axis 1 scores were used as

representative of vegetation composition at each of the sites and Axis 1 scores were

regressed against elevation.

Indicator, generalist and disjunctive plant taxa were allocated to one of Gentry’s (1982)

four paleogeographic categories: (1) Laurasian, (2) Amazonian-centered, (3) Andean-

centered, and (4) Miscellaneous. A chi-square contingency table test was run to analyze

the relationship between these categories and their corresponding vegetation belts.

Equations were constructed to describe the relationship between elevation and

temperature (temporal average for each meteorological station) to determine the lapse

rate. To predict the value of the chosen parameters for our study sites according to their

elevation, an interpolation was performed for sites located between 2 masl (Puerto

Barrios) and 2100 masl (Suiza Continental) in elevation, and an extrapolation was

performed for sites with elevations higher than Suiza Continental.

2.3 Results 

2.3.1 Ordination and grouping of sites and plant taxa

A linear regression of the DCA Axis 1 scores of sites and their elevation showed a

significant correlation (r2=0.53, p<0.01), and sites were ordered from lowlands to

Page 48: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

36

highlands (Figure 2.1 and 2.2). Based on the DCA diagram I identified the three

expected vegetation belts according to their elevation and related Axis 1 site scores

(Figure 2.2): (a) Lowland Rain Forest below 1000 masl, (b) Lower Montane Rain Forest

between 1000 and 1800 masl and (c) Montane Cloud Forest above 1800 masl. Axis 1

scores of indicator taxa combined against their average elevation, showed a significant

correlation to elevation (r2=0.80, p<0.0001; Figure 2.2 A).

2.3.2 Climate-elevation-species relationships

Although there was considerable variation in climatic variables between and within

meteorological stations, I found a strong linear relationships between temperature and

elevation (r2= 0.65–0.87) (Figure 2.3). Based on these estimations, the environmental

lapse rate of temperature is approximately 0.5°C/100 masl, close to the expected

theoretical value of 0.6°C /100 masl. Variations in correlations may be due to one, or a

combination, of two factors: (1) highly localized weather variability, and (2) insufficient

size of climatic data and/or missing data points. Based on my equations of climate

parameter by elevation, I could identify temperature ranges associated with each of the

three identified vegetation belts.

According to my climate data model for the elevation ranges associated with Lowland

Rain Forest, mean annual temperature (MAT) ranges from 19.9-25.3°C, maximum

absolute temperature (TXx) ranges from 32.2-36.1°C, and minimum absolute temperature

(TNn) ranges from 6.8- 12.3°C. For Lower Montane Rain Forest, MAT ranges from 16.7-

19.9°C, TXx ranges from 32.0-33.2°C, and TNn ranges from 5.2-6.8°C. For Montane

Page 49: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

37

0

0

40 80

20

40

60

80

Axis 1

Axis 2

MCF

MCF

LRF

LMRF -W

MCF LMRF - E

LMRF - E

LRF

MCF

Chelemhá

LRF

Tinajas- E

Lach

Chin

Tin1

Tin2

Tin3

Scruz

Tam

Tac

Che Pur1

Pur2

Pur3Pur4

Pur5

Tin4

Tin5

Tin6

Flo

Bvta

Che3

Che2Che1

Che4

0

0

MCF Purulhá

LRF

- E

LRF

LRF Tinajas

Tin1

Tin2

Tin3 Che Pur1

Pur2

Pur3Pur4

Pur5

Tin4

Tin5

Tin6

Flo

Bvta

Che4

0

0

40 80

20

40

60

80

Axis 1

Axis 2

MCF

MCF

LRF

LMRF -W

MCF LMRF - E

LMRF - E

LRF

MCF

Chelemhá

LRF

Tinajas- E

Lach

Chin

Tin1

Tin2

Tin3

Scruz

Tam

Tac

Che Pur1

Pur2

Pur3Pur4

Pur5

Tin4

Tin5

Tin6

Flo

Bvta

Che3

Che2Che1

Che4

0

0

MCF Purulhá

LRF

- E

LRF

LRF Tinajas

Tin1

Tin2

Tin3 Che Pur1

Pur2

Pur3Pur4

Pur5

Tin4

Tin5

Tin6

Flo

Bvta

Che4

Figure 2.1. Detrended Correspondence Analysis diagram of Las Verapaces sites along the first two DCA axes. Sites enclosed by ovals represent groups identified in the Hierarchical Cluster Analysis (HCA). LRF= Lowland Rain Forest sites, LMRF= Lower Montane Rain Forest sites, MCF= Montane Cloud Forest. W= West, E= East.

Page 50: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

38

y=3.85x+243

r2=0.80

y=3.04x+497.2

r2=0.53

y=3.85x+243

r2=0.80

y=3.04x+497.2

r2=0.53

Figure 2.2. Linear regression of Detrended Correspondence Analysis (DCA) Axis 1 scores (raw scores) of indicator plant taxa (A) and sites (B) against elevation values per site along the Las Verapaces gradient. LRF= Lowland Rain Forest, LMRF= Lower Montane Rain Forest, MCF= Montane Cloud Forest. The dashed square in panel B refers to Tin5 MCF sampling unit.

Page 51: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

39

Figure 2.3. Linear regression curves of temperature (°C) variables collected from meteorological stations from Central and Northern Guatemala. A) TXx = maximum absolute temperature, B) TNn = minimum absolute temperature, and C) MAT= mean annual temperature. Diamonds represents the mean temporal value for the period reported in the station, and dots represent the temporal variation over the length of the record. MAT is taken from 6 stations, and TXx and TNn from 7 stations (see Figure 1.2. for stations locations).

Elevation (m)

0 500 1000 1500 2000 2500

TXx

22

24

26

28

30

32

34

36

38

40

42

Elevation (m)

0 500 1000 1500 2000 2500

MA

T

16

18

20

22

24

26

28

30

Elevation (m)

0 500 1000 1500 2000 2500

TNn

0

5

10

15

20

25

30

r2 =0.79

r2 =0.87

r2 =0.65

Elevation (m)

0 500 1000 1500 2000 2500

TXx

22

24

26

28

30

32

34

36

38

40

42

Elevation (m)

0 500 1000 1500 2000 2500

MA

T

16

18

20

22

24

26

28

30

Elevation (m)

0 500 1000 1500 2000 2500

TNn

0

5

10

15

20

25

30

Elevation (m)

0 500 1000 1500 2000 2500

TXx

22

24

26

28

30

32

34

36

38

40

42

Elevation (m)

0 500 1000 1500 2000 2500

MA

T

16

18

20

22

24

26

28

30

Elevation (m)

0 500 1000 1500 2000 2500

TNn

0

5

10

15

20

25

30

r2 =0.79

r2 =0.87

r2 =0.65

°C°C

°C

Elevation (m)

0 500 1000 1500 2000 2500

TXx

22

24

26

28

30

32

34

36

38

40

42

Elevation (m)

0 500 1000 1500 2000 2500

MA

T

16

18

20

22

24

26

28

30

Elevation (m)

0 500 1000 1500 2000 2500

TNn

0

5

10

15

20

25

30

r2 =0.79

r2 =0.87

r2 =0.65

Elevation (m)

0 500 1000 1500 2000 2500

TXx

22

24

26

28

30

32

34

36

38

40

42

Elevation (m)

0 500 1000 1500 2000 2500

MA

T

16

18

20

22

24

26

28

30

Elevation (m)

0 500 1000 1500 2000 2500

TNn

0

5

10

15

20

25

30

Elevation (m)

0 500 1000 1500 2000 2500

TXx

22

24

26

28

30

32

34

36

38

40

42

Elevation (m)

0 500 1000 1500 2000 2500

MA

T

16

18

20

22

24

26

28

30

Elevation (m)

0 500 1000 1500 2000 2500

TNn

0

5

10

15

20

25

30

r2 =0.79

r2 =0.87

r2 =0.65

°C°C

°C°C

°C

Page 52: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

40

Cloud Forest, MAT ranges from 13.9-16.1°C, TXx ranges from 26.7-30.4°C, and TNn

ranges from 4.0-4.9°C.

2.3.3 Role of landscape position and watershed topography

Although elevation appears to be an important factor controlling plant taxa differences

between sites, the multivariate analyses show that other factors are important as well. The

arrangement of sites along the DCA axes responds possibly to landscape position which

was further confirmed through the HCA dendrogram (Figure 2.1). Lachuá and Chinajá

are both Lowland Rain Forest sites yet they show a separation on the ordination diagram

likely due to topographical factors (i.e. Lachuá flatlands versus Sierra Chinajá).

According to HCA, sampling units from Tinajas watershed were separated according to

elevation in the three vegetation belts. Lower Montane Rain Forest sites were separated

in two main groups according to their geographical location: east and west. The east

group consisted of the Tucurú sampling units and the west group included Tactic, Santa

Cruz and Tamahú sites. The Montane Cloud Forest sites were allocated into sub-groups

as a function of their location in three different ridges separated by valleys (Figure 1.2):

Sierra de Las Minas, Sierra Chuacús, and Sierra Yalijux.

2.3.4 Species assemblages and indicator species

Using regression of DCA Axis scores against elevation for each taxon, I sorted the

original 794 angiosperm plant taxa into 26 indicator, 20 generalist, and 9 disjunctive plant

taxa whose distributions were zonal, continuous and discontinuous (respectively) along

the elevational gradient. These categories allowed me to more clearly correlate vegetation

Page 53: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

41

with elevation, geographical conditions, and in turn, with climate. The remaining taxa

(739) did not present a unique DCA Axis 1 score, as many plant taxa shared the same

score. The identified indicator taxa were related to the elevational range of one of three

vegetation belts (Table 2.3). In terms of generalist taxa (Table 2.4) there were 14 taxa

common across Lowland Rain Forest and Lower Montane Rain Forest, 11 taxa common

across Lower Montane Rain Forest and Montane Cloud Forest, and 2 taxa across all three

vegetation belts. Disjunctive taxa (Table 2.5) were distributed in both Lowland Rain

Forest and Montane Cloud Forest.

2.3.5 Biogeographical affinities

There is an increase in Laurasian and Andean indicator taxa, and a decrease in

Amazonian taxa, when moving from Lowland Rain Forest to Montane Cloud Forest

(Figure 2.4). The generalist taxa common to Lowland Rain Forests and Lower Montane

Rain Forests are all Amazonian-centered taxa. Andean-centered and Laurasian generalist

taxa are only common between Lower Montane Rain Forest and Montane Cloud Forest

(Figure 2.4). Andean-centered taxa co-dominate the disjunctive taxa with Amazonian-

centered taxa, and to a much lesser extent, the unassigned taxa to a particular origin.

According to the chi-square contingency test, the frequencies observed of biogeographic

categories along vegetation zones are not at all likely explained by chance (Χ2= 35.00, df

=8, p<0.0001).

Page 54: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

42

2.3.6 Study Limitations: sampling effects

The sites included in this study can be separated into two groups according to their

sampling effort: high intensity sampling, representing a detailed collection of plants

(understory, subcanopy and canopy layers) along the spatial variability of an

environmental gradient (Purulhá, Chelemhá, Lachuá and Chinajá) and low intensity

sampling (Tucurú, Tamahú, Santa Cruz, Tactic, and Rio Tinajas). Many studies indicate

that sampling effort is directly related to species richness and diversity (Shen et al.,

2003). Low intensity sampling studies (e.g. mainly focused on forest inventories) are

likely to result in the collection of mostly abundant and generalist species than rare and

specialist species (Pitman et al., 2001). Most of the generalist plant taxa were found in the

Lower Montane Forest which includes exclusively the low intensity sampling sites

(Figure 2.2b).

The pattern found in the Las Verapaces region could be affected by differences in the

research objective of each study, experimental designs, and sampling efforts (Otypková

and Chytry, 2006). This limitation is significant to recognize because most of the

information on vegetation is in low intensity format for Guatemala (i.e. as it is over the

rest of the tropics) (Mathewson, 2006). Nevertheless, after being made aware of the

possible caveats and weaknesses, I found that the combination of information from both

low and high sampling effort studies allowed me to differentiate three vegetation belts

and explain their delineation based on elevation and climate, landscape position and

watershed topography, and biogeographic origin (Figure 2.1 and 2.2).

Page 55: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

43

Vegetation belt Family Biogeographic originLowland Rain ForestGenipa sp. RUBIACEAE ANSpondias mombim ANACARDIACEAE AMZTabebuia sp. BIGNONIACEAE AMZ

Lower Montane Rain ForestCedrela pacayana MELIACEAE AMZHeliocarpus mexicanus TILIACEAE AMZInga sp. FABACEAE AMZPerymenium grande ASTERACEAE ANSaurauia belisensis ACTINIDIACEAE LAU

Montane Cloud ForestBegonia oaxacana BEGONIACEAE ANCavendishia guatemalensis ERICACEAE ANCentropogon cordifolius CAMPANULACEAE ANClethra suaveolens CLETHRACEAE LAUErigeron karvinskianus ASTERACEAE ANFuchsia microphylla ONAGRACEAE ANLobelia nubicola CAMPANULACEAE ANMiconia aeruginosa MELASTOMATACEAE ANMiconia glaberrima MELASTOMATACEAE ANOcotea sp. LAURACEAE AMZOreopanax liebmanii ARALIACEAE ANPassiflora sexflora EUPHORBIACEAE AMZPhoradendron sp. LORANTHACEAE ANPsychotria parasitica RUBIACEAE ANRhynchosia sp. FABACEAE AMZStyrax argenteus STYRACACEAE LAUSynardisia venosa MYRSINACEAE ANWeinmannia pinnata CUNIONIACEAE AN

Table 2.3 Indicator plant taxa for the three vegetation belts along the Las Verapaces region, selected from DCA axis scores for species (see text for details). AN= Andean-centered, AMZ=Amazon-centered, LAU= Laurasian.

Page 56: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

44

Vegetation belts Family Biogeographic originLRF-LMRFBursera simaruba BURSERACEAE AMZCecropia peltata CECROPIACEAE AMZCeiba pentandra BOMBACACEAE AMZParathesis vulgata MYRSINACEAE ANTerminalia amazonia COMBRETACEAE AMZVirola sp. MYRYSTICACEAE AMZVochysia guatemalensis VOCHYSIACEAE AMZ

LMRF-MCFBillia hippocastanum HIPPOCASTANACEAE LAUBrunellia mexicana BRUNELLIACEAE ANDendropanax leptopodus ARALIACEAE ANEngelhardtia guatemalensis JUGLANDACEAE LAUEupatorium semialatum ASTERACEAE ANHedyosmum mexicanum CHLORANTHACEAE LAULiquidambar styraciflua HAMMAMELIDACEAE LAUMyrica cerifera MYRICACEAE LAUPersea donnell-smithii LAURACEAE AMZQuercus crispifolia FAGACEAE LAUQuercus sp. FAGACEAE LAU

LRF-LMRF-MCFClusia sp. CLUSIACEAE ANMollinedia guatemalensis MONIMIACEAE AN

Table 2.4 Generalist plant taxa for the Las Verapaces region, as determined by DCA axis scores for species (see text for details). AN= Andean-centered, AMZ=Amazon-centered, LAU= Laurasian.

Page 57: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

45

Table 2.5. Disjunctive plant taxa distributed in Lowland Rain Forest and Montane Cloud Forest in the Las Verapaces region generated from DCA axis scores for species (see text for details). AN= Andean-centered, AMZ=Amazon-centered, LAU= Laurasian.

Disjunctive taxa Family Biogeographic origin

Clidemia capitellata MELASTOMATACEAE AN Conyza bonariensis ASTERACEAE AN Dendropanax arboreus ARALIACEAE AN Lasciacis divaricata POACEAE Unassigned Matayba oppositifolia SAPINDACEAE AMZ Ocotea eucuneata LAURACEAE AMZ Peperomia cobana PIPERACEAE AN Phoebe sp. LAURACEAE AMZ Pouteria campechiana SAPOTACEAE AMZ

Page 58: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

46

Figure 2.4. A) Indicator (in one vegetation belt) and B) generalist (across two vegetation belts) taxa separated according to their biogeographic origin along the Las Verapaces region vegetation belts (increasing in elevation from left to right) (for indicator taxa Χ2= 35.00, df =8, p<0.0001). LRF=Lowland Rain Forest, LMRF=Lower Montane Rain Forest, MCF=Montane Cloud Forest, AN= Andean-centered, AMZ=Amazon-centered, LAU= Laurasian.

0%

20%

40%

60%

80%

100%

LRF LMRF MCF

LAUANAMZ

A

0%

20%

40%

60%

80%

100%

LRF LMRF MCF

LAUANAMZ

A

0%

20%

40%

60%

80%

100%

LRF-LMRF LMRF-MCF LRF-LMRF-MCF

LAUANAMZ

B

0%

20%

40%

60%

80%

100%

LRF-LMRF LMRF-MCF LRF-LMRF-MCF

LAUANAMZ

B

Page 59: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

47

2.4 Discussion  

2.4.1 Elevation and climate

Climatic dynamics along elevation is possibly the first-order explanatory factor in

understanding modern-day vegetation trends in our study region in Guatemala, as is

found with other Latin American countries (Quintana-Ascencio and González-Espinosa,

1993; Gerold et al., 2008) and generally world-wide (Hemp, 2006; Kappelle et al., 1995).

The arrangement of sites along Axis 1 in the DCA diagram shows a clear relationship

between floristic composition and elevation (Figure 2.2). Elevation-species interactions

are the result of the environmental lapse rate; in other words, changes in climate

associated with distance from sea level (Figure 2.3). Temperature data for each

vegetation belt generally corresponds to Holdridge’s Life Zones in Guatemala (De La

Cruz, 1982) that are themselves defined mostly as a function of climate. After reviewing

information for vegetation belts across Central America and Mexico (Islebe and

Kappelle, 1994; Islebe and Velázquez, 1994), elevation and its correlation with

temperature variability is found to be a common factor for differentiation of vegetation

belts.

Within a given elevation, my data highlight some unexpected differences in vegetation

composition, indicating that maybe broad-scale climatic changes dictated by the

environmental lapse rate are only part of the explanation for plant species turn-over

through space and time. Studies of tropical forests in Hawai'i (Crausbay and Hotchkiss,

2010), Venezuelan Andes (Cuello and Cleef, 2009) and the Chihuahuan Borderlands

(Poulos and Camp, 2010) indicate that factors such as strong moisture gradients,

Page 60: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

48

topography, and incident solar radiation are important to define changes in vegetation.

Lachuá’s dissimilarity to Chinajá (Figure 2.1) within the Lowland Rain Forest belt, for

example, may be due to variations in local microclimate. On the one hand, Chinajá

vegetation is located in an isolated topographic feature (e.g. approximately 500 m higher

than Lachuá) of the relatively flat lowlands of the Chixoy watershed and therefore will

likely exhibit a distinct microclimate. On the other hand, Lachuá vegetation is located

close to the Sierra Chamá foothills (~170 masl) and therefore exposed to increased

moisture due to orographic effects on precipitation. My climate data were unable to

capture this climatic variability likely because of the lack of spatial coverage of

meteorological stations in this region.

The climatic uniqueness of the Sierra Chinajá and Lachuá has resulted in the presence of

some plant taxa distributed normally in higher elevations at the Montane Cloud Forest.

Cloud forest microclimate and niche variability has possibly allowed highland species to

establish at locations outside of their expected distributional range, in this case in lower

elevations (Table 2.4). These species are found jointly in Chinajá and Lachuá, both at the

foothills of Sierra Chamá.

2.4.2 Watershed topography and landscape position

The other Montane Cloud Forest sites (Purulhá and Chelemhá) are most likely

differentiated in terms of their location in different mountain ranges: the Sierra Chuacús

and Sierra Yalijux, respectively (Figure 1.2). The low altitude mountain passes (valleys)

found between Purulhá and Chelemhá may function as a modern-day physical barrier for

Page 61: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

49

biological dispersal as montane species would have to migrate below the minimum

altitude defining cloud forests (~ 400-1400 masl) to reach to the other side. Alternatively,

these valley bottom habitats act as corridors for biota that have adapted physiological

tolerances to lower elevation conditions (Schmitt et al., 2010). I believe the separation of

Tinajas 6 site from the other highland sites (Purulhá, Chelemhá) is probably due to the

fact that the site's vegetation data are based solely on forest inventories that lack

taxonomic specificity. In other words, the underlying explanation for the Tinajas 6 site

containing mostly generalist taxa (Billia hippocastanum, Dendropanax leptopodus) for

Lower Montane Rain Forest and Montane Cloud Forest is most likely a reflection of

differences in sampling intensity.

Analysis of sites as a function of their watershed location indicates that the landscape

position of sampling sites with respect to local relief (topography) may be another

important element in explaining plant community composition in the Guatemalan

lowlands. The Lachuá and Chinajá Lowland Rain Forest sites are more probably

differentiated from their lowland counterpart, Tinajas, because the former sites are found

in the Chixoy watershed and the latter in the Polochic watershed (Figure 2.1). These two

watersheds are separated by steep highland mountains, the Sierra Chamá and the east

portions of Sierra Yalijux. Steep mountain divides may be currently acting as physical

barriers to species migrations between adjacent watersheds. Janzen (1967) was the first to

recognize the ecological importance of steep elevational (climatic) gradients that tend to

be more common in tropical mountain passes than they are in mountainous regions at

higher latitudes (i.e. temperate and boreal regions).

Page 62: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

50

DCA results indicate that longitudinal differentiation between western and eastern groups

in the Lower Montane Rain Forest correlate in part to topographical differences (Figure

1.2) and according to national descriptive maps (i.e. topographic and climatic), they may

also correlate to variations in temperature and precipitation (MAGA, 2004). The latter

response could not be tested due to the limitations of my climatic database. In the western

group, topography is influenced most by the narrow and higher elevation valleys at the

Cahabón watershed. In the eastern group, topography is characterized as wider and lower

elevated valleys located in the Polochic Watershed. The HCA dendrogram indicates that

although the Tamahú site is located in the eastern group, because of its geographical

proximity to the border of the Cahabón watershed, its vegetation is more similar to the

western group’s vegetation community than to that of the eastern group (i.e. an indication

of spatial autocorrelation).

2.4.3 Paleogeography and current vegetation biogeography

Analyses of data also indicate a possible role for biogeographic origin in explaining

vegetation distribution (Figure 2.4). Amazonian-centered taxa found in Central America

occupy ecologically important niches as lowland forest dominants with wide-ranging

distributions, and Andean-centered taxa dominate the humid foothills and mid-elevational

ranges (Gentry, 1982). Laurasian taxa are important in ecological terms because they are

dominant canopy members in montane forests, becoming more dominant as elevation

increases (Hammel and Zamora, 1990). In southeastern Mexico, Laurasian taxa were

found to increase with elevation, likely as a result of adaptations to climate-related

Page 63: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

51

disturbances such as the risk of night frosts and desiccation arising from strong winds

(Quintana-Ascencio and González-Espinosa, 1993).

At this time, however, I believe it is premature to speculate on modern-day vegetation

trends in the Las Verapaces region (e.g. Amazonian, Andean, and Laurasian) or in the

rest of Guatemala arising from paleoclimate-forest dynamics. Glacial and inter-glacial

cycles of the Pleistocene are known to cause mixing of lowland and highland plant taxa

(Hooghiemstra and Van der Hammen, 2004), as well as to create a deterrent (barrier) to

vegetation migration between two points on the landscape (Terrab et al., 2008). Both

mixing and separation of plant taxa during the Pleistocene could explain the present

vegetation pattern found in the Las Verapaces elevational gradient (Figure 2.4). Studies

on the population dynamics of Scarabaeoidea (dung beetles) in Guatemala have identified

locations within montane cloud forests containing endemics that likely resulted from

Pleistocene paleoclimatology (Schuster, 2006). Because the delineation of Scarabaoidea

communities as a function of elevation (Schuster et al., 2000) is very similar to my

proposed vegetation belts, it makes the connection to Pleistocene dynamics all that more

enticing. Strong similarities and redundancy between flora and fauna distributions are

good indications of the importance of historical biogeographic processes in explaining

modern-day species distributions (Jones and Kennedy, 2008)

The regional geological history of Guatemala as it relates to mountain building in North

and South America may contribute to the explanation of why I have found a combination

of plant taxa with different biogeographic origins occupying different ecological niches

Page 64: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

52

(Figure 2.4). Laurasian taxa typical of high elevation locations, for example, are found in

highland sites (Sauraia belisensis, Clethra suaveolens). For the most part, generalized

vegetation patterns across Central America were laid down in the Miocene period, when

Nuclear Central America was known as Proto-Central America (Raven and Axelrod,

1974; Graham, 1999). Orogenic processes in eastern-southern Mexico and central-

southern Guatemala (Padilla, 2007) likely promoted dispersion of Laurasian taxa

throughout the newly originated Guatemalan highlands. Amazonian-centered taxa

(Spondias mombin, Tabebuia sp.) dispersed into lowland regions via "island hopping"

over the Proto-Antillean Mountain Chain, both before and after the Pliocene closing of

the Central American Land Bridge approximately 3 million years ago (My). The physical

connection of Central America to South America also facilitated migration of Andean-

centered taxa (Oreopanax liebmanii, Weinmannia pinnata) into the foothills and

highlands in Guatemala where they currently dominate in the Montane Cloud Forest

(Table 2.3). Migration of Andean-centered taxa occurred sometime after Andean

orogenesis, beginning around 5 My BP (Antonelli et al., 2009).

2.4.4 Conservation biology and disjunctive taxa

Montane cloud forests are quickly becoming the focus of international conservation as

both their ecological and societal services are now being highly recognized (Vargas-

Rodríguez et al., 2010). Already, indications that cloud forests are experiencing change,

whether due directly to humans via land-use change or indirectly through climate change,

have been identified in India (Murugan et al., 2009), Mexico (Martínez et al., 2009) and

Central America (Colwell et al., 2008). Models predict that deforestation of lowland

Page 65: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

53

rainforest causes a lowering of cloud base heights, in turn promoting reductions in

atmospheric moisture in the upper reaches of cloud forest (Nair et al., 2003). From a

conservation perspective, this study containing lists of plant species composition in

Guatemalan cloud forests will be important; as it is important to know what species are

there now so to have a benchmark from which to ascertain potential species turnover in

the future. The fact that some Montane Cloud Forest taxa have been observed at lower

altitudes in Purulhá and Chelemhá indicate that Chinajá and Lachuá forests have unique

habitat conditions (i.e. canopy microclimate) that have provided a critical refuge for

cloud forest plant species.

Andean-centered forest taxa such as Clidemia capitellata (Melastomataceae), Conyza

bonariensis (Asteraceae), and Dendropanax arboreus (Araliaceae) that are typically

located in the montane cloud forests were also found in lowland rain forests, indicating an

important link between both vegetation belts. Other taxa along our montane Chinajá site

also include non-plant taxonomic groups such as dung beetles, birds and bats (Bonham,

2006). In some cases the disjunctive pattern that we observed in understory vegetation

such as Lasiacis divaricata is more likely the result of insufficient sampling size due

mostly to the fact that the forestry surveys were primarily focused on canopy tree species.

Some Amazon-centered plant taxa from the Lauraceae (e.g. Ocotea and Phoebe) family

have extended their distribution from typical lowland habitats to high elevation

conditions (Chanderbali et al., 2001). Mixing of species increases regional diversity (i.e.

gamma diversity) as a response to the presence of multiple habitats (i.e. alpha and beta

Page 66: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

54

diversity) each with unique physiographic and ecological features (Emmerson et al.,

2001).

2.5. Chapter summary  

In this chapter three vegetation belts have been identified as expected and described in

terms of changes in plant communities along elevation. Other factors were found to

complementarily explain the relationship plant taxa-elevation, such as variability in

climatic parameters (i.e. temperature related), watershed location and topography, and

biogeographic origin. A list of 794 angiosperm plan taxa was generated based on the

collation of a data base of local vegetation inventories in the Las Verapaces region. This

list contains information about the elevation distribution of each taxon in different

watersheds according to the location of the vegetation inventory.

Based on a Detrended Correspondence Analysis (DCA), plant taxa with unique scores

along ordination Axis 1 were separated according to their distribution in elevation ranges

that corresponded three vegetation belts: Lowland Rain Forest (170-1000 masl), Lower

Montane Rain Forest (1000-1800 masl), and Montane Cloud Forest (1800-2500 masl).

The selection of these plant taxa is useful in identifying indicators for each vegetation

belt, or for generalists with a wider elevational distribution preference over no more two

neighbouring vegetation belts (Tables 2.3 and 2.4).

The application of identifying the correspondence between vegetation belts and indicators

and generalist taxa, is to know how to identify in calibration and paleoecological studies

Page 67: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

55

the vegetation source of correspondent pollen taxa. Relating the prevalence of different

biogeographic origins (Amazonian, Andean, and Laurasian) and plant taxa provides the

linkage to understand the ecological characteristics of vegetation belts, which explains to

great extent the pollen source area and representations of pollen spectra of a location (i.e.

dispersion of pollen depends greatly on dispersal syndromes). In this sense Tables 2.3

and 2.4 are linked to Table 3.1.

Page 68: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

56

Chapter 3:

Modern pollen rain in the north-central Guatemalan lowlands and highlands

3.1 Introduction  

The correlation of modern pollen rain to landscape features is an important first step in

understanding the interpretation of paleoecological pollen signals at either local or

regional scales. Defined mostly from Northern Hemisphere paleoecological studies,

pollen observed in mid-sized to larger lakes represents mostly regional vegetation,

whereas pollen in observed in small basins represents local vegetation. As landscapes

become more open in character, the pollen signal from both smaller and larger basins

becomes more similar (Conedera et al., 2006; Lynch, 1996; Prentice, 1985). Thus, basin

size, range of pollen dispersal, and patterns of vegetation cover are factors that can

influence the mix of pollen found at any one sampling location. Moreover, other factors

such as topography, atmospheric conditions (i.e. prevailing wind circulation),

sedimentation rates and mode of sediment transport (Brown et al., 2008; Bunting et al.,

2004) may also control the mix of accumulating pollen.

In temperate regions (i.e. mid-latitude) it has been found that pollen content from

sediments of open basins has a more regional vegetation signal than a local signal,

because the pollen source area allows more deposition of wind-dispersed pollen (Fægri

Page 69: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

57

and Iversen, 1989). Higher proportions of local vegetation have has been found in pollen

collected in forest interior surfaces, such as bryophyte polsters (i.e. moss polsters), where

short-distance dispersed pollen (i.e. animal pollinated) is found in greater amounts. Pollen

dispersal syndrome directly influences pollen source area and indirectly influences the

way pollen grains are trapped in different habitats and reservoirs (Bush and Rivera,

1998). The probability that bryophyte polsters will trap airborne pollen is low relative to

sediment samples from mid- and large-sized basins, where the effectiveness of vegetation

barriers decreases as distance increases from the shoreline (Conedera et al., 2006; Fægri

and Iversen, 1989; Lynch, 1996). Therefore, pollen content is different according to the

pollen sources, and in this case pollen signal of bryophyte polsters may indicate which

local pollen types are absent from surface sediments (Wilmshurst and McGlone, 2005).

This is of special interest for calibration because surface sediments represent the best

analogue for samples collected from cores where fossil assemblages are extracted from

and from where landscape evolution is inferred.

The over-arching objectives in conducting this research are two-fold: (1) To quantify the

relationship between modern pollen rain and local-to-regional features of the natural

landscape in two sites along a north-to-south elevational gradient (i.e. 170 to ~ 2000

masl) in the Las Verapaces region in Central Guatemala (Figure 3.1). The two sites are

located at the Lachuá lowlands and Purulhá highlands, in the Lowland Rain Forest belt,

and the Lower Montane Rain Forest-Montane Cloud Forest ecotone, respectively (see

Chapter 2). The Las Verapaces region was chosen because of its complex mosaic of

biophysical settings (Avendaño et al., 2007, MAGA, 2001) and archaeological sites

Page 70: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

58

Figure 3.1. Location of Guatemala in Central America. Las Verapaces Region is enclosed in rectangle. 1= Lachuá lowlands, 2= Purulhá highlands.

MéxicoBelize

Honduras

NicaraguaEl Salvador

Costa Rica

Panama

Guatemala

Pacific Ocean

Caribbean Sea

Gulf of Mexico1

2México

Belize

Honduras

NicaraguaEl Salvador

Costa Rica

Panama

Guatemala

Pacific Ocean

Caribbean Sea

Gulf of Mexico1

2

Page 71: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

59

(Dillon, 1977; Ichon et al., 1996). (2) To provide much-needed data for the Las

Verapaces region, an area that to date is under-explored in terms of modern pollen

calibration and paleoecology relative to the number of studies focusing on the northern

Guatemalan lowlands (Binford and Leyden, 1987; Curtis et al., 1996; Hillesheim et al.,

2005; Wahl et al., 2007a). This study represents the second modern pollen rain

calibration conducted in Guatemala and is one of the few studies in the Mesoamerica

region across Southeast Mexico, Guatemala, Belize and Honduras. For these reasons, this

study represents an important contribution for paleoecological study of tropical

ecosystems in general.

Specifically, I address the following three research questions: (1) What are the

differences between pollen spectra represented in bryophyte polsters and surface

sediments? How does the first inform me about the latter? (2) Is the modern pollen rain in

bryophyte polsters and surface sediments representative of local or regional vegetation?

(3) Which pollen taxa are reliable indicators of environmental conditions or vegetation

zonation along the study gradient?

 

3.2 Methods  

3.2.1 Bryophyte polster pollen sampling

Bryophyte polsters were collected in the interior of minimally-disturbed forest habitats,

located far enough inside the forest (more than 250 m) to avoid “edge-effects” (Bush and

Rivera, 1998). Bryophyte polster samples from Lachuá lowlands (hereafter just Lachuá)

Page 72: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

60

were taken every 50 m along a 200 m transect in two locations in the interior of forests

(n=10) (elevation ~170 masl) (interior of LLNP) (Figure 1.3). Bryophyte polsters

samples from Purulhá highlands (hereafter just Purulhá) were collected along a 2 km

transect (10 samples spaced 200 m apart) (elevation ~1700-1800 masl) (interior of

BUCQ) (Figure 1.3). A bryophyte polster sample comprised of the amalgamation of

bryophytes cushions found in a 5 m radius, were stored and labeled in plastic Ziploc bags.

3.2.2 Surface sediment pollen sampling

Surface sediments samples of 1.0 cm in length were extracted from cores taken using a

Livingstone corer and stored in plastic Ziploc bags. Lachuá samples are from Lachuá

Lake and Tortugas Ponds; and Purulhá samples from Chichoj Lake and the Cahabón

River Floodplain. Surface sediments from Lachuá samples were collected in three

locations (L1, L2 and L3) near the Lachuá lakeshore because the ideal location (lake

centre) was too deep to core (200 m) (Figure 1.3). I chose sites that were located away

from stream inflow and outflows to minimize disturbance of sediments. Sample L1 was

located approximately 2-3 m from the shore, sample L2 approximately 20 m from the

shore, and sample L3 was located in a lakeside wetland. Approximately 5 km northeast

from Lachuá Lake one extra core was sampled from the Tortugas Ponds shore at Salinas

de los Nueve Cerros Regional Park (sample Sa2). The pond is about 200 m in diameter

and is completely surrounded by high canopy (40 m) lowland rainforest (Cajas, 2009).

Surface sediments from the Purulhá were sampled in several Fincas (Villa Trinidad,

Patal, and Chisiguan) along the floodplains (600 m to 1 km wide) of the headwaters of

Page 73: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

61

the Cahabón River (elevation range 1450-1560 masl), close to the towns of Purulhá

(samples P1 and P4) and Tactic (sample T1) (Figure 1.3). Another sample was taken

from a marsh adjacent to the heavily-polluted Lake Chichoj (sample J1) (47.6 ha) near

the town of Santa Cruz Verapaz (elevation 1390 masl) (Sánchez, 1994). As much caution

as possible was taken in choosing samples from locations where disturbance from

incoming rivers, landslides, or human activities was at a minimum.

3.2.3 Identification of pollen source areas

Pollen source area is considered "local" when the plant is reported in the local vegetation

inventory, "regional" when the vegetation source is located in a neighboring elevational

vegetation belt, and is considered “extra-regional” when the plant is separated more than

one vegetation belt (Chapter 2). Since I do not have modern pollen rain samples from

the Lower Montane Rain Forest (intermediate vegetation belt between Lachuá and

Purulhá), Lachuá pollen is regional when found in Purulhá because in terms of pollen

signal I considered Lowland Rain Forest and Lower Montane Rain Forest closely related,

an assumption based on Domínguez-Vázquez et al. (2004) (See Chapter 2 for definition

of vegetation belts). Due to their biogeographical affinity, arboreal pollen taxa from

Lachuá are named tropical and from Purulhá they are considered temperate. In the case of

Abies and Alnus, I consider them extra-regional for Lachuá because their plant stands are

found two vegetational zones higher, but for Purulhá they are regional. In the case of the

widely distributed Pinus, it is an indicator of highland temperate vegetation, independent

of its lowland populations (P. caribea). In order to create Table 3.1, identified pollen

types corresponding to plant taxa listed in Tables 2.3 and 2.4 were automatically

Page 74: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

62

assigned to one or two vegetation belts. When pollen types (i.e. genus) are not found in

the latter tables, their allocation to lowlands or highlands vegetation, or a vegetation belt

not covered in the Las Verapaces (i.e. due to its location in higher elevation) was based

on revision of Latin American pollen and vegetation literature (Gentry 1982, Marchant,

2002; Domínguez -Vázquez et al. 2004).

3.2.4 Modern pollen laboratory work

Samples for pollen analysis were processed under standardized acetolysis procedures to

remove organic matter and cellulose, as well as to concentrate pollen grains (Fægri and

Iversen, 1989). Pollen counting was completed on a 200 grain per sample basis when

possible, of which at least 100 pollen grains were from arboreal taxa. Pollen

concentration was calculated based on the addition of exotic Lycopodium clavatum spore

tablets. A total of ten bryophyte polsters and four surface sediments samples were

counted in each of Lachuá and Purulhá making a total of twenty (20) bryophyte polsters

and eight (8) surface sediment samples.

Pollen grain identification was done using regional and local pollen reference collections

obtained from Colombia (Hooghiemstra, 1984), Barro Colorado Island in Panama

(Roubik and Moreno, 1991), Lachuá (Barrientos, 2006) and generally for the Neotropics

(Bush and Weng, 2007). Pollen identification was aided by the Pollen Reference

Collection from the Neotropical Research Unit from the Department of Geography,

University of Leicester, England. Fixed slides were stored in the reference collection at

Page 75: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

63

the Paleoecology Laboratory of the Department of Geography at the University of

Toronto (Canada).

3.2.5 Modern pollen rain statistical calibration

The pollen sum included arboreal and non-arboreal taxa, which were identified to family

and genus level. Unknowns, spores and aquatics (e.g. Cyperaceae) were not included in

the pollen sum (Fægri and Iversen, 1989) and their abundance was measured as a ratio in

relation to the total pollen sum calculated per sample. Arboreal pollen (AP) and non-

arboreal pollen (NAP) percentages were calculated per site and pollen reservoir

(bryophyte polster and surface sediment) to represent local landscape vegetation cover.

Additionally, for each sample the contribution of pollen provenance (i.e. local or

regional) and pollen dispersal syndrome were identified. Dispersal syndrome included the

following: (1) zoophilous (animal dispersed), (2) anemophilous (wind dispersed), and (3)

ambiphilous (combination of both).

For local analysis at each site, species abundance matrices were built to compare

bryophyte polsters and surface sediments. In contrast, for regional analysis a common

matrix based on shared taxa between sites was built. Detrended correspondence analysis

(DCA) was used to visualize samples according to similarity of their pollen assemblages

and their probable arrangement as a function of environmental gradients (Jongman et al.,

1995). Analysis was complemented with a factor analysis (FA) using Varimax rotation in

order to isolate pollen types (factors) that explain the largest amount of variance (with

minimal loss of information) (May, 1974). The pollen type’s scores over FA gradients

Page 76: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

64

(named "factors") complemented the explanation of gradients found in DCA axes.

Summary pollen diagrams were plotted based on local and regional analyses. PC-Ord

(McCune and Mefford, 2006) and PAST (Hammer et al., 2001) were statistical packages

used for multivariate analysis, and C2 (Juggins, 2003) for building pollen diagrams.

Information on pollen dispersal syndrome per taxon was derived from local vegetation

studies (Ávila, 2004; Cajas, 2009; García, 1998).

 

3.3 Results 

56 pollen types were identified at least to family and genus taxonomic level at the Las

Verapaces region (Table 3.1). Pollen types were compared to information presented in

Tables 2.2 and 2.3 in order to link them to corresponding vegetation belt(s). Plant

indicator taxa (Table 2.2) correspond to one vegetation belt because of their specificity,

while generalist taxa correspond to two belts (Table 2.3). Based on criteria found in

bibliographic revisions of Latin American pollen studies, some pollen types were

interpreted to represent in general “lowlands” (i.e. Sapotaceae) or “highlands” (i.e.

Urticaceae) vegetation, when they represented more vegetation belts (i.e. Lowland Humid

Forest, see Table 2.1) that the ones covered in the Las Verapaces region vegetation

chapter (Chapter 2), or in the case that the pollen type had a wide altitudinal range

distribution either in lowlands (i.e. Celtis) or highlands (i.e. Pinus). Pollen types from

vegetation belts not found in the Las Verapaces region were collected (see Table 2.1),

such as Abies from Montane Mixed Forest, and Alnus from Montane Mixed Forest and

Subalpine Forest belt.

Page 77: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

65

Lachuá species richness is higher than Purulhá (45 pollen types versus 31), with a total of

20 pollen types shared between sites (Figure 3.2). Bryophyte polsters contain higher

arboreal pollen (AP) type richness than surface sediments at both locations.

3.3.1 Lachuá modern pollen spectra

AP content is dominant in both bryophyte polsters and surface sediments at Lachuá, and

with few exceptions, consists of mostly local provenance and zoophilous taxa (Figure 3.3

and Table 3.2). Once the over-represented Pinus is removed from the AP data matrix, the

DCA diagram shows a general separation between bryophyte polsters and surface

sediments along Axis 1, with the exception of surface sediment sample Sa2 which is

separated along Axis 2 (Figure 3.4). Sa2 is segregated from other surface sediment

samples because its pollen content has the highest abundances of local taxa (Bursera,

Psychotria, Spondias, and Trema) and has the only record of Inga for surface sediments.

Surface sediments are dominated by the local entomophilous Celtis and highland

anemophilous Pinus. Sample L2 has the higher abundances of the temperate highland

anemophilous Abies and Myrica, and is the only sample that contains local

entomophilous Mimosa. Surface sediments at L3 have the highest abundance of Ilex

(highland zoophilous taxon) and some local entomophilous taxa such as Myrtaceae,

Sapium, Solanaceae and Terminalia. The pollen assemblage in sample L1 is a mix of taxa

found at sites L2 and L3.

The dominant taxa in bryophyte polsters are the local entomophilous Solanaceae and

highland Pinus. Many local entomophilous taxa (e.g. Bombacaceae) are only found in

Page 78: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

66

bryophyte polsters (although poorly represented at ~1%). Various local entomophilous

(Brosimum and Terminalia), local anemophilous (Alchornea) and anemophilous

temperate (Alnus, Pinus, and Quercus) taxa have a wide representation across both pollen

reservoirs.

Asteraceae is similarly distributed in bryophyte polsters and surface sediments; Poaceae

is more abundant in the former, and Zea is poorly represented in both (~1%) (Figure

3.3). Trilete spores are consistently more abundant in bryophyte polsters, while monolete

spores (with the exception of sample L3) and aquatic pollen are similarly represented

across both pollen reservoirs.

3.3.2 Purulhá modern pollen spectra

The AP fraction is higher than non-arboreal pollen (NAP) in bryophyte polsters, and local

and anemophilous pollen fractions are higher than regional and zoophilous in both

bryophyte polsters and surface sediments (Table 3.3). Local anemophilous Hedyosmum

and Quercus are the most abundant AP taxa and are highly represented in bryophyte

polsters (Figure 3.5). With respect to AP and NAP combinations, DCA analysis shows a

separation between bryophyte polsters and surface sediments along Axis 1 (Figure 3.6).

Some regional taxa are represented in both pollen reservoirs (e.g. highland Alnus, and

lowland Celtis). Local (Ilex) and lowland taxa (Alchornea and Myrsinaceae) are found

only in bryophyte polsters relative to surface sediments. Temperate taxa abundances,

such as Abies, Pinus and the local Myrica, are similar in representation between

bryophyte polsters and surface sediments. Sample N4 presents the highest abundance of

Page 79: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

67

Cecropia. Most of lowland taxa (in particular, those with entomophilous dispersal

syndrome) were poorly represented in both polsters and surface samples in highlands.

Poaceae is the most abundant NAP type and is more abundantly represented in surface

sediments, while Asteraceae (second most abundant NAP type) is similarly represented in

both pollen reservoirs. Surface sediment sample P4 is the only sample where the

disturbance-related taxon, Alternanthera, is found. Trilete spores are more abundant in

bryophyte polsters, while monolete spores and aquatic pollen types are dominant in

surface sediments.

3.3.3 Las Verapaces regional modern pollen spectra

Fifteen AP and five NAP taxa are common in both Lachuá and Purulhá (Figure 3.2),

with most of these taxa having their highest relative abundance either where plant stands

are reported from local inventories, or according to their elevational range of distribution

(Table 3.1). When analyzing shared AP-types, there is a clear separation along DCA

Axis 1 between the bryophyte polsters of Lachuá and Purulhá. Surface sediments are

located in the middle of Axis 1 (with some separation along Axis 2) and slightly

separated according to their location (Figure 3.7). The exception to this includes samples

L1, Sa2, and P4 surface sediments that are placed closer to their polster counterparts.

Lachuá surface sediment (L3) and Purulhá bryophyte polster (N8) are isolated on Axis 2

because they have the highest values of Ilex, which is neither found in bryophyte polsters

at Lachuá or surface sediments at Purulhá. When surface sediments samples are

compared at both Lachuá and Purulhá, there is a clear segregation between lowlands and

Page 80: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

Table 3.1. Pollen types and their % range in abundance for bryophyte polsters (BP) and surface sediments (SS). Information about vegetation belt, plant habit, pollen dispersal syndrome (DS), and biogeographic origin (Biogeo) is provided, partially based on Table 2.2 and 2.3 (Chapter 2). Z= Zoophilous, W= anemophilous, A= Ambophilous (Z and W). L-BP to H-SS include percent pollen abundances. L= Lachuá lowlands, H=Purulhá highlands. Plant habit codes: T=Tree, S=Shrub, H=Herb. Vegetation belt codes (see Table 2.1): LRF= Lowland Rain Forest, LMRF= Lower Montane Rain Forest, MCF= Montane Cloud Forest, MMF= Montane Mixed Forest, SAF= Sub-Alpine Forest, *= Undefined vegetation belt, ?= Unassigned origin.

Pollen taxa

Genus Family Vegetation belt

Habit

DS

Biogeo

L-BP

L-SS

H-BP

H-SS

Acacia Fabaceae Lowlands T-S Z AMZ 2.70 1.61 Alchornea Euphorbiaceae Lowlands T W AMZ 0.8-8.5 1.6-2.6 0.5-2.1 Anthurium Araceae Lowlands H Z AN 10.80 Araliaceae Lowlands S Z AN 0.9 1.60 Arecaceae Lowlands S Z AMZ 1.1-10.9 0.9-3.22 Bignoniaceae Lowlands T Z AMZ 0.8-1.8 Bombacaceae Lowlands T Z AMZ 0.7-1.6 Boraginaceae Lowlands S Z LAU 0.80 Brosimum Moraceae Lowlands T Z AMZ 5.2-20 1.6-4.2 0.6-3.6 Celtis Ulmaceae Lowlands T Z LAU 1.5-20.1 1.6-14.4 0.5-3.1 0.90 Combretaceae/Melastomataceae Lowlands T Z AMZ/AN 0.8-9.1 0.8-1.6 Euphorbiaceae Lowlands T A AMZ 0.9-1.6 1.2-2.6 Fabaceae Lowlands T Z AMZ 1.8-8.1 0.80 Mimosa Fabaceae Lowlands T Z AMZ 0.60 Malpighiaceae Lowlands T-S Z AMZ 1.42 Moraceae Lowlands T W AMZ 1.8-12.1 0.90 0.7-1.1

Myrsinaceae Lowlands T W AN 0.5-6.6

Page 81: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

69

Table 3.1 continued.

Pollen taxa

Genus Family Vegetation belt

Habit

DS

Biogeo

L-BP

L-SS

H-BP

H-SS

Myrtaceae Lowlands T Z AN 1.10 4.80 Pachira Bombacaceae Lowlands T Z AMZ 0.8-1.5 Piper Piperaceae Lowlands S Z AMZ 0.9-4.2 0.8-5.3 Psychotria Rubiaceae Lowlands S Z AN 1.1-4.1 4.8-16.7 Rubiaceae Lowlands S Z AN 0.8-6 0.80 1.60 Salvia Lamiaceae Lowlands T-S Z LAU 2.70 Sapium Euphorbiaceae Lowlands T Z AMZ 0.8-0.9 6.50 Sapotaceae Lowlands T Z AMZ 0.8-16.4 0.6-1.7 Solanaceae Lowlands T-S Z AN 0.8-24.1 0.6-9.7 Trema Ulmaceae Lowlands T Z LAU 0.9-3.6 15.80 Ulmaceae Lowlands T W LAU 0.9 Verbenaceae Lowlands H Z ? 0.8-2.3 1.3-6.5 Spondias Anacardiaceae LRF T Z AMZ 0.7-2.3 6.10 Bursera Burseraceae LRF-LMRF T Z AMZ 0.9-3.3 11.40 0.9-1.3 Cecropia Cecropiaceae LRF-LMRF T W AMZ 0.9-8.1 2.50 0.5-56.4 Terminalia Combretaceae LRF-LMRF T Z AMZ 2.7-16.4 0.6-9.7 0.5-1.5 0.70

Page 82: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

70

Table 3.1. continued. Pollen taxa Genus Family Vegetation belt H DS Biogeo L-BP L-SS H-BP H-SS

Inga Fabaceae LMRF T Z AMZ 0.7-3.4 2.60 Hedyosmum Chloranthaceae LMRF-MCF T W LAU 0.8-1.8 0.80 3.3-31.3 1.2-5.9 Myrica Myricaceae LMRF-MCF S W LAU 0.8-4.1 2.6-42 0.9-10 1.9-6.1

Quercus Fagaceae LMRF-MCF T W LAU 0.7-8.1 0.8-2.6 10.3-60.9 3.1-16.7

Abies Pinaceae MMF T W LAU 0.8-1.5 3.4-9.6 0.7-7 1.2-8.5 Alnus Betulaceae MMF-SAF T W LAU 1.7-9.9 0.6-4.2 0.8-2.2 1.2-1.3 Ericaceae Highlands S W AN 1.7-4.2 Conifer6 Pinaceae Highlands T W LAU 0.8 Pinales Pinaceae Highlands T W LAU 0.7-3.6 0.7 Pinus Pinaceae Highlands T W LAU 9-35.1 1.6-46.6 2.2-16.6 3.7-14.8 Urticaceae Highlands H W AN 2.20

Page 83: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

71

Table 3.1 continued.

Pollen taxa

Genus Family

Vegetation belt

H

DS

Biogeo

L-BP

L-SS

H-BP

H-SS

Alternanthera Amaranthaceae * H Z ? 0.8-3.2 4.90 Amaranthaceae/Chenopodiaceae * H Z ? 0.9-1.7 0.80 0.5-0.9 0.6-1.3 Asteraceae * H Z AN 1.6-6.9 0.6-4.8 0.8-20.3 11.8-15.9

Cyperaceae * H AMZ 1.4-7.6 0.8-4.8 0.6-5.1 70.1-140.3

Peperomia Piperaceae * H Z AMZ 0.9-3.3 1.7-1.8 Piperaceae * S Z AMZ 0.6-2.6 0.7-0.9 Poaceae * H W ? 1.6-6.5 0.8-1.6 0.9-41.6 14.7-65.6 Polygonum Polygonaceae * H Z ? 1.4-24.5 Zea Poaceae * H W AMZ 0.8-0.9 1.70 0.5-0.9 0.7-1.9

Trilete spores * H ? 56.1-96.3

18.9-89.2

80.9-94.9 14.5-45.5

Monolete spores * H ? 3.7-43.9 10.8-81.1 5.1-19 54.5-85.5

Page 84: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

72

CaCbCcCeCdRaRbRcReRd

Sa2L1L2L3P1P4T1J1N1N2N3N4N5N6N7N8N9

N10

BP Lachua

SS Lachua

SS Purulha

BP Purulha

0

Alchorn

ea

0 20

Brosim

um

0 20

Bursera

0 20 40 60

Cecrop

ia

0 20

Celtis

0 20

Morac

eae

0

Rubiac

eae

0 20

Termina

lia

0 20 40

Hedyo

smum

0 20

Ilex

0 20 40

Myrica

0 20 40 60

Quercu

s

0

Abies

0

Alnus

0 20 40

Pinus

12 36 60 84 108

AP

-0.2 0.2 0.5 0.8 1.1

FA1

Tropical trees and shrubs Temperate trees and shrubs

Figure 3.2. Pollen diagram from Lachuá and Purulhá based on bryophyte polster (BP) and surface sediment (SS) samples. Ca to Rd Lachuá BP, and Sa2 to L3 Lachuá SS. P1 to J1 Purulhá SS, and N1 to N10 Purulhá BP. + = rare taxa appearing at <1%. AP=Arboreal pollen, FA1=Factor Analysis first component.

Percent pollen abundance

Page 85: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

73

CaCbCcCeCdRaRbRcReRd

Sa2L1L2L3P1P4T1J1N1N2N3N4N5N6N7N8N9

N10

BP Lachua

SS Lachua

SS Purulha

BP Purulha

0

Alterna

nthera

0

Amaranth

acea

e / C

heno

podia

ceae

0 20

Asterac

eae

0 20 40 60

Poace

ae

0

Zea

0 20 40 60

Trilete

0 20 40

Monole

te

0 30 60 90 120 150

Cypera

ceae

0 120 240 360 480 600

Pollen

conc

entra

tion (

x100

0 grai

ns/cm

3)

0 30 60 90

NAP

-0.2 0.2 0.5 0.8 1.1

FA1

Herbs Pteridophytes

Figure 3.2. Continued. NAP=Non-arboreal pollen, FA1=Factor Analysis first component.

Percent pollen/spore abundance

Page 86: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

74

CaCbCcCeCdRaRbRcReRd

Sa2L1L2L3

BP

SS

0

Acacia

0

Alchorn

ea

0

Araliac

eae

0 20

Arecac

eae

0

Bignon

iacea

e

0

Bomba

cace

ae

0

Boragin

acea

e

0 20

Brosim

um

0 20

Bursera

0

Cecrop

ia

0 20

Celtis

0

Combre

tacea

e / M

elasto

matace

ae

0

Fabac

eae

0

Inga

0 20 40

Solana

ceae

0

Malpigh

iacea

e

0

Mimos

a

0 20

Morace

ae

0

Myrtac

eae

0

Pachir

a

0 20

Psych

otria

0

Rubiac

eae 1

0 20

Rubiac

eae 2

0

Salvia

0Sap

ium

0 20

Sapota

ceae

0

Spond

ias

0 20

Termina

lia

0 20

Trema

0

Conife

r 6

0

Hedyo

smum

0

Ericac

eae

0 20

Ilex

0 20 40

Myrica

0

Quercu

s

0

Abies

0

Alnus

0 20 40

Pinus

0 20 40 60 80 100

AP

-1.0 0.0 1.0 2.0 3.0

DCA 1

0.0 1.0 2.0 3.0

DCA1 (-P

inus)

Tropical trees and shrubs Temperate trees and shrubs

Figure 3.3. Lachuá pollen diagram based on bryophyte polster (BP) and surface sediment (SS) samples. + = rare taxa appearing at <1%.

Percent abundance

Page 87: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

75

Ca

Cb

Cc

Ce

Cd

Ra

Rb

Rc

Re

Rd

Sa2

L1

L2

L3

BP

SS

0

Alterna

nthera

0

Ama/Che

n

0 20

Anthuri

um

0

Asterac

eae

0

Pepero

mia

0

Piper

0

Poace

ae

0

Verben

acea

e0

Zea0

Cypera

ceae

0 20 40

Trilete

0 20 40

Monole

te

0 16 32 48 64 80

Pollen

conc

entra

tion (

x100

0 grai

ns/cm

3)

0 4 8 12 16 20

NAP

Herbs Pteridophytes

Figure 3.3. Continued. AP=Arboreal pollen, NAP=Non-arboreal pollen, DCA1= Detrended Correspondence Analysis first axis.

Percent abundance

Page 88: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

76

Samples AP NAP Local Highlands Z W

Ca 81 19 78 22 72 28Cb 100 0 84 16 78 22Cc 88 12 86 14 84 16Cd 96 4 72 28 68 32Ce 86 14 75 25 59 41Ra 94 6 77 23 68 32Rb 93 7 42 58 30 70Rc 90 10 59 41 50 50Rd 90 10 74 26 67 33Re 89 11 52 48 30 70

Sa2 92 8 85 15 66 34L1 91 9 38 62 34 66L2 98 2 56 44 56 44L3 87 13 74 26 70 30

Sample AP NAP Local Lowlands Highlands A W Z

N1 79 21 83 10 7 4 90 7N2 65 35 82 5 12 1 95 4N3 86 14 67 25 7 18 77 5N4 88 12 92 5 3 1 98 1N5 77 23 75 22 4 15 84 1N6 49 51 78 14 8 11 84 6N7 62 38 71 19 10 8 86 5N8 69 31 82 14 3 2 66 32N9 83 17 89 5 6 2 95 3N10 94 6 90 8 2 0 94 6

P1 26 74 55 15 30 10 85 5P4 41 59 86 5 10 0 95 5T1 18 82 79 7 14 7 93 0J1 38 62 72 6 22 4 94 2

Table 3.2. Lachuá bryophyte polsters (Ca-Re) and surface sediments (Sa2-L3) samples. Local and highlands (regional and extra-regional) refers to percentages of arboreal pollen (AP) spectra. NAP= Non-arboreal pollen, Z= Zoophilous, W=Anemophilous.

Table 3.3. Purulhá bryophyte polsters (N1-N10) and surface sediments (P1-J1) samples. Local, lowlands, and highlands refers to percentages of arboreal pollen (AP) spectra. NAP= Non-arboreal pollen, A=Ambophilous, Z= Zoophilous, W=Anemophilous.

Page 89: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

77

Figure 3.4. Lachuá DCA Q-mode ordination diagrams of AP data with Pinus removal. (+) represent bryophyte polsters, and diamonds surface sediments.

Page 90: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

78

N1

N2

N3

N4

N5

N6

N7

N8

N9

N10

P1

P4

T1

J1

BP

SS

0 20

Eupho

rbiac

eae

0 20 40

Hedyo

smum

0 20

Ilex

0

Myrica

0 20 40 60

Quercu

s

0

Urticac

eae

0 20

Pinus

0

Pinales

0

Abies

0

Alnus

0Alch

ornea

0

Brosim

um

0

Bursera

0 20 40 60

Cecrop

ia

0

Celtis

0

Morace

ae

0

Myrsina

ceae

0

Rubiac

eae

0

Termina

lia

0

Ulmac

eae

16 32 48 64 80 96

AP

0 100 200 300

DCA1

Temperate trees and shrubs Tropical trees and shrubs

Figure 3.5. Purulhá pollen diagram based on bryophyte polster (BP) and surface sediment (SS) samples. + = rare taxa appearing at <1%.

Percent abundance

Page 91: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

79

N1

N2

N3

N4

N5

N6

N7

N8

N9

N10

P1

P4

T1

J1

BP

SS

0

Amaranth

acea

e / C

heno

podia

ceae

0 20

Asterac

eae

0

Piperac

eae?

0 20 40 60

Poace

ae

0

Zea

0 20

Polygo

num

0 30 60 90 120 150Cyp

erace

ae

0 20 40 60

Trilete

0 20 40

Monole

te

0 120 240 360 480 600

Pollen

conc

entra

tion (

x100

0 grai

ns/cm

3)

0 30 60 90

NAP

Herbs Aquatics Pteridophytes

Figure 3.5. Continued. AP=Arboreal pollen, NAP=Non-arboreal pollen, DCA1= Detrended Correspondence Analysis first axis.

Percent abundance

Page 92: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

80

Figure 3.6. Purulhá DCA Q-mode ordination diagrams based on AP and NAP data. (+) represent bryophyte polsters, and diamonds surface sediments

Page 93: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

81

and highlands pollen along DCA Axis 1.

At Lachuá, temperate Alnus and Pinus have high abundances in both pollen reservoirs.

Abies is similarly represented in surface sediments from Lachuá and Purulhá, and in

bryophyte polsters at Purulhá (Table 3.1). Highland taxa, Quercus and Hedyosmum, are

abundant in both bryophyte polsters and surface sediments at Purulhá. Pollen taxon

Myrica has similar abundances in Lachuá and Purulhá (both polsters and surface

sediments), with the exception of L2 where Myrica reaches its highest representation.

Cecropia plant stands are found in both Lachuá and Purulhá (i.e. indicator species for

disturbance-edge effects) and Cecropia pollen is over-represented in one sample of

bryophyte polsters at Purulhá.

Asteraceae and Poaceae taxa have their highest abundances at Purulhá. Poaceae pollen is

more abundant in surface sediments than in bryophyte polsters. At both Lachuá and

Purulhá, Amaranthaceae/Chenopodiaceae and Zea are rare (Table 3.1). Alternanthera has

similarly low abundances in Lachuá and Purulhá surface sediments. Trilete spores are

generally more abundant in bryophyte polsters of Lachuá and Purulhá like in surface

sediment samples of Lachuá. In contrast, monolete spores have their highest abundance

in Purulhá surface sediments. Aquatics (i.e. Cyperaceae) are likely over-represented

relative to the pollen types included in the pollen sum in surface sediments from Purulhá.

Results of factor analysis indicate pollen types that explain the maximum amount of

variance along positive and negative trends of the ordination gradients (Table 3.4). For

Page 94: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

82

Lachuá, the variance in positive trend is explained by Pinus and Myrica; with the

negative variance trend explained by Solanaceae and Sapotaceae. The trend represented

by these groups of pollen taxa, correspond most likely the pollen source area, the former

highlands and the latter lowlands. Once Pinus (over-represented) is removed, the

following other pollen types explain the variation: (1) Celtis, Brosimum, Terminalia and

Sapotaceae along a positive trend, and (2) Ilex and Trema explains the negative trend. For

Purulhá, the main explanatory taxa are the following: Quercus and Ilex in a positive

trend, and Euphorbiaceae and Abies in a negative trend. These two trends in Purulhá may

indicate small scale gradients, because apparently they do not reveal an environmental

gradient between forest interior (e.g. bryophyte polster) and open landscape (i.e. surface

sediment). The main taxa for both Lachuá and Purulhá correspond to the lowlands to

highlands environmental gradient: Quercus, Hedyosmum, Asteraceae, and Cecropia

along a positive trend, and Celtis and Brosimum along a negative trend.

Page 95: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

83

Figure 3.7. Las Verapaces DCA Q-mode diagram AP shared data. Lachuá surface sediments are indicated by circles, and bryophyte polster samples are enclosed by the continuous line polygon. Purulhá surface sediments are indicated by discontinuous line squares, and bryophyte polster samples are enclosed by the discontinuous line polygon. See tables 3.2. and 3.3 for codes.

Page 96: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

84

Table 3.4. Factor Analysis scores for pollen types with highest amount of variance. + or – indicates direction of magnitude along factors. Location Taxon Factor 1 Factor 2 Factor 3

Brosimum

+2.4

Celtis +4.6

Lachuá

Terminalia +1.6 Sapotaceae +1.3 Solanaceae -5.7 Poaceae +6.9 Asteraceae +1.6

Quercus +4.3 Ilex +1.0

Purulhá

Euphorbiaceae -0.6 Abies -0.4 Myrica -2 Pinus -3.3 Cecropia +4.4

Quercus +5.4 Hedyosmum +4.1 Cecropia +1.1

Las Verapaces

Asteraceae +1.4 Pinus +6.0 Celtis +2.1 Brosimum +1.9 Ilex -1.0 Poaceae +6.9 Asteraceae +1.6 Hedyosmum -0.7

Page 97: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

85

3.4 Discussion  

3.4.1 Relevance of dispersal syndrome for pollen assemblages in bryophyte

polsters and surface sediment

Taxonomic content, which in turn is tightly correlated with dispersal syndrome and

pollen source area, contributes in differentiating bryophyte polsters and surface sediments

at both Lachuá and Purulhá (Tables 3.2 and 3.3). Because local plant taxa in Purulhá are

mostly temperate and therefore anemophilous (i.e. rich pollen producers), AP spectra are

almost entirely local. In an analysis of modern pollen rain where moss polsters were

collected in the same sites where vegetation cover was recorded using a Braun-Blanquet

scale in the Guatemalan western highlands (2800-3800 masl) and Volcanic Chain (3000-

4000 masl) (Islebe and Hooghiemstra, 1995), pollen spectra showed widespread over-

representation of anemophilous taxa (e.g. Pinus, Alnus, and Quercus). The explanation is

because at these elevations, anemophily is the dominant pollen dispersal syndrome.

Lachuá's AP spectra are more representative of local tropical provinces that contain more

zoophilous plant taxa (i.e. poor pollen producers). Islebe and Hooghiemstra (1995) found

that in spite that zoophilous pollen taxon poor abundance could under-estimate the local

abundance of the plant, their presence correspond well with the elevational vegetation

zone associated with the plant itself (e.g. Buddleja pollen found only in the subalpine

forest belt at 3400-4000 masl). This is probably because short-distance pollen dispersal

results in more accurate representation of local vegetation. Pollen taxa from highlands

still contribute in a major way to Lachuá pollen spectra, because highlands pollen is

largely more adapted for airbone dispersion than lowlands pollen. On the other hand,

Page 98: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

86

representation of lowlands pollen taxa in highlands is minimal because of the poor

abilities of lowlands pollen for long-distance dispersal.

Regional analysis of Lachuá data indicates that bryophyte polsters and sediment samples,

L1 (close to lakeshore) and Sa2 (small basin), are similar in that they share at least some

percentage of local pollen (Figure 3.4). In general, pollen trapped in bryophyte polsters

travels shorter distances from within the surrounding forest (Fægri and Iversen, 1989),

and to a much lesser extent, traps pollen that is airborne (i.e. transported great distances)

or is the consequence of wind friction created by forest canopy gaps. Sediment sample

Sa2, was collected from a small basin surrounded by high canopy forest; therefore its

similarity in pollen signal to bryophyte polsters is not surprising because the surrounding

high canopy forest likely acted as a barrier to long-distance dispersal. Despite this, Sa2

shows partial separation from bryophyte polsters when the full AP spectrum is analyzed

(Figure 3.4). Lachuá vegetation analysis shows that the "landscape unit" (i.e.

homogeneous biological and geomorphological area) where Sa2 is located (Salinas de los

Nueve Cerros) is different due to its unusual hilly topography within the generally flat

landscape of Lachuá. Landscape topography has been shown to influence composition of

vegetation communities (Cajas, 2009) and therefore to influence pollen source.

Even though surface sediments samples were not collected from the exact center of the

Lachuá Lake (i.e. ideal sampling location), their long-distance dispersed AP content is

sufficient to produce a typical highland signal (15-62%). Local analysis of Lachuá shows

a clear separation of both pollen reservoirs once the over-represented Pinus is removed

Page 99: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

87

(Figure 3.4). As with many regions world-wide, Pinus is notorious for overshadowing

signals from local pollen taxa (for discussion about Pinus pollen see section 3.4.5).

Because Purulhá’s surface sediments were collected in mid-sized basins located in

deforested landscapes, pollen spectra would be expected to have a major regional pollen

content (Sugita et al., 1999). Nevertheless, because zoophilous pollen dispersal syndrome

dominates, Purulhá’s regional component from the lowland is minimal (zoophilous; 5-

15%) and mostly local (anemophilous; 55-86%) (Table 3.2). The higher content of

lowlands pollen in bryophyte polsters in comparison to surface sediments could be

explained in terms of differential preservation. Forest interior conditions where bryophyte

polsters are found allow for better preservation of pollen (i.e. less dessication under a

canopy cover) (Vermoere et al., 2000). Surface sediments from the small basin Chichoj

Lake and Cahabón river floodplain correspond to lentic (still water ecosystem) and lotic

(flowing water ecosystem) environments, respectively (Brown et al., 2007), yet

surprisingly their pollen spectra shows a degree of similarity (Figure 3.6). The similarity

in pollen collection in lentic and lotic environments is likely due to the energy

environments in which the sediments were deposited (i.e. both are low energy

floodplains). In addition, their location within a deforested landscape results in overall

low AP values.

3.4.2 Influence of land-use change on pollen source

Deforestation rates are currently high in Purulhá highlands, thus the expected high AP

sediment signal in basins and lakes surrounded by forest is not possible to be assessed

Page 100: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

88

(e.g. no forest surrounds any possible sediment reservoir candidate for paleoecological

research). Despite this limitation, forested conditions should be identified if pollen

spectra from sedimentary records are similar to modern pollen spectra from bryophyte

polsters (i.e. high AP values). Other studies in tropical highlands with other types of

reservoirs have differentiated between forest, grasslands, and open spaces based on

different pollen spectra and associated forest taxa contribution (Kennedy et al., 2005;

Olivera et al., 2009).

From Quintana Roo in Mexico, Islebe et al. (2001) analyzed pollen rain from moss

polsters along a disturbance gradient which included lowland forest, disturbed forest, and

secondary vegetation. The pollen data provided a clear signal for the three vegetation

types because they cover large areas in the region, and a list of 15 indicator taxa was

selected based on their overall good representation in the pollen spectra. In contrast,

bryophyte polsters and surface sediments from Lachuá reflect the local forested condition

because of their high AP values (14,500 ha of forest on Lachuá Lake National Park),

which is similar to other lowland pollen analyses in the tropics where forest cover

conditions are similar (Behling and Negrelle, 2006; Batthacharya et al. 2011). In contrast,

Dominguez-Vásquez et al. (2004) found a dominant allochtonous (i.e. anemophilous)

signal in the Lacandon Lowland Rain Forest in Chiapas, Mexico. A greater allochtonous

contribution in the Lacandon lowlands pollen spectra may be a response to higher

deforestation rates, because as openness increases in landscapes long-distance dispersed

pollen input increases (Lynch, 1996; Gaillard et al., 2008; Hellman et al., 2009), which in

the case of lowlands scenarios correspond to highlands anemophilous pollen taxa.

Page 101: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

89

3.4.3 Tropical and extra-tropical generalities

The present study and others (Bush and Rivera, 1998; Bush and Rivera, 1991) indicate

that it is not possible to generalize pollen differences in bryophyte polsters from surface

sediments in tropical regions like it is done for northern temperate and boreal latitudes

(Bush, 1995). In extra-tropical climates of the Northern Hemisphere, generalizations

concerning the primary presence of short-distance dispersed pollen in bryophyte polsters

and long-distance dispersed pollen in surface sediments are acceptable (Fægri and

Iversen, 1989). Tropical environments are completely different because dispersal

syndrome is more important than basin-size generalizations in long- and short-distance

dispersal. In tropical lowland environments, for example, short-distance dispersed pollen

refers mainly to zoophilous taxa, whereas long-distance dispersed pollen mainly to

anemophilous pollen. In contrast, in tropical highland environments short-distance

dispersed pollen (i.e. local input) will contain anemophilous pollen. Long-distance

dispersed pollen (i.e. regional input) in tropical highland environments will contain

pollen from both anemophilous pollen from higher elevations and lowland zoophilous

taxa (although not aerodynamically designed, some are transported by upslope winds).

Elevational gradients in the tropics do not conform to generalities made for northern

temperate regions (Janzen, 1967).

In three out of four existing modern pollen rain studies for the Maya region (Islebe and

Hooghiemstra, 1995; Islebe et al., 2001; Domínguez-Vázquez et al., 2004), pollen is

captured by moss polsters in different scenarios, lowland and highland settings, and all

contain an over-representation of long-distance dispersed taxa (anemophilous taxa) and

Page 102: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

90

an under-representation of short-distance dispersed taxa (mostly entomophilous taxa).

This in turn interfered with the overall ability to detect typical "regional" versus "local"

vegetation signals. Broadly-dispersed tropical and temperate anemophilous taxa in

tropical regions, such as the Amazon Forest, reflect a less heterogeneous and diverse

landscape. This relationship poses a limitation to differentiating ecosystem types in the

pollen record (Bush et al., 2001). To overcome difficulties associated with anemophilous

taxa, Gosling et al. (2009) have stressed that more attention should be placed on

identifying and differentiating pollen abundances and accumulation rates for ecosystems.

Nonetheless, to achieve this it requires extensive spatial and detailed temporal sampling.

3.4.4 Identifying indicator taxa and vegetation associations

The pollen collection sites represented changes along the elevational gradient from

Lachuá to Purulhá (Figure 3.7 and Table 3.4), as has been found in other pollen studies

in Latin America (Weng et al., 2004; Weng et al., 2007). For their study region in

Guatemala, Islebe and Hooghiemstra (1995) concluded that moisture gradients have an

important role in explaining variation in pollen assemblages. In an elevational gradient

(from 130 to 1191 masl) in the Chiapas Lacandon Forest in Mexico, modern pollen

spectra collected from moss polsters (Domínguez-Vázquez et al., 2004) indicate high

overlapping of lowland rain forest and lower montane rain forest vegetation zones, and

montane rain forest and pine-oak forest respectively. The majority of pollen types from

Lachuá and Purulhá are generalist since they represent mainly lowlands and highlands

vegetation, while few represent a specific vegetation belt (Table 3.1). This representation

pattern in the Las Verapaces region is similar to the analysis of Dominguez-Vásquez et

Page 103: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

91

al. (2004), which indicates that pollen taxa represent broader and less specific

environmental ranges, possibly in response to the taxonomic resolution of genus and

family. According to results from Chapter 2 (Table 2.2 and 2.3) a few pollen types can

be associated to specific vegetation belts, such as the case of Spondias for Lowland Rain

Forest, and Inga for Lower Montane Rain Forest, and although not part of the Las

Verapaces Region, Abies for Mixed Montane Rain Forest. In a smaller spatial scale,

Batthacharya et al. (2011) were able to differentiate lowlands ecosystem types (i.e.

upland, bajo, and riparian forests) in Northeast Belize based on changes in abundance of

pollen types in relationship to ecological preferences of the correspondent source plant

taxa.

Gradients interpreted from DCA Axes and factor analysis identify indicator taxa for

Lachuá and Purulhá. Pollen from Celtis is associated with Lowland Rainforest in both

pollen reservoirs and is known to indicate high canopy transition (medium- to large-size

trees) in an early- to mid-succession phase following human disturbance (Marchant et al.,

2002). In Guatemala, plant stands of C. trinervia have been recorded in Lowland Humid

Forest from the Petén (Standley and Steyermark, 1946), an area which receives less

precipitation than the Lachuá Lowland Rain Forest (Standley, 1958). Celtis has thus

adapted to drier conditions and therefore could be found from lowland vegetation up to

lower montane rain forest. Such a trend has already been reported in pollen studies from

Costa Rica (Islebe and Hooghiemstra, 1997). The large tree, Brosimum has similar

distribution preferences as Celtis plant stands because it too has been reported in Lowland

Page 104: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

92

Humid Forest (Cascante and Estrada, 2001) (see Table 2.1) and Lowlands Rain Forest

(Chapter 2), such as the lowland rain forests from Lachuá (Table 3.1).

Terminalia pollen is a generalist taxon characteristic of Lowland Rain Forest and Lower

Montane Rain Forest (Chapter 2). I also identify Sapotaceae pollen as representative of

Lachuá Lowland Rain Forest, an observation similarly made in other areas of Latin

America (Marchant et al., 2002). In my study, both Hedyosmum and Quercus indicate

Lower Montane Rain Forest and Montane Cloud Forest vegetation, a distribution also

found elsewhere in the tropics (Domínguez-Vázquez et al., 2004). At Purulhá, the values

of these two taxa are lower in surface sediments than in bryophyte polsters most likely

because of proximity of forest stands to polsters and because in surface sediments, values

are diluted by high abundance of non-arboreal pollen (NAP).

The abundance of Quercus in bryophyte polsters (10-61%) is higher than those recorded

by Islebe and Hooghiemstra (1995) (3-16%), whom sampled at higher elevations (above

3000 masl) where oak is naturally less abundant (i.e. too cold). Pollen abundances of

Hedyosmum found in bryophyte polsters in the present study (3-31%) are similar to

values documented by a study in southern Peru (15-65%), at elevations between 1600-

2000 masl (Weng et al., 2004). In contrast, Islebe and Hooghiemstra (1995) did not find

Hedyosmum pollen at higher elevations (> 3000 masl) in Guatemala. Islebe and

Hooghiemstra (1995) found relatively high concentrations of Abies and Alnus at

elevations higher than 3000 masl (30 and 40%, respectively) yet I report a maximum

abundance of 10% for both because of their relative absence in forests in my study region

Page 105: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

93

(Chapter 2). Ilex pollen appears to be behaving as an outlier, because it is unclear why

this zoophilous highland taxon is as abundant in Lachuá’s surface sediments as in

Purulhá’s bryophyte polsters, and yet is totally absent in Purulhá’s surface sediments.

Differential preservation could in part explain this seemingly odd distribution, but the

explanatory mechanism remains speculative. According to Behling et al. (1999), the

ecological significance of finding Ilex within a paleoecological context remains uncertain.

Analysis of pollen in bryophyte polsters and surface sediments aided in identifying

indicator and generalist plant taxa at Las Verapaces region, showing partially the results

from the inventory study of vegetation belts (Chapter 2). Lowland taxa such as Bursera,

Inga, Spondias and Trema, and highland taxon such as Myrica, are reported in current

forest inventories (Chapter 2) but they are not statistically relevant in the ordination

gradients formed for DCA and Factor Analysis based on pollen in this study.

Nevertheless, they are qualitatively key taxa whose importance remains in their

associated presence-absence in the pollen spectra. Weng et al. (2004) have suggested that

in order to maximize detecting environmental changes for tropical studies qualitative

information such as presence/absence data should become more prominent in

palynological studies.

3.4.5 Interpretation of Pinus pollen

Overrepresentation of Pinus taxon is discussed separately because its abundance has to be

read carefully due to its high dispersion ability (Bohrerova et al., 2009). The location of

the sampling point is therefore critical to understand what Pinus pollen percentages

Page 106: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

94

reflect. Pinus is more abundant in bryophyte polsters and surface sediments from Lachuá

than from Purulhá possibly because of a mixture of pollen sources from different

populations of Pinus species (Figure 3.2). Natural Pinus populations are established in

different highland regions around Lachuá, in the northwest at the Chiapas highlands of La

Selva Lacandona (Breedlove, 1981), in the west and southwest at the Sierra de los

Cuchumatanes (Islebe et al., 1995), and Pinus caribea populations north of Lachuá and to

the northeast in Belize (Bridgewater et al., 2006). Pinus pollen has been considered as an

indicator of highlands vegetation where abundances can reach up to 90-95% (above 2500

masl) which inform me more about larger scale scenarios (i.e. across different regions)

(Islebe and Hooghiemstra, 1995).

Nevertheless, as was found in Chapter 5, Pinus pollen percentages of the last ca. 2000

yrs in Lachuá lowlands have never been as abundant as modern pollen rain analysis

shows. High abundance of Pinus pollen in Lachuá may represent the general increasing

environmental deterioration of Mesoamerican forests in Southeast Mexico and

Guatemala, because it is known that Pinus is a successful colonizer of disturbed areas. A

related factor to be considered is the extensive Pinus plantations established in lowlands

and highlands in the recent years in Guatemala as part of governmental reforestation

programs (Gaillard, 2003). This probably created a bias in the modern pollen rain

(Behling and Negrelle, 2006), which urges the necessity to develop more modern pollen

rain studies in the region. A recent study by Battacharya (2011) in a pine savanna

(characterized by the presence of Pinus caribea and Quercus) shows similar Pinus pollen

Page 107: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

95

percentages to the ones found in Lachuá (up to ca. 40%), which possibly explains how

Pinus pollen could reflect local Pinus populations.

One factor that currently precludes geographical discrimination of pollen provenance of

Pinus populations is that highland Pinus species and P. caribea pollen cannot be

differentiated because they form part of the same subgenus (i.e. Diploxylon). More

palynological work with Pinus and other pollen types (e.g. Combretaceae and

Melastomataceae) is required to overcome these taxonomic limitations in order to

identify more accurately pollen source areas, geographical provenance, and ecological

preferences.

3.5. Chapter summary  

The purpose of modern pollen rain calibrations developed in this chapter is to understand

better the meaning of the pollen spectra of surface sediments, as they represent the best

analogue for sedimentary records. Calibrations of modern pollen rain of bryophyte

polsters and surface sediments from Lachuá lowlands and Purulhá highlands revealed the

importance of geographical context and related vegetation. Results from Chapter 2 were

the basis to determine pollen source areas of pollen types found in modern pollen

reservoirs, as Table 3.1 indicates. The pollen assemblage in Lachuá lowlands is

dominantly zoophilous because the associated vegetation is mainly of tropical

biogeographic origin (Amazonian and Andean). This is the reason why lowlands pollen is

poorly represented in the highlands pollen assemblage, because zoophilous pollen taxa

Page 108: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

96

disperse mostly over short distances. On the contrary, because the pollen assemblage in

Purulhá highlands is dominantly anemophilous (adapted for airborne dispersal) and of

temperate biogeographic origin (Laurasian), they are relatively more abundant in the

Lachuá lowlands pollen assemblage.

In general terms, surface sediments in Lachuá lowlands have similar pollen spectra than

the one found in bryophyte polsters, with the exception that the latter contained higher

abundances of forest interior taxa (e.g. Brosimum, Celtis, and Terminalia) and some

additional taxa (e.g. Bignoniaceae and Salvia). In the Lachuá lowlands, in both types of

depositional environments (polsters and surface sediments), high arboreal pollen content

was linked to the high remaining forest cover of the area (ca. 50%). In Purulhá

highlands, bryophyte polsters and surface sediments pollen spectra are different due to

the fact that the former were collected in forested conditions (e.g. high percentages of

Hedyosmum and Quercus), and the latter in a more open landscape. The non-arboreal

pollen content of bryophyte polsters and surface sediments reflected the degree of forest

cover where pollen reservoirs were collected, low for the former and high for the latter.

Combined analysis of pollen spectra of Lachuá and Purulhá showed the clearest

elevational differentiation when comparing bryophyte polsters, and lesser when

comparing surface sediments. The explanation for this pattern could be that surface

sediments from Lachuá lowlands have a significant representation of highland pollen,

which needs careful attention when interpreting fossil pollen spectra.

Page 109: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

97

The application of modern pollen rain calibration from bryophyte polsters and surface

sediments is developed in the following chapters on the paleoecology of Lachuá lowlands

and Purulhá highlands. Data matrices of modern pollen rain from both types of pollen

reservoirs are compared with fossil pollen assemblages from different levels in the cores

L-3 in Lachuá lowlands and P-4 in Purulhá highlands. These comparisons are the basis

for determining analog or non-analog environmental conditions along the temporal frame

covered in each core. Pollen types included in Table 4.4 were relevant in developing

paleoecological reconstructions in Chapter 4 and Chapter 5. The contributions of

Chapter 2 and Chapter 3 in understanding the importance of the relationships between

geographical context, vegetation biogeography, environmental conditions, and pollen

spectra resulted in better interpretations of the paleoecology of Lachuá lowlands and

Purulhá highlands.

Page 110: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

98

Chapter 4:

Late-Holocene History of a Highland Floodplain in Las Verapaces, Guatemala

4.1 Introduction  

Sedimentary records from lakes and peat deposits are one of the most relied upon tools

for paleoecological research, in part because aquatic environments are generally more

stable (i.e. continuous accumulation through time) and less disturbed than riverine

environments (Larsen and Macdonald, 1993; Birks, 2005; Brown et al., 2007). In

circumstances where lacustrine and peat deposits are rare, records from floodplains,

terraces and alluvial fans have been instead studied (Cheng, 2011; Gandouin et al., 2006;

Gandouin and Ponel, 2010). Paleoenvironmental information retrieved from river

floodplains can reflect in some cases floodplain communities, and to a lesser degree

upland vegetation communities (Solomon et al., 1982; Xia et al., 2002; Zazula et al.,

2006). Regional pollen however, is generally represented in floodplain sedimentary cores

(Qinghai et al., 1996). Floodplain sedimentary records offer a unique opportunity to study

the paleoecology of high energy systems (i.e. riparian plant communities) as well as their

successional dynamics related to flood events (Pokorny et al., 2000) and disturbances

such as fire (Gagnon, 2009).

It was the objective of my study to retrieve fossil pollen spectra from the headwaters of

the Cahabón River floodplain in the Las Verapaces highlands region of central

Guatemala (Figure 1.2 and 1.3). Due to the dominant karstic geology of the Las

Page 111: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

99

Verapaces region (Fourcade et al., 1999), aquatic ecosystems are mostly dominated by

sinkholes or “cenotes” where sediments are minimal to non-existent. Only a few small

lakes have been reported in the region as most have been either naturally drained through

karstic bedrock or anthropogenically disturbed (Castañeda, 1995). This study represents

the first paleoecological study (spanning the past ~2,400 years) of a region located in a

floodplain environment in highland Mesoamerica. Archaeological studies indicate the

existence of numerous small Maya centers in the Las Verapaces (i.e. Carcha, Sakajut,

Chican, Pasmolon) (Arnauld, 1978; Arnauld, 1987; Arnauld, 1997; Sharer and Sedat,

1987) that date from the Pre-Classic through to the Spanish Conquest.

The research goals of the following study are twofold: (1) to reconstruct the

paleoenvironmental history of a river floodplain located in the Las Verapaces highlands

for the past ~2400 years, with methods based on pollen and loss-on-ignition (LOI)

analysis, and (2) to make regional comparisons between other records in Mesoamerica,

with special emphasis on highland ecosystems. This research builds on previous highland

studies of vegetation biogeography (Chapter 2) and modern pollen calibrations

(Chapter 3).

4.2 Methods  

4.2.1 Core Sampling and Laboratory Work

With the use of a Livingstone Corer, I extracted a ~1.5 m core (labeled as P-4) on the

north side of the Cahabón River headwaters floodplain (Figure 1.3). The location of my

Page 112: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

100

core was cautiously selected in order to avoid any disturbance from modern hydrological

modifications on the Cahabón River floodplain, landslides on the slopes or incoming or

outflowing rivers. Coring was stopped once it became impossible to continue due to

stiffness of sediments.

The top centimeter of P-4 core was separated for modern pollen rain calibration (Chapter

3), and the rest was wrapped in aluminum foil and enclosed in a PVC pipe. A 1 ml sub-

sample was taken every 5 cm along P-4 core. Sub-samples were stored in Ziploc bags,

and the core’s stratigraphy was qualitatively described. After proper labeling, the core

and the subsamples were stored in a cold room.

According to the Department of Geography Protocol (No. 010) of the University of

Leicester, samples were pre-treated overnight with pyrophosphate, followed by standard

acetolysis procedure and heavy liquid separation with the use of bromoform (Fægri and

Iversen, 1989). Exotic Lycopodium spore tablets were added as markers to calculate

pollen concentration. Pollen counting was completed to 200 grains per sample when

possible (Lytle and Wahl, 2005). Pollen sum included arboreal and non-arboreal taxa that

were identified to family and genus level. Unknowns, spores and aquatics (Cyperaceae)

were not included in the pollen sum (Fægri and Iversen, 1989) and their abundance was

measured as a ratio in relationship to the total pollen sum per sample.

Arboreal pollen (AP) and non-arboreal pollen (NAP) percentages were calculated to

represent local landscape vegetation cover. The Loss of ignition (LOI) protocol used at

Page 113: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

101

the Paleoecology Laboratory of the University of Toronto (Heiri et al., 2001) was applied

for each subsample where pollen was analyzed (see above), to calculate the organic

(550°C), inorganic (950°C), and silicate (% left) contribution to the sediment sample

(estimated 2% error in the measurement). Only the LOI at 550°C is presented and

referred as "LOI". A bulk sample from the bottom level (145 cm in depth) of core P-4

(Purulhá) was radiocarbon dated to estimate the time span of the core, and three

additional samples (25, 50 and 70 cm in depth) were dated to develop models of sediment

accumulation rates. Dates were calibrated through the use of the IntCal04 curve from

CALIB 6.0 (Stuiver et al., 2005).

4.2.2 Core data analysis

Pollen counts were tabulated for pollen types and core levels (sub-samples), including for

the surficial level, for comparison with the modern pollen rain data presented in Chapter

3. Arboreal and non-arboreal pollen types were included in the pollen sum, excluding

aquatics (e.g. Cyperaceae), pteridophytes spores, and unknowns. For pollen

concentration, all counts were included. Principal component analysis (PCA) (Shi, 1993)

was performed with the statistical package PAST (Hammer et al., 2001) to describe

changes of pollen spectra along the core including common and rare taxa (Figure 4.3). A

second PCA was done (Figure 4.6) for the comparison of fossil pollen spectra of sampled

levels of the P-4 core and modern pollen spectra from bryophyte polsters and surface

sediments from Purulhá highlands (Chapter 3). The software C2 (Juggins, 2003) was

used to construct a stratigraphic diagram (based on depth measured in cm and calibrated

time scale) (Figure 4.1) according to the information on vegetation belts and pollen types

Page 114: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

102

presented in Table 3.1. Vegetation belts included in Figure 4.1 are Lower Montane Rain

Forest-Montane Cloud Forest and Mixed Montane Forest (see Table 2.1 and Table 3.1),

which comprised the sum of the percentages of Hedyosmum, Myrica and Quercus; and

Abies and Alnus, respectively. Complementary information for stratigraphic diagrams

(Figure 4.1 and 4.4), included relative abundance of arboreal (temperate trees and

shrubs) and non-arboreal pollen content, aquatics, pteridophytes spores, LOI (loss-on-

igition), sedimentation rate, and PCA axes scores. The stratigraphic diagram was divided

into a priori zones according to cultural periods defined for Mesoamerica (i.e. see

Introduction, section 1.7).

Equations for sedimentation rates were calculated based on sediment thickness (in cm)

per number of years between two identified dates. An analysis based on nine regional

studies from Mexico and Central America and our data spanning the Preclassic to

colonial times (Almeida et al., 2005; Carrillo-Bastos et al., 2010a; Conserva and Byrne,

2002; Dull et al., 2010; Figueroa-Rangel et al., 2008; Islebe and Hooghiemstra, 1997;

McNeil et al., 2010; Wahl et al., 2006) was performed to compare our calculated

sedimentation rates with values found in other Mesoamerican highland and lowland sites.

A total of 65 levels (radiocarbon dates) from the nine sites were included in my analysis,

which were allocated into elevation and cultural period categories. The sites used in this

regional comparison were placed into groups according to elevation: 0-500 m (n=20),

500-1000 m (n=16), 1000-2000 m (n=9), and 2000-3100 m (n=20); and cultural periods:

Preclassic (n=26), Classic (n=10), Postclassic (n=13), and Colony (n=3). Levels that

cover Pleistocene (n=4) and Archaic times (n=9) (5000-10000 yrs BP) were excluded

Page 115: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

103

from my analysis. Based on non-parametric boxplots (Tukey, 1977; Hyndman and Shang,

2010), P-4 sedimentation rate values were plotted in order to explore if they behaved as

outlier values for the groups were they belonged. The analysis was complemented with a

Kruskal-Walis test to test the coherence of created elevation and cultural period groups

(Kruskal and Wallis, 1952). P-4 sedimentation rate values are located in the 1000-2000 m

ranges, and across four cultural periods (Preclassic, Classic, Postclassic, European

conquest and Colonial Guatemala).

4.3 Results 

4.3.1 Stratigraphical description

The P-4 sediment core is characterized by an alternation of different tones of gray fine

grained sediment from the base at 144 cm up to 30 cm in depth (Table 4.1). Brown fine

grained sediment are found in between 30 and 5 cm depths, and dark brown organic

matter in the top 5 cm. LOI values steadily increase from 9 to 13% between the Late-

Preclassic and Late-Classic period (144-45 cm in depth) (Figure 4.1). At the time of the

Terminal Classic and onset of the Postclassic, there is a decrease in LOI to 9% (45-40 cm

core interval). A two-fold increase in LOI to 16% occurs during the next 300 years; LOI

values remain high (15-22 %,) during the Colonial to present-day cultural period.

4.3.2 Chronological control and sedimentation rates

The oldest age of 2390 yrs BP (all ages are reported as calibrated years before 1950 AD)

corresponds to sediments at the bottom of the core (144 cm in depth). Subsequent

Page 116: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

104

radiocarbon analyses indicate ages of 2060 yrs BP at 70 cm, 1510 yrs BP at 50 cm, and

150 yrs BP at 25 cm (Figure 4.2) (Table 4.2). There are two critical inflection points in

the age-depth curve, at 70 cm of depth, and above 25 cm of depth (Figure 4.2).

According to such a pattern, different periods of changing sedimentation rates (i.e.

changing of energy at time of deposition, sediment type, or source area) can be

postulated.

The lower phase from 2390 yrs BP until 2060 yrs BP (144-70 cm core interval) records a

relatively rapid sedimentation rate of 0.22 cm yr-1 (corresponding cultural period of the

Late Preclassic to Terminal Preclassic). The next phase, which includes dates from 2060

until 1510 yrs BP (70-50 cm core interval), has a marked reduction in sedimentation rate

to 0.036 cm yr-1 (corresponding cultural period of the late Preclassic to the middle

Classic). The following phase includes dates from 1510 until ~150 yrs BP (50-25 cm core

interval) and has an even more marked reduction in sedimentation rate to 0.018 cm yr-1

(corresponding cultural period of the middle Classic to end of the Postclassic, and the

start of Colonial Guatemala). The upper phase, from ~150 yrs BP to modern-day (25-0

cm core interval) shows a relative increase in sedimentation rate to 0.17 cm yr-1 (cultural

period of Colonial Guatemala).

The sedimentation rates found for P-4 core (0.22, 0.036, 0.018, and 0.17 cm yr-1) are not

outliers in the groups they belonged according to elevation and age (Figure 4.3). There is

no statistically significant difference in rate of sediment accumulation among the sites in

the regional comparison when they are divided by cultural period (Kruskal-Walis H=,

Page 117: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

105

p>0.01); and in relation to elevation, ranges 500-1000 and 1000-2000 m ranges are

considered a group, and 0-500 and 2000-3100 m another group (Kruskal-Walis H= 32.87,

p<0.01). The exploration of sedimentation patterns and differences in relationship to

environmental and cultural factors needs further discussion in the future, but is beyond

the scope of this thesis.

4.3.3 Description of pollen diagram

The variability found along the first factor of principal component analysis (PCA1) and

the first component of factor analysis (F1) (Figure 4.1), corresponds to the pollen zones

that were identified a priori based on cultural periods. According to trends in the PCA

Axis 1 analysis, the main variability in pollen assemblages occurs as a function of

changes in abundance of Asteraceae (Figure 4.4). PCA Axis 2 scores are linked to

variation in the dominances of Pinus and Alternanthera, Quercus, Poaceae and

Polygonum, alternating with dominance of Amaranthaceae/Chenopodiaceae (Figure 4.4).

The presented pollen diagram (Figure 4.5) is based on common and rare taxa identified

to at least to the level of genus or family. Tropical lowlands pollen taxa are extremely

rare and do not show a clear trend in the P-4 core. Pollen zones were closely related to the

four cultural periods identified to our region: Pre-Classic, Classic, Post-Classic, Colonial

and modern-day Guatemala. For the core interval between 144 and 90 cm (representing

the period from 2390 to 2150 yrs BP) there was no pollen or spore preservation.

Page 118: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

106

Table 4.1. P-4 core stratigraphic sequence.

Stratigraphy of core P-4 taken from the Cahabón Floodplain, Purulhá Depth (cm)

Description

0-5 5-30 30-32 32-39 39-44 44-49 49-86 86-93 93-105 105-111 111-119 119-144

Dark brown organic material Light brown organic material Dark gray fine grained sediment Medium gray fine grained sediment Bright gray fine grained sediment Medium gray fine grained sediment with oxide (red) spots Brown-greyish fine grained sediment Dark gray fine grained sediment Medium gray fine grained sediment Dark gray fine grained sediment Bright gray fine grained sediment Bright grey-green fine grained sediment with black laminations

Table 4.2. AMS radiocarbon dates, calibrates age ranges from dated sediment from P-4 core of the Cahabón Floodplain, Purulhá highlands. Bold numbers in brackets are the calibrated dates. All radiocarbon dated material is from bulk samples.

Depth (cm)

Lab No

13C/12C

14C yrs BP

Age range 2σ and Median Age (in

brackets) (cal yrs BP)

25

Beta-281243

-24.6

150±40

42-(151)-284

50

Beta-281244

24.3

1510±40

1313-(1394)-1517

70

Beta-281245

-24.3

2060±40

1926-(2029)-2133

144

GrA-40112

---

2390±35

2341 –(2422)- 2683

Page 119: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

107

Zone 1: Pre-Classic (2390-1650 yrs BP)

Pollen is preserved starting at 2150 yrs BP (90 cm). This zone is characterized by a

dominance of non-arboreal pollen (78-90%) consisting mostly of Asteraceae (48-76%);

with Asteraceae remaining at high values (55%) until the end of the Postclassic. Other

non-arboreal taxa include Poaceae (1-17%), Amaranthaceae/Chenopodiaceae (4-21%)

and Zea (1-3%) which remain at relatively low to medium values throughout the zone.

Cyperaceae (27-80%), Polygonum (0-4%), trilete spores (6-23%) and Monolete spores

(0-14%) all reach their minimum values in this zone. Pinus is present at intermediate

values (3-11%), similar to Hedyosmum (0-6%) and Quercus (1-8%), with Myrica

reaching its maximum abundance (0-5%) in the entire core. Regional taxa for the

modern-day Cahabón floodplain, Abies and Alnus, are not present at this time.

Zone 2: Classic (1650-1240 yrs BP)

This zone has relatively decreasing pollen abundances of Pinus (4-6%), Hedyosmum (1-

3%), Myrica (0-3%) and Quercus (0-4%). Alnus (1%) and Malphigiaceae (2%) make

their first appearance along the core. Non-arboreal pollen dominates this zone (87-89%),

once again consisting mainly of Asteraceae (46-71%). In comparison to Zone 1, Poaceae

(6-29%), Polygonum (0-9%), Cyperaceae (60-165%), and trilete spores (9-48%) increase

in abundance, while Amaranthaceae/Chenopodiaceae (4-9%) decrease. Zea (0-4%) and

monoletes spores (4-17%) have similar values across Zone 1 and Zone 2.

Page 120: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

108

Zone 3: Post-Classic (1240-420 yrs BP)

Total arboreal pollen shows a subtle decrease to between 5 and 13%; comprising mainly

Pinus (2%), Hedyosmum (2-6%), Quercus (1-2%), Alnus (0-1%) and Malphigiaceae

(1%). Myrica is absent. Similar to Zone 2, Asteraceae is the dominant taxon (55-56%)

showing only small variations in concentration between the zones. Other taxa that show

small decreases in concentration include Poaceae (4-16%), whereas Cyperaceae (19-

20%) decreases the most substantially. Amaranthaceae/Chenopodiaceae (1-25%), trilete

spores (22-112%) and monolete spores (47-97%) all show an increase in abundance.

Polygonum (3-5%) and Zea (3-5%) abundance remains the same across the previous zone

to Zone 3.

Zone 4: Colonial to modern-day Guatemala (420 yrs BP-Present)

This zone is characterized by a co-dominance of Pinus and Asteraceae with maximum

values at 30 and 35%, respectively. It is only at the most recent time that Pinus shows a

modest decrease (to 12%). In general, Quercus shows stability with low values (~1%)

along most of the core. Whereas Pinus shows decrease in the uppermost sample, Quercus

shows an abrupt increase (18%). Abies (modern-day regional species; Chapter 4) appears

for the first time in this zone, with a trend towards increasing abundance through time

(i.e. from 0 to 4% by the top of the core). The observed three-fold increase in arboreal

pollen (from between 10-22% to between 32-49%) is due to increased presence of Pinus,

Quercus and Abies. Hedyosmum shows relatively stable presence (2-7%) at the start of

the zone, and then begins to decrease (2%) towards the modern-day time. Myrica (0-3%)

Page 121: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

109

and Alnus (0-1%) show a similar pattern where their populations are relatively stable in

the lower sections of the core, then begin to decrease towards the upper part of the core.

By the onset of Colonial Guatemala, Asteraceae begins to show a sharp drop in

dominance (17-36%) showing even a lesser value at the present (13%) (Chapter 3). Zone

4 shows Cyperaceae increasing to its maximum value (from 49 to 226%), followed by a

sharp decrease to (58%) in the upper-most sample. In contrast, Polygonum shows a

steady and gradual increase (2-17%) until the modern day (26%). Poaceae shows an

increase in abudance to its peak (4-26%) during this zone, but right at present day, the

abundance of Poaceae returns to values more characteristic of previous zones (16%).

Amaranthaceae/Chenopodiaceae shows small decreases (3-13%) towards present day, a

pattern also observed for trilete spores (10-36%) and monolete spores (10-32%). At the

start of Zone 4, Zea abundance begins to decrease (0-1%) but then begins to recover (2%)

near present day. The most significant pollen signal in this zone is the first appearance of

the disturbance indicator, Alternanthera (1 - 11%). According to the PCA ordination

(Figure 4.4), the major division in the core in terms of the pollen assemblages is between

the Guatemala zone and pre-Guatemala zones.

Page 122: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

110

0

10

20

30

40

50

60

70

80

90

100

110

120

130

140

Depth

(cm)

Guatemala

Postclassic

Classic

Preclassic

0 20 40

Pinus

0 20

LMRF-M

CF

0

Mixed M

ontan

e Fore

st

0 20 40 60 80 100

Herbs

0 10 20 30 40 50

AP

0 4 8 12 16

Pollen

conc

entrati

on (x

1000 g

rains

/cm3)

0 8 16 24

LOI 5

50

0.0 0.1 0.2 0.3

Sed. r

ate (c

m/yr)

-80 0 80 160

PCA 1

100

350600850

110013501600

1850

2100

2350

Cal yrs

BP

Figure 4.1. P-4 core paleoecological diagram taken from Cahabón River Flooplain. AP = Arboreal pollen; LOI550 = Loss on ignition at 550°C (expressed as % of dry mass); PCA1 = Axis 1 scores from principal components analysis. Sedimentation rate is shown in cm/yr. Mayan cultural periods are shown. The Guatemala zone represents modern day Republic of Guatemala. Vegetation belts are composed of pollen indicator taxa (see Table 3.1 and section 2.4.3 for calculations). LMRF= Lower Montane Rain Forest, MCF= Montane Cloud Forest.

Percent abundance

Page 123: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

111

Depth (cm)

0 20 40 60 80 100 120 140

Cal

enda

r age

(cal

yr B

P)

0

500

1000

1500

2000

2500

Figure 4.2. Graph showing depth (cm) vs. calendar age (cal yrs BP) of sediments from core P-4 taken from the Cahabón River floodplain, as determined by 4 radiocarbon dates (black diamond symbols, see Table 4.2). Asymmetric bars indicate ±2σ in cal. yrs.

Page 124: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

112

Figure 4.3. Sedimentation rate (cm/yr) values from elevation ranges (masl) across Mesoamerica, as determined from a regional comparison from nine published studies (65 radiocarbon samples). Small squares= outliers.

Page 125: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

113

Modern

G5

G4

G3

G2

G1

PT3

PT2PT1

C4

C3C2

C1PC6

PC5

PC4PC3

PC2

PC1

-32 -16 16

PCA1

-20

-10

10

PC

A2

Figure 4.4. Principal Component Analysis (PCA) of sampled levels from core P-4. PCA1= First principal component, PCA2= second principal component. PC= preclassic levels, C=classic levels, PT=postclassic levels, G= Guatemala zone, modern= surface sediment (0-1 cm) for core P-4.

Page 126: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

114

0

10

20

30

40

50

60

70

80

90

100

110

120

130

140

Depth

(cm)

Guatemala

Postclassic

Classic

Preclassic

Arecac

eae

Hedyo

smum

Ilex

Malphig

iacea

e

Myrica

20

Quercu

s

20 40

Pinus

Abies

Alnus

20 40 60 80

Asterac

eae

20

Amaranth

acea

e /Che

nopo

diace

ae20

Poace

ae

20Alte

rnanth

eraZea Alch

ornea

Brosim

umBurs

eraCelt

isCom

bretac

eae /

Mela

stomata

ceae

Myrtac

eae

0 20

Polygo

num

0 50 100 150 200 250

Cyper

acea

e

0 30 60 90 120

Trilete

0 24 48 72

Monole

te

0 20 40

AP

100

350600850

110013501600

1850

2100

2350

Cal yrs

BP

Temperate trees and shrubs Herbs Tropical trees and shrubs Aquatics Pteridophytes

Figure 4.5. Pollen percentage diagram of P-4 core from the Cahabón River floodplain. Rare taxa appearing at <1% are indicated by a "+" symbol. .

Percent abundance

Page 127: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

115

4.4 Discussion  

4.4.1 Ancient lacustrine-like conditions at the floodplain?

The development of the Cahabón River Floodplain at 2390 yr BP is characterized by a

relatively higher sedimentation rate than in the present, which based on other studies in

similar types of settings, could mean a higher energy regime related to fluvio-lacustrine

(river to lake transition) environments (Figure 4.1). A range of 5-9% for LOI during this

time period (2390 until 2170 yrs BP) at P-4 core is similar to what was recorded in the

environments of a paleolake that once existed in the basin of Bogota from the late-

Pliocene through the Pleistocene (Torres et al., 2005). LOI values around 10% found in

lacustrine conditions indicate the provenance of the sediments mainly from swampy

environments. It is possible that core P-4 is located at what was once a swamp (i.e. light

gray fine grained sediments) in the remnants of a lacustrine environment (Shuman, 2003).

The location of core P-4 corresponds to the headwaters of the Cahabón River watershed,

at the top of a plateau where the drainage divide is located (Figure 4.6). The relatively

flat terrain and the location of the floodplain are factors that may help explain the

existence of a paleolake, but nevertheless particle size analysis is needed to permit more

conclusive inferences. The fine grained sediments suggest a location with low energy

and not close to a main channel where sediments are in general coarse-grained (Bridge,

2003). In the Bogota Basin swamp environments with fine grained sediment in a fluvio-

lacustrine hydrological context with LOI values from around 10% (Torres et al., 2005),

are interpreted as episodes of high rates of bioturbation due to high levels of biological

activity in organic-rich mud. This in turn may help to explain the absence of pollen

grains observed in our P-4 core during this time (144 to 90 cm) as a result of bioturbation

Page 128: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

116

disturbance. Further exploration about the possible existence of formerly paleolacustrine

conditions in the floodplain is needed (e.g. particle size analysis), because our speculation

is greatly based only on the Torres et al. (2005) study.

An alternative explanation for the absence of palynomorphs in the lower section of the

core could be that this was a period of rapid sedimentation due to increased erosion or

flows in the river, either due to natural (e.g. higher precipitation due to a wetter climate,

or any other related hydrological changes) or cultural factors (e.g. increased erosion due

to deforestation), preventing in the end the accummulation and deposition of pollen and

spores into the sediments. The absence of palynomorphs in the lower section of the core

makes it difficult to infer what the vegetation cover could have been during this time

period (2390 until 2150 yrs BP). Although differences exist, sedimentation rates from the

bottom (Preclassic period) and the top (Guatemala zone) belong to higher ranges (0.22

and 0.17 cm yr-1), which suggests that sedimentary conditions may have been relatively

similar. In this area today, economic human activities have resulted in a high

deforestation rate leaving most of the floodplain valley floor and slopes with scarce

vegetation cover (MAGA, 2006). By analogy, the same low vegetation cover could be

inferred for the beginning of our core; however, there is insufficient evidence to indicate

whether or not the cause of low vegetation cover was due to natural or cultural

circumstances. Nevertheless, a cultural cause may help explain the high sedimentation

rate caused from slope erosion into the floodplain (Thieme, 2001; Charlton, 2008). The

first Preclassic agriculturalists in Mesoamerica have been associated with evidence of the

highest rates of soil erosion and degradation in the region, mainly due to the

Page 129: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

117

VP

CH

CX

SL

P‐4

Figure 4.6. Cahabón River Floodplain. The core P-4 is located in the headwaters of the Cahabón River (running eastwards). Rivers are irregular black thick lines. Dark grey polygons represent natural reserves (mainly Montane Cloud Forest). Archaeological sites are indicated by circles: VP= Valparaiso, CH= Chican, CX= Cerro Xucaneb, SL=Sulin. See Figure 1.4 for elevation references.

Page 130: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

118

trial and error of people learning the consequences of land-use change (Beach et al.,

2006).

The lower-most pollen sample in the P-4 core (at 90 cm) shows a low percentage of

arboreal pollen and high values of herbs. The bottom of P-4 core could be indicating

intense land use change and conversion to agricultural uses (e.g. Zea and Cyperaceae).

The inferred lacustrine swampy environment was probably disturbed at this time, as a

slight increase in LOI (9 to 13%) possibly represent land use modifications in the

floodplain. The fact that pollen appears simultaneously strengthens the possibility that

LOI increase is due to the floodplain stabilization.

4.4.2 Evolution of Mayan land management at the Cahabón Floodplain

The accompanying changes in fluvial parameters may have been influenced by

(culturally-induced) hydraulic management of the floodplain (i.e. leading to the presence

of fine-grained sediments that are darkin in colour and higher in organic matter) (Table

4.1, see interval 49-86 cm). Approximately 340 years after the first appearance of pollen

in P-4 core (2040 yrs BP), the sedimentation rate decreases several-fold (from 0.22 to

0.036 cm yr-1), possibly indicating that early Mayans evolved along this time some form

of soil conservation practice that helped to decrease rates of soil erosion. It is possible

then that progressive agriculturally-related fallow debris may have increasingly

influenced the slight increase of LOI from 9 to 13% during those three centuries. Changes

in PCA1 values suggest that vegetation dynamics reflect approximately the conditions in

the riparian zone before the sedimentation rate decreased, in between the end of the

Page 131: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

119

Preclassic and start of the Classic (Figure 4.1). For many years, archaeologists have

shown evidence that the early Mayans incorporated terraces and raised fields in their

agricultural management plans, ultimately leading to reduced soil erosion over the time

the region was in agricultural use (Beach, 2003; Beach et al., 2009). Raised fields were

useful in modulating soil-water conditions, where channels in between the field

functioned to regulate water tables resulting in permanent flooded conditions (e.g. rise in

water table) (Turner and Harrison, 1981; Scarborough, 1991; Beach et al., 2011). The

sedimentation rate in core P-4 remained relatively low for roughly 1600 years (e.g. from

0.036 to 0.018 cm yr-1), until the end of the Postclassic period, when presumably

dispersion of Mayan populations resulted in abandonment of established agricultural

plots.

Arboreal components (Hedyosmum, Myrica, and Quercus) from Lower Montane Rain

Forest (i.e. located most likely along the valley bottom) and Montane Cloud Forest (i.e.

located most likely along the valley slopes) may be an indication of recovery from

agricultural disturbance, since their presence in the floodplain environment increases

towards the late-Preclassic (i.e. rising 3 to 11%). The first appearance of Zea pollen

during the late-Preclassic with values lower than 1% supports the development of

agriculture in the floodplain, because generally Zea pollen disperses close to its source

(McNeil et al., 2010). Progressive increment in Zea pollen percentages greater than 1%

suggest possibly that agriculture was developing in a wider area along the floodplain and

more intensively, as has been reported in other Mesoamerican locations (Wahl et al.,

2007). Currently the major land use at the floodplain is cattle pasture with some

Page 132: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

120

agricultural plots (personal observation). This is reflected in modern pollen rain (Chapter

3) where values of Zea pollen are higher in surface sediments (>1%) than in bryophyte

polsters (<1%) (Figure 3.5), because the latter were collected in forest interiors far away

from the floodplain (ca. 5 km). PCA ordination of modern and fossil AP pollen spectra

shows how bryophyte polsters do not overlap with sediments over PCA1, while modern

surface sediments overlap with Preclassic, Classic, and Postclassic sediments because

possibly the landscape was similar in openness (Figure 4.7). Analysis of multiple cores

along the study site (e.g. Cahabón Catchment) will in the future allow a more complete

view of the evolution of the floodplain.

Present day AP values range from 16 to 32% which corresponds to the current deforested

and open landscape at the Cahabón River Floodplain, but with scattered forest remnants

in the valley slopes (Figure 3.5 and Figure 4.8). In comparison the 11% arboreal content

at the late-Preclassic could be signalling even less presence of continuous forest cover,

and more open shrubland near the floodplain, most likely on valley slope environments.

Pollen abundance below 5% for Myrica throughout the core supports this hypothesis (van

der Hammen and Hooghiemstra, 2003). Approximately 90 ky yrs BP, higher abundance

pollen values (around 30%) at Lake Fuquene in Colombia indicate the presence of

Myrica shrub forest surrounding the lake. In the Las Verapaces region today at

elevational ranges from ~1400-2000 masl, Myrica pollen is an indicator of open

landscapes and humid grounds (Marchant et al., 2002) (i.e. 2-6% in surface sediments,

see Chapter 3) which are similar to the values observed during the Preclassic (i.e. 1-4%).

In west-central Mexico today, montane forest taxon Quercus has remained non-dominant

Page 133: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

121

in the pollen record, even when it has maintained a stable presence during fully-forested

conditions (Figueroa et al. 2008), but in this case, the presence of Myrica pollen

throughout the core supports a more open structure vegetation.

The existence of an open environment during the late-Preclassic is highly supported by

the presence of Asteraceae pollen, a disturbance-related family of vegetation. Asteraceae

pollen dominates the floodplain until the end of the Postclassic when agriculture ceased.

Other pollen taxa that indicate localized disturbance include Amaranthaceae

/Chenopodiaceae and Poaceae. The latter pollen types have opposing patterns of

maximum values (r = -0.52 p = 0.02), possibly related to different phases of post-

agricultural vegetation succession. Present day Poaceae pollen has greater values in the

floodplain (16 to 66%) than in the past (3 to 28%), evidence for the current major land

use as pasture lands (e.g. grasses). Lower present day values of Asteraceae (<16%),

support the idea that land use was different in the past (e.g. agriculture) (Figure 3.5 and

4.7). The floodplain could have exhibited high water table levels until the middle-Classic

(supported by a change in PCA1 values in Figure 4.5), since low values of Cyperaceae

indicates flooded environments (e.g. deep waters). Although Cyperaceae represents

azonal vegetation of aquatic environments, it has been found that its preferred

establishment conditions are from shallow waters at shore locations in lakes and

floodplains, but not deep waters (Van’t Veer and Hooghiemstra, 2000). The increase of

water table and periodic floodings in the floodplain could be the result of Mayan

hydraulic management that included the production of canals, water reservoirs, and raised

fields for agriculture (Beach, 2003). A rise in water table may have precluded the

Page 134: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

122

successful colonization of other aquatic plants that prefer shallower shore conditions such

as Polygoynum and members of the Apiaceae, Juncaceae and Typhaceae families (Berrio

et al., 2002). Another explanation for the decrease in Cyperaceae involves removal

related to anthropogenic use and management (Macia and Balslev, 2000). Ordinations

including NAP modern data show that similarly high percentages of Cyperaceae are

found both during the late-Classic and modern times, possibly because both time periods

are characterized by minimal minimum hydraulic or agricultural management, and

therefore flooding events (Figure 4.7). Archaeological work in the floodplain is needed

in order to be more conclusive about the existence of such agricultural structures.

Analysis of quantity and size of macroscopic and microscopic charcoal throughout the

core will complement hypotheses on the use of the site for Mayan agriculture.

Pinus values from the Preclassic are lower than at present, suggesting the presence of an

agricultural regime during the Preclassic period on the floodplain. Pinus is generally one

of the first trees to colonize open areas and is traditionally appreciated as a pioneer

species in vegetation succession (Conserva and Byrne, 2002). The fact that Pinus does

not ever seem to increase in abundance could be evidence that Pinus was under Mayan

management (i.e. fuelwood and ceremonial uses), which prevented its colonization in

agricultural fields and environs (e.g. valley slopes). During the Preclassic, Classic, and

Postclassic, the floodplain generally remains an agricultural center, characterized by

surrounding open landscapes with isolated shrubby patches of Myrica interspersed

throughout the environment. Some minor changes in arboreal content, however, may be

signalling small regional changes in landscape structure.

Page 135: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

123

Figure 4.7. Principal Component Analysis (PCA) of modern pollen rain samples from Purulhá highlands and sampled levels from core P-4. Triangles represent bryophyte polster modern samples, diamonds are modern surface sediments, and (+) are sedimentary records.

Page 136: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

124

Arboreal pollen values gradually increase through to the late-Preclassic (from 12 to 18%),

probably due to a decrease in agricultural activity associated with the Mayan Preclassic

collapse (Arnauld, 1987; Dahlin et al., 1987). According to modern pollen calibration

(Chapter 3), Poaceae pollen may be highly represented (i.e. less than 20%) in the

background signal from continuous forest because it is always found in low values.

Cyperaceae and Polygonum (notably successful in shallow water shore environments)

increase at the transition, likely as a result of lowering water-table levels on the

floodplain caused by temporary abandonment of strict water management practices (PCA

1 values slightly increase). It has been observed in Pacific Ocean coastal pollen records in

Guatemala that as mangroves decrease due to environmental changes (sea level drop

5500 yrs BP), aquatic plants such as Cyperaceae increase (Neff et al., 2006). In the

floodplain Asteraceae values decrease as the temporary and partial halt to large-scale

agriculture allows colonization from other herb species. Amaranthaceae-Chenopodiaceae

pollen increase at this time indicating onset of secondary succession on some but not all

plots as Zea pollen remains present at this time. From the onset of the Classic until the

end of the Postclassic period, arboreal pollen values once again decrease, likely due to a

reestablishment of large-scale agricultural practices in the floodplain.

Page 137: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

125

Figure 4.8. Cahabón River and its floodplain. Series of 3 arrows indicate river flow direction. N= North, masl=meters above sea level. Taken by J.C. Berrio © 2006.

4.4.3 Classic-Postclassic transition and its effects on floodplain management

Based on the apparent absence of Zea pollen in the P-4 core at the transition between the

Classic and Postclassic period, it is likely that large-scale agriculture is temporarily

locally abandoned (reflected as an abrupt change in PCA1 values). From the

anthropogenic point of view, absence of Zea pollen in Copán at the Classic-Postclassic

transition has been interpreted as a temporary abandonment of local agricultural land and

as temporary migration to nearby locations, and not necessarily a complete abandonment

(McNeil et al., 2010). However, this reconstruction has to be cautiously interpreted since

Page 138: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

126

the Classic-Postclassic transition samples have relatively low pollen concentration in P-4

core (Figure 4.1).

Recognizing that the Classical period in my study has few samples, careful interpretation

is needed when reconstructing the floodplain scenarios at this time. However, pollen

evidence indicates that possibly the start of the Classic (1780-1510 yrs BP) is

characterized by the stable abundance of agriculturally-related pollen (Zea). This possibly

determined the concomittant increase of Asteraceae that ecologically replaces (i.e.

successfully out-competes) Amaranthaceae /Chenopodiaceae. Towards the end of the

Classic (late-Classic to Terminal Classic, 1510-1240 yrs BP), Cyperaceae increases

markedly which suggests a possible abrupt drop in the water-table level of the floodplain

(i.e. because Cyperaceae is extremely successful in shallow water environments) due to

temporal abandonment of hydraulic management of agricultural terraces and raised fields.

This agricultural interruption is overlain by a background signal of decreasing arboreal

pollen. Geological and geomorphological analyses (e.g. particle size analysis) of the

floodplain will allow in the future a more complete reconstruction of the fluvial dynamics

at the headwaters of the Cahabón River.

An arid climate event associated with an increase in solar activity is identified around ca.

1200 yrs BP by Hodell (2007) using stable isotope and lithological evidence from Lake

Punta Laguna in the Yucatán Peninsula. Titanium evidence, used as a proxy for the

strength of the hydrological cycle, taken from the Cariaco Basin in Venezuela suggests

that this drying event was widespread as three centennial periods of reduced rainfall were

Page 139: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

127

reconstructed at that site for ~1190, 1140, and 1090 yrs BP (Haug et al., 2003). The P-4

core could be interpreted as indicating that agricultural practices were abandoned at the

floodplain possibly related to a drought related event. A decrease in LOI values from 12

to 9% at the Classic-Postclassic transition (60 cm depth) could suggest a slight decrease

in organic matter contribution to overall sedimentation. This decrease in LOI is supported

by increments in percentages of Cyperaceae and Poaceae (PCA1 values), as both expand

when water table lowers.

However, the occurrence of a drought is not likely for the Cahabón River Floodplain, as

this lowering in the water table could be due to cultural management (e.g. abandonment

of agricultural terraces) since the percentages of Cyperaceae and Poaceae at the Classic-

Postclassic transition are similar to present day when no drastic drought has been

registered (Figure 4.7). It is possible that agricultural practices were temporarily

transferred to a different location along the floodplain as a regular practice, as suggested

by pollen reconstructions by McNeil (2010) for Rio Amarillo in the Copán Valley

(Honduras). Analysis of stable oxygen isotopes is needed to permit more conclusive

inferences about drought occurrences in the Las Verapaces highlands.

Climate as a causal element for the Mayan collapse during the terminal Classic is still a

contentious issue (Aimers, 2007; Powell, 2008). Agriculture at the Cahabón floodplain

during the Classic-Postclassic transition may have been temporarily abandoned (i.e. as

Zea pollen absence suggests) but the reappearance of Zea pollen during the Postclassic

indicates that agriculture is likely re-established. The recovery of agricultural activities at

Page 140: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

128

the floodplain seems to contradict theories postulating a complete and widespread

collapse of Mayan societies across Mesoamerica (Arnauld, 1988; Borgstede and Mathieu,

2007). On the other hand, if the temporary disappearance of Zea represents a counting

artefact for P-4 core, continuous agricultural activity is supported during the Classic-

Postclassic transition. However, changes in other pollen taxa (e.g. Aquatics and

Asteraceae) during this transition support temporally local abandonment of the Cahabón

floodplain.

The hypothetical abandonment of the floodplain at this transitional period (Classic-

Postclassic) allowed a change in vegetation succession to take place, possibly due to a

halt in hydraulic management operations which led water levels to decrease abruptly. By

the time that agriculture is re-established at the Postclassic (indicated by the reappearance

of Zea pollen), Cyperaceae reaches the lowest values (from 138 to 29%) as an indication

of a higher water-table related to artificial flooding (e.g. deeper waters). The marked

decrease in sedimentation rate from 0.036 to 0.018 cm yr-1 during the Postclassic supports

the continuity of soil conservation practices in the Cahabón floodplain (Figure 4.1),

although possibly with less complex hydraulic management (i.e. less water volumes in

channels). A two-fold increase in LOI (9 to 16%) suggests secondary vegetation

succession taking place at the floodplain environs (i.e. increase of organic matter input),

as pollen from Asteraceae and Amaranthaceae/Chenopodiaceae increase temporarily at

this time. Cyperaceae is being replaced by trilete spores (increase from 42 to 107%) and

monolete spores (increase from 18 to 41%). Based on modern day calibration (Figure

3.5), the former may indicate for the Postclassic a temporary forest recovery (i.e.

Page 141: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

129

scattered forest remnants) because trilete spores may represent tree fern taxa found in

forested environments (e.g. tree ferns from Cyatheaceae family) (Van’t Veer and

Hooghiemstra, 2000), while the latter supports disturbance due to landscape management.

Archaeological evidence from the terminal Classic in the Maya highlands at Las

Verapaces indicates a vigorous creativity and imagination at the ceramics production

level (Arnauld, 1987), which possibly means that although activities at the urban centers

may have not halted nor declined, they may have temporarily at the agricultural centers,

such as the Cahabón floodplain.

4.4.4 European conquest and possible climatic variability

The end of the Postclassic is clearly characterized by a change in pollen assemblages in

the P-4 paleoecological record (over 270 yrs, during the period 420 to 150 yrs BP) (PCA1

values show a marked change). A critical increase in sedimentation rates (from 0.018 to

0.17 cm yr -1) may be indicating that soil conservation practices have been abandoned and

that agriculturally-related structures such as terraces have been removed. According to

historical documents from Catholic Dominic Missionaries (Godoy, 2006), approximately

eight cities (e.g. including Cobán and Purulhá) in the Las Verapaces region were located

along the Cahabón Floodplain, all established in a 30-year period from ca. AD 1544 to

1574. This 30-year period falls within the 400-year period characterized by the decrease

of Zea pollen as a result of the onset of the European Conquest (~400 yrs BP). Due to

decreasing water-table levels on the floodplain, Cyperaceae once again becomes a

dominant pollen type, trilete spores decrease possibly due to diminishing forest cover,

and monoletes spores maintained relativey high values due to high levels of disturbance.

Page 142: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

130

Asteraceae co-dominates with the disturbance related taxon Alternanthera (Marchant et

al., 2002), which appears for the first time along the core. Although not an exotic species

in Latin America (i.e. found in Amazonian pollen records dating to 8 ky BP), the

presence of Alternanthera is related to abrupt land-use change, abandonment of

traditional agricultural practices in the floodplain and the installment of different

European management regimes (i.e. Colonial period and modern times) (Behling et al.,

2001).

A critical change in land management at the time of the European conquest is supported

by other patterns observed in the P-4 core. Increases in Hedyosmum pollen may show

development (succession) of the lower montane and montane forest during the 16th

century, while Quercus pollen decreases temporarily (close to 1%). It is possible that

selective forest management is being practiced, because while Quercus is extracted for

timber and fuelwood (Ramírez-Marcial et al., 2001; Ramírez-Marcial, 2003),

Hedyosmum stands appears to have been left unaffected based on P-4 pollen record.

Pinus pollen increases several-fold during this time period and this could be explained by

the fact that Pinus is a pioneer in disturbed conditions (Richardson, 2000). Similarly,

based on pollen records, pioneering colonization by Pinus has been suggested during the

same period at Laguna Azteca in Central Mexico (Conserva and Byrne, 2002). On the

other hand, in a 4200-yr paleoecological study at Sierra Manantlan Biosphere Reserve in

the West-Central Mexican highlands, under low to null human disturbance, Pinus

colonization responded positively to intervals of aridity (Figueroa-Rangel et al., 2008).

During the 20th century until the present, Pinus pollen decreases (reflected in a change of

Page 143: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

131

PCA1 values) (Figure 4.1) probably due to extraction of pine trees for timber use, and

because of land use change in the floodplain to cattle pastures and agriculture, which are

now a more serious threat for forest species conservation than climate change (van

Zonneveld et al., 2009).

Indicator pollen type from higher elevation (mixed montane forest) Abies briefly appears

for the first time in the sedimentary record at the onset of the European conquest (422 yrs

BP), and remains present until the 20th century (Figure 4.1). Appearance of Abies in the

P-4 pollen record matches approximately solar minima events experienced at the time of

the Little Ice Age ca. 300-400 yrs BP (Helama et al., 2009). A Mesoamerican regional

drop in temperature, associated with drier conditions has been identified in sedimentary

records from the Yucatán lowlands (Hodell et al., 2005, 2007), and lakes Zempoala and

Quila in Central Mexican highlands (Almeida et al., 2005). On the other hand, in Lago

Verde, Los Tuxtlas in Mexico (Lozano-García et al., 2010) humid conditions 300-400 yrs

BP promoted an increase in abundance of pollen of upland vegetation. Core P-4 and other

highland pollen records may support differential effects of the Little Ice Age cooling

event, in general terms increasing aridity in the lowlands and increasing humidity in the

highlands. Temporary decreases of Abies and Myrica pollen (~300-100 yrs BP) are

possibly explained by increases in regional and local timber harvesting and land clearing

by Spanish colonizers, respectively (Islebe and Hooghiemstra, 1995; Andersen et al.,

2006). Land clearing at the Cahabón valley slopes, led possibly to Myrica’s niche

occupation by Pinus. Despite of the currently deforested landscape conditions (~60%

forest removal) at the Cahabón floodplain, there seems to be an increase at present in

Page 144: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

132

pollen indicative of the Montane Cloud Forest and Mixed Montane Forest vegetation

belts, possibly due to recent conservation efforts (CECON, 1999; Jolon-Morales, 2007).

 

4.5. Chapter Summary 

Changes in fossil pollen spectra in the P-4 core from the Cahabón River Floodplain

corresponded to transitions between Mayan cultural periods and the start of the European

conquest (ca. 350 yrs BP). The geographical location of the headwaters of the Cahabón

River, where the P-4 core was collected, fulfills geomorphological conditions for the

possible existence of a paleolake in the plateau where currently the floodplain exists. LOI

values (ca. 10%) and the non-preservation of pollen support the paleolake explanation.

Agricultural practices are inferred from the appearance of pollen, which is mainly non-

arboreal (e.g. Asteraceae and Zea), together with slight increases in LOI values (>10%)

and low values of Cyperaceae pollen (which prefers shallow water in shore

environments). Based on this evidence it is believed that Mayan agriculture was

developed during the transition from lacustrine to floodplain conditions. Sedimentation

rates on the floodplain decreased many-fold ca. 360 yrs after the first pollen appearance,

which is possibly evidence for the development of agricultural terraces that eventually

dimished and controlled locally soil erosion.

During ca. 2000 yrs of agriculture there is evidence of a temporal abandonment of

agriculture in the floodplain during the Classic-Postclassic transition, because temporarily

Zea pollen disappears, Cyperaceae and Poaceae pollen increase (e.g. less flooding of

Page 145: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

133

terraces), and LOI slightly decreases (12 to 9%). Effects of the occurrence of the known

Mesoamerican megadroughts are not obvious in the floodplain record presented here,

since the temporary increase of indicators of dryer conditions (e.g. Cyperaceae and

Poaceae) have values similar to current ones according to modern pollen rain calibrations

when no droughts are registered. The appearance of Abies pollen in sedimentary record at

the end of the Postclassic ca. 350 yrs BP could be an evidence of regional cooling

conditions related to the Little Ice Age cold event.

Agriculture re-establishment during the Postclassic (e.g. Cyperaceae and Poaceae pollen

decrease, and Zea pollen reappears) is completely halted by the European conquest, as

evidence of disturbances is registered in the sedimentary record, including: abrupt

increase of sedimentation rates (0.018 to 0.17 vm yr-1), appearance of disturbance related

pollen taxa (e.g. Alternanthera), increase of Cyperaceae pollen due to change in land use

(e.g. lower water table for non-agricultural land use, such as cattle pastures), and increase

of Pinus pollen; Pinus colonizes areas that have been cleared for timber extraction and

the development of European colonies. The major change in the sedimentary record is

related to environmental disturbances after the European conquest over ca. 300 yrs, never

seen before in 2000 yrs of history.

Page 146: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

134

Chapter 5:

The Lachuá Lowlands Rain Forest in Guatemala: 2,000 yrs of forested landscape?

5.1 Introduction  

Along the banks of the Chixoy River at the foothills of the Sierra Chama and south of the

Petén Lowlands, there was a city known as Salinas de los Nueve Cerros (Figure 1.2 and

1.3). It was located in a unique landscape, at the intersection between highlands and

lowlands, and was a city known to play an important role as salt producer since Preclassic

times. The city epicentre is located around the Tortugas and Nueve Cerros hills, where

salt extraction took place from an exposed salt dome (Andrews, 1983; Woodfill, 2012).

The city is located close to the floodplain of the Chixoy River, a likely location for

agricultural activity. Salinas de los Nueve Cerros probably reached its climax during the

late-Classic and was abandoned during the Postclassic; its abandonment was likely in

response to the reduced important of salt production during the Terminal Classic, when

major cities no longer required large quantities of salt (Arroyo, 1994). Perhaps due to its

reliance on salt as an economic resource, the surrounding forest was likely less disturbed

than cities that relied on large-scale agriculture. Scientific expedition reports dating to the

the 16th century, indicate that the area around Salinas de los Nueve Cerros was likely

covered in dense and continuous forest. By ca. 1950, the region still had a forest cover

close to 100%, although currently it has decreased to ~50% (Avendaño et al., 2007).

Page 147: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

135

Archaeological studies indicate that land use intensity (i.e. agriculture) increased towards

the epicentres of Mayan cities (Johnson et al., 2007), leaving outskirts with less urban

development, however in some locations urban and rural land use intensity was similar

(Beach et al., 2009). Agriculture was developed using a variety of strategies, such as

stone boxes, terraces, swidden techniques (i.e. slash and burn), including crop rotation

and forest management (e.g agroforestry) (Demarest, 2005). Agroforestry involved the

combination of agriculture and management (silviculture) of beneficial trees (e.g. food,

medicines, tools, construction, ceremonial), in a strategy that imitated forest structure and

distribution patterns, in the so-called Maya forest gardens (Ford and Nigh, 2009; Ross,

2011). Silviculture was probably more intense than agriculture in furthest points from city

epicentres as population density decreased.

The overall objective of this study is to examine a sediment record spanning 2000 years

from a wetland located next to Lake Lachuá, approximately 5 km southwest from Salinas

de los Nueve Cerros and within the Lake Lachuá National Park. This region is of interest

because it lies in a transition zone between the lowlands and highlands of Mayan

occupation and has not yet been studied. The Lake Lachuá National Park is currently the

last remnant of tropical rain forest in the Franja Transveral del Norte region. The past 60

years has seen much disturbance of natural forests due to (1) colonization of displaced

populations, (2) introduction of export cash crops, and (3) most recently oil exploration.

Since this region has a rich history of human-environment interactions spanning pre-

Hispanic to post-Colonial times, I expect that most changes to vegetation will have some

Page 148: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

136

relation to human land management and not to climate change. Calculations of the

approximate age of the existing remnants of tropical rain forest at Lake Lachuá National

Park should enrich discussions of forest conservation, in particular, the baselines

necessary for successful conservation of biological diversity.

5.2 Methods  

5.2.1 Core Sampling and Laboratory Work

With the use of a Livingstone Corer, I extracted a ~0.5 m core from the Lachuá Lowlands

in a wetland located at the northeast section of the Lachuá Lake (labeled as L-3), at

approximately 10 m from the shore (Figure 1.3). The location of L-3 core was cautiously

selected in order to avoid any disturbance from modern hydrological modifications,

landslides or incoming or outflowing rivers. Coring was stopped once it became

impossible to continue due to stiffness of sediments.

The top centimeter of each core was separated for modern pollen rain calibration

(Chapter 3), and the rest was wrapped in aluminum foil and enclosed in a PVC pipe. A 1

ml sub-sample was taken every 2.5 cm along L-3 core. Sub-samples were stored in

Ziploc bags, and the core’s stratigraphy was qualitatively described. After proper

labeling, the core and the subsamples were stored in a cold room.

According to the Department of Geography Protocol (No. 010) of the University of

Leicester, samples were pre-treated overnight with pyrophosphate, followed by standard

acetolysis procedure and heavy liquid separation with the use of bromoform (Fægri and

Page 149: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

137

Iversen, 1989). Exotic Lycopodium spore tablets were added as markers to calculate

pollen concentration. Pollen counting was completed to 200 grains per sample when

possible (Lytle and Wahl, 2005). Pollen sum included arboreal and non-arboreal taxa that

were identified to family and genus level. Unknowns, spores and aquatics (Cyperaceae)

were not included in the pollen sum (Fægri and Iversen, 1989) and their abundance was

measured as a ratio in relationship to the total pollen sum per sample.

Arboreal pollen (AP) and non-arboreal pollen (NAP) percentages were calculated to

represent local landscape vegetation cover. The Paleoecology Laboratory Loss of ignition

(LOI) protocol from the University of Toronto (Heiri et al., 2001) was applied for each

subsample where pollen was analyzed (see above), to calculate the organic (550°C),

inorganic (950°C), and silicate (% left) contribution to the sediment sample (estimated

2% error in the measurement). Only the LOI at 550°C is presented and referred as "LOI";

it is expressed as percent of dry mass In the L-3 core, a bulk sample from the lowest level

(47.5 cm in depth) and one additional sample (22.5 cm in depth) were radiocarbon dated

to develop models of sediment accumulation rates. Dates were calibrated through the use

of the IntCal04 curve from CALIB 6.0 (Stuiver et al., 2005).

5.2.2 Core data analysis

Pollen counts were tabulated for pollen types and core levels, including for the surficial

level, for comparison with the modern pollen rain data presented in Chapter 3. Arboreal

and non-arboreal pollen types were included in the pollen sum, excluding aquatics (e.g.

Cyperaceae), pteridophytes spores, and unknowns. For pollen concentration, all counts

Page 150: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

138

were included. Principal component analysis (PCA) was performed with the statistical

package PAST (Hammer et al., 2001) to describe changes in pollen spectra along the core

including common and rare taxa. The software C2 (Juggins, 2003) was used to construct

a stratigraphic diagram (based on depth measured in cm and calibrated time scale) based

on relative abundance of arboreal (temperate trees and shrubs) and non-arboreal pollen

types, aquatics, pteridophytes spores, LOI (loss-on-igition), sedimentation rate, and PCA

axes scores. The stratigraphic diagram was divided into zones according to cultural

periods defined for Mesoamerica (i.e. see Introduction, section 1.6).

An analysis based on nine regional studies and my data (Mexico and Central America)

spanning the Preclassic to colonial times (Almeida et al., 2005; Carrillo-Bastos et al.,

2010a; Conserva and Byrne, 2002; Dull et al., 2010; Figueroa-Rangel et al., 2008; Islebe

and Hooghiemstra, 1997; McNeil et al., 2010; Wahl et al., 2006) was performed in order

to compare my calculated sedimentation rates with values found in other Mesoamerican

highland and lowland scenarios. The sites used in this regional comparison were placed

into groups according to elevation (0-500 m, 500-1000 m, 1000-2000 m, and 2000-3100

m) and cultural periods (Preclassic, Classic, Postclassic, and Colony), to determine if the

sedimentation rates in core L-3 were expected or outlier values for the groups to which

they belonged (see section 2.4.3 for further explanation). The analysis was

complemented with a Kruskal-Walis test to test the coherence of created elevation and

cultural period groups. The sedimentation rate values for core L-3 are compared to other

sites in the same elevational range (0-500 m), and across four cultural periods.

Page 151: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

139

5.3 Results 

5.3.1 Stratigraphic description

The L-3 core is composed mostly of a homogeneous dark peat with some fine-grained

inorganic inclusions and scattered wood fragments, with no clear stratigraphic

differentiation. Wood fragments are identified along the core at different depths.

LOI values found at the bottom of the L-3 core are characteristic of wetland

environments (~ 80-90%), with variability in LOI observed along the core likely being

representative of human activity during known cultural periods of the Maya (Figure 5.1).

LOI between 47.5-45 cm core interval decreases gradually from 92 to 86%, during the

late-Preclassic period (i.e. core time interval 1835-1750 yrs BP). At the Preclassic-Classic

transition, LOI values decrease once again from 86-85% to 79% in approximately a span

of 90 years (time interval 1750-1580 yrs BP). LOI values remain relatively stable (75-

81%) for the remainder of the Classic period (1580- 1070 yrs BP). There is a notable

decrease in LOI values down to 68% shortly after the Classic-Postclassic transition (~90

yrs), but as it stands as a single point, it should be carefully interpreted. Decreasing LOI

indicates a relatively reduced contribution of organic matter into the sediment. During the

Postclassic and European Conquest-Colonization, LOI values remain stable around 82-

86%.

Page 152: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

140

Table 5.1. AMS radiocarbon dates, calibrates age ranges from dated sediment from P-4 core of the Cahabón Floodplain, Purulhá. Bold numbers in brackets are the calibrated dates. All radiocarbon dates material is from bulk samples.

Depth (cm)

Lab. No

13C/12C

14C yrs BP

Age range 2σ and

Median Age (cal yrs BP)

22.5

Beta-281242

-27.8

990±40

795-(902)-964

47.5

GrA-40111 --- 1835±30 1705-(1773)-1864

5.3.2 Chronological control and sedimentation rates

The oldest age of 1835 yrs BP corresponds to sediments found at the bottom of the L-3

core at a depth of 47.5 cm (Table 5.1). The second date taken from the core is 990 yrs BP

and is found at a depth of 22.5 cm. The depth versus age model is very close to linear (r2=

0.99) although significance of this relationship will be tested when more radiocarbon

dates are obtained. Based on the L-3 chronological model, a relatively slow accumulation

rate of 0.026 cm yr-1 is observed, a value similar to the slow sedimentation rates observed

in the P-4 core (0.018 and 0.036 cm yr-1) during the time of Mayan agricultural

management of the Cahabón floodplain (see Chapter 4). The accumulation rate found for

the L-3 core (0.026 cm yr-1) is not an outlier in the groups they belonged according to

elevation and age (Figure 4.3). For further information about regional Mesoamerican

analysis of sedimentation rate values, see section 4.3.2.

Page 153: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

141

0

5

10

15

20

25

30

35

40

45

Depth

(cm)

Guatemala

Postclassic

Classic

Preclassic

0

Asterac

eae

0

Poace

ae

0

Ama/Che

no

0

Zea

0 24 48 72

Monole

tes

0 16 32 48 64

Triletes

0

Cypera

ceae

0 20 40 60 80 100

LOI

0 50 100 150

Pollen

conc

entra

tion (

x100

0 grai

ns/cm

3)

-24 -12 0 12 24

PCA1

-12 -4 4 12 20

PCA2

200

400

600

800

1000

1200

1400

1600

1800

Cal yrs

BP

Herbs Pteridophytes

Figure 5.1. L-3 core paleoecological diagram taken from a wetland next to Lake Lachuá. AP = Arboreal pollen (%); LOI550 = Loss on ignition at 550°C; PCA1 and PCA2= Axis 1 and 2 scores from principal components analysis. Mayan cultural periods are shown. The Guatemala zone represents modern day Republic of Guatemala. Pollen concentration (x1000 grains/cm3).

Percent abundance

Page 154: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

142

5.3.3 Description of pollen diagram

The variability found along the first and second axes of my principal component analysis

(PCA1 and PCA2) is consistent with the a priori pollen zones based on Mayan cultural

periods (Figure 5.2). Based on PCA1, primary variability is explained by a change in

vegetation composition from Solanaceae to Combretaceae/Melastomataceae, where

Solanaceae dominates from the late Preclassic until the Classic-Postclassic transition, and

then is replaced by Combretaceae/Melastomataceae (which in turn dominates until 100

years ago). There are other secondary changes in vegetation succession observed in

PCA1 that are explained below (see pollen zones). PCA2 is governed mostly by

variability in regional and local pollen rain, more specifically, by Ilex and a group of

tropical pollen types (e.g. Terminalia, Sapium), respectively.

I present in the pollen diagram common and rare taxa identified at least to genus and

family (Figure 5.3). Zones were closely related to four cultural periods: Terminal Pre-

Classic, Classic, Post-Classic, and Colonial to modern-day Guatemala. The two bottom

levels represent the Terminal Preclassic, where the Preclassic-Classic transition had no

pollen content. The non-arboreal pollen contribution is relatively low overall in the L-3

core (<10%).

Page 155: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

143

Figure 5.2. Principal Component Analysis (PCA) of sampled levels from core L-3. PCA1= scores along first principal component axis, PCA2= scores along second principal component axis. PC= preclassic levels, C=classic levels, PT=postclassic levels, G= Guatemala zone, Modern= represents surface sediment (0-1 cm).

Page 156: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

144

Zone 1: Terminal Preclassic (1835-1750 yrs BP)

Information about vegetation from the Terminal Preclassic comes from the second last

level, since both the bottom and the Preclassic-Classic transition levels contain no

preserved pollen. Despite the fact that the bottom level at 1835 yrs BP has no preserved

pollen, the highest LOI value (92%) at L-3 core is observed at this level, and may be

related to a higher vegetation cover in the wetland (i.e. less opened-up landscape).

Possibly oxidation processes at both levels due to temporarily dessication (e.g. lower

water levels) resulted in null preservation of pollen.

Solanaceae dominates the pollen assemblage from this single sample dating to the

Preclassic (32%) to the late-Classic, with other taxa such as Alchornea and Spondias co-

dominating at ~5%. Other tropical taxa such as Psychotria, Combretaceae

/Melastomataceae, Bursera, Rubiaceae and Terminalia are present at lower values (1-

7%); while Arecaceae, Brosimum, Caesalpinaceae, Celtis, Malphigiaceae, Myrtaceae,

Pachira, Sapium, Sapotaceae, and Trema are absent when compared to modern samples.

At 1750 yrs BP (end of the Preclassic), the temperate taxon Myrica co-dominates with

Solanaceae at 26% and shows a positive correlation for the rest of the core. Regional taxa

Pinus and Quercus are present at intermediate values (6 and 4%, respectively), Alnus and

Hedyosmum at low values (1%), and Hyeronima and Abies are absent during this period.

In fact, Hyeronima, Abies and Alnus are regularly absent at locations along the core.

Although Ilex is at low values (2%) in this period, it will soon become an important taxon

for explaining variability along the core.

Page 157: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

145

0

5

10

15

20

25

30

35

40

45

Depth

(cm)

Guatemala

Postclassic

Classic

Preclassic

0 20 40

Solana

ceae

0 20 40

Comb/M

elas

0 20

Myrtac

eae

0

Alchorn

ea

0

Rubiac

eae

0

Spond

ias

0

Termina

lia

0

Sapium

0 20

Psych

otria

0

Arecac

eae

0

Malphig

iacea

e

0

Bursera

0

Sapota

ceae

0

Piper

0

Celtis

0

Brosim

um

0

Trema

0

Bomba

cace

ae0

Caesa

lpina

ceae

0 20Ile

x0

Hedyo

smum

0

Hyero

nima

0 20

Myrica

0

Quercu

s

0 20

Pinus

0

Abies

0

Alnus

0 20 40 60 80 100

AP

-24 -12 0 12 24

PCA1

-12 -4 4 12 20

PCA2

200

400

600

800

1000

1200

1400

1600

1800

Cal yrs

BP

Tropical trees and shrubs Temperate trees and shrubs

Figure 5.3. Pollen percentage diagram of L-3 core from a wetland next to Lake Lachuá. + = taxa appearing at <1%. Mayan cultural periods are shown. The Guatemala zone represents modern day Republic of Guatemala.

Percent abundance

Page 158: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

146

Non-arboreal pollen contribution is low (4%) and is characterized by Asteraceae and

Poaceae pollen. Cyperaceae is present at intermediate values (3%) and only increases

significantly at the end of the Classic and in modern times (8%). Monolete and trilete

spores show relatively intermediate values (34 and 23 %, respectively) during the late-

Preclassic.

Zone 2: Classic (1750-1070 yrs BP)

Although Solanaceae dominates during the Classic (32%), it shows a subtle decreasing

trend (down to 25%) up to the Terminal Classic (1070 yrs BP) when it is replaced by

Combretaceae/Melastomataceae. Combretaceae/Melastomataceae experiences a decrease

in abundance during the middle Classic (from 9 to 6%), but progressively increases

towards the Terminal Classic (18%). Some arboreal taxa decrease in abundance at the

Terminal Classic, behaving similarly to Solanaceae. These taxa include Alchornea (8-

5%), Psychotria (16-1%), and Rubiaceae (5-2%). Myrtaceae increases during the Classic

(15-4%) and abruptly decreases at the Classic-Postclassic transition or shortly thereafter.

The regional taxon Ilex (9-3%) follows a similar increasing and abruptly decreasing

trend.

Regional taxa, Myrica and Pinus, show a decreasing trend from the onset of the Classic

towards the Postclassic. Alnus appears during the early-Classic (1%) and slightly

increases at the Classic-Postclassic transition (~2%). Other arboreal taxa are scarcely

present, although Arecaceae, Bursera, Celtis, Malphigiaceae, Sapotaceae and Trema

appear in minor abundances during the Classic. Herb species (Asteraceae and Poaceae)

Page 159: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

147

and Cyperaceae are also scarce until the Classic where they appear in minor abundance.

Zea appears only once in a very low abundance (1%) during the early-Classic, and

reappears (~2%) continuously until the modern period (110 yrs BP). Abundances of

monolete spores are relatively low, while trilete spores have their maximum relative

values during this period. Both show a sharp peak at the Classic-Postclassic transition (as

they did during the late-Preclassic).

Zone 3: Postclassic (1070-440 yrs BP)

Solanaceae (~22%) and Combretaceae/Melastomataceae (~23%) co-dominate during the

early Postclassic, and then the former progressively decreases until the present.

Combretaceae/Melastomataceae has a bell shaped dominance curve during the

Postclassic, reaching a maximum (33%) at the middle-Postclassic (770 yrs BP).

Following Combretaceae/Melastomataceae decreases until the late-Postclassic (22%),

remaining nevertheless as dominant taxon during cultural period. Other arboreal tropical

taxa such as Myrtaceae, Alchornea, Rubiaceae, and Spondias are present at low and

intermediate values (2-14%) without showing a stable pattern. Some arboreal taxa,

Terminalia, Sapium, Malpighiaceae and Psychotria, are consistently present in the core

but at lower values (1-4%). The remaining arboreal taxa, such as Arecaceae,

Bombacaceae, Bursera, Celtis, Sapotaceae and Trema, are present in few levels at low

values (~1-3%) as they are during the Classic period. Brosimum pollen is present at low

values (~1%) for the only time in the L-3 core.

Page 160: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

148

Temperate taxa like Myrica, Quercus and Pinus are generally less abundant during the

Postclassic. In contrast, Ilex shows an abrupt increase during the early Postclassic

(represented by a change in PCA2 values), although overall it generally maintains

intermediate values throughout the core. Hyeronima reappears and increases during the

Postclassic (1-4%), whereas Alnus reappears and remains relatively stable (1-2%) during

the Postclassic. Herbs and Cyperaceae maintain a similar pattern from the Classic and are

not abundant (<4%). Monolete spores gradually increase during the Postclassic, while

trilete spores decrease markedly in abundance.

Zone 4: Colonial to Modern Guatemala (440 yrs BP-Present)

Solanaceae and Myrtaceae maintain intermediate values (7-12% and 4-13%, respectively)

in the last ca. 400 yrs. Combretaceae/Melastomataceae increases at the onset of the

Colonial period (from 22-32%) and then decreases towards the present (15-6%).

Alchornea, Rubiaceae, and Spondias show a decreasing tendency towards the present (ca.

7 to 1%), whereas Terminalia, Sapium and Psychotria show an increasing trend towards

the present (ca. 2 to 9%). Tropical arboreal taxa like Bursera, Sapotaceae, Celtis, and

Malpighiaceae are present at low values (<2%). Trema slightly increases its abundance

up to the present-day (1 to 2%).

Ilex, Pinus and Myrica progressively increase towards the present-day (~ 3 to 15%),

whereas Hyeronima stays stable until it disappears from the modern-day sediments. Alnus

is present only at low values (~1%) during the Colonial period and then disappears, and

Quercus slightly increases towards the present-day (1 to 3%). Asteraceae and Cyperaceae

Page 161: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

149

maintain low values at 2-5% and 1-8%, respectively. Poaceae begins at a low value and

decreases at the present. Zea appears for the second time ca. 110 yrs BP. Monolete and

trilete spores disappear at 220 yrs BP and then monolete spores show an abundance peak

110 yrs BP and trilete spores present their lowest relative values.

5.4 Discussion  

5.4.1 Role of cultural management and vegetation succession

Pollen information from our L-3 core is likely supporting the existence of Mayan forest

management practices close to Lachuá Lake at the outskirts of Salinas de los Nueve

Cerros (Figure 5.4). Pollen of arboreal taxa that have been identified as important tree

species to the Maya (e.g. Psychotria, Spondias, Terminalia) have been found in L-3 core,

in addition to high arboreal pollen percentages (as much as 80% or more abundance) that

reflect a densely-forested portion of the landscape. At the Classic-Postclassic transition,

vegetation succession likely reflected cultural activities. For example, dominance of

Solanaceae pollen in L-3 core from the Preclassic to the Classic indicates managed forest

because this pollen type is an indicator of shrubs and small trees taxa of secondary

succession (e.g. Cestrum, Lycianthes, Lycium) (Marchant et al., 2002). Some of these

taxa had economic use (e.g. Capsicum, Solanum). The eventual dominance of trilete

spores over monolete spores is possibly indicative of progressive successful forest

management since the former benefits from lower levels of disturbance, as modern pollen

calibration from Lake Lachuá National Park suggests that trilete spores are indicators of

closed canopy forest (Figure 3.4 and Figure 5.5). Monolete spores were found in lower

Page 162: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

150

Figure 5.4. Location of the ancient Mayan city of Salinas de los Nueve Cerros on the banks of the Chixoy River, Alta Verapaz, Guatemala. Blue shaded area represents current known maximum spatial extension of Salinas de los Nueve Cerros (Woodfill 2011). Gray polygon represents Lachuá Lake National Park. Irregular black lines represent rivers. Inverse dark triangle represents L-3 core location next to Lachuá Lake (light gray polygon). Figure modified from Dillon (1977) and Woodfill (2011).

Page 163: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

151

abundances in forest interior conditions according to modern pollen calibrations (Figure

3.4).

Salinas de los Nueve Cerros was a key salt producing center for many important Mayan

cities including the Petexbatún and Tikal kingdoms, and remained so until the end of the

Classic (Wright, 2005; Bachand, 2010). Decreasing Solanaceae pollen during the

Postclassic could be reflective of city abandonment (as reflected by a change of PCA1

values, Figure 5.1), where all major economic activities including cropping, selective

tree extraction and forestry ceased to occur due to a drop in the economic importance of

salt demand from larger cities facing major societal transformations (e.g. Tikal). The

process of abandonment at the Classic-Postclassic transition has been identified in

neighbouring Mayan cities of the Petexbatún Region, which is the location where it is

believed the Terminal Classic so called collapse started (Demarest, 2006).

Once forest management was possibly halted by the Classic-Postclassic transition,

vegetation succession was allowed to occur naturally, eventually resulting in dominance

of old-growth forest species such as Combretaceae/Melastomataceae (e.g. Combretum)

(Marchant et al., 2002) as seen in the L-3 core. Trends in pollen abundance within the

L-3 core support this hypothesis. Combretaceae/Melastomataceae pollen has also been

shown to increase at the onset of the Postclassic in other sites nearby, including Laguna

Naja (800 masl) located in the Mexican Lacandon Forest (Domínguez-Vázquez and

Islebe, 2008). The shift in Combretaceae/Melastomataceae dominance took

approximately 200 years after the onset of the Postclassic (PCA1 values notably reflect

Page 164: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

152

change), wherein populations remained dominant for the next ca. 600 years, only

decreasing first 400 yrs BP and then abruptly during the last 150 yrs BP.

Information from our L-3 core strongly supports the possible existence of Maya forest

gardens located at the outskirts of the city-centre, Salinas de los Nueve Cerros. I found

pollen from taxa known to be planted by the Maya, including Terminalia, Spondias,

Psychotria, Myrtaceae, Rubiaceae, Sapotaceae, Arecaceae; and used by the Maya for

fuelwood, construction material, medicines, food (i.e. fruit, nuts), as well as latex

extraction (Ross and Rangel, 2011). The L-3 core contains these taxa in high abundance

during the Classic period, the heyday of Maya civilization (Schele and Freidel, 1990;

Freidel et al., 1993). Mayan gardens were planted so that the overall structure mimicked

the horizontal and vertical dimensions of a natural forest (Ford and Nigh, 2009). When

agroforestry likely ceased during the Classic-Postclassic transition, changes are recorded

in the L-3 core among the preferred Mayan garden tree species; some of them declined

abruptly and others disappeared for a short period. Nevertheless most of forest garden

taxa observed in the L-3 core persisted through Postclassic and Colonial times. The

occurrence of a detectable vegetation structure reflecting the composition of Mayan

forest gardens has been found elsewhere in ancient Mesoamerican sites (Ford, 2008;

Ross, 2011).

Forestry management may have been successful in reducing soil erosion, as stable LOI

values in the L-3 core suggest a continuous forest cover close to the wetland. Other

lowland and highland Mesoamerican locations have generally had similar sedimentation

Page 165: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

153

rates as was found for L-3 core (Islebe and Hooghiemstra, 1997; Figueroa-Rangel et al.,

2008; Carrillo-Bastos et al., 2010), which is relatively slow (0.026 cm yr-1) and consistent

with forested conditions surrounding the wetland that prevent high inputs of eroded

materials into the peat. In tropical regions, LOI values above 75% (and even as low as

20%) have been traditionally described as containing wetland-marsh environments

(Berrio et al., 2002; Torres et al., 2005).

During the early-Postclassic when silvicultural practices were probably abandoned in

Salinas de los Nueve Cerros, LOI abruptly decreased from 78 to 68% indicate decreasing

organic matter contribution into the sediment load. Diminishing LOI values are probably

explained by decreasing organic matter contribution (i.e. increase of clastic material to

sediment load) (Shuman, 2003), due to deforestation as a consequence of temporary

forest gap openings. Similar patterns have been found in ancient peatlands in Georgia

near the Black Sea, where aeolian input increases in the sediment load due to more open

landscape conditions (de Klerk et al., 2009). The LOI decrease at this single level

(Classic-Postclassic transition) in L-3 core has been reported in a core from Laguna

Tamarindito in the Petexbatún region, with a drop from ca. 85 to 60% (Dunning et al.,

1998). Around the time of the city abandonment, changes in pollen abundances (from

disappearance to reduced values) of forest garden species in Core L-3 indicate a probable

abrupt increase in the extraction of valuable plant taxa (i.e. during socially unstable

transitional times). Monolete spores abruptly increase possibly reflecting temporary

invasion of forest gaps. Shortly after (~ 100 years later), LOI increases to values above

80%, indicating forest recovery wherein late-successional tree taxa such as

Page 166: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

154

Combretaceae/Melastomataceae begin to dominate the pollen record. In the literature

regarding the neotropics, the problem of separating Melastomataceae (e.g. Clidemia for

lowlands and Miconia for highlands) (see Tables 2.1 and 2.3) from Combretaceae (e.g.

Combretum) pollen taxa has been discussed (Marchant et al., 2002), but it is believed that

possibly the former could be in higher abundances during early forest succession and

progressively replaced by the latter as succession develops (Pascarella et al., 2007).

Nevertheless, the Combretaceae/Melastomataceae pollen type has been found to be

representative of mature forests and seasonally inundated forest in Belize (Bhattacharya

et al., 2011) and South America (Gosling et al., 2009), such as the forest that surrounded

the wetland where Core L-3 was collected up until ca. 100 yrs BP when disturbances

related to salt extraction increased (Figure 5.3) (Dillon, 1979).

5.4.2 Baseline for forest conservation at Lachuá lowlands region

Co-dominance of pollen from Combretaceae/Melastomataceae and Solanaceae lasted for

approximately 200 yrs after the Classic-Postclassic transition (PCA1 values remain

relatively constant), likely indicating that some economic activities such as silviculture

and salt production remained, but soon began to decrease gradually up to the complete

abandonment of Salinas de los Nueve Cerros. Salinas de los Nueve Cerros is believed to

have been inhabited by scattered populations after its abandonment because its name

appears in colonial historical documents dating from 1625 AD (Godoy, 2006). It is

registered in such documents that salt extraction by Europeans started around 1626 AD,

which matches the decrease in Combretaceae/Melastomataceae (Figure 5.3, ca. 500 yrs

BP) probably due to related natural resources extraction in the nearby area (e.g. timber).

Page 167: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

155

Salinas de los Nueve Cerros was identified as a critical location for salt production, an

economy that the Conquistadors felt they needed control to “pacify” local inhabitants.

Salt extraction was suspended for ca. 100 yrs due to local revolts, and resumed in the

early 1700’s until the 20th century (Dillon, 1979; Woodfill, 2012).

Despite continued activity at the Salinas salt mine, the surrounding vegetation shows

signs of recovery (ca. 300 yrs BP) because the pollen record is dominated by

Combretaceae/Melastomataceae. Similar pollen taxa were found to increase in abundance

during the Postclassic in the nearby tropical rain forest region of the Mexican Lacandon

Forest (e.g. Combretaceae /Melastomataceae and Myrtaceae) (Domínguez-Vázquez and

Islebe, 2008) and in the less humid Mirador Basin in Northern Petén (i.e. Combretaceae

/Melastomataceae) (Wahl et al., 2006). Forest in Lachuá environs likely remained

without intense anthropogenic management for approximately 800 yrs after its

abandonment during the Postlassic (PCA1 values remain relatively stable until ca. 150

yrs BP, Figure 5.1). Nevertheless, relatively low values of trilete spores and higher

values of monolete spores (McNeil et al., 2010), suggest some degree of disturbance.

Arboreal pollen from Combretaceae/Melastomataceae, Myrtaceae, Alchornea, Rubiaceae,

and Spondias that can tolerate low disturbance regimes begin to show decreases in

abundance during the last 150 yrs BP. On the contrary, Solanaceae benefit from

intermediate disturbance regimes as L-3 pollen pollen record indicates an increase, more

similar to what is found in modern times in bryophyte polsters (Figure 5.5). The present

day pattern of Solanaceae and Combretaceae/Melastomataceae pollen (Figure 3.3) could

Page 168: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

156

be associated with the fact that the Lachuá Lake National Park was recently established

(1974 AD) (Monzón, 1999), so most likely the forest is still recovering from recent

disturbances occurring during the 20th century. Forest structure that prevailed from 770-

100 years BP changed dramatically during the 20th century, due to economic activities

that involved natural resources extraction, including salt production, oil prospecting

during the 1970’s, 1980's and 2000's, and arrival of displaced populations since the

1950's (Avendaño et al., 2007).

The remnant of tropical rain forest protected currently at Lake Lachuá National Park

(14,500 ha area) since the mid 1970’s is possibly more similar to the one that prevailed

during Classic times when forest gardens dominated land use at the outskirts of Salinas

de los Nueve Cerros (Figure 5.1 and 5.3). Present day PCA1 values are similar to PCA1

values during the Late Classic, suggesting similar disturbance levels.

A baseline to return to a “healthier” forest status (e.g. restoration) before recent

disturbances in the Lachuá environs could be considered to be the forest condition that

prevailed for approximately 800 years after Salinas de los Nueve Cerros was abandoned.

However, the core L-3 time span does not provide any reference for whether this 800-

year condition has an analog at earlier times before the Preclassic colonization at Lachuá.

The previous forest condition composed mostly of forest gardens lasted approximately

600-700 years was maintained under silvicultural principles which required deep

knowledge to imitate forest structure (Gomez-Pompa, 1991; Pyburn, 1998; Fedick,

2010). Although current arboreal composition is similar to the one from the Classic

Page 169: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

157

-32 -16

PCA1

-20

-10

10

20

PC

A2

Figure 5.5. Principal Component Analysis (PCA) of modern pollen rain samples from Lachuá lowlands and sampled levels from core L-3. (+) represent bryophyte polsters modern samples, triangles are modern surface sediments, and inversed filled triangles are sedimentary records.

Page 170: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

158

period, silvicultural practices are at the moment not being used and therefore current

forest management and disturbances may be leading forest into a new equilibrium state

with no previous analog (Whitehouse, 2010).

5.4.3 Lachuá Lowlands in the context of Mesoamerican Holocene paleoecology

The location of the L-3 core at the intersection between lowlands and highlands is

important in order to cross-correlate Mesoamerican regional paleoecology (Islebe and

Leyden, 2006) (Figure 3.1). The Lachuá lowlands are located in a wet (high

precipitation) zone that extends from Izabal at the Guatemalan Caribbean coastline to the

Uxpanapa region in the Gulf of Mexico; this contrasts with the much lower rates of

precipitation in the Yucatán Peninsula and the Petén lowlands (Imbach et al., 2010). The

high precipitation in this zone is believed to have been consistent in the long term

(Wendt, 1989), because it is hypothesized that partly due to resultant wetter conditions,

relicts (refugia) of tropical rain forest species were held during the last glacial cycle of

the Pleistocene.

For Mesoamerican sites located outside the zone of high precipitation, a series of

droughts occuring throughout the late Holocene (i.e. related to either insolation

variability or migration of the Intertropical Convergence Zone) have been found to match

important transitional cultural periods in Mesoamerica (i.e. Mayan Terminal Classic, the

Toltecs, the Aztecs and Spanish Conquest). These periods of drought have also been

correlated with extensive measurements from Cariaco Basin in Venezuela that also show

periods of low rainfall. Late-Holocene sites that indicate signs of drought in the paleo-

Page 171: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

159

record are located in Central Mexico (Stahle et al., 2011), Northern and Southern

Yucatán Peninsula (Hodell et al., 2001; Carrillo-Bastos et al., 2010), and the Petén region

in Guatemala (Islebe and Leyden, 2006; Gill et al., 2007). Changes in vegetation inferred

from our L-3 sediment core do not support continued periods of droughts, because at

levels in the L-3 core where droughts are expected to be observed, there are no changes

in abundance of NAP pollen that could benefit from dryer conditions, such as Asteraceae,

Poaceae, and Amaranthaceae /Chenopodiaceae (Leyden et al., 1998; Wahl et al., 2006;

Wahl et al., 2007). Nevertheless, more high resolution exploration is needed in future

studies in the Lachua region, since temporal resolution of the L-3 core may not be

adequate to detect decade-long droughts that have been reconstructed for some sites in

the Mesoamerican lowlands.

The time of major variability within the pollen composition from L-3 core is at the

Classic-Postclassic transition (PCA1 values cross the zero threshold which indicates

opposing trends) and indicates culturally-driven dynamics and not natural climate forcing

(Figure 5.1 and 5.3). The changes in vegetation composition reflected in the L-3 core are

likely directly related to cultural management of Mayan forest gardens. Non-arboreal taxa

that tend to indicate occurrence of droughts (i.e. Asteraceae or Poaceae) do not show

consistent increases at the time of cultural transitional periods (Wahl et al., 2006). Pollen

and LOI information from the L-3 core both support a paleoclimate hypothesis of no

extended periods of drought, an observation that may be explained by the site's location

in the Izabal to Uxpanapa "wet belt". Other locations in this wet belt in Mexico have

shown evidence of humid conditions during expected dry conditions, such as the

Page 172: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

160

Terminal Classic droughts in Laguna Atezca (Conserva and Byrne, 2002), and during the

Little Ice Age in Los Tuxtlas (Lozano-García et al., 2007). The use of other proxies, such

as oxygen isotopes, is needed in order to develop more conclusive inferences about the

occurrence of droughts in the Lachuá lowlands and nearby Petexbatún region.

There is considerable controversy surrounding the hypothesis that heavy deforestation by

the Mayans directly led to drying microclimate and ultimately the demise of Mayan

civilization. In contrast, evidence is surfacing from Copán in Honduras that indicates that

the low disturbance effects of the Mayan environmental management regime (Beach et

al., 2006) did not always lead to deforestation effects sufficient to have catalyzed the

Mayan Collapse (Fedick, 2010; McNeil et al., 2010). LOI values from core L-3 indicate

that possibly the management of Mayan forest gardens did not have a negative impact on

the environment, since wetland conditions were maintained in the region over centuries.

Nonetheless, further sedimentological analysis is needed to support our interpretations.

Forest management at Salinas de los Nueve Cerros did not result in a deforested

landscape, not even during its phase of maximal development during the Terminal

Classic. This is in stark contrast for Petén cities like Tikal where evidence for

deforestation is prevalent (Lentz and Hockaday, 2009) and supported by pollen data and

other proxies from different authors (Islebe and Leyden, 2006). To further support the

"no deforestation" hypothesis, sediment cores should be taken from the Salinas de los

Nueve Cerros city epicentre and nearby locations to differentiate between urban versus

rural land use change and correlating environmental impacts. More site-specific

approaches like the one applied in this study are likely more appropriate to

Page 173: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

161

paleoecological-archaeological research because geographical heterogeneity is generally

considered more relevant in explaining cultural and environmental idiosyncrasies

(Aimers, 2007; Emery and Thornton, 2008; Beach et al., 2009; Demarest, 2009) than

assuming homogeneous responses across large expanses of landscapes and regions

(Powell, 2008).

5.5 Chapter summary 

The fossil pollen spectra from the L-3 core indicate three major phases along ca. 2000 yrs

in the environs of the Lachuá Lake at the outskirts of Salinas de los Nueve Cerros. The

first phase spans over ca. 700 years of development of Mayan forest gardens (i.e. forestry

practices) mainly during the Classic, to its eventual abandonment at the Classic-

Postclassic transition ca. 1100 yrs BP. Main pollen taxa related to forest management and

economic uses are Solanaceae (related possibly to medium disturbance levels), and

Bursera, Myrtaceae, Sapium, Spondias, and Terminalia, respectively. The effects of

registered regional droughts at the Classic-Postclassic transition are discarded because

pollen taxa associated to dryer conditions (e.g. Poaceae and Cyperaceae) show no

significant increases relative to what is registered in the modern pollen rain.

The second phase is related to an increase in percentages of Combretaceae

/Melastomataceae pollen at the onset of the Postclassic, and to its eventual dominance in

the fossil pollen spectra during the next ca.700 years. Combretaceae/Melastomataceae

pollen in the L-3 core sedimentary record is associated with the prevailing forest

conditions after Salinas de los Nueve Cerros was abandoned due to a cease in major salt

Page 174: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

162

production at the end of the Classic. The Mayan forest garden pollen “signal” is

maintained during the remains of this second phase until recent times (ca. 150 yrs BP) as

modern pollen rain calibration from L-3 core suggests. A minor drop in

Combretaceae/Melastomataceae pollen abundance ca. 500 yrs BP may be indicative of

the reactivation of salt production by European settlers, which may have disturbed the

forests at some degree since Solanaceae pollen abundance remains relatively constant

since then (ca. 10%). An increase in monolete and a decrease in trilete spores support the

changes observed in Combretaceae/Melastomataceae and Solanaceae pollen.

The recent phase dates back ca. 150 yrs when a decrease in Combretaceae

/Melastomataceae pollen percentages suggests an increase in the salt production by

European colonists in the region, and probably the extraction of other natural resources

(e.g. timber). In general terms, disturbance levels observed during the last 150 yrs in the

region are similar to the one observed during the Classic period according to PCA (which

reflect changes in abundance and composition of pollen spectra). The main difference

between recent times (150 yrs BP) and the Classic period, is that in the latter forest

management was well planned under complex forestry principles.

Page 175: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

163

Chapter 6:

Conclusions

Page 176: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

164

6.1 What are the factors that explain vegetation distribution along the 

Las Verapaces environmental gradient and what taxa can be used as 

"indicator species"?  

Indicator plant taxa which had discrete distribution allowed me to delineate three

vegetation belts, which represent changes in vegetation communities along the Las

Verapaces elevational gradient in the Central Guatemalan Highlands and Lowlands:

Lowland Rain Forest, Lower Montane Rain Forest, and Montane Cloud Forest.

Generalist taxa smoothed the delineation of vegetation belts because of their continuous

distribution, and montane taxa that were distributed in lowlands informed me of the

existence of montane-like habitats beyond their expected elevation range (disjunctive

taxa).

The collation of unpublished vegetation inventories was effective as it was possible to

identify explanatory factors, such as elevation which in combination with temperature

variability (based on a temperature database) are the main criteria for vegetation belt

delineation. Other factors such as landscape position in topographically-controlled

drainage divides, and biogeographic origin provided complementary explanations.

Landscape position within a watershed and topographic variability (i.e. geomorphology

and underlying bedrock controls) influence vegetation distribution through their

relationship with dispersal processes and localized microclimate (physical and

physiological barriers). Patterns in the distributions of plant taxa along the Las Verapaces

gradient possibly reflect in part the biogeographic origin of taxa. Plant biogeography

Page 177: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

165

integrates vegetation responses (i.e. physiological tolerances) to variability in elevation

and climate, with local relief determining whether or not an area is acting as a dispersal

corridor or barrier.

Iti is hoped that the research approach used in this study of the Las Verapaces gradient

serves as a model for future research in other parts of Guatemala as well as neighboring

regions in Mesoamerica. With future climate change and enhanced anthropogenic

disturbance of natural landscapes, there is a growing need for baselines from which to

compare future changes in vegetation communities. The present study contributes to this

objective. Moreover, the categorization of indicator, generalist and idiosyncratic taxa

permit more efficient and rigorous analysis of other meta-data bases, enabling better

decisions about conservation priorities and design.

Page 178: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

166

6.2 Can paleoecological calibrations for fossil pollen be constructed 

from a comparison of modern pollen rain from surface sediments and 

bryophyte polsters?  

In tropical regions, pollen spectra found in pollen reservoirs depend mainly on the

geographical location (i.e. lowlands or highlands) as it determines vegetation type and

related pollen production and dispersal mechanisms. Biogeographic origin of plants from

the highlands is mainly temperate or Laurasian, and therefore the major pollen dispersal

mechanism is anemophily; lowland plants have mainly zoophilous pollen dispersal

syndromes, because their origin is tropical or Amazonian. In spite of the fact that

anemophilous pollen can reach longer distances, analysis of bryophyte polsters from the

lowlands shows that zoophilous and local pollen taxa have in general a higher input than

in surface sediments, where zoophilous and anemophilous inputs are generally more

even. Pollen input in surface sediments and bryophyte polsters from highlands is

dominantly local and anemophilous, while input from lowlands due to wind transporation

is minimal. In general, some pollen taxa present in bryophyte polsters are “silent” in

surface sediments, because the former contains more pollen types from forest interiors

(i.e gravity and trunk space components, Faegri and Iversen, 1989).

Pollen spectra from small basins could have a higher local pollen input (especially if

surrounded by a high dense canopy vegetation) than mid to large sized basins (less barrier

effect from surrounding vegetation). Based on the collected information, it is believed

that if surface sediments are collected in a landscape that is forested to a large degree,

Page 179: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

167

their pollen assemblages would be comparable to those found in bryophyte polsters from

forest interiors (i.e. high arboreal pollen content).

A preliminary modern pollen rain calibration has been developed in this study between

vegetation and bryophyte polsters and surface sediments, as a means to understand better

the pollen signal from the latter as it represents the best analogue for fossil pollen spectra

found in sedimentary records. The present calibration study is important because it covers

an unexplored important region in Guatemala; these data can be linked to the northern

Petén lowlands, and the Las Verapaces lowland and highlands, with the rest of

Guatemala and Mesoamerica in terms of palynological and paleoecological analyses.

Page 180: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

168

6.3 What are the major vegetation changes recorded in the highland 

core from the Las Verapaces region?  

Paleoecological methods based on pollen and loss-on-ignition, aided in the reconstruction

of the paleoenvironmental history of the Cahabón river floodplain for the past ~2400

years. At the oldest date reported for the P-4 core (2390 yrs BP), possible ancient

lacustrine-like conditions are reconstructed for the floodplain, specifically a shoreline

environment where due to high rates of decomposition and oxidation, pollen absence is

explained. Initial agricultural exploration by Mayan populations at these earlier times

during the Preclassic could explain the higher sedimentation rate (0.25 cm yr-1) which

decreases (0.017 cm yr-1) once land management techniques minimized soil erosion. The

P-4 pollen record indicates agricultural activities (e.g. Zea and Asteraceae presence) at

the Cahabón river floodplain almost uninterruptedly for ca. 1700 yrs. One possible

explanation is that agriculture is interrupted first, temporarily at the Classic-Postclassic

transition (e.g. Terminal Classic) without having a significant impact on the culturally-

established floodplain dynamics; and later, completely at the European conquest and

colonization (e.g. Asteraceae pollen decrease), which marked a dramatic change in the

local vegetation dynamics (e.g. Pinus colonization) and floodplain sedimentation regime

(increase from 0.017 to 0.17 cm-1).

Nevertheless, in this mainly culturally driven sequence of vegetation changes, the

appearance of Abies, a higher elevation pollen taxon at the time of the “Little Ice Age”

(300-400 yrs BP), indicates the possibility of some vegetation change in response to

Page 181: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

169

decreased solar activity, as seen in other locations in Mesoamerica (Lozano-García et al.,

2010). The linkage between lowering of water table and climatic forcing, such as the

occurrence of a drought at the Classic-Postclassic transition is temporarily discarded,

because Cyperaceae and Poaceae percentages increase to similar values observed in

present day when no major droughts are registered. Most likely temporary abandonment

of floodplain terraces for agriculture explains a possibly lower water table, since

maintenance is responsible of resultant water level rising. Recorded droughts in

Mesoamerica in particular for the Terminal Classic are not regionally synchronous, since

there are locations with no clear evidence of droughts such as in the Las Verapaces

highlands. However, more exploration is needed in this region to have a more conclusive

explanation about the existence of an agricultural centre in the Cahabón River floodplain.

Paleoecological exploration of highland environments in Mesoamerica is expanding, the

relevance of connections between lowland and highland Maya chiefdoms in cultural

evolution are better understood. In the face of non-existent lacustrine environments in

Las Verapaces highlands, such as it is in many geographical locations globally,

paleoecological analysis of riverine sediments in this study strengthens the use of

alternative reservoirs of paleorecords as a means to reconstruct natural and cultural

evolution of past landscapes.

Page 182: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

170

6.4 What are the major vegetation changes recorded in the lowland core 

from the Las Verapaces region?  

The examination of a ca. 2000-year paleorecord indicates that vegetation changes in the

Lachuá region could be closely related to the history of the Mayan city of Salinas de los

Nueve Cerros. The L-3 core location holds the history of land management that took

place at the city's outskirts, where high abundance of pollen belonging to beneficial trees

taxa (e.g. Myrtaceae and Spondias) indicates that forest gardens could have been the

dominant land use. Cultural management of forest gardens determined vegetation

succession during the Preclassic and Classic cultural periods; these systems remained

mainly in a secondary succession stage, as indicated by the dominance of Solanaceae

pollen. At the Terminal Classic when the abandonment of Salinas de los Nueve Cerros

started, later succesional vegetation stages take place as inferred from

Combretaceae/Melastomatacae pollen co-dominance to an eventual complete dominance

for ca. 800 years after abandonment. It is not until ca. 150 yrs BP that vegetation changes

once again probably in the face of a different disturbance regime that involves major

natural resource extraction. In this case, Solanaceae pollen increases and pollen of

Combretaceae/Melastomataceae decreases.

Landscape evolution in the Lachuá lowlands over the development of Salinas de los

Nueve Cerros, including the Guatemalan colonial period, is determined mainly by

cultural factors. At the resolution level applied for the L-3 core, there is no clear evidence

of local occurrence of droughts reported elsewhere in Mesoamerica, especially at the

Page 183: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

171

Classic-Postclassic transition. Changes in vegetation are believed to be more a response

to the abandonment of cultural management practices (i.e. forestry) and not to occurrence

of droughts since pollen benefited by drier conditions showed no change in their

abundances (e.g. Poaceae). Nevertheless, more high resolution paleoecological analysis

with more proxies (e.g. oxygen isotopes) is needed in future studies in the Lachuá region

to test hypotheses related to decade-long droughts in Mesoamerica.

It is critical to recognize vegetation succession in the long term, and not only analysis

based on forest cover percentage calculations because it may lead to unrealistic

interpretations in conservation biology (e.g. AP values). Despite the fact that arboreal

pollen percentages indicate that at the physiognomic level, Lachuá lowlands landscape

forest cover remains at values above 80% for ca. 2000 yrs, vegetation succession

indicates dynamics related to different management and disturbance regimes. The

question related to “What is natural?” is important to answer in order to understand

landscape natural and cultural variability and to incorporate it into conservation

management practices aided by paleoecological analysis.

Page 184: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

172

6.5 What is the role of natural variability and cultural factors related to 

the Maya Civilization in the evolution of landscapes in the Las Verapaces 

Region?  

Las Verapaces Region is located at an important transitional region between the Northern

Lowlands and the North Central Highlands in Guatemala. The pollen record of the last

ca. 2000 yrs BP has not shown any evidence of vegetation dynamics (i.e. succession)

driven by climate variability, especially at the time of the hypothesized droughts at the

Terminal Classic or Early Postclassic. Reduction in precipitation at that time period has

been hypothesized as caused by alterations in the latitudinal migration of the Intertropical

Convergence Zone (ITCZ), or ultimately by an increased aridity effect as a consequence

of intense land clearing, agricultural activities and high rates of deforestation,

respectively. Contrary to the latter hypothesis, Las Verapaces landscapes have an

important multi-centennial cultural imprint of successful Mayan management. Pollen

records indicate in on one hand the existence of agriculture at the Cahabón River

Floodplain, with possible soil conservation practices that led to the successful

establishment of an agricultural center (i.e. low sedimentation rates); and on the other

hand, a Mayan forest garden at the outskirts of Salinas de los Nueve Cerros in the Lachuá

region. In these scenarios, cultural factors possibly played an important role in the

evolution of landscapes, coupled with the influence of relatively climatic stable

conditions. Regarding natural factors, the lowlands to highlands gradient in the Las

Verapaces is located in a high precipitation and humidity envelop (i.e. known as the

Page 185: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

173

Uxpanapa wet belt), a fact that contributes to its climatological stability, and thus in part

to its enormous biological diversity.

However, more exploration is needed in the Las Verapaces Region, in terms of collecting

longer Holocene records to examine landscape evolution at a longer temporal scale.

Inclusion of locations inside and outside of the “Uxpanapa wet belt” in future studies in

the Mesoamerican region will provide basis to explain thoroughly landscape evolution.

Inclusion of more paleoecological proxies (e.g. oxygen isotopes and macro and

microscopic charcoal) in future studies will enhance the understanding and the possibility

of testing hypotheses related to landscape evolution in terms of natural and cultural

factors.

Pollen records, LOI measurements, and sedimentation rates examined in this thesis

provided key information to support the idea that sustainable anthropogenic management,

if well planned, could enhance natural resources conservation. Lessons learned from the

paleoecology of the lowlands and highlands in the Las Verapaces include understanding

the negative effects that the European conquest and colony had on landscape dynamics

through drastic natural resource extraction. Even in the face of climatic and

environmental stable conditions, non-planned, non-measured and non-sustainable natural

resources management represents a threat to the conservation of biological diversity and

cultural legacies.

Page 186: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

174

References

Aguilar, E., T. Peterson, P. Ramírez, R. Frutos, J. Retana, M. Solera, J. Soley, I. González-García, R. Araujo, A. Santos, V. Valle, M. Brunet, L. Aguilar, L. Álvarez, M. Bautista, C. Castañón, L. Herrera, E. Ruano, J. Sinay, E. Sánchez, G. Hernández, F. Obed, J. Salgado, J. Vázquez, M. Baca, M. Gutiérrez, C. Centella, J. Espinosa, D. Martínez, B. Olmedo, C. Ojeda, R. Núñez, M. Haylock, H. Benavides, and R. Mayorga. (2005). Changes in precipitation and temperature extremes in Central America and northern South America, 1961-2003. Journal of Geophysical Research, 110.

Aimers, J. (2007). What Maya collapse? Terminal Classic variation in the Maya Lowlands. Journal of Archaeological Research, 15.

Alonso, N. (1999). Propuesta de manejo forestal para el municipio de Tamahú, Alta Verapaz. Centro Universitario del Norte. Universidad de San Carlos, Cobán, Alta Verapaz.

Alley, R. B., Marotzke, J., Nordhaus, W. D., Overpeck, J. T., Peteet, D. M., Pielke, R. A., Pierrehumbert, R. T., Rhines, P. B., Stocker, T. F., Talley, L. D., and Wallace, J. M. (2003). Abrupt Climate Change. Science, 299, 2005-2010

Almeida, L., Hooghiemstra, H., Cleef, A. and Van Geel, B. (2005). Holocene climatic and environmental change from pollen records of lakes Zempoala and Quila, central Mexican highlands. Review of Palaeobotany and Palynology, 136, 63-92

Andersen, U. S., Prado, J. P., Sorensen, M., and Kollmann, J. (2006). Conservation and utilisation of Abies guatemalensis rehder (Pinaceae) - an endangered endemic conifer in Central America. Biodiversity and Conservation 15, 3131-3151.

Andrews, A. (1983). Maya salt production and trade. The University of Arizona Press, Tucson.

Andrews, A. P. (1984). Long-Distance Exchange among the Maya - a Comment on Marcus. American Antiquity, 49, 826-828.

Anselmetti, F. S., Ariztegui, D., Hodell, D. A., Hillesheim, M. B., Brenner, M., Gilli, A., McKenzie, J. A., and Mueller, A. D. (2006). Late Quaternary climate-induced lake level variations in Lake Peten Itza, Guatemala, inferred from seismic stratigraphic analysis. Palaeogeography Palaeoclimatology Palaeoecology, 230, 52-69.

Antonelli, A., Nylander, J., Persson, C., and Sanmartin, I. (2009). Tracing the impact of the Andean uplift on Neotropical plant evolution. PNAS, 106, 9749-9754.

Arnauld, M. (1978). Habitat et société préhispaniques en Alta Verapaz occidentale, Guatemala: étude archéologique et ethnohistorique. Journal de la Société des Américanistes, 65, 41-62.

Arnauld, M. (1987). Regional ceramic development in the Northern Highlands, Alta Verapaz, Guatemala: Classic and Postclassic material. Maya Ceramics (ed. by P. Rice and R. Sharer). BAR International series, Oxford, Great Britain.

Arnauld, M. (1988). Ceramic Units from Southwestern Alta Verapaz, Guatemala. Ceramica de Cultura Maya, 15, 79-83.

Page 187: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

175

Arnauld, M. (1997). Relaciones interregionales en el Area Maya durante el Postclásico en base a datos arquitectónicos. In: X Simposio de Investigaciones Arqueológicas en Guatemala, (ed. by J. Leporte and H. Escobedo), pp. 117–131. Museo Nacional de Arqueología y Ethnología, Guatemala.

Arroyo, B. (1994). El proyecto Nueve Cerros, un ejemplo de la arqueologia de rescate: Ventajas y desventajas. VII Simposio de Investigaciones Arqueológicas en Guatemala (ed by J. Laporte and H. Escobedo), pp. 188-198. Ciudad de Guatemala.

Avendaño, C., García, M., Cajas, M. and De León, K. (2007). Dinámica del uso de la tierra y conservación de los Recursos Naturales en la Eco-región Lachuá. Escuela de Biología, Universidad de San Carlos de Guatemala, Guatemala, pp 73.

Ávila, R. (2004). Estudio base para el programa de monitoreo de la vegetación en la Zona de Influencia del Parque Nacional Laguna Lachuá. Escuela de Biología. Universidad de San Carlos de Guatemala, Guatemala.

Bachand, B.R. (2010). Onset of the early Classic period in the Southnern Maya Lowlands: New evidence from Punta de Chimino, Guatemala. Ancient Mesoamerica, 21, 21-44.

Barrientos, M. (2006). Atlas Palinológico de las plantas más abundantes de la Región Lachuá. Tesis Licenciatura en Biología. Escuela de Biología, Facultad de Ciencias Químicas y Farmacia. Universidad de San Carlos de Guatemala, Guatemala.

Bhattacharya, T., Beach, T., and Wahl, D. (2011). An analysis of modern pollen rain from the Maya lowlands of northern Belize. Review of Palaeobotany and Palynology 164, 109-120.

Beach, T. and Dunning, N.P. (1995). Ancient Maya terracing and modern conservation in the Petén rain forest of Guatemala. Journal of Soil and Water Conservation, 50, 138-145.

Beach, T., Luzzadder-Beach, S., Dunning, N., Hageman, J., and Lohse, J. (2003). Upland Agriculture in the Maya Lowlands: Ancient Maya Soil Conservation Northwestern Belize. The Geographical Review, 92, 372-397.

Beach, T., Dunning, N., Luzzadder-Beach, S., Cook, D.E. and Lohse, J. (2006). Impacts of the ancient Maya on soils and soil erosion in the central Maya Lowlands. Catena, 65, 166-178.

Beach, T., Luzzadder-Beach, S., Dunning, N. and Cook, D. (2008). Human and natural impacts on fluvial and karst depressions of the Maya Lowlands. Geomorphology, 101(1-2), 308-331.

Beach, T., Luzzadder-Beach, S., Dunning, N., Jones, J., Lohse, J., Guderjan, T., Bozarth, S., Millspaugh, S. and Bhattacharya, T. (2009). A review of human and natural changes in Maya Lowland wetlands over the Holocene. Quaternary Science Reviews, 28, 1710-1724.

Beach, T., Luzzadder-Beach, S., Terry, R., Dunning, N., Houston, S., and Garrison, T. (2011). Carbon isotopic ratios of wetland and terrace soil sequences in the Maya Lowlands of Belize and Guatemala. Catena 85, 109-118.

Behling, H., Berrio, J. C., and Hooghiemstra, H. (1999). Late Quaternary pollen records from the middle Caqueta river basin in central Colombian Amazon. Palaeogeography Palaeoclimatology Palaeoecology 145, 193-213.

Page 188: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

176

Behling, H., Keim, G., Junk, W. and de Mello, J. (2001). Holocene environmental changes in the Central Amazon Basin inferred from Lago Calado (Brazil). Palaeogeography Palaeoclimatology Palaeoecology, 173, 87-101.

Behling, H., Negrelle, R. (2006). Vegetation and pollen rain relationship from the tropical Atlantic rain forest in southern Brazil. Brazilian Archives of Biology and Technology, 49, 631-642.

Bennington, J. B., Dimichele, W. A., Badgley, C., Bambach, R. K., Barrett, P. M., Behrensmeyer, A. K., Bobe, R., Burnham, R. J., Daeschler, E. B., Van Dam, J., Eronen, J. T., Erwin, D. H., Finnegan, S., Holland, S. M., Hunt, G., Jablonski, D., Jackson, S. T., Jacobs, B. E., Kidwell, S. M., Koch, P. L., Kowalewski, M. J., Labandeira, C. C., Looy, C. V., Lyons, S. K., Novack-Gottshall, P. M., Potts, R., Roopnarine, P. D., Stromberg, C. A. E., Sues, H. D., Wagner, P. J., Wilf, P., and Wing, S. L. (2009). Critical Issues of Scale in Paleoecology. Palaios, 24, 1-4.

Beniston, M. (2005). Natural Forcing of the Climate System. In: Climatic Change and its Impacts. An Overview Focusing on Switzerland. (K. A. Publishers, Ed.), pp. 53-72. Advances in Global Change Research. Springer Netherlands, United States of America.

Berger, A. (1989). Pleistocene climatic variability at astronomical frequencies. Quaternary International, 2, 1-14.

Berrio, J. C., Boom, A., Botero, P. J., Herrera, L. F., Hooghiemstra, H., Romero, F., and Sarmiento, G. (2001). Multi-disciplinary evidence of the Holocene history of a cultivated floodplain area in the wetlands of northern Colombia. Vegetation History and Archaeobotany, 10, 161-174.

Berrio, J.C., Hooghiemstra, H., Marchant, R. and Rangel, O. (2002). Late-glacial and Holocene history of the dry forest area in the south Colombian Cauca Valley. Journal of Quaternary Science, 17, 667-682.

Bertoldi G., Notarnicola C., Leitinger G., Endrizzi S., Zebisch M., Della Chiesa S. and Tappeiner U. (2010). Topographical and ecohydrological controls on land surface temperature in an Alpine catchment. Ecohydrology, 3, 189-204

Bettis III, E.A., Benn, D.W. and Hajic, E.R. (2008). Landscape evolution, alluvial architecture, environmental history, and the archaeological record of the Upper Mississippi River Valley. Geomorphology, 101 (1-2), 362-377.

Bhattacharya, T., Beach, T. and Wahl, D. (2011). An analysis of modern pollen rain from the Maya lowlands of Northern Belize. Review of Palaeobotany and Palynology, 164, 109-120.

Binford, M., and Leyden, B. (1987). Ecosystems, Paleoecology and Human Disturbance in Subtropical and Tropical America. Quaternary Science Reviews, 6, 115-128.

Birks, H. and Birks, H. (2003). Reconstructing Holocene climates from pollen and plant macrofossils. In: A. Mackay, R. Battarbee, J. Birks and F. Oldfield (Ed), Global Change in the Holocene. Oxford University Press, pp. 342-357.

Birks, H. (2005). Fifty years of Quaternary pollen analysis in Fennoscandia 1954-2004. Grana, 44, 1-22.

Bjune, A., Birks, H.J.B. and Seppa, H. (2004). Holocene vegetation and climate history on a continental-oceanic transect in northern Fennoscandia based on pollen and plant macrofossils. Boreas, 33, 211-223.

Page 189: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

177

Bohrerova, Z., Bohrer, G., Cho, K.D., Bolch, M.A., Linden, K.G. (2009). Determining the viability response of pine pollen to atmospheric conditions during long-distance dispersal. Ecological Applications, 19, 656-667.

Bonham, C. (2006). Biodiversity and conservation of Sierra Chinajá: A rapid assessment of biophysical, socioeconomic, and management factors in Alta Verapaz, Guatemala. University of Montana.

Borgstede, G. and Mathieu, J.R. (2007). Defensibility and settlement patterns in the Guatemalan Maya highlands. Latin American Antiquity, 18, 191-211.

Bradley, R.S. (1999). Paleoclimatology: Reconstructing climates of the quaternary. Academic Press, San Diego, CA.

Bradley, R.S. and England, J.H. (2008). The Younger Dryas and the Sea of ancient ice. Quaternary Research, 70, 1-10.

Breedlove, D. (1981). Introduction to the Flora of Chiapas. In: Flora of Chiapas. Part I. (D. Breedlove, Ed.). The California Academy of Sciences, San Francisco.

Bridge, J. (2003). Rivers and floodplains: forms, processes, and sedimentary record. Blackwell Science, Oxford, pp 491.

Bridgewater, S. Harris, D., Whitefoord, C., Monro, A. Penn, M. Sutton, D., Sayer, B., Adams, B., Balick, M., Atha, D., Solomon, J. and Holst, B. (2006). A preliminary checklist of the vascular plants of the Chiquibul Forest, Belize. Edinburgh Journal of Botany, 63, 269-321.

Brook, E.J. (2009). Palaeoclimate. Atmospheric carbon footprints? Nature Geoscience, 2, 170-172.

Brown, A.G., Carpenter, R.G. and Walling, D.E. (2007). Monitoring fluvial pollen transport, its relationship to catchment vegetation and implications for palaeoenvironmental studies. Review of Palaeobotany and Palynology, 147, 60-76.

Brown, A. G., Carpenter, R. G., and Walling, D. E. (2008). Monitoring the fluvial palynomorph load in a lowland temperate catchment and its relationship to suspended sediment and discharge. Hydrobiologia, 607, 27-40.

Bunting, M., Gaillard, M.J., Sugita, S., Middleton, R. and Brostrom, A. (2004). Vegetation structure and pollen source area. The Holocene, 14(5), 651-660.

Bush, M. B. (1995). Neotropical Plant Reproductive Strategies and Fossil Pollen Representation. American Naturalist 145, 594-609.

Bush, M., and Rivera, R. (1998). Pollen dispersal and representation in a neotropical rain forest. Global Ecology and Biogeography, 7, 379-392.

Bush, M.B., Rivera, R. (2001). Reproductive ecology and pollen representation among neotropical trees. Global Ecology and Biogeography, 10, 359-367.

Bush, M. B., and Weng, C. Y. (2007). Introducing a new (freeware) tool for palynology. Journal of Biogeography, 34, 377-380.

Cajas, M. (2009). Análisis de la heterogeneidad geoecológica y la diversidad biológica en el Parque Nacional Laguna Lachuá Cobán, Alta Verapaz. Tesis Licenciatura en Biología. Escuela de Biología, Facultad de Ciencias Químicas y Farmacia. Universidad de San Carlos de Guatemala, Guatemala.

Campbell, I. (1999). Quaternary pollen taphonomy: examples of differential redeposition and differential preservation. Palaeogeography, Palaeoclimatology, Palaeoecology, 149, 245-256.

Page 190: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

178

Cao, X., Xu, Q., Jing, Z., Tang, J. Li, Y. and Tian, F. (2010). Holocene climate change and human impacts implied from the pollen records in Anyang, central China. Quaternary International. Holocene Vegetation Dynamics and Human Impact, I: Agricultural Activities and Rice Cultivation in East Asia, 227(1), 3-9.

Carrillo-Bastos, A., Islebe, G.A., Torrescano-Valle, N. and Emilia Gonzalez, N. (2010). Holocene vegetation and climate history of central Quintana Roo, Yucatán Peninsula, Mexico. Review of Palaeobotany and Palynology, 160, 189-196.

Cascante, A., Estrada, A. (2001). Floristic composition and structure of the premontane moist forest in the central valley of Costa Rica. Revista de Biología Tropical, 49, 213-225.

Castañeda, C. (1995). Sistemas Lacustres de Guatemala. Recursos que mueren. Colección Estudios, 1. Editorial Universitaria, Ciudad de Guatemala.

Castañeda, C. (1997). Estudio Florístico del Parque Nacional Laguna Lachuá, Alta Verapaz, Guatemala. Tesis Ingenieria Agronoma, Universidad de San Carlos de Guatemala.

CECON. 1999. Plan Maestro Biotopo Universitario “Mario Dary Rivera” para la conservación del quetzal 2000-2004. Centro de Estudios Conservacionistas, Universidad de San Carlos de Guatemala. pp 119.

Chanderbali, A. S., van der Werff, and S. S. Renner. (2001). Phylogeny and Historical Biogeography of Lauraceae: Evidence from the Chloroplast and Nuclear Genomes. Annals of the Missouri Botanical Garden, 88, 104-134.

Chase, A. F., Chase, D. Z., and Smith, M. E. (2009). States and Empires in Ancient Mesoamerica. Ancient Mesoamerica, 20, 175-182.

Chase, M., Kristian III, W., Lynam, A., Price, M. Rotenberry, J. (2000). Single species as indicators of species richness and composition in California coastal sage scrub birds and small mammals. Conservation Biology, 14(2), 474-87.

Cheng, J. W. (2011). Evolution of terraces I–III along the Anning River, western Sichuan, based on pollen records and terrace structure. Science China Earth Sciences 54, 127-135.

Conserva, M.E. and Byrne, R. (2002). Late Holocene vegetation change in the Sierra Madre Oriental of central Mexico. Quaternary Research, 58, 122-129.

Charlton, R. (2008). Fundamentals of fluvial geomorphology. Routledge, London, pp 234.

Chiang, J.C.H. and Bitz, C.M. (2005). Influence of high latitude ice cover on the marine Intertropical Convergence Zone. Climate Dynamics, 25(5), 477-496.

Chinchilla, E. (1984). Los jades y las sementeras. Guatemala, Tipografía Nacional. Clark, D. (2007). Detecting tropical forests' responses to global climatic and atmospheric

change: current challenges and a way forward. Biotropica, 39(1), 4-19. Colwell, R. K., Brehm, G., Cardelús, C., Gilman, A.C. and Longino, J. T. (2008). Global

warming, elevational range shifts, and lowland biotic attrition in the wet tropics. Science. 322: 258-261.

CONAP. (2002). Informe Institucional Consejo Nacional de Areas Protegidas (CONAP), Guatemala.

Conedera, M., Tinner, W., Crameri, S., Torriani, D., and Herold, A. (2006). Taxon-related pollen source areas for lake basins in the southern Alps: an empirical approach. Vegetation History and Archaeobotany, 15, 263-272.

Page 191: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

179

Crausbay, S. D. and S. C. Hotchkiss. (2010). Strong relationships between vegetation and two perpendicular climate gradients high on a tropical mountain in Hawai'i. – Journal of . Biogeography. 37, 1160-1174.

Crutzen, P. (2006). The "Anthropocene". In: E. Ehlers and T. Krafft (Editors), Earth System Science in the Anthropocene. Springer Berlin Heidelberg, pp. 13-18.

Cuello, A. N. L. and A. M. Cleef. (2009). The forest vegetation of Ramal de Guaramacal in the Venezuelan Andes. Phytocoenologia. 39, 109-156.

Curtis, J. H., Hodell, D. A., and Brenner, M. (1996). Climate variability on the Yucatan Peninsula (Mexico) during the past 3500 years, and implications for Maya Cultural Evolution. Quaternary Research 46, 37-47.

Dahlin, B.H., Quizar, R. and Dahlin, A. (1987). Linguistic divergence and the collapse of Precassic civilzation in southern Mesoamerica. American Antiquity, 52, 367-382.

de Klerk, P., Haberl, A., Kaffke, A., Krebs, M., Matchutadze, I., Minke, M., Schulz, J. and Joosten, H. (2009). Vegetation history and environmental development since ca 6000 cal yr BP in and around Ispani 2 (Kolkheti lowlands, Georgia). Quaternary Science Reviews, 28, 890-910.

De La Cruz, J. (1982). Clasificación de zonas de vida de Guatemala nivel de reconocimiento. INAFOR.

Dambrine, E. et al. (2007). Present forest biodiveristy patterns in France related to former Roman agriculture. Ecology, 88(6): 1430-1439.

Davidar, P., Puyravaud, J.P. and Leigh, E.G. (2005). Changes in rain forest tree diversity, dominance and rarity across a seasonality gradient in the Western Ghats, India. Journal of Biogeography, 32(3): 493-501.

Delhon, C., Thiébault, S. and Berger, J. (2009). Environment and landscape management during the Middle Neolithic in Southern France: Evidence for agro-sylvo-pastoral systems in the Middle Rhone Valley. Quaternary International. Rhythms and Causalities of the Anthropisation Dynamic in Europe between 8500 and 2500 cal BP: Sociocultural and/or Climatic Assumptions, 200(1-2): 50-65.

Demarest, A. (1997). The Vanderbilt Petexbatún Regional Archaeological Project 1989-1994: Overview, History, and Major Results of a Multidisciplinary Study of the Classic Maya collapse. Ancient Mesoamerica, 8(02): 209-227.

Demarest, A. (2005). Ancient Maya: The Rise and Fall of a Rainforest Civilization (Case Studies in Early Societies). Cambridge University Press.

Demarest, A. (2006). The Petexbatún Regional Archaeological Project: A Multidisciplinary Study of the Maya Collapse. Vanderbilt University Press, USA.

Demarest, A. (2009). Maya archaeology for the twenty-first century: The progress, the perils, and the promise. Ancient Mesoamerica, 20, 253-263.

Demarest, A., Rice, P. and Rice, D. (2004). The Terminal Classic in the Maya Lowlands: Collapse, Transition, and Transformation. University Press of Colorado, Boulder, 676 pp.

Dergachev, V.A. and van Geel, B. (2004). Large-scale periodicity of climate change during the Holocene. In: E.M. Scott, A.Yu. Alekseev and G. Zaitseva (eds) Impact of the environment on human migration in Eurasia . NATO Science Series, IV. Earth and Environmental Sciences 42: 159-183.

Diamond, J. (2005). Collapse: How societies choose to fail or succeed. Viking, New York.

Page 192: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

180

Diamond, J. (2009). Archaeology: Maya, Khmer and Inca. 461(7263): 479-480. Díaz , H.F. and Stahle, D.W. (2007). Climate and cultural history in the Americas: An

overview. Climatic Change, 83(1-2): 1-8. Dillon, B. (1977). Salinas de los Nueve Cerros, Alta Verapaz, Guatemala. Preliminary

Archaeological Investigations. Ballena Press Studies in Mesoamerican art, archaeology and ethnohistory, 2. Ballena Press, Socorro, New Mexico.

Dillon, B. (1979). The Archaeological ceramics of Salinas de los Nueve Cerros, Alta Verapaz, Guatemala. University of California, Berkeley.

Dillon, B. (1990). Proyecto de rescate de los vasijones de Salinas de los Nueve Cerros, Alta Verapaz. 1414.

Domínguez-Vázquez, G., Islebe, G. A., and Villanueva-Gutiérrez, R. (2004). Modern pollen deposition in Lacandon forest, Chiapas, Mexico. Review of Palaeobotany and Palynology 131, 105-116.

Domínguez -Vázquez, G. and Islebe, G. (2008). Protracted drought during the late Holocene in the Lacandon rain forest, Mexico. Vegetation History and Archaeobotany, 17, 327-333.

Dull, R. A., Nevle, R. J., Woods, W. I., Bird, D. K., Avnery, S., and Denevan, W. M. (2010). The Columbian Encounter and the Little Ice Age: Abrupt Land Use Change, Fire, and Greenhouse Forcing. Annals of the Association of American Geographers 100, 755-771.

Dunning, N., Beach, T. and Rue, D. (1997). The paleoecology and ancient settlement of the Petexbatún region, Guatemala. Ancient Mesoamerica, 8(2): 255-266.

Dunning, N., Rue, D.J., Beach, T., Covich, A. and Traverse, A. (1998). Human-Environment interactions in a tropical watershed: The paleoecology of Laguna Tamarindito, El Petén, Guatemala. Journal of Field Archaeology, 25, 139-151.

Elliott, M., Fisher, C., Nelson, B., and Molina, R. (2010). Climate, Agriculture, and cycles of occupation over the last 4000 yr in southern Zacatecas, Mexico. Quaternary Research 74, 26-35.

Emmerson, M. C., Solan, M., Emes, C., Paterson, D., Raffaelli, D. (2001). Consistent patterns and the idiosyncratic effects of biodiversity in marine ecosystems. Nature. 411: 73-77.

Emery, K.F. and Thornton, E.K. (2008). A regional perspective on biotic change during the Classic Maya occupation using zooarchaeological isotopic chemistry. Quaternary International, 191, 131-143.

Erwin, D.H. (2009). Climate as a Driver of Evolutionary Change. Current Biology, 19, R575-R583.

Fægri, K. and Iversen, J. (1989). Textbook of pollen analysis. John Wiley and Sons, Chichester, 328 pp.

Fedick, S.L. (2010). The Maya Forest: Destroyed or cultivated by the ancient Maya? Proceedings of the National Academy of Sciences of the United States of America, 107, 953-954.

Figueroa-Rangel, B.L., Willis, K.J. and Olvera-Vargas, M. (2008). 4200 years of pine-dominated upland forest dynamics in west-central Mexico: Human or natural legacy? Ecology, 89, 1893-1907.

Ford, A. (2008). Dominant plants of the Maya forest and gardens of El Pilar: Implications for paleoenvironmental reconstructions. Journal of Ethnobiology, 28, 179-199.

Page 193: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

181

Ford, A. and Nigh, R. (2009). Origins of the Maya forest garden: Maya resource management. Journal of Ethnobiology, 29, 213-236.

Fourcade, E., Piccioni, L., Escriba, J., and Rosselo, E. (1999). Cretaceous stratigraphy and palaeoenvironments of the Southern Petén Basin, Guatemala. Cretaceous Research, 20, 793-811.

Fowler, W.R., Jr., Demarest, A.A., Michel, H.V., Asaro, F. and Stross, F. (1989). Sources of Obsidian from El Mirador, Guatemala: New Evidence on Preclassic Maya Interaction. American Anthropologist, 91(1), 158-168.

Freidel, D., Schele, L. and Parker, J. (1993). Maya Cosmos. Three thousand years on the Shaman's path. Harper Paperbacks, USA.

Freycon, V., Krencker, M., Schwartz, D., Nasi, R., Bonal, D. (2010). The impact of climate changes during the Holocene on vegetation in northern French Guiana. - Quaternary Research, 73, 220-225.

Fuller, D., Sato, Y., Castillo, C., Qin, L., Weisskopf, A., Kingwell-Banham, E., Song, J., Ahn, S., Etten, J. (2010). Consilience of genetics and archaeobotany in the entangled history of rice, Archaeological and Anthropological Sciences. Springer Berlin / Heidelberg, pp. 115-131.

Fuller, D.Q., Qin, L., Zheng, Y., Zhao, Z., Chen, X., Hosoya, L., Sun, G. (2009). The Domestication Process and Domestication Rate in Rice: Spikelet Bases from the Lower Yangtze. Science, 323(5921), 1607-1610.

Gabriel, J., Reinhardt, E., Peros, M., Davidson, D., van Hengstum, P. and Beddows, P. (2009). Palaeoenvironmental evolution of Cenote Aktun Ha (Carwash) on the Yucatán Peninsula, Mexico and its response to Holocene sea-level rise, Journal of Paleolimnology. Springer Netherlands, 42(2), 199-213.

Gagnon, P. (2009). Fire in floodplain forests in the southeastern USA: Insights from disturbance ecology of native bamboo, Wetlands. Springer Netherlands, pp. 520-526.

Gaillard, C. (2003). Plantaciones Forestales: Oportunidades para el desarrollo sostenible, Facultad de Ciencias Ambientales y Agrícolas, Universidad Rafael Landivar, Ciudad de Guatemala.

Gaillard, M., Sugita, S., Bunting, M., Middleton, R., Brostrom, A., Caseldine, C., Giesecke, T., Hellman, S., Hicks, S., Hjelle, K., Langdon, C., Nielsen, A., Poska, A., von Stedingk, H., Veski, S. and Members, P. (2008). The use of modelling and simulation approach in reconstructing past landscapes from fossil pollen data: a review and results from the POLLANDCAL network. Vegetation History and Archaeobotany, 17, 419-443.

Gandouin, E., Maasri, A., Van Vliet-Lanoe, B., and Franquet, E. (2006). Chironomid (Insecta: Diptera) Assemblages from a Gradient of Lotic and Lentic Waterbodies in River Floodplains of France: A Methodological Tool for Paleoecological Applications. Journal of Paleolimnology 35, 149-166.

Gandouin, E., and Ponel, P. (2010). Rivers. In "Changing Climates, Earth Systems and Society." (J. Dodson, Ed.), pp. 161-175. International Year of Planet Earth. Springer Netherlands.

Page 194: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

182

García, B. (1998). Estudio del dosel de la selva nublada del Biotopo Universitario para la Conservación del Quetzal "Lic. Mario Dary Rivera". Tesis Licenciatura Biología, Escuela de Biología, Facultad de Ciencias Químicas y Farmacia. Universidad de San Carlos de Guatemala, Guatemala.

García, M. (2001). Estructura y composición florística de los estratos arbustivo y arbóreo en la Zona de Influencia del Parque Nacional Laguna Lachuá entre las comunidades Santa Lucia y Rio Tzetoc. Informe EDC. Escuela de Biología, Facultad de Ciencias Químicas y Farmacia, Universidad de San Carlos de Guatemala, Guatemala.

Garrido, J., (2009). Grupos arquitectónicos asociados a Salinas de Los Nueve Cerros, Cobán, Alta Verapaz, Guatemala. Mesoweb Publications.

Gentry, A. (1982). Neotropical Floristic Diversity: Phytogeographical Connections between Central and South America, Pleistocene Climatic Fluctuations, or an Accident of the Andean Orogeny? Annals of the Missouri Botanical Garden, 69, 557-593.

Gerold, G., Schawe M. and Bach, K. (2008). Hydrometeorologic, Pedologic and Vegetation Patterns along an Elevational Transect in the Montane Forest of the Bolivian Yungas. Die Erde 139(1-2), 141-168.

Gill, R.B., Mayewski, P.A., Nyberg, J., Haug, G.H. and Peterson, L.C. (2007). Drought and the Maya collapse. Ancient Mesoamerica, 18(02), 283-302.

Godoy, W. (2006). De la historia de La Verapaz. Casa del Arte, Coban, Alta Verapaz. Gómez-Pompa, A. (1991). The forest gardens of the Maya. Association for Tropical

Biology. Annual Meeting of the Association for Tropical Biology Held at the Aibs (American Institute of Biological Sciences) Annual Meeting of Scientific Societies; San Antonio, Texas, USA, August 4-8, 1991. No Pagination. Association for Tropical Biology: St. Louis, Missouri, USA. Paper,

Gosling, W.D., Mayle, F.E., Tate, N.J. and Killeen, T.J. (2009). Differentiation between Neotropical rainforest, dry forest, and savannah ecosystems by their modern pollen spectra and implications for the fossil pollen record. Review of Palaeobotany and Palynology, 153, 70-85.

Graham, A. (1999). The Tertiary History of the Northern Temperate Element in the Northern Latin American Biota. - American Journal of Botany, 86, 32-38.

Graham, A. (2006). The history of the vegetation of Guatemala: Cretaceous and Tertiary. In: E. Cano (Editor), Biodiversidad de Guatemala. Universidad del Valle de Guatemala, Guatemala.

Granados, P. (2001). Ictiofauna de la Laguna Lachuá, Cobán, Alta Verapaz. Tesis Licenciatura Biología, Escuela de Biología, Facultad de Ciencias Químicas y Farmacia. Universidad de San Carlos de Guatemala, Guatemala.

Haase-Schramm, A., Bohm, F., Eisenhauer, A., Garbe-Schonberg, D., Dullo, W.C. and Reitner, J. (2005). Annual to interannual temperature variability in the Caribbean during the Maunder sunspot minimum. Paleoceanography, 20, 8.

Hammel, B. E. and N. A. Zamora. (1990). Nyssa-Talamancana (Cornaceae), an Addition to the Remnant Laurasian Tertiary Flora of Southern Central-America. Brittonia, 42, 165-170.

Page 195: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

183

Hammer, Ø., Harper, D.A.T. and Ryan, P.D. (2001). PAST: Paleontological Statistics Software Package for Education and Data Analysis. Palaeontologia Electronica, 4, 9 pp.

Harris, D.R. (2003). Climatic change and the beginnings of agriculture: The case of the Younger Dryas Evolution on Planet Earth. In: L.J. Rothschild and A.M. Lister (Editors). Academic Press, London, pp. 379-394.

Harvey, C. A., Komar, O., Chazdon, R., Ferguson, B., Finegan, B., Griffith, D., Martínez-Ramos, M., Morales, H., Nigh, R., Soto-Pinto, L., van Breugel, M., and Wishnie, M. (2008). Integrating agricultural landscapes with biodiversity conservation in the Mesoamerican hotspot. - Conservation Biology, 22, 8-15.

Haug, G., Gunther, D., and Peterson, L. (2003). Climate and the collapse of maya civilization. Science 299, 1731-1735.

Heiri, O., Lotter, A.F. and Lemcke, G. (2001). Loss on ignition as a method for estimating organic and carbonate content in sediments: reproducibility and comparability of results. Journal of Paleolimnology, 25, 101-110.

Helama, S., Timonen, M., Holopainen, J., Ogurtsov, M.G., Mielikainen, K., Eronen, M., Lindholm, M. and Merilainen, J. (2009). Summer temperature variations in Lapland during the Medieval Warm Period and the Little Ice Age relative to natural instability of thermohaline circulation on multi-decadal and multi-centennial scales. Journal of Quaternary Science, 24, 450-456.

Hellman, S., Gaillard, M., Bunting, J. and Mazier, F. (2009). Estimating the Relevant Source Area of Pollen in the past cultural landscapes of southern Sweden - A forward modelling approach. Review of Palaeobotany and Palynology, 153, 259-271.

Hemp, A. (2006). Continuum or zonation? Altitudinal gradients in the forest vegetation of Mt. Kilimanjaro. Plant Ecology, 184, 27-42.

Higgins, S., Nathan, R. and Cain, M. (2003). Are long-distance dispersal events in plants usually caused by nonstandard means of dispersal? Ecology, 84, 1945–1956.

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., and Schnurrenberger, D. W. (2005). Climate change in lowland Central America during the late deglacial and early Holocene. Journal of Quaternary Science, 20, 363-376.

Hodell, D. A., Curtis, J. H., Jones, G. A., Higueragundy, A., Brenner, M., Binford, M. W., and Dorsey, K. T. (1991). Reconstruction of Caribbean climate change over the past 10,500 years. Nature 352, 790-793.

Hodell, D.A., Brenner, M., Curtis, J.H. and Guilderson, T. (2001). Solar forcing of drought frequency in the Maya lowlands. Science, 292, 1367-1370.

Hodell, D.A., Brenner, M. and Curtis, J.H. (2005). Terminal Classic drought in the northern Maya lowlands inferred from multiple sediment cores in Lake Chichancanab (Mexico). Quaternary Science Reviews, 24, 1413-1427.

Hodell, D. A., Brenner, M., Curtis, J. H., Medina-Gonzalez, R., Ildefonso-Chan, E., Albornaz-Pat, A., and Guilderson, T. P. (2005). Climate change on the Yucatán Peninsula during the Little Ice Age. Quaternary Research, 63(2), 109-121.

Page 196: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

184

Hodell, D.A., Brenner, M. and Curtis, J.H. (2007). Climate and cultural history of the northeastern Yucatán Peninsula, Quintana Roo, Mexico. Climatic Change, 83, 215-240.

Holbourn, A., Kuhnt, W., Regenberg, M. Schulz, M. Mix, A. and Andersen, N. (2010). Does Antarctic glaciation force migration of the tropical rain belt? Geology, 38(9), 783-786

Hooghiemstra, H. and Van der Hammen, T. (2004). Quaternary Ice-Age dynamics in the Colombian Andes: developing an understanding of our legacy. Philosophical Transactions of the Royal Society of London Series B-Biological Sciences, 359(1442), 173-180.

Hooghiemstra, H., Wijninga, V. and Cleef, A. (2006). The paleobotanical record of Colombia: Implications for biogeography and biodiversity. Annals of the Missouri Botanical Garden, 93, 297-324.

Horrocks, M., Nichol, S. D'Costa, D., Augustinus, P. Jacobi, T., Shane, P. and Middleton, A. (2007). A Late Quaternary Record of Natural Change and Human Impact from Rangihoua Bay, Bay of Islands, Northern New Zealand. Journal of Coastal Research, 592-604.

Hunter, J.P. (1998). Paleoecology meets ecology on questions of scale. Trends in Ecology and Evolution, 13(12), 478-479.

Hurtado, L. (2008). Dinámicas agrarias y reproducción campesina en la globalización: El caso de Alta Verapaz, 1990-2007. FandG Editores, Guatemala.

Hyndman, R.J. and Shang, H.L. (2010). Rainbow Plots, Bagplots, and Boxplots for Functional Data. Journal of Computational and Graphical Statistics, 19, 29-45.

Ichon, A., Douzant-Rosenfeld, D. and Usselmann, P. (1996). La Cuenca Media del Río Chixoy. Ocupación Prehispánica y problemas actuales. Cuadernos de Estudios Guatemaltecos 3. Escuela de Historia, Universidad de San Carlos de Guatemala, 213 pp.

Imbach, P., Molina, L., Locatelli, B., Roupsard, O., Ciais, P., Corrales, L. and Mahe, G. (2010). Climatology-based regional modelling of potential vegetation and average annual long-term runoff for Mesoamerica. Hydrology and Earth System Sciences, 14, 1801-1817.

Innes, J.B., Zong, Y., Chen, Z., Chen, C., Wang, Z. and Wang, H. (2009). Environmental history, palaeoecology and human activity at the early Neolithic forager/cultivator site at Kuahuqiao, Hangzhou, eastern China. Quaternary Science Reviews, 28(23-24), 2277-2294.

Islebe, G. and M. Kappelle. (1994). A phytogeographical comparison between subalpine forests of Guatemala and Costa Rica. Feddes Repert, 105, 73-87.

Islebe, G. A. and A. Velázquez. (1994). Affinity among mountain ranges in Megamexico: A phytogeographical scenario. Vegetatio, 115, 1-9.

Islebe, G., Velázquez, A., and Cleef, A. (1995). High elevation coniferous vegetation of Guatemala. A phytosociological approach. Vegetatio 116, 7-23.

Islebe, G.A. and Hooghiemstra, H. (1997). Vegetation and climate history of montane Costa Rica since the last glacial. Quaternary Science Reviews, 16, 589-604.

Islebe, G., Villanueva-Gutierrez, R., Sánchez-Sánchez, O. (2001). Relación lluvia de polen-vegetación en selvas de Quintana Roo. Boletín de la Sociedad Botánica de Mexico, 69, 31-38.

Page 197: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

185

Islebe, G. and Leyden, B. (2006). La Vegetación de Guatemala durante el Pleistoceno terminal y Holoceno. Biodiversidad de Guatemala (ed. by E. Cano), pp. 15-23. Universidad del Valle de Guatemala, Guatemala.

Janzen, D. H. (1967). Why Mountain Passes Are Higher in Tropics. American Naturalist, 101, 233-249.

Johnson, K.D., Wright, D.R. and Terry, R.E. (2007). Application of carbon isotope analysis to ancient maize agriculture in the Petexbatún region of Guatemala. Geoarchaeology-an International Journal, 22, 313-336.

Jolon-Morales, M. R. 2007. Análisis de Vacíos y Omisiones para el Sistema Guatemalteco de Áreas Protegidas. Informe Final Consultoría. Guatemala: TNC. pp 132.

Jones, A. W. and R. S. Kennedy. (2008). Evolution in a tropical archipelago: comparative phylogeography of Philippine fauna and flora reveals complex patterns of colonization and diversification. Biological Journal of the Linnean Society, 95, 620-639.

Jongman, R., Braak, C. and Van Tongeren, O. (1995). Data analysis in community and landscape ecology. Cambridge University Press, Cambridge, United Kingdom.

Joosten, J.H.J. (1995). Between Diluvium and Deluge: The origin of the Younger Dryas concept. Geologie En Mijnbouw, 74, 237-240.

Jouzel, J., Masson-Delmotte, V., Cattani, O., Dreyfus, G., Falourd, S., Hoffmann, G., Minster, B., Nouet, J., Barnola, J. M., Chappellaz, J., Fischer, H., Gallet, J. C., Johnsen, S., Leuenberger, M., Loulergue, L., Luethi, D., Oerter, H., Parrenin, F., Raisbeck, G., Raynaud, D., Schilt, A., Schwander, J., Selmo, E., Souchez, R., Spahni, R., Stauffer, B., Steffensen, J. P., Stenni, B., Stocker, T. F., Tison, J. L., Werner, M., and Wolff, E. W. (2007). Orbital and Millennial Antarctic Climate Variability over the Past 800,000 Years. Science 317, 793-796

Juggins, S. (2003). C2 Data Analysis. In. University of Newcastle, Newcastle. Kappelle, M., Vanuffelen, J.G., and Cleef, A.M. (1995). Altitudinal Zonation of Montane

Quercus Forests Along 2 Transects in Chirripo National-Park, Costa-Rica. Vegetatio, 119, 119-153.

Kappelle, M. (1996). Los Bosques de Roble (Quercus) de la Cordillera de Talamanca, Costa Rica. Biodiversidad, Ecología, Conservación y Desarrollo. Instituto Nacional de Biodiversidad (INBio). Santo Domingo de Heredia, Costa Rica.

Kennedy, L., Horn, S. and Orvis, K. (2005). Modern pollen spectra from the highlands of the Cordillera Central, Dominican Republic. Review of Palaeobotany and Palynology, 137, 51– 68

Kerhoulas, N.J. and Arbogast, B.S. (2010). Molecular systematics and Pleistocene biogeography of Mesoamerican flying squirrels. Journal of Mammalogy, 91(3), 654-667.

Kessler, M. (2000). Altitudinal Zonation of Andean Cryptogam Communities. Journal of Biogeography, 27(2), 275-282.

Kitahara, F., Mizoue, N. and Yoshida, S. (2009). Evaluation of data quality in Japanese National Forest Inventory, Environmental Monitoring and Assessment. Springer Netherlands, pp. 331-340.

Page 198: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

186

Klein, E. Lavigne, C., Foueillassar, X., Gouyon, P. and Laredo, C. (2003). Corn Pollen Dispersal: Quasi-Mechanistic Models and Field Experiments.Ecological Monographs, 73 (1), 131-150.

Kokorowski, H.D., Anderson, P.M., Mock, C.J. and Lozhkin, A.V. (2008). A re-evaluation and spatial analysis of evidence for a Younger Dryas climatic reversal in Beringia. Quaternary Science Reviews, 27(17-18), 1710-1722.

Knapp, S. and G. Davidse (2006). Flora of Guatemala revisited. Biodiversidad de Guatemala. E. Cano. Guatemala, Universidad del Valle de Guatemala. No.1, 25-47.

Kramer, A., Herzschuh, U., Mischke, S. and Zhang, C. (2010). Late glacial vegetation and climate oscillations on the southeastern Tibetan Plateau inferred from the Lake Naleng pollen profile. Quaternary Research, 73(2), 324-335.

Kruskal, W. and Wallis, W. (1952) Use of ranks in one-criterion variance analysis. Journal of the American Statistical Association, 47, 583-621.

Kunes, P., Pokorny, P. and Sida, P. (2008). Detection of the impact of early Holocene hunter-gatherers on vegetation in the Czech Republic, using multivariate analysis of pollen data, Vegetation History and Archaeobotany. Springer Berlin / Heidelberg, pp. 269-287.

Larsen, C.P.S. and Macdonald, G.M. (1993). Lake morphometry, sediment mixing and the selection of sites for fine resolution paleoecological studies. Quaternary Science Reviews, 12, 781-792.

Latorre, C., Betancourt, J. and Arroyo, M. (2006). Late Quaternary vegetation and climate history of a perennial river canyon in the Río Salado basin (22°S) of Northern Chile. Quaternary Research. 65, 450–466.

Lentz, D. and Hockaday, B. (2009). Tikal timbers and temples: ancient Maya agroforestry and the end of time Journal of Archaeological Science, 36, 1342-1353.

Leyden, B., Brenner, M. and Dahlin, B.H. (1998). Cultural and climatic history of Coba, a lowland Maya city in Quintana Roo, Mexico. Quaternary Research, 49, 111-122.

Leyden, B. (2002). Pollen evidence for climatic variability and cultural disturbance in the Maya Lowlands. Ancient Mesoamerica 13, 85-101.

Li, Y., Wu, J., Hou, S., Shi, C., Mo, D., Liu, B., and Zhou, L. (2010). Palaeoecological records of environmental change and cultural development from the Liangzhu and Qujialing archaeological sites in the middle and lower reaches of the Yangtze River. Quaternary International. Holocene Vegetation Dynamics and Human Impact, I: Agricultural Activities and Rice Cultivation in East Asia, 227(1), 29-37.

Lisiecki, L.E. (2010). Links between eccentricity forcing and the 100,000-year glacial cycle. Nature Geosciences, 3(5), 349-352.

López, A. (2009). Diversidad de flora en la Reserva Natural Privada Chelemhá y su zona de influencia, Tucurú, Alta Verapaz. Tesis Licenciatura Biología, Escuela de Biología, Facultad de Ciencias Químicas y Farmacia. Universidad de San Carlos de Guatemala, Guatemala.

Page 199: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

187

Lorimer, C.G. (2001). Historical and ecological roles of disturbance in eastern North American forests: 9,000 years of change. Wildlife Society Bulletin, 29(2), 425-439.

Lozano-García, M., Caballero, M., Ortega, B., Rodríguez, A. and Sosa, S. (2007). Tracing the effects of the Little Ice Age in the tropical lowlands of eastern Mesoamerica. Proceedings of the National Academy of Sciences of the United States of America, 104, 16200-16203.

Lozano-García, S., Caballero, M., Ortega, B., Sosa, S., Rodríguez, A. and Schaaf, P. (2010). Late Holocene palaeoecology of Lago Verde: Evidence of human impact and climate change in the northern limit of the neotropics during the late formative and classic periods. Vegetation History and Archaeobotany, 19, 177-190.

Lynch, E.A. (1996). The ability of pollen from small lakes and ponds to sense fine-scale vegetation patterns in the Central Rocky Mountains, USA. Review of Palaeobotany and Palynology, 94, 197-210.

Lytle, D.E. and Wahl, E.R. (2005). Palaeoenvironmental reconstructions using the modern analogue technique: the effects of sample size and decision rules. Holocene, 15, 554-566.

Macia, M.J. and Balslev, H. (2000). Use and management of totora (Schoenoplectus californicus, Cyperaceae) in Ecuador. Economic Botany, 54, 82-89.

MAGA. (2001). Mapa de Clasificación de Reconocimiento de los Suelos: por departamento. Ministerio de Agricultura, Ganadería y Alimentación (MAGA), Guatemala.

MAGA. (2004). Atlas temático de las cuencas hidrográficas de la República de Guatemala. Ministerio de Agricultura, Ganadería y Alimentación, Guatemala.

MAGA. (2006). Mapa de Cobertura del Suelo y Uso de la Tierra de la República de Guatemala, escala 1:50,000. Ministerio De Agricultura, Guatemala.

Marchant, R., Almeida, L., Behling, H., Berrio, J.C., Bush, M., Cleef, A., Duivenvoorden, J., Kappelle, M., De Oliveira, P., de Oliveira, A.T., Lozano-García, S., Hooghiemstra, H., Ledru, M.P., Ludlow-Wiechers, B., Markgraf, V., Mancini, V., Paez, M., Prieto, A., Rangel, O. and Salgado-Labouriau, M.L. (2002). Distribution and ecology of parent taxa of pollen lodged within the Latin American Pollen Database. Review of Palaeobotany and Palynology, 121, 1-75.

Marchant, R., Cleef, A., Harrison, S. P., Hooghiemstra, H., Markgraf, V., van Boxel, J., Ager, T., Almeida, L., Anderson, R., Baied, C., Behling, H., Berrio, J. C., Burbridge, R., Bjorck, S., Byrne, R., Bush, M., Duivenvoorden, J., Flenley, J., De Oliveira, P., van Geel, B., Graf, K., Gosling, W. D., Harbele, S., van der Hammen, T., Hansen, B., Horn, S., Kuhry, P., Ledru, M. P., Mayle, F., Leyden, B., Lozano-García, S., Melief, A. M., Moreno, P., Moar, N. T., Prieto, A., van Reenen, G., Salgado-Labouriau, M., Schabitz, F., Schreve-Brinkman, E. J., and Wille, M. (2009). Pollen-based biome reconstructions for Latin America at 0, 6000 and 18 000 radiocarbon years ago. Climate of the Past 5, 725-767.

Markgraf, V., Whitlock, C., Anderson, R.S. and García, A. (2009). Late Quaternary vegetation and fire history in the northernmost Nothofagus forest region: Mallín Vaca Lauquen, Neuquen Province, Argentina. Journal of Quaternary Science, 24(3), 248-258.

Page 200: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

188

Martínez, M. L., Pérez-Maqueo, O., Vázquez, G., Castillo-Campos, G., García-Franco, J., Mehltreter, K., Equihua, M. and Landgrave, R. (2009). Effects of land use change on biodiversity and ecosystem services in tropical montane cloud forests of Mexico. Forest Ecology Management, 258, 1856-1863.

Mathewson, K. (2006). A century and counting: Geographical research on Guatemala in historical perspective. Geoforum, 37, 15-30.

May, R. (1974). Progresses in R- and Q-Mode Analysis: Correspondence Analysis and its Application to the Study of Geological Processes: Discussion. Canadian Journal of Earth Sciences, 11, 1494-1497

McCune, B. and M. J. Mefford. (2006). PC-ORD. Multivariate Analysis of Ecological Data. MjM Software, Gleneden Beach, Oregon, U.S.A.

McKey, D., Rostain, S., Iriarte, J., Glaser, B., Birk, J., Holst, I., and Renard, D. (2010). Pre-Columbian agricultural landscapes, ecosystem engineers, and self-organized patchiness in Amazonia. Proceedings of the National Academy of Sciences, 107(17), 7823-7828.

McNeil, C.L., Burney, D.A. and Burney, L.P. (2010). Evidence disputing deforestation as the cause for the collapse of the ancient Maya polity of Copán, Honduras. Proceedings of the National Academy of Sciences of the United States of America, 107, 1017-1022.

McWethy, D.B., Whitlock, C., Wilmshurst, J.M., McGlone, M.S. and Li, X. (2009). Rapid deforestation of South Island, New Zealand, by early Polynesian fires. The Holocene, 19(6), 883-897.

Mendoza, B. (2005). Total solar irradiance and climate. Fundamentals of Space Environment Science, 35, 882-890.

Mercuri, A.M. (2008). Human influence, plant landscape evolution and climate inferences from the archaeobotanical records of the Wadi Teshuinat area (Libyan Sahara). Journal of Arid Environments, 72(10), 1950-1967.

Merkel, U., Prange, M. and Schulz, M. (2010). ENSO variability and teleconnections during glacial climates. Quaternary Science Reviews, 29, 86-100.

Mollinedo, L. (2002). Propuesta de manejo forestal sostenible para el municipio de Tactic, Alta Verapaz con base a la cualificación y cuantificación del recurso bosque. Centro Universitario del Norte. Universidad de San Carlos de Guatemala, Cobán, Alta Verapaz.

Monzón, R. (1999). Estudio general de los recursos agua, suelo y del uso de la tierra del Parque Nacional Laguna Lachuá y su Zona de Influencia, Cobán, Alta Verapaz. Tesis Ingenieria Agrónoma, Universidad de San Carlos de Guatemala, Guatemala.

Mörner, N.A. (2010). Solar Minima, Earth's rotation and Little Ice Ages in the past and in the future: The North Atlantic-European case. Global and Planetary Change Quaternary and Global Change: Review and Issues Special issue in memory of Hugues FAURE, 72(4), 282-293.

Muñoz, S.E. and Gajewski, K. (2010). Distinguishing prehistoric human influence on late-Holocene forests in southern Ontario, Canada. The Holocene, 20(6), 967-981.

Murray, M., Scoffield, R., Galan, C., Villamil, C. (2007). Airborne pollen sampling in a wildlife reserve in the south of Buenos Aires province, Argentina. Aerobiologia, (23), 107–117.

Page 201: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

189

Murugan, M., Shetty, P., Anandhi, A., Ravi, R., Alappan, S., Vasudevan, M., Gopalan, S. (2009). Rainfall changes over tropical montane cloud forests of southern Western Ghats, India. Current Science. 97, 1755-1760.

Nair, U. S., Lawton, R., Welch, R. and Pielke, R. (2003). Impact of land use on Costa Rican tropical montane cloud forests: Sensitivity of cumulus cloud field characteristics to lowland deforestation. Journal of Geophysical Research-Atmospheres, 108.

Neff, H., Pearsall, D., Jones, J., Arroyo de Pieters, B. and Freidel, D. (2006). Climate change and population history in the Pacific Lowlands of Southern Mesoamerica. Quaternary Research, 65.

Nielsen, A., and Van Odgaard, B. (2005). Reconstructing land cover from pollen assemblages from small lakes in Denmark. Review of Palaeobotany and Palynology, 133, 1-21.

Olivera, M. Duivenvoorden, J., Hooghiemstra, H. (2009). Pollen rain and pollen representation across a forest-páramo ecotone in northern Ecuador. Review of Palaeobotany and Palynology 157, 285–300

Olofsson, J. and Hickler, T. (2008). Effects of human land-use on the global carbon cycle during the last 6,000 years, Vegetation History and Archaeobotany. Springer Berlin / Heidelberg, pp. 605-615.

Ortega-Gutiérrez, F., Solari, L., Ortega-Obregón, C., Elias-Herrera, M., Martens, U., Moran-Ical, S., Chiquin, M., Keppie, J., Torres, R., and Schaaf, P. (2007). The Maya-Chortis Boundary: A tectonostratigraphic approach. International Geology Review 49, 996-1024.

Otypková, Z. and Chytry, M. (2006). Effects of plot size on the ordination of vegetation samples. Journal of Vegetation Science, 17, 465-472.

Padilla, R. J. (2007). Evolución geológica del sureste mexicano desde el Mesozoico al presente en el contexto regional del Golfo de México. Boletín de la Sociedad Geológica Mexicana, 59, 19-42.

Palala, Y. (2000). Propuesta de manejo forestal para el municipio de Santa Cruz Verapaz, Alta Verapaz. Centro Universitario del Norte. Universidad de San Carlos de Guatemala, Cobán, Alta Verapaz.

Partel, M., Helm, A., Reitalu, T., Liira, J. and Zobel, M. (2007). Grassland diversity related to the Late Iron Age human population density. Journal of Ecology, 95(3), 574-582.

Pascarella, J.B., Aide, T.M. and Zimmerman, J.K. (2007). The demography of Miconia prasina (Melastomataceae) during secondary succession in Puerto Rico. Biotropica, 39, 54-61.

Paz, H. (2001). Propuesta de manejo forestal para el municipio de Tucurú, Alta Verapaz. Centro Universitario del Norte. Universidad de San Carlos de Guatemala, Cobán, Alta Verapaz.

Peterson, L.C., Overpeck, J.T., Kipp, N.G. and Imbrie, J. (1991). A High-Resolution Late Quaternary Upwelling Record from the Anoxic Cariaco Basin, Venezuela. Paleoceanography, 6(1), 99-119.

Pitman, N., Terborgh, J., Silman, M., Nuñez, P., Neill, D., Ceron, C., Palacios, W., and Aulestia, M. (2001). Dominance and distribution of tree species in Upper Amazonian Terra Firme forests. Ecology 82, 2101-2117.

Page 202: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

190

Pokorny, P., Klimesova, J., Klimes, L. (2000). Late Holocene history and vegetation dynamics of a floodplain alder carr: A case study from eastern Bohemia, Czech Republic, Folia Geobotanica. Springer Netherlands, pp. 43-58.

Popescu, S.-M., Biltekin, D., Winter, H., Suc, J., Melinte-Dobrinescu, M., Klotz, S., Rabineau, M., Combourieu-Nebout, N., Clauzon, G., and Deaconu, F. (2010). Pliocene and Lower Pleistocene vegetation and climate changes at the European scale: Long pollen records and climatostratigraphy. Quaternary International. Plio-Pleistocene Correlation and Global Change, 219(1-2), 152-167.

Poulos, H. M. and A. E. Camp. (2010). Topographic influences on vegetation mosaics and tree diversity in the Chihuahuan Desert Borderlands. Ecology, 91, 1140-1151.

Powell, E.A. (2008). Do civilizations really collapse? (Jared Diamond). Archaeology, 61, 18-56.

Prentice, I. C. (1985). Pollen Representation, Source Area, and Basin Size - toward a Unified Theory of Pollen Analysis. Quaternary Research, 23, 76-86.

Pyburn, K.A. (1998). Smallholders in the Maya lowlands: Homage to a garden variety ethnographer. Human Ecology, 26, 267-286.

Qinghai, X., Xiaolan, Y., Chen, W., Lingyao, M., Zihui, W. (1996). Alluvial Pollen on the North China Plain. Quaternary Research, 46, 270-280.

Quintana-Ascencio, P. and M. González-Espinosa. (1993). Afinidad fitogeográfica y papel sucesional de la flora leñosa de los bosques de pino-encino de los Altos de Chiapas, México. Acta Botánica Mexicana, 21, 43-57.

Ramírez-Marcial, N. (2003). Survival and growth of tree seedlings in anthropogenically disturbed Mexican montane rain forests. Journal of Vegetation Science, 14, 881-890.

Ramírez-Marcial, N., González-Espinosa, M. and Williams-Linera, G. (2001). Anthropogenic disturbance and tree diversity in Montane Rain Forests in Chiapas, Mexico. Forest Ecology and Management, 154, 311-326.

Rapp, D. (2010). Assessing Climate Change. Temperatures, Solar Radiation and Heat Balance. Environmental Sciences, Springer Praxis Books. Praxis Publishing, UK. 410 p.

Raven, P. H. and D. I. Axelrod. (1974). Angiosperm Biogeography and Past Continental Movements. Annals of the Missouri Botanical Garden, 61, 539-673.

Rice, P.M., Michel, H.V., Asaro, F. and Stross, F. (1985). Provenience Analysis of Obsidians from the Central Petén Lakes Region, Guatemala. American Antiquity, 50(3), 591-604.

Richardson, D. 2000. Ecology and biogeography of Pinus. Cambridge University Press. United Kingdom.

Rokaya, M.B., Münzbergová, Z. and Timsina, B. (2010). Ethnobotanical study of medicinal plants from the Humla district of western Nepal. Journal of Ethnopharmacology, 130(3), 485-504.

Ross, N.J. (2011). Modern tree species composition reflects ancient Maya "forest gardens" in Northwest Belize. Ecological Applications, 21, 75-84.

Ross, N.J. and Rangel, T.F. (2011). Ancient Maya Agroforestry Echoing Through Spatial Relationships in the Extant Forest of NW Belize. Biotropica, 43, 141-148.

Roubik, D. W., and Moreno, J. (1991). Pollen and spores of Barro Colorado Island. Monogr. Syst. Bot.Missouri Bot Gard.

Page 203: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

191

Rowley-Conwy, P. (2009). Human Prehistory: Hunting for the Earliest Farmers. Current Biology, 19(20): 948-949.

Ruddiman, W.F. (2003). The Anthropogenic Greenhouse Era Began Thousands of Years Ago, Climatic Change. Springer Netherlands, pp. 261-293.

Rzedowski, J. (2006). Vegetación de México. Comisión Nacional para el Conocimiento y Uso de la Biodiversidad, México.

Sánchez, J., (1994). Determinación de nitrógeno total, fósforo total y metales pesados en macrofitas acuáticas de la especie Eichornia crassipes (lirio acuático) de la Laguna Chichoj, Alta Verapaz, para sugerir su posible utilización como purificadores biológicos y fertilizante agrícola. Licenciatura en Química, Universidad de San Carlos de Guatemala, Ciudad de Guatemala, 66 pp.

Sarmiento, G., Gaviria, S., Hooghiemstra, H., Berrio, J.C. and Van der Hammen, T. (2008). Landscape evolution and origin of Lake Fúquene (Colombia): Tectonics, erosion and sedimentation processes during the Pleistocene. Geomorphology, 100(3-4), 563-575.

Scarborough, V. L., and Gallopin, G. G. (1991). A Water Storage Adaptation in the Maya Lowlands. Science 251, 658-662.

Scharf, E. (2010). Archaeology, land use, pollen and restoration in the Yazoo Basin (Mississippi, USA), Vegetation History and Archaeobotany. Springer Berlin / Heidelberg, pp. 159-175.

Scharf, E. (2010). A statistical evaluation of the relative influences of climate, vegetation, and prehistoric human population on the charcoal record of Five Lakes, Washington (USA). - Quaternary International. 215, 74-86.

Schmitt, C. B., Manfred, D., Sebsebe, D., Ib, F., and Hans Juergen, B. (2010). Floristic diversity in fragmented Afromontane rainforests: Altitudinal variation and conservation importance. - Applied Vegetation Science, 13, 291-304.

Schele, L. and Freidel, D. (1990). A forest of kings. Morrow, New York. Schueler, S. and Schluenzen, K. (2006). Modeling of oak pollen dispersal on the

landscape level with a mesoscale atmospheric model. Environmental Modeling and Assessment, 11, 179-194.

Schuster, J. (2006). Passalidae (Coleoptera) de Mesoamérica: diversidad y biogeografía. In: E. Cano (Editor), Biodiversidad de Guatemala. Universidad del Valle de Guatemala, Guatemala, pp. 379-392.

Schuster, J., Cano, E. and Cardona, C. (2000). Un método sencillo para priorizar la conservación de los bosques nubosos de Guatemala, usando Passalidae (Coleoptera) como organismos indicadores. Acta Zoológica Mexicana. 80, 197-209.

Sharer, R. and Sedat, D. (1987). Archaeological investigations in the northern Maya Highlands, Guatemala. Interaction and the development of Maya Civilization., Hanover, Pennsylvania, 495 pp.

Shaw, J.M. (2003). Climate change and deforestation: Implications for the Maya collapse. Ancient Mesoamerica, 14(1), 157-167.

Shen, T., Chao, A., and Lin, C. (2003). Predicting the number of new species in further taxonomic sampling. Ecology, 84, 798-804.

Page 204: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

192

Seager, R., Kushnir, Y., Nakamura, J., Ting, M. and Naik, N. (2010). Northern Hemisphere winter snow anomalies: ENSO, NAO and the winter of 2009/10. Geophysical Research Letters, 37, 6.

Shi, G.R. (1993). Multivariate data-analysis in paleoecology and paleobiogeography - a review. Palaeogeography Palaeoclimatology Palaeoecology, 105, 199-234.

Shuman, B. (2003). Controls on loss-on-ignition variation in cores from two shallow lakes in the northeastern United States. Journal of Paleolimnology, 30, 371-385.

Simona, M.F., Gretherc, R., de Queirozd, L., Skemae, C., Penningtone, R., and Hughesa, C. (2009). Recent assembly of the Cerrado, a neotropical plant diversity hotspot, by in situ evolution of adaptations to fire. PNAS, 106(48), 20359-20364.

Smith, M. and Demarest, A. (2001). Review: Climatic change and the Classic Maya Collapse: The return of catastrophism. Latin America Antiquity, 12(1), 105-107.

Smith, N., Mori, S., Henderson, A., Stevenson, D., and Heald, S. (2004). Flowering plants of the Neotropics. Princeton University Press and The New York Botanical Garden.

Solanki, S., Usoskin, I., Kromer, B., Schussler, M. and Beer, J. (2004). Unusual activity of the Sun during recent decades compared to the previous 11,000 years. Nature, 431, 1084-1087.

Solomon, A. M., Blasing, T. J., and Solomon, J. A. (1982). Interpretation of Floodplain Pollen in Alluvial Sediments from an Arid Region. Quaternary Research, 18, 52-71.

Southgate, E. 2010. Herbaceous plant communities on the Delaware River floodplain, New Jersey, during the Mid-Holocene. The Journal of the Torrey Botanical Society, 137(2), 252-262.

Stahle, D.W., Villanueva Díaz , J., Burnette, D.J., Cerano Paredes, J., Heim, R.R., Jr., Fye, F.K., Soto, R.A., Therrell, M.D., Cleaveland, M.K. and Stahle, D.K. (2011). Major Mesoamerican droughts of the past millennium. Geophysical Research Letters, 38

Standley, P., Steyermark, J. (1946). Flora of Guatemala, 24 Part IV. Chicago Natural History Museum, Chicago.

Standley, P.a.S., J. 1958 Flora of Guatemala: Plan of the Flora. Fieldiana, Bot, 24-, I-II. Stuiver, M., Reimer, P.J. and Reimer, R.W. (2005). CALIB 5.0. WWW program and

documentation Sugita, S. (1993). A model of pollen source area for an entire lake surface. Quaternary

Research. 39, 239-244. Terrab, A., Schonswetter, P., Talavera, S., Vela, E. and Stuessy, T. (2008). Range-wide

phylogeography of Juniperus thurifera L., a presumptive keystone species of western Mediterranean vegetation during cold stages of the Pleistocene. Molecular Phylogenetics and Evolution, 48, 94-102.

Thieme, D.M. (2001). Historic and possible prehistoric human impacts on floodplain sedimentation, North Branch of the Susquehanna River, Pennsylvania, USA. River Basin Sediment Systems: Archives of Environmental Change, 375-403.

Tietjen, B., Jeltsch, F., Zehe, E., Classen, N., Groengroeft, A., Schiffers., K., and Oldeland, J. (2010). Effects of climate change on the coupled dynamics of water and vegetation in drylands. Ecohydrology, 3, 226-237.

Page 205: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

193

Torres, V., Vandenberghe, J. and Hooghiemstra, H. (2005). An environmental reconstruction of the sediment infill of the Bogota basin (Colombia) during the last 3 million years from abiotic and biotic proxies. Palaeogeography Palaeoclimatology Palaeoecology, 226(1-2), 127-148.

Torrescano, N. and Islebe, G.A. (2006). Tropical forest and mangrove history from southeastern Mexico: a 5000 yr pollen record and implications for sea level rise. Vegetation History and Archaeobotany, 15(3), 191-195.

Tot, C. (2000). Caracterización de los recursos naturales suelo, agua y flora en la subcuenca del Río Tinajas, Reserva de la Biosfera Sierra de las Minas. Centro Universitario del Norte. Universidad de San Carlos de Guatemala, Cobán, Alta Verapaz.

Traverse, A. (1994). Sedimentation of palynomorphs and palynodebris: an introduction. In "Sedimentation of Organic Particles." (A. Traverse, Ed.), pp. 1-8. Cambridge University Press, Cambridge.

Tukey, J. (1977). Exploratory Data Analysis. Addison-Wesley. Turner, B. L., and Harrison, P. D. (1981). Prehistoric Raised-Field Agriculture in the

Maya Lowlands. Science 213, 399-405. Twiddle, C. L., and Bunting, M. J. (2010). Experimental investigations into the

preservation of pollen grains: A pilot study of four pollen types. Review of Palaeobotany and Palynology 162, 621-630.

Valsecchi, V., Carraro, G., Conedera, M. and Tinner, W. (2010). Late-Holocene vegetation and land-use dynamics in the Southern Alps (Switzerland) as a basis for nature protection and forest management. The Holocene, 20(4), 483-495.

Van Buren, M. (2010). The Archaeological Study of Spanish Colonialism in the Americas. Journal of Archaeological Research, 18, 151-201.

Van't Veer, R., and Hooghiemstra, H. (2000). Montane forest evolution during the last 650 000 yr in Colombia: a multivariate approach based on pollen record Funza-I. Journal of Quaternary Science 15, 329-346.

Van Zonneveld, M., Jarvis, A., Dvorak, W., Lema, G. and Leibing, C. (2009). Climate change impact predictions on Pinus patula and Pinus tecunumanii populations in Mexico and Central America. Forest Ecology and Management, 257, 1566-1576.

van der Hammen, T., and Hooghiemstra, H. (2003). Interglacial-glacial Fuquene-3 pollen record from Colombia: an Eemian to Holocene climate record. Global and Planetary Change 36, 181-199.

Vanderpoorten, A., Gradstein, S. and Devos, N. (2010). The ghosts of Gondwana and Laurasia in modern liverwort distributions. Biological Reviews, 85, 471-487.

Vargas-Rodríguez, Y. L., Platt, W., Vázquez-García, J. and Boquín, G. (2010). Selecting Relict Montane Cloud Forests for Conservation Priorities: The Case of Western Mexico. Natural Areas Journal. 30: 156-173.

Veen, P., Fanta, J., Raev, I., Biriş, I., de Smidt, J. and Maes, B. (2010). Virgin forests in Romania and Bulgaria: results of two national inventory projects and their implications for protection. Biodiversity and Conservation, 19, 1805-1819.

Vermoere, M., Vanhecke, L., Waelkens, M., and Smets, E. (2000). A comparison between modern pollen spectra of moss cushions and Cundill pollen traps - Implications for the interpretation of fossil pollen data from Southwest Turkey. Grana 39, 146-158.

Page 206: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

194

Viau, A.E. and Gajewski, K. (2009). Reconstructing Millennial-Scale, Regional Paleoclimates of Boreal Canada during the Holocene. Journal of Climate, 22(2), 316-330.

Villar, L. (1998). La flora silvestre de Guatemala. Universidad de San Carlos de Guatemala, Guatemala.

Vogler, A., Eisenbeiss, H., Aulinger-Leipner, I. and Stamp, P. (2009). Impact of topography on cross-pollination in maize (Zea mays L.). European Journal of Agronomy, 31, 99-102.

Wahl, D., Byrne, R., Schreiner, T. and Hansen, R. (2006). Holocene vegetation change in the northern Petén and its implications for Maya prehistory. Quaternary Research, 65, 380-389.

Wahl, D., Schreiner, T., Byrne, R. and Hansen, R. (2007). A paleoecological record from a late classic Maya reservoir in the north Petén. Latin American Antiquity, 18, 212-222.

Wainwright, J. (2008). Can modelling enable us to understand the role of humans in landscape evolution? Geoforum, 39(2), 659-674.

Weiser, A. and Lepofsky, D. (2009). Ancient Land Use and Management of Ebey's Prairie, Whidbey Island, Washington. Journal of Ethnobiology, 29(2), 184-212.

Weiss, C. and Brunner, U. (2010). Lacustrine to palustrine sediments from the Sirwah Oasis, Yemen sedimentological implications for the local paleoecology and archeology. Neues Jahrbuch fur Geologie und Palaontologie - Abhandlungen, 257(2), 129-145.

Wendt, T. (1989). Las Selvas de Uxpanapa, Veracruz-Oaxaca, México: Evidencias de refugios florísticos cenozoicos. Anales del Instituto de Biología, UNAM, Serie Botánica 58, 29-54.

Weng, C., Hooghiemstra, H. and Duivenvoorden, J. (2007). Response of pollen diversity to the climate-driven altitudinal shift of vegetation in the Colombian Andes. Proceedings of the National Academy of Sciences of the United States of America, 362, 253-262.

Weng, C., Bush, M.B. and Silman, M.R. (2004). An analysis of modern pollen rain on an elevational gradient in southern Peru. Journal of Tropical Ecology, 20, 113-124.

Weyl, L. (1980). "Geology of Central America." Gebruder Borntraeger. Whitehouse, N. (2010). Conservation lessons from the Holocene record in "natural" and

"cultural" landscapes. Restoration and history. The search for a usable environmental past (ed. by M. Hall), p. 329. Routledge, New York.

Williams, A.N., Ulm, S., Goodwin, I.D. and Smith, M. (2010). Hunter-gatherer response to late Holocene climatic variability in northern and central Australia. Journal of Quaternary Science, 25(6), 831-838.

Willis, K.J. and Birks, H.J.B. (2006). What Is Natural? The Need for a Long-Term Perspective in Biodiversity Conservation. Science, 314(5803), 1261-1265.

Wilmshurst, J., and McGlone, M. (2005). Origin of pollen and spores in surface lake sediments: Comparison of modern palynomorph assemblages in moss cushions, surface soils and surface lake sediments. Review of Palaeobotany and Palynology, 136, 1-15.

Page 207: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

195

Woodfill, B., Monterroso, M., Dillon, B., Valle, J., Mijangos, B., Burgos, W., Velásquez, A., Wolf, M., Tox Tiul, C. (2012). Salinas de Los Nueve Cerros: Nuevos aportes de una ciudad Maya “Industrial” entre el Altiplano y las Tierras Bajas. XXV Simposio de Investigaciones Arqueológicas en Guatemala, Guatemala City.

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.

Xia, Z., Chen, G., Zheng, G., Chen, F., and Han, J. (2002). Climate background of the evolution from Paleolithic to Neolithic cultural transition during the last deglaciation in the middle reaches of the Yellow River. Chinese Science Bulletin 47, 71-75.

Ye, L., Zhang, Y.A., Mei, Y., Nan, P. and Zhong, Y. (2010). Pollen-inferred vegetation and environmental changes since 16.7 ka BP at Balikun Lake, Xinjiang. Chinese Science Bulletin, 55(22), 2373-2379.

Yu, X., Zhou, W., Liu, X., Xian, F., Liu, Z., Zheng, Y. and An, Z. (2010). Peat records of human impacts on the atmosphere in Northwest China during the late Neolithic and Bronze Ages. Palaeogeography, Palaeoclimatology, Palaeoecology, 286(1-2), 17-22.

Zazula, G. D., Schweger, C. E., Beaudoin, A. B., and McCourt, G. H. (2006). Macrofossil and pollen evidence for full-glacial steppe within an ecological mosaic along the Bluefish River, eastern Beringia. Quaternary International 142, 2-19.

Zhao, Y., Chen, F., Zhou, A., Yu, Z. and Zhang, K. (2010). Vegetation history, climate change and human activities over the last 6200 years on the Liupan Mountains in the southwestern Loess Plateau in central China. Palaeogeography, Palaeoclimatology, Palaeoecology, 293(1-2), 197-205.

Zizumbo-Villarreal, D. and Colunga-García, P. (2010). Origin of agriculture and plant domestication in West Mesoamerica, Genetic Resources and Crop Evolution. Springer Netherlands, pp. 813-825.

Zonneveld, M., Jarvis, A., Dvorak, A., Lema, G., and Leibing, C. (2009). Climate change impact predictions on Pinus patula and Pinus tecunumannii populations in Mexico and Central America. Forest Ecology and Management 257, 1566-1576.

Page 208: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

196

Appendices  

Page 209: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

197

Appendix 2.1. Indicator, generalist, and disjunctive plant checklist.

Plant Taxa Axis 1 2466 2300 2200 2200 2100 2100 2000 1900 1900 1800 1800 1650 1500 1400 1258 1106 1048 1000 1000 600 400 200 170

Spondias mombin 109Tabebuia sp. 176Genipa sp. 226Inga sp. 257Saurauia belisensis 264Heliocarpus mexicanus 265Cedrela pacayana 266Perymenium grande 268Weinmannia pinnata 363Miconia aeruginosa 365Oreopanax liebmanii 373Psychotria parasitica 390Centropogon cordifolius 436Cavendishia guatemalensis 437Jocotillo 440Fuchsia microphylla 445Miconia glaberrima 445Styrax argenteus 452Lobelia nubicola 466Synardisia venosa 479Clethra suaveolens 498Phoradendron sp. 500Erigeron karvinskianus 526Passiflora sexflora 527Begonia oaxacana 555Ocotea sp. 557Rhynchosia sp. 655

Indi

cato

r T

axa

Page 210: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

198

Appendix 2.1. continued. Idiosyncratic refers to Disjunctive Taxa.

Che4 Che3 Tin6 Pur5 Che2 Pur4 Pur3 Pur2 Che1 Pur1 Tin5 Tac Scruz Tin4 Che Flo Tam Bvta Tin3 Tin2 Chin Tin1 LachAxis 1 2466 2300 2200 2200 2100 2100 2000 1900 1900 1800 1800 1650 1500 1400 1258 1106 1048 1000 1000 600 400 200 170

Vochysia guatemalensis 153Terminalia amazonia 168Bursera simarouba 172Ceiba pentandra 190Parathesis vulgata 199Cecropia peltata 244Virola sp. 245Dendropanax leptopodus 269Mollineda guatemalensis 308Billia hippocastanum 312Engelhardtia guatemalensis 317Brunellia mexicana 330Persea donnell-smithii 333Liquidambar styraciflua 337Clusia sp. 368Hedyosmum mexicanum 384Quercus sp. 395Quercus crispifolia 442Myrica cerifera 512Eupatorium semialatum 549

Dendropanax arboreus 135Lasciacis divaricata 156Ocotea eucuneata 277Phoebe sp. 293Matayba oppositifolia 315Peperomia cobana 336Pouteria campechiana 338Clidemia capitellata 430Conyza bonariensis 499

Gen

eral

ist T

axa

Idio

sync

ratic

Tax

a

Page 211: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

199

Appendix 3.1. Pollen types found on modern pollen calibrations and fossil pollen spectra from cores P-4 and L-3. Associated plant taxa and uses by ancient Mayan populations are shown. Information about vegetation belt (VB) is provided based on Table 2.2 and 2.3 (Chapter 2). * indicate know usage by modern Mayan populations.

Pollen taxa VB Genus

Family Associated plant taxon/taxa

Uses

Acacia Fabaceae Acacia spp. Medicine, forage, construction Alchornea Euphorbiaceae Anthurium Araceae

Araliaceae Dendropanax arboreus Food, ornamental, forage, medicine

Arecaceae Acrocomia mexicana Food Attalea cohune Food, construction, medicine Chamaedora spp. Food

Low

land

s

Cryosophila stauracanhta Construction. medicine Sabal morrisiana Construction

Bignoniaceae Tabebuia rosea Medicine, timber, ornamental Bombacaceae Ceiba pentandra Medicine, timber, ritual Quararibea Food Pseudobombax Ritual

Boraginaceae Cordia sp. Food, medicine

Page 212: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

200

Appendix 3.1. continued Pollen taxa

VB Genus Family Associated plant taxon/taxa Uses

Brosimum Moraceae Brosimum alicastrum Food, medicine, forage, ritual Celtis Ulmaceae Celtis iguanaea Food Combretaceae Combretum Melastomataceae Clidemia Euphorbiaceae Cnidoscolus aconitifolius Food Manihot esculenta Food, medicine Fabaceae Lonchocarpus castilloi Ritual

Pachyrhizus erosus Food Phaseolus lunatus Food Mimosa Fabaceae Mimosa spp. Medicine, fuel Malpighiaceae Byrsonima crassifolia Food, medicine, apiculture L

owla

nds

Moraceae Castilla elastica Latex Pseudolmedia spuria Food

Myrsinaceae Myrsine sp. Myrtaceae Psidium guayaba Food Pimienta dioica Food

Pachira Bombacaceae Pachira aquatica Food, medicine, construction Piper Piperaceae Piper amalago Medicine

Psychotria Rubiaceae Psychotria chiapensis Medicine Rubiaceae Alseis yucatanensis Wood

Hamelia axillaris Medicine

Page 213: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

201

Appendix 3.1. continued

Pollen taxa VB Genus Family Associated plant taxon/taxa Uses

Salvia Lamiaceae Salvia coccinea Sapium Euphorbiaceae

Sapotaceae Crysophyllum mexicanum Food, medicine Manilkara zapota Food, medicine, latex Pouteria campechiana Food Solanaceae Solanum Food, medicine, forage, ritual L

owla

nds

Cestrum Medicine, ornamental Capsicum Food Trema Ulmaceae Trema micrantha Food Ulmaceae

Verbenaceae Vitex gaumeri Ornamental, fuel, forage, construction

LRF Spondias Anacardiaceae Spondias mombin, S. purpurea, S. radlkoferi Medicine, construction, food

Bursera Burseraceae Bursera simaruba Medicine, ritual Burseraceae Protium copal1 Ritual LRF-LMRF Cecropia Cecropiaceae Cecropia sp. Medicine, timber Terminalia Combretaceae Terminalia amazonia Construction

Page 214: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

202

1 Important plant for ancient and modern Maya, but not found in the pollen record. Appendix. 3.1. continued.

Pollen taxa VB Genus Family Associated plant taxon/taxa Uses

LMRF Inga Fabaceae Inga spp. Hedyosmum Chloranthaceae Hedyosmum mexicanun Food LMRF-MCF Myrica Myricaceae Myrica cerifera Medicine* Quercus Fagaceae Quercus sp. Fuel MMF Abies Pinaceae Abies guatemalensis MMF-SAF Alnus Betulaceae Alnus acuminata,

A, jorullensis

Ericaceae Arbutus sp,Cavendishia guatemalensis

Conifer6 Pinaceae Ritual Pinales Pinaceae Ritual Pinus Pinaceae Pinus caribaea, P. oocarpa Ritual

Hig

hlan

ds

Urticaceae Phenax, Pilea, Urera, Urtica Medicine

Page 215: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

203

Appendix 3.1. continued

Pollen taxa

Genus Family Associated plant taxon/taxa Uses

Alternanthera Amaranthaceae Alternanthera sp. Amaranthaceae Amaranthus Food Chenopodiaceae Chenopodium ambrosioides Food Asteraceae2 Medicine* Cyperaceae Cyperus esculentus Food Eleocharis caribaea Peperomia Piperaceae Piperaceae Poaceae Food Polygonum Polygonaceae Food Zea Poaceae Zea mays Food Trilete spores Monolete spores

Microgramma lycopodioides, Acrostichum aureum Food

2 Uncertainty about native plant taxa

Page 216: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

204

POLLEN TYPE P4 P4-1 P4-2 P4-3 P4-4 P4-5 P4-6 P4-7 P4-8 P4-9 P4-10 P4-11 P4-12 P4-13 P4-14 P4-15 P4-16 P4-17 P4-18Abies 4 5 3 1 0 8 0 0 0 0 0 0 0 0 0 0 0 0 0Alnus 0 0 1 1 0 1 1 0 0 1 1 0 0 1 0 0 0 0 0Hedyosmum 6 7 10 16 4 12 13 1 1 4 1 1 4 6 7 3 2 4 0Ilex 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0Myrica 2 5 1 0 1 3 0 0 0 2 1 4 3 3 6 5 5 1 1Pinus 11 72 54 62 56 53 21 1 7 4 6 8 6 12 9 4 9 14 17Quercus 17 9 5 1 2 2 6 1 2 1 4 1 11 5 4 3 6 7 4Alternanthera 5 4 11 30 21 20 0 0 0 0 0 0 0 0 0 0 0 0 0Amaranthaceae/Chenopodiaceae 0 13 11 13 10 29 5 15 13 17 10 6 23 22 14 18 14 15 21Asteraceae 12 45 77 59 50 60 138 44 88 81 62 83 83 88 111 139 145 162 151Poaceae 15 21 29 10 56 8 5 2 8 41 11 12 5 10 4 11 16 29 7Zea 2 0 1 0 1 0 5 3 0 0 3 4 4 5 1 2 2 0 1Alchornea 0 2 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0Arecaceae 0 4 5 0 0 2 0 0 3 7 0 0 1 0 0 3 3 1 1Brosimum 0 0 0 1 1 0 0 0 0 0 0 0 0 1 0 0 0 0 0Bursera 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0Combretaceae/Melastomataceae 0 0 0 3 0 0 0 0 0 1 0 0 0 0 0 0 0 0 2Celtis 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0Malphigiaceae 0 0 0 0 0 2 0 0 1 0 0 0 0 1 0 0 0 0 0Myrtaceae 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0Pollen sum 101 222 217 206 206 208 202 70 128 172 99 135 142 156 157 194 210 237 205Cyperaceae 54 159 452 458 462 471 150 20 176 270 63 90 58 48 62 109 74 158 76Polygonum 25 35 9 8 4 8 8 3 5 13 0 16 2 2 1 5 7 4 0Trilete 20 26 46 64 50 56 83 75 54 19 16 27 15 34 23 10 28 43 23Monolete 24 35 19 141 46 67 22 29 23 9 8 7 13 17 8 15 13 5 7

Appendix 4.1. Pollen counts (raw) from P-4 core.

Page 217: Natural and cultural landscape evolution during the … · Natural and cultural landscape evolution during the Late Holocene in North Central Guatemalan Lowlands and Highlands

205

Appendix 5.1. Pollen counts (raw) from L-3 core. POLLEN TYPE L3‐0 L3‐1 L3‐2 L3‐3 L3‐4 L3‐5 L3‐6 L3‐7 L3‐8 L3‐9 L3‐10 L3‐11 L3‐12 L3‐13 L3‐14 L3‐15 L3‐16 L3‐18Abies 0 1 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0Alchornea 1 4 13 14 12 8 9 10 11 7 6 10 10 7 16 16 6 8Alnus 0 0 0 1 2 2 0 4 3 1 1 2 3 0 0 1 2 1Arecaceae 4 10 4 3 1 4 6 3 1 2 0 0 0 3 2 1 3 0Bombacaceae 1 1 0 1 0 1 1 1 1 0 1 0 0 0 3 1 0 0Brosimum 0 0 0 0 0 1 1 0 0 0 0 0 0 0 0 0 0 0Bursera 0 3 4 0 0 1 4 3 0 3 1 1 2 2 1 1 1 7Caesalpinaceae 0 2 0 0 0 0 1 1 0 0 0 0 0 0 0 0 1 0Celtis 2 0 0 0 2 0 3 1 1 1 0 0 2 0 0 7 2 0Combretaceae/Melastomataceae 7 19 61 63 43 48 51 65 38 45 21 16 10 8 11 18 13 11Hedyosmum 0 0 0 0 0 0 1 0 0 0 0 1 0 0 0 1 1 2Hyeronima 0 0 8 5 4 5 8 1 4 5 0 0 0 0 1 0 0 0Ilex 21 16 17 19 17 15 17 12 7 37 9 13 9 9 11 5 9 3Malphigiaceae 2 1 0 4 5 5 4 6 6 1 3 2 3 2 1 0 1 0Myrica 10 7 10 9 13 5 5 11 3 6 6 14 8 16 21 34 18 42Myrtaceae 9 10 27 8 20 28 11 12 13 9 12 18 22 20 18 7 1 0Pinus 7 5 6 0 5 3 11 6 4 10 10 7 10 4 9 4 18 9Piper 4 0 0 0 3 0 0 0 0 0 0 0 0 0 0 0 0 0Psychotria 7 6 6 8 3 5 8 3 5 2 1 4 6 9 17 32 1 3Quercus 4 1 4 0 2 4 0 0 0 3 0 5 3 0 4 0 1 6Rubiaceae 1 4 4 12 10 4 13 10 6 2 2 7 6 7 10 5 3 2Sapium 11 6 3 4 7 6 5 4 4 0 0 1 0 3 2 1 1 0Sapotaceae 0 0 0 3 2 2 2 1 2 2 2 2 4 3 5 1 2 0Solanaceae 15 10 20 14 21 24 13 29 37 44 30 39 37 27 45 46 43 51Spondias 3 4 1 11 9 12 16 9 7 7 3 0 3 3 5 5 4 7Terminalia 8 4 6 5 5 5 3 4 2 6 0 5 2 1 1 2 1 2Trema 3 0 0 0 1 0 0 1 1 0 0 1 0 1 2 0 1 0Amaranthaceae/Chenopodiaceae 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0Asteraceae 4 6 4 3 5 4 3 1 4 4 2 1 1 2 1 1 1 2Poaceae 1 4 3 7 8 4 2 2 6 3 4 4 2 5 0 8 6 5Zea 0 2 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0Cyperaceae 11 2 1 9 2 1 6 5 4 3 5 13 2 4 1 5 2 5Pollen sum 125 126 201 195 200 196 198 200 166 201 114 153 143 132 186 198 140 161Monoletes 24 87 57 80 72 54 56 30 31 64 50 26 18 12 16 16 55 10Triletes 3 2 6 10 8 3 8 8 9 48 97 53 63 25 11 6 38 21